Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Jul 28.
Published in final edited form as: Adv Drug Deliv Rev. 2020 Jul 28;156:40–64. doi: 10.1016/j.addr.2020.07.020

POLYMER NANOMEDICINES

Jindřich Kopeček 1,2, Jiyuan Yang 1
PMCID: PMC7736172  NIHMSID: NIHMS1616028  PMID: 32735811

Abstract

Polymer nanomedicines (macromolecular therapeutics, polymer-drug conjugates, drug-free macromolecular therapeutics) are a group of biologically active compounds that are characterized by their large molecular weight. This review focuses on bioconjugates of water-soluble macromolecules with low molecular weight drugs and selected proteins. After analyzing the design principles, different structures of polymer carriers are discussed followed by the examination of the efficacy of the conjugates in animal models and challenges for their translation into the clinic. Two innovative directions in macromolecular therapeutics that depend on receptor crosslinking are highlighted: a) Combination chemotherapy of backbone degradable polymer-drug conjugates with immune checkpoint blockade by multivalent polymer peptide antagonists; and b) Drug-free macromolecular therapeutics, a new paradigm in drug delivery.

Keywords: Polymeri nanomedicines, macromolecular therapeutics, polymer-drug conjugates, cancer, nanomedicine, checkpoint inhibition, multivalent polymer peptide antagonists, drug-free macromolecular therapeutics

1. Introduction

There is no distinct classification of the term macromolecular therapeutics in the literature, but we shall focus on water-soluble polymer – drug conjugates and on macromolecular systems that hold therapeutic activity and do not contain low molecular weight drugs. Some reports in the literature use the term „nanoparticles“ for both vesicular carriers and water soluble macromolecular therapeutics. However, there is a clear distinction in size, shape, and flexibility of the two types of carriers.

The research on water-soluble macromolecules for drug delivery started more than 60 years ago. Jatzkewitz used a glycylglycine spacer to bind mescaline to poly(vinylpyrrolidone) (PVP) [1], Ushakov’s group conjugated several antibiotics to PVP and studied biological properties of the conjugates [24]. De Duve discovered the lysosomotropism of macromolecules [5] and Ringsdorf published a first comprehensive model of the potential of water-soluble polymers as drug carriers [6]. Maeda revealed the enhanced permeability and retention (EPR) effect – the enhanced deposition of macromolecules in solid tumors [7]. The early research of polymer-drug conjugates was reviewed in 1977 [8]. The interdisciplinary field developed dramatically due to the advantages of polymer-drug conjugates vs. free drugs [9]. The possibility to manipulate pharmacokinetics, biorecognition, distribution and subcellular location, as well as therapeutic efficacy was the driving force.

Numerous reviews on macromolecular therapeutics have been published [1031]. This review focuses on the design and structure of macromolecular drug carriers, important research achievements that are prerequisites for future clinical translation and novel directions, including combination chemotherapy and immunotherapy and drug-free macromolecular therapeutics.

2. Design of macromolecular therapeutics

The major rationale for the use of water-soluble polymers as drug carriers is the potential to improve the efficacy of low molecular weight drugs. One of the main distinction of conjugates vs. free drugs is the different mechanism of cell entry [5,32]. Macromolecular therapeutics are composed from a suitable polymer carrier, spacer between the macromolecular carrier and drug, and optionally a targeting moiety [21]. Synthetic nondegradable macromolecules need to have the whole molecular weight below the renal threshold to ensure the elimination of the carrier by glomerular filtration [21,28]. Biodegradable carriers such as poly(amino acids) or polysaccharides need to preserve their degradability after covalently binding drugs. The structure modification (by drug attachment) impairs the formation of the enzyme-substrate complex resulting in decreased degradability [33,34]. When the major degradation mechanism is pure hydrolysis, the substitution of the degradable chain may not substantially impair degradability as in conjugates based on poly(β,L-malic acid) [35,36].

In addition to molecular weight, structure of the carrier, architecture, mechanism of drug release, and subcellular location impact the efficacy of polymer-drug conjugates.

2.1. Spacers

Spacers used in attachment of drugs to synthetic polymer carriers need to remain attached to the carrier in the blood stream but are cleaved off the carrier by enzymes in the lysosomal compartment [3740]. Otherwise, the drug can be bound to the carrier via a hydrolytically cleavable bond to be released in late endosomes and lysosomes due to decreased pH in these subcellular compartments [41].

Self-immolative spacers are elongated spacers where the enzymatically sensitive bond is separated from the drug by a self-eliminating group. Rate controlling enzymatic cleavage is followed by fast 1,6-elimination as demonstrated in the design of cathepsin K sensitive spacers for the release of prostaglandin in osteoclasts [42]. Garrison and coworkers designed cathepsin S degradable spacers [4345]. Santi and coworkers developed non-enzymatically cleavable linkers where the rate of release based on β-elimination reaction can be manipulated in a wide range [46,47]. An interesting design for self-immolative spacers is a combination of aromatic azobonds with a 1,6-elimination unit suitable for colon delivery [48]. Following azobond reduction in the colon the drug (9-aminocamptothecin) is released in addition to p-aminobenzylalcohol and CO2 [49].

Notably, FDA approved recently ENHERTU® (trastuzumab deruxtecan) an antibody drug conjugate containing a tetrapeptide spacer GGFG, validating the concept of cathepsin B sensitive peptide spacers [50].

Spacer length is important for drug attachment. Proteases responsible for the cleavage of oligopeptide side chains terminated in drug have considerable large active sites which are capable of combining with a number of amino acid residues [51]. Generally, four amino acid residues toward the NH2 end of the substrate (i.e., toward the polymer backbone) take part in the interactions that determine the formation of the enzyme-substrate complex and ultimately the drug release [37,38]. Another factor is the steric hindrance of the polymer chain on the formation of enzyme-substrate complex. An oligopeptide attached to the end of the polymer chains accommodates into the enzyme active site easier than the same sequence attached to the polymer as a side chain [52].

Spacer length can be important also in the attachment of targeting moieties. In a prostate cancer model, DUPA (3-(1,3-dicarboxypropyl)-ureido]pentanedioic acid) attached to polymer carrier via a longer (dodecaethyelene glycol, 46 Å) spacer reaches easier the PSMA (prostate-specific membrane antigen) binding site via a ~ 20 Å funnel shaped tunnel than conjugates with a short spacer [53].

Coiled-coil forming peptides were used for stable non-covalent attachment of drugs [54,55] or recombinant scFv fragments [56,57] to polymeric carriers. A triheptad or tetraheptad peptide was attached to the polymeric carrier. The complementary peptide sequence is terminated in drug or scFv fragment. Decreased pH in the lysosomes causes the detachment of peptide and release of the modified drug.

2.2. Molecular weight of synthetic polymer carriers

Macromolecules accumulate into solid tumors by the EPR effect (see 3.1). The efficacy of the EPR process is molecular weight dependent – deposition of conjugates in solid tumor increases is directly proportional to molecular weight [58,59]. However, to ensure biocompatibility, the molecular weight of non-degradable synthetic macromolecules need to be below the renal threshold (about 50 kDa for random coil conformation of hydrophilic macromolecules). Consequently, first HPMA copolymer – doxorubicin (DOX) conjugates evaluated in clinical trials had a molecular weight of 28 kDa [6062]. To keep polymer carriers biocompatible and long-circulating, degradable bonds can be inserted into the polymer backbone as described in 2.4.1.2.

2.3. Architecture

Molecular architecture (Figure 1) can exert an important effect on the biological properties of polymer-drug conjugates [25,63]. Several studies have described advanced methods for the synthesis of linear, star, graft, and branched copolymers [64,65] and evaluated the impact of architecture on the rate of drug release, IC50 doses, biodistribution and in vivo efficacy [66,67]. Star-like N-(2-hydroxypropyl)methacrylamide (HPMA) copolymers can be synthesized by conjugating semitelechelic HPMA copolymer-drug conjugates to different generation (G2, G3, G4) PAMAM dendrimers [68].

Figure 1.

Figure 1.

Different architectures of polymer conjugates. Adapted from ref. [25].

The conclusions of comparative studies are frequently difficult to generalize because the activity of conjugates depends on several factors, including molecular weight, architecture, hydrodynamic volume, type of bond used for drug attachment (enzymatically or hydrolytically cleavable), and rate of cellular uptake. Frequently, conjugates that were compared (e.g., linear vs. star) have different molecular weight as well as different size. For example, Nakamura et al. compared star-like PAMAM dendrimer-based HPMA copolymer-pirarubicin conjugate (400 kDa, 26 nm) with linear HPMA copolymer-pirarubicin conjugate (39 kDa, 8.2 nm; hydrazone bond used in both conjugates) [69]. The tumor growth inhibitory activity of the star-like conjugate was higher than that of the linear one in S-180 tumor-bearing mice. However, due to the higher molecular weight of star-like conjugate it is difficult to distinguish the impact of architecture from the molecular weight dependent accumulation in solid tumors [69]. Šírová et al. evaluated the efficacy of combination treatment of murine EL4 lymphoma using linear HPMA copolymer-doxorubicine (DOX) and -docetaxel (DTX) conjugates (28 kDa) as well as star DOX and DTX conjugates (250 kDa; all hydrolytically cleavable bonds). Whereas the combination of linear conjugates resulted in additive treatment effects, the star-like DTX conjugate decreased the treatment efficacy of the star DOX conjugate [70]. Sadekar et al. compared biodistribution of PAMAM dendrimers and linear HPMA copolymer (29 kDa) in tumor bearing mice using 125I labeled macromolecules. Hydroxyl terminated generation 5, 6, and 7 dendrimers were taken up by the liver and kidney more than the HPMA copolymer that was mainly excreted into the urine [71].

Architecture impacts the cellular internalization – linear HPMA copolymer-meso-tetra(4-carboxyphenyl)porphyrin) (MTCP) conjugate possessed higher internalization rates as well as light induced cytotoxicity than MCTP attached to hyperbranched amine terminated PAMAM dendrimer [72].

Generally, one can assume if the drug is bound to the polymer carrier via a hydrolytically cleavable bond the rate of drug release (and corresponding activity) will not be substantially different when comparing linear and hyperbranched structures. However, for enzymatically cleavable spacers complex architectures may result in impairing the formation of the enzyme-substrate complex, and slower drug release with concomitant decrease in activity.

Structural factors have an important impact on the efficacy of polymer-drug conjugates. Different chemical structures of polymer backbones are discussed below.

2.4. Structure of polymer carrier

Different structures have been used as drug carriers. The factors important in the selection of a particular macromolecular structure are hydrophilicity (to keep the conjugate soluble after covalent attachment of hydrophobic drugs), biodegradability (to achieve long intravascular half-life) and biocompatibility (appropriate host response). The main polymer carriers are discussed in the following paragraphs.

2.4.1. HPMA copolymers

2.4.1.1. First generation of HPMA copolymer carriers

Before using polyHPMA [73,74] as drug carrier, the biocompatibility of the homopolymer was established. Early studies revealed that polyHPMA (working name Duxon) was nontoxic, apyrogenic, did not exhibit any effect on LEP and HeLa cells cultures, and did not initiate formation of antibodies in mice [75,76]. 14C-Labeled polyHPMA (Mw=28,000) was eliminated from the organism of rabbits following intravenous administration [77]. PolyHPMA prolonged the survival of semiallogeneic skin grafts in mice and rats and did not inhibit the growth of lymphocytes in tissue culture [78,79]. Additionally, no effect of polyHPMA on the activation of porcine complement was observed [80]. Interestingly, a recent study on intracellular polymerization of HPMA in HeLa cells mentioned remarkable biocompatibility up to 250 mM of HPMA inside the cells [81].

Studies of HPMA copolymer-drug conjugates were reviewed numerous times [810,13,18,23,24,26,28,31]. Just briefly, the first generation of HPMA copolymer-drug conjugates with nondegradable backbone was active on numerous animal models of cancer [82]. The drugs were attached to the polymer backbone via lysosomally degradable oligopeptide spacers [3739,42] or hydrolytically cleavable (e.g., hydrazone) bonds [41]. Several conjugates were evaluated in clinical trials [6062,8386]; the results clearly demonstrated a significant decrease of non-specific toxicity – the MTD of PK1 (HPMA copolymer-DOX conjugate) was 320 mg/m2 (of DOX equivalent; MTD for free DOX is 60–80 mg/m2) and no congestive heart failure was observed despite doses up to 1680 mg/m2 [60]. In phase II of PK1 (dose 280 mg/m2) the difference of PK1 efficacy [62] when compared to free DOX was smaller than in animal models [8791]. Similar observations were made in other clinical trials [8486]. See detailed analysis in ref. [10].

It was hypothesized that the reason for lower efficacy in humans was the short blood circulation time of the low molecular weight HPMA copolymer-drug conjugates. The intravascular half-life was sufficient in animal models where the tumor is up to 10% of body weight. In humans, a prolonged concentration gradient between vasculature and tumor is needed to ensure abundant extravasation of the conjugate into the tumor and efficacy [10,18,23]. This observation initiated the design of backbone degradable, 2nd generation conjugates described in the next paragraph.

2.4.1.2. Backbone degradable, long-circulating HPMA copolymer carriers

To address the discrepancy between circulation time and biocompatibility of nondegradable carriers, new backbone degradable, long circulating HPMA copolymers were designed, synthesized, and evaluated (Figure 2). These carriers contain an enzymatically degradable sequence both in the main chain and in the spacer between polymer backbone and drug. They are synthesized by a combination of living radical polymerization and click reactions [9294]. In particular, RAFT (reversible addition-fragmentation chain transfer) polymerization in the presence of a bifunctional chain transfer agent (peptide2CTA) is used to prepare a degradable HPMA diblock copolymer using a scalable, one step process [94]. Peptide2CTA (Nα,Nε-bis(4-cyano-4-(phenylcarbonothioylthio)pentanoylglycylphenylalanylleucylglycyl)lysine) contains an oligopeptide (GFLGKGLFG) degradable sequence flanked by two dithiobenzoate groups. Monomer inserts at both positions with the same rate, producing a diblock copolymer with a very narrow molecular weight distribution in one scalable synthetic step. Further expansion of molecular weight can be achieved by click reactions, producing tetrablock and hexablock copolymers. This is a platform technology and numerous drugs can be attached to these carriers (Figure 2).

Figure 2.

Figure 2.

Long-circulating, backbone degradable HPMA copolymer-drug conjugates. (A) Structure of non-degradable (1st generation) and backbone degradable conjugates. (B) Synthesis of backbone degradable conjugates by RAFT copolymerization of HPMA and polymerizable derivative of drug containing degradable GFLG sequence (MA is methacryloyl). Bifunctional chain transfer agent (Peptide2CTA) is composed from a lysosomally degradable sequence GFLGKGLFG flanked by two dithiobenzoate groups. This permits the synthesis of degradable diblock copolymers with narrow polydispersity in one, scalable synthetic step. Post polymerization click reaction produces multiblock copolymers.

To optimize the molecular weight of the multiblock backbone degradable copolymers, diblock (Mw~100 kDa), tetrablock (Mw~200 kDa), and hexablock (Mw~300 kDa) HPMA copolymers containing gemcitabine (GEM) or paclitaxel (PTX) were used in the treatment of human ovarian carcinoma A2780 xenografts in mice [95]. Notably, degradable diblock conjugates 2P-GEM and 2P-PTX inhibited tumor growth more successfully than multiblock conjugates [95]. It is known that higher molecular weight enhances the accumulation of conjugates in solid tumors [59]. However, other factors, such as the conformation of the macromolecule [66,96,97], association of side-chains by hydrophobic interactions (formation of unimolecular micelles), or random association of flexible chains by “point-point” contacts [98100], influence the rate of enzymatic drug release [37,38], penetration of the solid tumor [101] and, ultimately, the efficacy.

Similar approach based on RAFT polymerization and incorporation of enzymatically degradable sequences into the polymer backbone was used by Luo’s [102104] and Garrison’s labs [4345,105]. Pechar et al. used different chemistry, they prepared multiblock copolymers by interfacial polycondensation of telechelic PEG (2 kDa) and oligopeptide diamine from glutamic acid and lysine. The oligopeptide sequence contained GFLG-DOX side chains. The conjugate was cleavable by cathepsin B and possessed anticancer activity [106]. Additionally, multiblock HPMA copolymers were synthesized via polyaddition of homotelechelic polymer diazides with azo-compounds containing two alkyne groups [107]. Yang et al. synthesized an enzyme-sensitive, alkyne-functionalized, chain transfer agent (CTA-GFLG-alkyne) for RAFT polymerization and copolymerization of HPMA. Post-polymerization modification with 4,4’-azobis(azidopropyl 4-cyanopentanoate) resulted in the formation of heterotelechelic HPMA copolymers containing terminal alkyne and azide groups. Chain extension via click reaction resulted in high molecular weight multiblock copolymers [92].

2.4.1.3. Comparison of first generation (non-degradable) and second generation (backbone degradable) HPMA copolymer – drug conjugates

Numerous in vivo data proved the higher therapeutic efficacy of 2nd generation conjugates when compared to 1st generation conjugates. These results are a consequence of dramatically improved pharmacokinetics and tumor accumulation [10]. Comparison of epirubicin (EPI) conjugates [non-degradable ~40,000 Da P-EPI vs. degradable diblock ~100,000 Da 2P-EPI) on human ovarian carcinoma A2780 xenografts (5 mg/kg EPI equivalent doses administered i.v. on days 0, 4, and 8) revealed that complete tumor regression was observed for 100 days in all five evaluated mice treated with 2P-EPI, whereas four mice (out of five) treated with P-EPI started fast tumor regrowth from day 35 (Figure 3) [108]. Measured up to the 1st generation conjugate (P-EPI, Mw~40kDa), 2P-EPI (Mw100 kDa) demonstrated outstandingly improved pharmacokinetics such as four-fold increase in terminal half-life (33.22±3.18 h for 2P-EPI vs 7.55±3.18 h for P-EPI) [108], which is primarily due to the increased molecular weight of the polymer carrier. Strikingly, complete tumor remission and long-term inhibition of tumorigenesis (100 days) were reached in the mice (n=5) treated with 2P-EPI (Figure 3). As a top product in our development pipeline, 2P-EPI (also called KT-1), was selected by Nanotechnology Characterization Laboratory (NCL) scientific review board, and has been evaluated with respect to physicochemical characterization, in vitro cytotoxicity, immunotoxicity, and cancer treatment efficacy [109]. 2P-EPI exhibited strongly superior antitumor activities over free EPI in various preclinical models including colon, lung, pancreatic, and breast cancers. Similar data were obtained with doxorubicin conjugates [110] and paclitaxel conjugates [111].

Figure 3.

Figure 3.

Treatment of nude mice bearing A2780 human ovarian carcinoma xenogratfs with HPMA copolymer-epirubicin (EPI) conjugates. Comparison of 1st generation conjugate (P-EPI) with backbone degradable, long circulating conjugate (2P-EPI); free EPI as control. The dose for each injection was 5 mg/kg EPI equivalent). (A) Dose schedule; Monitoring of tumor growth in conjugate treatment; Tumor volume after treatment with P-EPI and 2P-EPI at day 80 (saline and free EPI groups were terminated at day 20 due to large tumors). (B) Terminal half-life of EPI, P-EPI, and 2P-EPI. (C) Average body weights of mice in individual groups. Adapted from ref. [108].

Clinical treatment of cancer frequently uses combination of drugs that possess different mechanisms of action. One such combination is GEM + PTX. Based on a combination index study [112] synergistic doses were selected and mice bearing A2780 human ovarian carcinoma xenografts were treated with a combination of 2P-GEM and 2P-PTX. Combination of 1st generation conjugates (P-GEM + P-PTX) and combination of free drugs (GEM + PTX) were used as controls. The data showed that 2nd generation backbone degradable conjugates possessed prolonged blood circulation time, enhanced tumor accumulation, and improved anti-tumor efficacy as compared to 1st generation low-Mw conjugates and free drugs. The new 2nd generation conjugates were degradable in vivo and lacked noticeable systemic toxicity [113].

2.4.2. Poly(ethylene glycol)

Poly(ethylene glycol) (PEG) is the most widely used polymer in the clinics [30,114,115]. Its main application is the modification of proteins via one-point attachment using semitelechelic PEG derivates. The idea to modify proteins with PEG originated in the 1970s. Prof. Davies from Rutgers University has shown that such a modification results in prolonged blood circulation, no interactions with antibodies toward the native protein, better solubility, and resistance to proteolysis [116,117]. Numerous protein-PEG conjugates have been FDA approved [see Tables in refs. 30,114,115], starting with PEGylated adenosine deaminase (Adagen) [118] and PEGylated asparaginase (Oncaspar) [119]. Both conjugates contain multiple 5 kDa PEG molecules attached to ε-amino groups of lysines. Commercialized products include enzymes (uricase, phenylalanine ammonia lyase), interferons, granulocyte colony stimulating factor, factor VIII, and anti-tumor necrosis factor Fab’ fragment [30]. The binding chemistry improved dramatically, strategies to attach PEG specifically to cysteine, histidine, and tyrosine are available. Alternatively, engineered protein molecules (mutants) that incorporate a specific amino acid as an attachment point, for example cysteine in the Fab’ fragment of anti-TNF antibody. The molecular weight of PEG increased to 20 kDa (G-CSF conjugate), 30 kDa (epoetin conjugate) and branched 40 kDa (interferon α2a, anti-VEGF aptamer) [30,115].

The improved chemistry resulted in more homogeneous conjugates [120]. However, the modification of proteins with multiple molecules of PEG will impact biorecognition and/or activity. Modification of enzymes with numerous PEG molecules did not have an impact on the efficacy of cleavage of low molecular weight substrates, but substantial decrease of cleavage efficiency of high molecular weight substrates occurred [121]. Modification of vesicles with PEG inhibits cellular uptake and endosomal release. To surmount this challenge targeting moieties and reversible (cleavable) attachment of PEG are being used [122,123].

PEG has also been studied as a carrier in PEG-drug conjugates as well as in micelle-forming block copolymers that incorporate drugs in the core. The latter will be evaluated in a special chapter in this volume. Several PEG-drug (camptothecin, docetaxel, irinotecan, paclitaxel, SN-38) conjugates reached phase I/phase II clinical trials [30,114]. Recently, a four-arm PEG-irinotecan conjugate (Onzeald) reached Phase III clinical trials to treat breast cancer. The drug is bound to the PEG via a slowly hydrolyzable ester bond. Improved pharmacokinetics was observed but longer survival following Onzeald administration was not reached [124]. Current research involves long-circulating PEG-SN-38 conjugates; their accumulation in experimental solid tumor was correlated with μPET/CT imaging of PEG labeled with zirconium-89 [125].

There are numerous concerns related to the biocompatibility of PEG [126]. The formation of IgM antibodies against PEG results in accelerated blood clearance [127] of PEGylated liposomes [128] or nanoparticles [129]. Following second injection, complement activation occurs as a result of IgM binding to the surface of PEGylated vesicles. Additionally, PEGylated proteins are not efficient in the treatment of patients that have anti-PEG antibodies in their blood. Treatment of a substantial fraction of patients with acute lymphoblastic leukemia (ALL) was not efficient due to rapid clearance of PEG-asparaginase [130]. Another concern is the formation of cytoplasmic vacuoles in cortical tubular epithelial cells [131] and the above the renal threshold hydrodynamic volume of PEGs used; the latter may impair elimination from the organism [132].

McSweeney et al. [133] have shown in experiments on mice that pre-administration of 40 kDa PEG can saturate circulating anti-PEG antibodies and restore the prolonged circulation of stealth liposomes.

Alternatives to PEG for protein or vesicle modification

Polymers with various compositions have been studied as PEG replacement [120], including polyoxazolines [134], poly(N-vinyl-2-pyrrolidone) (PVP) [129], poly(N-acryloylmorpholine) (PAcM) [129], poly(N,N-dimethylacrylamide) (PDMA) [129], and polyHPMA [135,136]. Repeated administration of nanoparticles coated with PVP did not result in the ABC phenomenon and antibodies against PVP were not found [129]. Notably, modification of vesicular carriers with PEG alternatives (as well as with PEG) increases the intravascular half-life of modified carriers as well as decreases the adsorption of blood plasma proteins [136138]. Early results on nanosphere modification with semitelechelic polyHPMA [136] demonstrated molecular weight dependent increase of intravascular half-life in rats and decrease of protein adsorption (albumin, IgG, fibrinogen) on modified particles when compared to unmodified nanospheres [136]. Recent data on the interaction of HPMA copolymers with blood plasma proteins observed similar phenomena. Moreover, they contributed to the explanation of the mechanism of corona formation [137139].

Szoka and coworkers evaluated the effect of backbone structure on the induction of accelerated blood clearance (ABC) of polymer modified liposomes [140]. PVP, PDMA, HPMA, and PAcM modified liposomes did not demonstrate ABC in rats, whereas PEG and poly(2-methyl-2-oxazoline) modified liposomes exhibited strong ABC effect (Figure 4). Apparently, a follow-up study on the potential immunogenicity of polyoxazolines is warranted.

Figure 4.

Figure 4.

A) Elimination of a single dose of liposomes (100 nm) from blood of rats over 48 h (mean ± SEM, n = 3). B) Elimination of the second dose of liposomes from blood of rats administered one week after the first dose measured over 48 h after administration of the second dose (mean ± SEM, n = 3–4). PEG, poly(ethylene glycol); HPMA, poly[N-(2-hydroxypropyl)methacrylamide]; PMOX, poly(2-methyl-2-oxazoline); PVP, polyvinylpyrrolidone; PDMA, poly(N,N-dimethylacrylamide); PAcM, poly(N-acryloylmorpholine); molecular weight of polymers = 2.1 – 2.5 kDa. Reprinted from ref. [140].

Other macromolecular structures used in protein modification are zwitterionic peptides (alternating copolymers of E and K) [141] and polymers [142]. A novel approach, called “pepylation” is the modification of proteins with synthetic peptides [143].

Alternatively, proteins may be modified via multipoint attachment. For example, cobra venom acetylcholinesterase was modified with multifunctional HPMA copolymer. The copolymer-modified acetylcholinesterase demonstrated a substantially longer blood circulation time as well as higher stability to thermal inactivation when compared to the native enzyme [144].

Similarly, lateral attachment of multivalent HPMA copolymers to polyelectrolyte complexes of polycations (poly-L-lysine or polyethyleneimine) with DNA results in increased blood circulation time [145]. Seymour and coworkers demonstrated that multilateral modification of therapeutic viruses with HPMA copolymers provided protection against neutralization by antibodies and complement. In addition, the extended plasma circulation and the potential of targeting provide a real opportunity for intravenous delivery of gene and virotherapy agents to treat disseminated cancer [146]. Reversible multipoint coating was achieved with HPMA copolymers with thiazolidine-2-thione (TT) groups at side chain termini and attached to the polymer backbone via spacers containing hydrolytically cleavable hydrazone bonds [147].

2.4.3. Poly(malic acid)

Synthetic poly(β-malic acid) has been studied as a drug carrier in the 1990s [148]. In the last 15 years Ljubimova’s laboratory studies copolyesters derived from poly(β,L-malic acid) (PMLA) as a nanoconjugate platform in drug delivery [29,36]. PMLA is produced by biological fermentation from the slime of Physarium polycephalum. The unsubstituted PMLA backbone is degradable; it is metabolized into water and carbon dioxide. The complex conjugate used in numerous cancer models including glioma is called Polycefin. It is composed from the PMLA backbone, morpholino antisense oligonucleotides (AON; attached via disulfide bonds) targeting α4 and β1 chains of laminin-8 overexpressed in gliomas, monoclonal anti-transferrin antibody, and pH-sensitive membrane disrupting units [149,150]. The endosomolytic function of Polycefin has been optimized; both trilysine [151] and tritryptophan [152] units were evaluated. Obviously, a challenge for such multicomponent conjugates is the detailed analysis of each component. A quantitative analytical method composed of backbone cleavage, chromatography, and HPLC was developed [153].

The nanoconjugates were proven effective in breast cancer treatment. A conjugate using 12-mer peptide mimetic of trastuzumab was effective toward HER2-positive breast cancer model [154]. Another modification of the nanoconjugate composition was successful against triple negative breast cancer [155]. However, the main focus of the team is focused on brain tumors [36]. Attachment of Angiopep-2, a 19 amino acid peptide [156] recognized by the low-density lipoprotein receptor-related protein (LRP1) mediates the transcytosis of the conjugate across the blood brain barrier (BBB) due to high expression of LRP1 on capillary endothelial cells at the BBB surface [157]. Attachment of anti-CTLA-4 and anti-PD-1 checkpoint inhibitor antibodies to an angiopep-2 containing nanoconjugate was successful in the treatment of mice bearing intracranial GL261 glioblastoma. An increase of CD8+ T cells, NK cells and macrophages as well as a decrease of Tregs was observed in the brain tumor area. Survival of mice was significantly longer when compared to single antibody containing conjugates or free antibodies [158].

2.4.4. Poly(glutamic acid)

Synthetic poly(L-glutamic acid) (PG) was employed as a carrier of taxol (TXL) by Li et al. Excellent research and development efforts proved the efficacy of PG-PTX in numerous animal cancer models. PG-TXL possessed better antitumor activity than free paclitaxel, improved pharmacokinetics, and different mechanism of action. In addition to different pharmacokinetics and biodistribution, efficient uptake of PG-TLX by macrophages resulting in enhancement of necrotic tissue are important mechanistic phenomena associated only with macromolecular conjugate of PTX [159]. Cathepsin B was found to mediate selective proteolysis of the PG backbone and site-specific delivery and antitumor efficacy of PG-TLX [160,161]. Extensive preliminary data [reviewed in 159] provided a road map for clinical trials of PG-TLX (CT-2103, XyotaxR). The clinical formulation of CT-2103 had a molecular weight 48,000 and contained 37% paclitaxel by weight. No Cremophor or alcohol were needed and the solution of CT2103 was infused within about 30 minutes. Based on successful Phase I and Phase II studies, more than 1700 patients with NSCLC took part in Phase III studies. The results revealed similar efficacy with controls, but CT-2103 lessened side effects and provided a more convenient administration [159,162]. Interestingly, evaluation of a subset of premenopausal women with normal estrogen levels indicated superior survival following administration of CT-2103. This may relate to estrogen role in pulmonary physiology and enhanced cathepsin B levels [163,164].

Further research on PG as carrier focused in the design of noninvasive imaging agents [165], phthalocyanine conjugates for photodynamic therapy [166], and PG-doxycycline conjugates as fibril disrupters for potential amyloidosis treatment [167]. Recently, star-shaped polyglutamates were designed as new carriers [168]. Star-shaped polyglutamate – curcuminoid conjugate was designed for the treatment of acute kidney injury [169].

2.4.5. Cyclodextrin-based polymers

Davis and coworkers designed linear polymers that contain cyclodextrins in their main chain [170]. For the delivery of nucleic acids, they synthesized positively charged polymers [171] for the delivery of anticancer drugs (camptothecin, CPT) they synthesized neutral copolymers of cyclodextrin and PEG [172] (Figure 5).

Figure 5.

Figure 5.

Scheme of the synthesis of cyclodextrin-based polymer conjugate IT-101 (CRLX101). β-CD, beta-cyclodextrin; PEG-DISPA, poly(ethylene glycol) dipropanoic succinimide; CPT, camptothecin. Reprinted from ref. [170].

Following preclinical evaluation [173,174], the cyclodextrin polymer - CPT conjugate (first named IT-101, later CRLX101) was evaluated in several clinical trials [175, reviewed in 176]. The conjugate localized in human tumors and not in adjacent tissue following intravenous infusion [177]. Evaluation of immune response after IT-101 administration revealed decrease of adverse effects and modulation of immune responses resulting in enhanced drug efficacy [178]. Correlation of animal and human studies indicated that the behavior of CRLX101 in animals is translatable to humans [179].

2.4.6. Poly(2-oxazoline)s and poly(2-oxazine)s

Monomers 2-oxazoline and 2-oxazine are cyclic iminoethers. They can be polymerized by living cationic ring opening polymerization (CROP) initiated by strong alkylating agents, such as alkyl halides [180]. However, the synthesis of uniform high-molecular weight polymers is troubled by chain transfer. Recently, polymerization conditions were established that minimize side reactions and produce 50 kDa polymers with polydispersity 1.05 [180]. The use of polyoxazolines and polyoxazines as drug carriers has been reviewed [181]. Early work focused on protein-polyoxazoline conjugates [182], recent studies concentrated on conjugates with low molecular weight drugs, doxorubicin [183] and rotigotine [184]. The latter conjugate, poly(2-ethyloxazoline) with rotigotine bound via ester bonds, was evaluated in Phase I clinical trials as a potential therapeutic to treat Parkinson disease. Methods to control the kinetics of drug release from polyoxazoline conjugates were published [185,186]; this should encourage more studies on polyoxazoline-drug conjugates.

2.4.7. Other polymer structures

Numerous water-soluble polymer structures that have been used as carriers of drugs, including poly(vinyl pyrrolidone) [14,187], polysaccharides [188191], polypeptoids [192,193], and polyacetals [194]. Fleximer is a biodegradable polyacetal macromolecule (1-hydroxymethylethylene hydroxymethyl formal); its conjugate with camptothecin is being evaluated in clinical trials [195]. Mersana Therapeutics also developed a polymer (Fleximer)-based antibody-drug conjugate platform including XMT-1536, antibody-Fleximer-auristatin F-hydroxypropylamide conjugate targeting the sodium-dependent phosphate transport protein NaPi2b expressed on NSCLC adenocarcinoma and ovarian cancer [196].

3. Examples of efficacy of polymer-drug conjugates in animal models

Polymer-drug conjugates possess abundant advantages when compared to low molecular weight drugs [9]. The major gains from attaching drugs to polymer carriers are the different mechanism of cellular uptake and subcellular fate [21,33,197201], changed pharmacokinetics [113,174], accumulation in solid tumors [59,113], lessened non-specific toxicity of the conjugated drugs [60], surmounting multidrug resistance [91], modulation of cellular signaling and apoptotic pathways [202,203], different stress responses [204], multivalency of targeted conjugates [150,205,206], induction of immunogenic cell death and immunostimulating properties [14], enhanced efficacy in the treatment of cancer and other diseases [110,113,167,173,207210], and prevention of metastasis [211,212].

3.1. Enhanced permeability and retention (EPR) effect

Enhanced permeability and retention (EPR) effect was proposed by Maeda and coworkers to explain enhanced accumulation of macromolecules in solid tumors [7]. The phenomenon was attributed to fenestrations in tumor vasculature and impairment of lymphatic drainage, but numerous other factors contribute [213]. Indeed, the EPR effect was detected in plentiful animal models and became a leading scientific hypothesis behind the design of various nanomedicines. However, clinical trials revealed a complex situation and a large interpatient heterogeneity [214216]. The main problem in the translation of animal data to the clinics was the belief that EPR is a generally occurring phenomenon. In contrast to the development of antibodies or other targeted therapies, where patients have been selected based on the expression of the target antigen on cancer cells, no patient selection based on tumor susceptibility to EPR has been made in clinical trials with nanomedicines [60,62,217,218].

The incongruity between animal and human data on the treatment of solid tumors led to suggestions that the EPR effect does not work in humans [219,220]. Apparently, this conclusion is premature. Future research needs to take into account the fact that tumors are highly variable with large interpatient, intertumor and intratumor variability [218]. Inspiring approaches how to improve delivery of nanomedicines to solid tumors were suggested. Jain and coworkers have shown that small size nanomedicines (~12 nm) can take advantage of the normalization of tumor blood vessels [221]. Such size is typical for water-soluble macromolecular therapeutics, in contrast to the majority of nanoparticles. A recent review summarized factors that have the potential to improve patient outcomes, namely anti-angiogenic therapy, administration of tumor necrosis factor-alpha (TNF-α), angiotensin II receptor blockers, and ionizing radiation [222]. Additionally, NO, CO, and hyperthermia augment EPR effect [223227].

Indeed, basic research and clinical data provide hope for the future of macromolecular therapeutics [228]. Nearly 20 years ago, Harrington et al. have shown the successful accumulation of PEGylated liposomes in patients with progressed tumors. Simultaneously, a wide interpatient heterogeneity was observed [214]. CRLX101, a cyclodextrin-containing polymer conjugate of camptothecin (CPT), self-assembled into nanoparticles of about 30 nm was administered to nine cancer patients with gastric, gastroesophageal, or esophageal cancer. CPT was detected in the tumor of five patients 24 – 48 h post administration. Evidence of biological activity was detected as well as of intact deposition of CRLX101 in the tumor tissue [176]. Two Merrimack clinical trials provided evidence and quantitative data on the accumulation of nanomedicines in human metastatic tumors using positron emission tomography. The ferumoxytol iron nanoparticles uptake as detected by MRI was correlated with the tumor response to liposomal irinotecan [215,216].

These [214216] and other reports [218,229231] clearly indicate that it is of highest significance for the future success of nanomedicines to develop diagnostic tests to select patients that will benefit from the treatment as it is done routinely for molecularly targeted therapies. Identification of predictive “biomarkers” is crucial for the development of this treatment modality.

Clinical imaging including magnetic resonance imaging, positron emission tomography, single photon emission computerized tomography is an important tool that can accelerate the translation of nanomedicines into the clinic [229,232]. In particular, it can provide techniques for the selection of patients whose tumors will be EPR responsive. Generally, there are two approaches to the combination of therapeutic effects and imaging. One is to combine the nanomedicine with the imaging moiety in one conjugate [233]. This is valuable for basic research, but may not be favored for clinical use. The other approach is to develop diagnostic nanoconjugates [43,125,215,216,231,234,235] that will be used for the selection of patients, followed by treatment of selected patients.

Additional important topics to study are tumor-stromal interactions [236238], intratumoral pressure [239], and tumor penetration [101,240,241].

3.2. Targeting

There are several targeting terms used in the literature: passive targeting, (active) targeting, negative targeting, and subcellular targeting. Passive targeting is being used for the passive accumulation of macromolecular therapeutics in solid tumors. It is discussed in 3.1. Negative targeting is sometimes used to illustrate the fact that polymer-bound drugs possess a different body distribution than free drug and might avoid the major organ where free drug exerts toxic effects. For example, polymer-bound doxorubicin avoids heart tissue [242] the main toxic site for free doxorubicin [243].

Active targeting of polymer-drug conjugates can be achieved by attachment of cell-specific ligands, such as antibodies [88,244], antibody fragments [57,245247], affibodies [205], saccharides [248251], lectins [252], peptides [253255], and aptamers [256] resulting in increased biorecognition and cellular uptake [257261]. Combinatorial approaches, such as phage display [262] and synthetic peptide (one-bead-one-compound) libraries [263] are being used for identification of peptides complementary to a receptor, systematic evolution of ligands by exponential enrichment (SELEX) technique is being used for selection of aptamers [264]. An ideal target is expressed predominantly on the target cells and minimally in other locations. Upon binding of the targeting moiety to the relevant receptor the complex can internalize followed by subcellular trafficking via the endosomal/lysosomal route; it can be recycled recovering the surface receptor and releasing the ligand, or it can stay for an extended time at the surface. The internalization and subcellular trafficking processes can be manipulated by receptor crosslinking [265]. Multivalent ligands not only possess higher binding avidity [266], but they initiate internalization mediated by receptor crosslinking [267]. Very slowly internalizing CD20 and HER2 receptors will internalize by crosslinking-mediated endocytosis [268,269]. Crosslinking of PD-L-1 receptors on breast cancer cells by multivalent PD-L1 antagonist modified the subcellular fate of the PD-L1/antagonist complex; instead of recycling it back to the surface it biases the PD-L1 into the lysosomal route [270]. The design of macromolecular therapeutics for a particular receptor must match the receptor property [257].

Antibodies have been frequently used as targeting moieties. A widely studied field are antibody drug conjugates that reached clinics [271]. Recent designs incorporate polymers in the design of ADCs. To avoid using very toxic drugs (calicheamycins, auristatin monomethyl ester, IC50 <1 nM) that could do damage in off-site locations, semitelechelic polymer-drug conjugates are attached to antibodies. This permits to increase the drug to antibody ratio and employ less potent drugs, such as epirubicin [272,273].

Subcellular targeting relates to subcellular manipulation of drug fate. It is known that subcellular location impacts the efficacy of drugs [274]. Smart designs of conjugates permit endosomal escape if degradation in lysosomes needs to be avoided [275,276], basic studies on the passive [201] and chaperoned [277] transport through the nucleopore complex help the design of conjugates for nuclear targeting [278,279].

Mitochondrial targeting can be mediated by utilizing the negative mitochondrial potential and employing positively charged triphenylphosphonium groups as targeting moieties [201,280]. Triphenylphosphonium containing HPMA copolymer-bound photosensitizer, mesochlorin e6 (Mce6), enhanced cytotoxicity toward human ovarian SKOV3 cells as compared to non-targeted HPMA copolymer-Mce6 conjugates [280].

Nuclear targeting can be mediated by steroid hormone receptors. For example, when glucocorticoid receptor (GR) binds to its steroid ligand, it dimerizes and transfers into the nucleus. Rebuffat et al. used this approach to enhance the steroid-mediated transport of transfected DNA into the nucleus of GR positive cells [281]. This approach can be also used for the delivery of photosensitizer Mce6 into the nucleus. To this end, photosensitizer Mce6 was modified with cortisol (attachment moiety for GR) to produce Mce6-cortisol. The latter was attached to the HPMA copolymer via a lysosomally degradable GFLG spacer. The HPMA copolymer - Mce6-cortisol conjugate was incubated with 1471.1 cells transfected with GR labeled with green fluorescent protein, GFP. Following internalization of the HPMA copolymer conjugate by endocytosis, Mce6-cortisol was enzymatically released by cathepsin B inside the lysosomes, translocated to the cytoplasm, and bound to GR (mediated by cortisol); this initiated GR dimerization and nuclear transport of the modified photosensitizer [282].

3.3. Brain delivery

The majority of polymer-drug conjugates is designed for intravenous delivery. Brain delivery has an additional hurdle to overcome, i.e., the nanomedicine needs to cross the blood brain barrier (BBB) [283]. To enhance the transport across the BBB either physical methods, such as osmotic pressure [284] and focused ultrasound [285] are being used. Alternatively, the BBB can be circumvented by nasal administration [286].

Recent designs use biorecognition at the BBB for transport into the brain [29]. Transferrin receptor antibody has been used for many years to mediate transport of therapeutics into the brain [287,288]. An advanced design focuses on peptides that bind to receptors and transcytose bound cargo into the brain. LDLR (low-density lipoprotein receptor)-related protein (LRP)-1, a member of LDLR family binds Kunitz domain-related peptides and initiates transcytosis. Demeule et al. [156] designed 96 Kunitz domain-related peptides, evaluated their potential for transport through BBB and angiopep-2 (AP-2; TFFYGGSRGKRNNFKTEEY) was found to be the best. AP-2 was reported to highly improve therapeutic efficiency in brain targeting studies [289,290]. Recently, the Ljubimova group has used polymer-attached trastuzumab-mimetic anti-HER2 peptide to successfully target and treat HER2+ breast cancer and metastases in the brain, as well demonstrated AP-2-polymer privileged ability for enhancing BBB penetration [157]. AP-2 is also useful for brain tumor targeting as LRP-1 is overexpressed in human glioma cells [291] and in the treatment of brain metastases of breast cancer [292]. AP-2 successfully transcytosed across the BBB conjugates based on poly(β,L-malic acid) [151] and HPMA copolymers [293]. Interestingly, a recent Phase II clinical study of ANG1005, a conjugate of AP-2 with 3 molecules of PTX, demonstrated that all patients, particularly patients with leptomeningeal carcinomatosis, showed symptomatic improvements and prolonged overall survival compared to controls [294].

3.4. Multidrug resistance

Nanomedicines can surmount the efflux pump-type of multidrug resistance [295]. They are internalized in the cells by endocytosis [198] in contrast to diffusion of free low molecular weight drugs. Internalization in membrane limited organelles avoids interaction with P-glycoprotein (Pgp), an ATP-dependent efflux pump that constricts the transport of drugs into the cytoplasm [296]. Polymer-drug conjugates reach the lysosomal compartment located in the perinuclear region of the cell. Following enzymatically mediated release from the carrier they diffuse to the cytoplasm out of reach of the efflux pump. In vitro and in vivo comparison of DOX and HPMA copolymer-DOX conjugate in sensitive A2780 and DOX-resistant A2780/AD human ovarian carcinoma cells/xenografts validated the hypothesis. In contrast to free DOX the IC50 dose [297] and tumor treatment efficacy [91] of the polymer conjugate was very similar for both cell types. Additionally, chronic exposure of A2780 cells to HPMA copolymer-DOX conjugate did not induce resistance, whereas chronic exposure to DOX resulted in a new phenotype expressing the MDR1 gene that encodes Pgp [298]. Similarly, polymer bound mesochlorin e6 (Mce6) also did not induce multidrug resistance following chronic exposure to human ovarian carcinoma cells [299].

Another approach is the design of a HPMA copolymer that contains both DOX and Pgp inhibitor R121 bound via hydrolytically cleavable hydrazone bonds. The conjugate was efficient in the treatment of resistant P388/MDR and CT26 tumors [300]. Alternatively, anti-Pgp antibody targeted polymer bound photosensitizer Mce6 provided an efficient conjugate. An important design factor was the activity of polymer-bound Mce6 to produce singlet oxygen before internalization [244].

3.5. Combination therapy with polymer-drug conjugates

Treatment of cancer in the clinics is frequently performed with combination of drugs. Combinations of polymer-bound drugs have been evaluated for several decades [301]. The factors that were studied include combination of two polymer conjugates vs. polymer conjugate that contains both drugs covalently attached; sequence of administration (which drug first, or simultaneous administration), and dosing in synergistic concentrations. Another approach is the combination of polymer-drug conjugates (chemotherapy) with physical stimuli – focused ultrasound, photothermal, and photodynamic therapy (activation by light).

Satchi-Fainaro’s group found that treating murine model of mammary adenocarcinoma with a PGA conjugate containing two drugs, PTX and DOX (PGA-PTX-DOX), was more efficient than a mixture of two separate conjugates containing one drug each [302,303]. Obviously, the evaluation of a combination of two separate conjugates is easier since changing of dose ratio is easy; for conjugates containing two drugs a change in concentration ratio requests a new synthesis. One can assume that if the two drugs have similar hydrophobicity, then the biodistribution of respective conjugates will be similar and two conjugates may be advantageous. However, for drugs with different properties one conjugate should be evaluated.

Numerous combinations of polymer-drug conjugates were evaluated with enhanced efficacy of the combination when compared to the monotreatment [70,113,304]. An interesting combination treatment is the combination of two conjugates that target different cell subpopulations, cancer stem cells and differentiated cells. The eradication of stem cells would prevent their self-renewal, cancer recurrence, and metastasis. In a prostate cancer model, the stem cells could be downregulated by blocking the hedgehog pathway; the differentiated cells are sensitive to the action of clinically used docetaxel. Indeed, the combination of two HPMA copolymer conjugates, one containing cyclopamine, a hedgehog pathway inhibitor, the second containing docetaxel (DTX) were active in vitro [305] and demonstrated long-term tumor growth inhibition in nude mice bearing PC-3 prostate cancer xenografts. [306]. Similarly, a combination of a DTX conjugate with a conjugate containing GDC-0980, a dual PI3K/mTOR inhibitor, was effective in the PC-3 prostate cancer model [307].

Combination treatments that involve physical stimuli are efficient in cancer treatment

In combination chemotherapy with photodynamic therapy (PDT), polymer-bound chemodrugs are combined with polymer-bound photosensitizers (PS)s. The latter are not active in dark, but when irradiated with a characteristic wavelength the PS gets into its first excited singlet state. Intersystem crossing may convert the singlet state PS into its triplet state. Exchange of energy with ground state triplet state oxygen produces singlet oxygen (1O2) a reactive species indeed. The 1O2 has a short half-life in aqueous environment, it interacts with biological molecules close to the location where it originated. Photosensitizers are active both when covalently bound to the polymer carrier or when released inside the cell. However, the quantum yield of singlet oxygen formation is substantially higher when released from the carrier [301]. Consequently, enzymatically degradable spacers are preferred for PS attachment and PS that are activated with near infrared wavelengths are more efficient due to deeper penetration of light into the tissue. Combination of HPMA copolymer-DOX and HPMA copolymer-Mce6 (mesochlorin e6) conjugates was more active in the treatment of Neuro 2A neuroblastoma in A/J mice [301], treatment of human ovarian OVCAR-3 xenografts in female athymic mice than each of the individual therapies [308]; OV-TL-16 antibody targeted HPMA copolymer-DOX and - Mce6 conjugates were more active in the OVCAR-3 model than non-targeted conjugates [309].

Larson et al. used moderate (43 °C, 30 min) gold nanorod induced tumor hyperthermia to increase the delivery and efficacy of heat shock protein (HSP) targeted (mediated by peptide WDLAWMFRLPVG) HPMA copolymer-DTX and -aminohexylgeldanamycin conjugates in the treatment of DU145 prostate cancer in vitro and in DU145 tumor bearing mice. Hyperthermia induced the expression of cell surface HSP glucose regulated protein 78 kDa (GRP78) and synergistically enhanced the efficacy of the polymer-drug conjugates [310]. Using the same DU145 prostate cancer model in nu/nu mice the hyperthermia was induced by high intensity focused ultrasound. Enhancement of efficacy of HPMA copolymer-DTX was observed [311].

3.6. Potential immunogenicity/immunostimulation of polymer carriers

When evaluating polymer-drug conjugates both the cytostatic and immunomodulating effects have to be evaluated [24]. Due to the complex structure of these nanomedicines, test on the immunogenicity of the carrier, immunogenicity of the oligopeptide sequence used for drug attachment or self-assembly, the possibility that the drug attached to the side-chain termini may act as hapten need to be performed. Also, potential immunosuppression by depleting immune cells needs to be evaluated. When homopolymers are used for protein modification the situation is simpler – the immunogenicity of the polymer is the main factor. Evaluation of potential immunotoxicity of nanomedicines is an important part of research [312].

Detailed studies were performed for PEG [124129] and HPMA copolymers [14,313318]. Whereas administration of PEG results in the production of IgM antibodies, HPMA homopolymer is not recognized as an antigen and does not act as an immunogen. The attachment of oligopeptide spacers induces mild immunogenicity, but several orders of magnitude lower than bovine gamma-globulin [318]. Doxorubicin bound to GFLG spacers did not act as a hapten [318]. Interestingly, HPMA copolymer-DOX conjugate induces immunogenic cell death [24].

Notably, immunostimulatory properties of HPMA copolymer-drug conjugates were observed. Mice previously cured with HPMA copolymer-DOX conjugates survived cancer re-challenge without treatment [14]. Thus, these conjugates produce a double attack on cancer cells – cytotoxicity and immunostimulation. Importantly, clinical data revealed that similarly to experiments on animals both cytotoxic and immunostimulating role of HPMA copolymer-DOX-IgG conjugates was observed [313315]. Similarly, camptothecin conjugated to cyclodextrin-based polymer (IT-101) modulated the immune response to enhance drug efficacy [178].

HPMA copolymer-peptide conjugates described in Section 5 contain peptide grafts 38 amino acid residues long. Detailed studies in vitro and in vivo using both D- and L- peptides have shown no immune response to free peptides and their copolymer conjugates in vitro. In vivo (in BALB/c mice) minor production of antibodies was detected and no cellular response observed when administered with complete Freund’s adjuvant [319].

3.7. Clinical experience

Polymer-drug conjugates and similar nanomedicines (frequently named nanoparticles) possess numerous advantages over free drugs [9,10,320,321]. However, the clinical potential of conjugates with low molecular weight drugs has not been fully exploited [10,217,322]. Yet, data from activity on patients [163,164,313,314,323,324] and the development of long-circulating backbone degradable HPMA copolymer-drug conjugates [108,109,113,270] bode well for the future. As discussed in 3.1 an important part of speeding up the translation is the development of biomarkers that will permit patient stratification before clinical trials.

The major clinical use of polymer conjugates are the PEG-modified proteins. Numerous FDA approved conjugates [30,114,115] and the intensive research on PEG alternatives [129,140] demonstrate the success of this field. Generally, nanomedicines are progressing to clinics with an acceptable speed. Recent reviews [30,320,321,325,326] and opinions [327329] summarize the numerous clinical trials and FDA approved nanomedicines. Very recent successes include Onpattro®, Givlaari®, and Onivide®.

4. New directions in macromolecular therapeutics

In spite of several negative reports on the future of nanomedicine [219,220,330], we and others [327329,331333] strongly believe in the bright future of macromolecular therapeutics. Obviously, as any new science field that is well funded, nanomedicine experienced a period of inflated expectations. Using Gartner “hype cycle” terminology [334] nanomedicine is beyond the period of disillusionment and basic research by devoted scientists brought it back to a plateau of excellent productivity. This is one important aspect of future developments. Excellent progress will be made by scientists devoted to targeted basic research without the disturbing positive or negative hype.

Another important approach to macromolecular therapeutics will be the design of new paradigms based on the knowledge of mechanisms of physiological processes. The design of new paradigms is an extremely important part of scientific progress. However, it is important to avoid packaging old concepts in more complex approaches/structures [335] and elude superficiality [336]. Examples of two new designs, both based on employing receptor crosslinking, are presented below: a) improvement in combination chemotherapy and immunotherapy by designing a new multivalent anti PD-L1 antagonist; and b) design of drug-free macromolecular therapeutics.

4.1. Combination chemotherapy and immunotherapy with macromolecular therapeutics

Many chemotherapeutic drugs have significant immunosuppressive side effects, either directly, by inhibiting or killing effector cells, or indirectly, by provoking allergy or immune paralysis [337]. However, over the past decades, considerable evidence demonstrated cytotoxic drugs have also favorable function in the activation of the immune system [338]. There are two ways to elicit the immune system: some therapeutics can stimulate specific cellular responses that render tumor-cell death immunogenic, resulting in enrichment of T-cell infiltration in the tumor, whereas other drugs may have inhibition effects on regulatory T cell activity and foster an immunosuppressive tumor microenvironment, therefore improve immunotherapy response rate [339,340].

Immune checkpoint blockade (ICB) is a formidable strategy for cancer treatment. Immune checkpoints are regulators of immune activation. Tumor cells often overexpress immune checkpoint proteins to escape immunosurveillance and avoid attack by the immune system. T-lymphocyte-associated antigen 4 (CTLA-4) and programmed cell death protein 1 (PD-1) promote immune evasion and result in cancer immunoresistance [341,342]. The blockade of immune checkpoints, e.g., blockade of the interaction of programmed cell death ligand 1 (PD-L1 on cancer cells) with PD-1 on T cells using anti-PD-1 or anti-PD-L1 antibodies is a successful cancer treating strategy [343]. In addition to antibodies, small molecules, peptides, and macrocycles are efficient checkpoint inhibitors [344].

A myriad of factors contributes to the low response rate of immune checkpoint blockade (ICB). For example, tumor-infiltration of T lymphocytes is associated with prognostic implication. There is increasing evidence that patients with low effector T cell recruitment in tumor microenvironment fail to such immunotherapies. Therefore, to exploit ICB for cancer treatment, it is important to expand tumor-infiltrating lymphocytes (TIL). One of such strategies is immunogenic cell death induced by specific chemotherapy drugs.

Combining chemotherapies with immunotherapies creates the possibility of synergistic and/or additive effects and therefore, offers great promise in modulating tumor microenvironment and treating cancers successfully. Comparing with free drugs, polymer-drug conjugates have advantages in ‘priming’ tumors attributed to their long-circulation in blood stream and favorable accumulation in tumors. However, the combined approach is effective in inflamed “hot tumors” (lung, melanoma, liver, bladder, head and neck cancers); the treatment outcome is considerably poorer in non-inflamed “cold tumors” (estrogen receptor positive breast and prostate cancers) [345]. The distinction between hot and cold tumors relates to the amount of T cells inside tumor and expression of PD-1 and PD-L1. Fortunately, some anticancer drugs and polymer-drug conjugates can initiate immunogenic cell death (ICD) and convert cold tumors into warm or hot ones [270,338,346,347].

4.1.1. Immunogenic cell death by polymer-drug conjugates

Cytotoxic drugs not only exhibit chemotherapy effects on tumor cells, but they also affect the immune system to contribute to tumor regression. For example, anthracyclines such as doxorubicin and epirubicin induce tumor-specific immune response [338,339] by depleting regulatory T cells (Treg)s, stimulating dendritic cells (DC), and generating tumor immunogenic cell death (ICD). Cells which undergo ICD are characterized by pre-apoptotic translocation of calreticulin to cell surface and post-apoptotic release large amounts of ATP and high-mobility group box 1 (HMGB1) protein into the extracellular space [348,349]. The HMGB1 released from immunogenically dying tumor cells binds to several receptors such as Toll-like receptors 2 and 4, which are expressed on antigen-presenting cells, and functions as a “danger” signal to promote activation of DCs and boost antigen-specific T cell responses. Consequently, ICD can be utilized not only for killing cancer cells but also for “heating-up” the antitumor immunity, priming the tumor for consecutive immune checkpoint inhibition therapy.

Polymer-drug conjugates demonstrated their efficiency in inducing ICD. HPMA copolymer-doxorubicin conjugates demonstrated immunostimulating potential. In EL-4 lymphoma cancer model mice survived re-transplantation with cancer cells without further treatment [14,350]. Backbone degradable HPMA copolymer-epirubicin conjugate (2P-EPI; KT-1) triggered an increase of calreticulin expression and release of HMGB1 in 4T1 cells in vitro. In vivo in BALB/c mice bearing 4T1 tumors, 2P-EPI produced higher enhancement of calreticulin expression and HMGB1 release than free EPI (Figure 6). Importantly, 2P-EPI attracted considerably more CD8+ cells into the tumor than free EPI (Figure 6F). Both free EPI and 2P-EPI enriched PD-L1 expression on tumor cells (Figure 6G). The efficacy of 2P-EPI is supported by CD*+ T cells. Coadministration of anti-CD8+ antibodies considerably weakened 2P-EPI mediated tumor regression (compare Figures 6D and 6E) [270].

Figure 6.

Figure 6.

Immunogenic cell death of 4T1 murine breast cancer cells/ BALB/c mice bearing 4T1 tumors treated with backbone degradable, long circulating HPMA copolymer-epirubicin conjugate (2P-EPI). (A) Conversion of cold tumor into hot tumors as detected by calreticulin expression (B), high mobility group box 1 protein (HMGB1) release (C) and T-lymphocyte infiltration. (D) Tumor growth curves following treatment with saline, EPI, or 2P-EPI (KT-1); (E) Treatment with saline, or 2P-EPI (KT-1) plus anti-CD8 antibody (to disable CD8 positive T cells). (F) Tumor recruitment of CD8+ cells; (G) PD-L1 expression on tumor cells after chemotherapy with EPI or 2P-EPI (KT-1). Adapted from ref. [270].

Moon’s lab has shown ICD and robust T cell response in CT26 and MC38 colon carcinoma cells exposed to DOX containing high density lipoprotein-mimicking nanodiscs [351]. Priming tumors by ICD for subsequent checkpoint inhibition therapy seems to be effective when using nanomedicines of various structures and combinations, such as nab-paclitaxel [352], mitoxantrone combined with multilamellar lipid-hyaluronic acid adjuvant nanodepots [353], and Doxil [354].

4.1.2. Immune checkpoint blockade by multivalent polymer peptide antagonists (MPPA)

Combination of ICD inducing chemotherapy with immune checkpoint blockade using anti-PD-1 or anti-PD-L1 antibodies has shown marked clinical success in melanoma, prostate, kidney, and lung cancer. However, the response rate in breast cancer is low due to the immunologically “cold” characteristics including low amount of cytotoxic T lymphocytes inside the cancer tissue. Recent studies demonstrated active redistribution of PD-L1 back to the cell surface following application of anti-PD-L1 antibodies, suggesting traditional mAb-based approaches produce inefficient blockades of PD-L1 and lead to adaptive therapy resistance [355,356].

To generate a more effective PD-L1 elimination strategy the subcellular trafficking of PD-L1 needs to be changed from recycling to the lysosomal route. Wang et al. have shown that HIP1R (Huntingtin interacting protein 1 related protein) targets PD-L1 to lysosomal degradation (through a lysosomal targeting signal) and alters T cell mediated cytotoxicity [357]. Another approach to decrease the PD-L1 surface expression is to inhibit palmitoylation of the intracellular domain of PD-L1 by palmitoyltransferase ZDHHC3. Palmitoylation of proteins acts as a mechanism to regulate protein localization and function. Inhibition of PD-L1 palmitoylation decreased the PD-L1 expression on tumor cells [358]. Additionally, CMTM6 depletion prevents PD-L1 recycling and decreases PD-L1 surface expression [359].

We proposed a new design of a PD-1/PD-L1 inhibition based on multivalency and receptor crosslinking - multivalent polymer-peptide antagonist (MPPA). It is well known that receptor crosslinking manipulates subcellular trafficking of receptor-bound ligands to the endosomal/lysosomal route [265]. Peptides have become interesting alternatives to block PD-1/PD-L1 interactions [360]. Chang et al. developed a peptide-based antagonist to the PD-1/PD-L1 pathway using phage display. This antagonist (PPA-1, sequence NYSKPTDRQYHF) has been shown not only to possess a high binding affinity to PD-L1 in vitro, but also to effectively disrupt the PD-1/PD-L1 interaction both in vitro and in tumor-bearing mice [361]. MPPA is a linear HPMA copolymer containing multiple PPA-1 grafts. Such a construct will hyper-crosslink the cell surface PD-L1 receptors and efficiently traffic the complex to the lysosome for degradation by lysosomal enzymes. This results in decreased PD-L1 expression at cell surface and enhanced antitumor activity [270].

4.1.3. Efficacy of combination of 2P-EPI (KT-1) with MPPA

As an example of the efficacy of combination chemotherapy and immunotherapy the treatment of triple negative 4T1 breast cancer with 2P-EPI followed by MPPA is presented (Figure 7). The 4T1 syngeneic cancer model shares genomic features of basal-like breast cancer which is a non-immunogenic tumor. The treatment strategy involves (a) immunogenic chemotherapy with long-circulating backbone degradable epirubicin conjugate KT-1, and (b) PD-L1 degradation immunotherapy with multivalent HPMA polymer-peptide antagonist to PD-L1 (MPPA). First, KT-1 immunologically “heated up” 4T1 tumor mediated by ICD. In the second step, multivalent MPPA was administered. This approach resulted in 10/10 complete tumor regressions (Figure 7F). For comparison, treatment with 2P-EPI (KT-1) and anti-PD-L1 antibodies produced 8/10 complete tumor regressions (Figure 7C) and treatment with 2P-EPI (KT-1) and peptide PPA 2/10 complete tumor regressions (Figure 7F). Treatment with only immunotherapy, PPA or MPPA was not effective (Figure 7F). The excellent efficacy of the combination with MPPA is a consequence of different subcellular fate of PD-L1 following receptor crosslinking (Figure 7D). The decreased expression of PD-L1 after treatment with MPPA (Figure 7E) validates the hypothesis on the impact of receptor crosslinking on the efficacy of immunotherapy.

Figure 7.

Figure 7.

Combined chemotherapy and immunotherapy; design of multivalent polymer peptide PD-L1 antagonist (MPPA). (A) Scheme of PD-1/PD-L1 immune checkpoint blockade (ICB). (B) Treatment of BALB/c mice bearing 4T1 tumor with combination chemotherapy (EPI or 2P-EPI (KT-1)) and immunotherapy (anti-PD-L1 antibodies): schedule of dosing and tumor growth curves. (C) Individual tumor growth curves following administration of KT-1 and anti-PD-L1 antibodies; 8/10 complete tumor regressions. (D) MPPA changes the subcellular fate of PD-L1. Following binding of anti-PD-L1 antibodies the PD-L1 receptor recycles back to the surface. In contrast, binding of multivalent MPPA crosslinks the PD-L1 receptors and biases its subcellular fate into lysosomes, where they degrade. (E) Recovery of surface PD-L1 after treatments with anti-PD-L1 antibodies, PPA, or MPPA. (F) Individual tumor volumes for combination treatment of 4T1 BALB/c mice tumors with saline, PPA peptide, multivalent MPPA, 2P-EPI (KT-1) followed by PPA, and 2P-EPI (KT-1) followed by MPPA. Combination of 2P-EPI (KT-1) with MPPA produced 10/10 complete tumor regressions. Adapted from ref. [270].

4.2. Drug-free macromolecular therapeutics (DFMT)

4.2.1. First generation of DFMT

The concept of DMFT is a new paradigm in macromolecular therapeutics – treating diseases by biorecognition at cell surface and crosslinking of receptors without the need for low molecular weight drug [362,363]. Its concept emerged from the study of self-assembly of hybrid macromolecules composed of HPMA backbone and peptide grafts. Two pentaheptad peptides (CCE and CCK) with differing charge were bound to HPMA copolymer, respectively. Separately, the conformation of CCE and CCK is random coils, but their equimolar mixture forms antiparallel coiled-coil heterodimers and could serve as physical crosslinkers. Indeed, HPMA graft copolymers, P-(CCE)x and P-(CCK)y, self-assemble into hybrid hydrogels with a high degree of biorecognition [364].

Efforts to use the pair of peptides in biological processes logically focused on crosslinking of receptors – crosslinking of hydrogels is based on similar physicochemical phenomena. The CD20 receptor is one of the most reliable cell surface markers of B lymphocytes. It has a low rate of internalization; when crosslinked it relocates to lipid rafts, resulting in calcium influx, mitochondrial depolarization, and caspase 3 activation [365367]. The new DMFT system is composed of two nanoconjugates: a) bispecific engager: anti-CD20 Fab’ antibody fragment attached to CCE (Fab’-CCE); and b) crosslinking (effector) component: HPMA copolymer containing multiple grafts of the complementary peptide, CCK (P-(CCK)x). Fab’ was chosen over Ab to avoid crosslinking of CD20 bound Abs by immunocompetent cells.

The DFMT system translates the biomaterial biorecognition principles to nanomedicine [368]. Its three main features are: a) absence of a low molecular weight drug; b) pretargeting – the two-step treatment permits to manipulate the pharmacokinetics by separating the targeting modality from the effector modality; and c) multivalency – the second nanoconjugate is multivalent; its efficiency is valency dependent (Figure 8). The term DFMT is based on experimental observations that both parts of the system, the bispecific engager and the crosslinking effector do not initiate apoptosis when used individually. Only when both nanoconjugates colocalize at cell surface, resulting in receptor crosslinking, apoptosis is initiated.

Figure 8.

Figure 8.

Design of drug-free macromolecular therapeutics (DFMT). Exposing B cells to a conjugate of one motif with a CD20 targeting ligand (Fab’-motif1) decorates the cells with this motif. Further exposure of decorated cells to a macromolecule (synthetic polymer or human serum albumin) grafted with multiple copies of the complementary motif [P-(motif2)x or HSA-(motif2)x] results in receptor crosslinking and apoptosis induction [362,363].

The design is active both in vitro [368] and in vivo on disseminated Raji B cell lymphoma in CB17 SCID mice [369]. The structure of the crosslinking effector nanoconjugate was modified; HPMA copolymer was replaced by human serum albumin, HSA-(CCK)x and high apoptotic levels were obtained [370].

Advanced design employed complementary 25 base pair morpholino oligonucleotides instead of peptides, resulting in Fab’-MORF1 and P-(MORF2)x (or HSA-(CCK)x) nanoconjugates. Detailed studies of apoptotic mechanism in Raji B cells revealed that, following biorecognition of two nanoconjugates at cell surface, two pathways are involved in apoptosis initiation: a) CD20 crosslinking→calcium influx→mitochondrial depolarization; b) Bcl-2 inhibition→mitochondrial depolarization→cytochrome c release→caspase 3 activation [371]. These data are supported by visualization of biorecognition at cell surface by super resolution imaging [372,373]. The morpholino based system was very active with both effectors, based on HPMA copolymer [267,371] and on HSA [268]. DMFT overpowers rituximab resistance by amplifying surface CD20 crosslinking through enhancing surface CD20 expression following pretreatment with gemcitabine; by enhancing surface CD20 accessibility due to absence of Fc fragment; and by multivalency of P-(MORF2)x) [374].

Efficacy of DMFT in cells isolated from patients with various B cell malignancies

Apoptosis induction by DFMT in cells from 44 patients with different B-cell malignancies, chronic lymphocytic leukemia (CLL), diffuse large B cell lymphoma (DLBCL), marginal zone lymphoma (MZL), follicular lymphoma (FL), mantle cell lymphoma (MCL), and Burkitt’s lymphoma (BL) was investigated. DFMT induced apoptosis in 65.9% of patient samples [375]. High-risk mutations such as 13q14, 17p13 and 11q22 deletions, which are considered poor prognostic factors in CLL, did not hamper the therapeutic efficacy of DFMT treatment [375]. On the other hand, in patient samples with poor response to DFMT treatment, low CD20 expression level was observed. Pre-treatment with gemcitabine enhanced surface CD20 expression and restored the cell responsiveness to DFMT [374]. DFMT effectively increased apoptosis of tumor cells from patients with a variety of B cell malignancies, irrespective of genomic aberrations. The results, in agreement with our early data [267,376], suggest that DFMT is a capable modality for the treatment of various B cell malignancies.

4.2.2. Second generation of DFMT

Rituximab (RTX) and other anti-CD20 antibodies (ofatumumab and obinutuzumab (OBN)) dramatically improved treatment of Non-Hodgkin lymphoma (NHL) and CLL. On their own or in combination with chemotherapy (e.g. R-CHOP, a combination with cyclophosphamide, doxorubicin, vincristine, and prednisone) they produced better clinical outcomes [366,377]. However, frequent relapses and poor patient outcomes demonstrate the need for improved treatment strategies. To this end, the second generation of DFMT for the treatment of B-cell malignancies was designed.

Type I (RTX) and Type II (OBN) anti CD20 antibodies have distinctive patterns of binding to the CD20 receptor. RTX binds between CD20 tetramers resulting in accumulation in lipid rafts, calcium influx and caspase activation [377]. OBN binds within one tetramer with the conformation consistent with homotypic adhesion regions, steering to actin cytoskeleton remodeling and lysosome disruption [378,379]. The second generation DFMT enhances the activity of Type II OBN by triggering the apoptosis activation pathways of both types of antibodies. The system consists of two nanoconjugates: a) bispecific engager, OBN-MORF1 (OBN conjugated to one morpholino oligonucleotide MORF1); and b) a crosslinking (effector) component HSA-(MORF2)X (human serum albumin grafted with multiple copies of complementary MORF2). Modification of OBN with one MORF1 does not affect the binding of OBN-MORF1 to CD20 and following binding Type II effects occur. Further exposure to multivalent HSA-(MORF2)X results in receptor crosslinking and clustering the OBN-MORF1-CD20 complexes into lipid rafts and Type I effects occur (Figure 9). This new approach, called “clustered OBN (cOBN)” combines effects of both antibody types resulting in very high apoptotic readings.

Figure 9.

Figure 9.

Mechanism of apoptosis induction of second-generation drug-free macromolecular therapeutics (DFMT) (clustered obinutuzumab, cOBN): combination of Type I and II effects into one system). (A+B) Mechanism of action of cOBN. Following binding of OBN-MORF1 to CD20 Type II effects occur. Further exposure of cells with multivalent HSA-(MORF2)x results in crosslinking of CD20 receptors, clustering the OBN-MORF1-CD20 complexes into lipid rafts and onset of Type I effects. (C) Caspase activity after Raji cell treatment with RTX, cRTX (clustered rituximab), OBN, or cOBN. *P < 0.05, n.s, not significant, by students t-test. (D) Comparison of whole antibodies (RTX and OBN) and their Fab’ fragments after multivalent binding of HSA-(MORF2)10 to induce apoptosis. (E) Reactive oxygen species (ROS) generation after Raji cells were treated with RTX, cRTX, OBN, or cOBN. (G) Treatment of NRG mice bearing disseminated Raji B-cell lymphoma by clustered obinutuzumab (OBN-MORF1 followed by HSA-(MORF2)10). Treatment schedule (dosing: OBN (0.5 nmol), or cOBN (OBN-MORF1, 0.5 nmol→HSA-(MORF2)10 and 1.5 nmol MORF2, 5 h later); Paralysis-free survival; Residual Raji cells (human CD10+CD19+) in bone marrow (BM) following exposure to saline, OBN, or cOBN. BM cells isolated from native NRG mouse served as the negative control, and Raji cells served as the positive control. *P<0.05, **P<0.01, by students t-test. Adapted from ref. [380].

The design is scientifically novel and has great translational potential. Also, it provides a new paradigm for the design of macromolecular therapeutics applicable to other diseases beyond lymphomas [380].

In vivo activity of cOBN was tested on NOD/SCID-Rag1null γnull (NRG) mice bearing systemically disseminated Raji B-cell lymphoma. Vehicle (saline) treated mice became hind limb paralyzed at 19 days (median survival). OBN significantly extended the median survival to 57 days. Significantly, cOBN further prolonged the mice survival with the median survival of 83 days, and half of the mice were still paralysis-free after 100 days (Figure 9) without loss of body weight [380].

4.2.3. Other potential targets for DFMT

An excellent example of suitable targets for DMFT are death receptors 4 and 5 (DR4, DR5). The clinical inefficiency of therapeutic anti-DR5 antibodies stimulated the design of multivalent receptor crosslinking agonists [381386]. Such multivalent nanoconjugates induce apoptosis without the need for FcγR-based crosslinking in contrast to antibodies that require crosslinking for exhibition of activity. Successful examples of multivalent conjugates are: Crosslinking of DR5 by drug-free immunoliposomes decorated by multiple copies of anti-DR5 Ab (TRA-8) or Fab’ antibody fragments triggers the employment of the death-inducing signaling complex and initiates apoptosis [381]. Huet et al. prepared multivalent nanobody DR agonists. Increasing the valency of domains to tetramer and pentamer resulted in enhanced potency for cell killing. Multivalent DR5 nanobodies were efficient in the treatment of patient derived pancreatic and colon cancer models [382]. A nanoconjugate composed of six single chain TRAIL-receptor-binding domains (scTRAIL-RBD) is a potent inducer of apoptosis in vitro and exhibited dose dependent anticancer activity in mice bearing Colo205 xenografts [385]. The augmentation of biorecognition of multivalent conjugates is strong. Trimeric and hexameric forms of DR5-binding peptide attached to adamantane-based dendrons augmented affinity about 1500 × and 20,000 ×, respectively when contrasted to the monomer. Only multivalent constructs induced apoptosis in DR5 expressing cells [386].

DMFT has a potential with additional receptors that are slowly internalizing, such as CD45 [387], prostate stem cell antigen [388], carcinoembryonic antigen [389], CD79b [390], and CD22 [391]. DMFT is also suitable for the treatment of diseases where CD20 crosslinking and B cell depletion therapy is suitable, such as rheumatoid arthritis [392,393], systemic lupus erythematosus [394], multiple sclerosis [395] and organ transplantation [396].

DMFT based on murine anti-CD20 Fab’-MORF1 and P-(MORF2)X was effective in the treatment of collagen-induced rheumatoid arthritis in a mouse model [392].

4.2.4. E-selectin targeted drug-free macromolecular conjugates

Macromolecular therapeutics have a great potential to treat endothelial disorders in cancer and other diseases [397]. Efficient E-selectin binding peptides were identified and established their ability to manipulate biological processes in need of E-selectin interactions [398,399]. Notably, employing multivalent peptide – E-selectin interactions, the concept of drug-free therapeutics was applied also for prevention of metastasis and healing of microvascular and vascular inflammation.

David and coworkers used HPMA copolymer grafted with multiple copies of E-selectin binding peptide (ESBP; CDITWDQLWDLMK-CO-NH2) to demonstrate that this conjugate hindered metastasis development in vivo by interrupting the interaction between cancer and endothelial cells [400]. In two independent experiments it was demonstrated that the drug-free conjugate was as efficient as the ESBP containing conjugate with drug also attached. P-ESBP (P is the HPMA copolymer backbone) was as efficient as the conjugate containing dexamethasone (P-ESBP-DEX) in the treatment of microvascular and vascular inflammation and in the inhibition of development and progression of atherosclerosis in atherosclerotic ApoE(−/−) mice [401]. Similarly, P-(A5G27)x an HPMA copolymer grafted with peptide A5G27 (RLVSYNGIIFFLK) binding to CD44v3 and CD44v6, inhibited cancer cell migration [211] and was as proficient in tumor growth inhibition as the P-(A5G27)x-paclitaxel conjugate [402].

5. Beyond cancer

Polymer-drug conjugates are suitable for the treatment of non-cancerous diseases, such as musculoskeletal and infectious diseases [403]. Examples of such nanomedicines follow:

5.1. Bone delivery

Nanomedicines are an important part of bone therapeutics development. Bone and bone diseases can be targeted by nanomedicines containing acidic oligopeptides, bisphosphonates, tetracycline, and chelating compounds [403405]. The location of binding to bone depends in the crystallinity of hydroxyapatite. D-aspartic acid octapeptide (D-Asp8) would recognize resorption sites in skeletal tissues, whereas alendronate would direct bound cargo to both bone resorption and formation sites. This selectivity relates to the strength of the binding force between targeting moiety and bone as determined by atomic force microscopy [406].

HPMA copolymer – prostaglandin E1 (PGE1) conjugates targeted by D-Asp8 were designed for the treatment of osteoporosis. The PGE1 was bound to the carrier via a self-immolating spacer containing a cathepsin K sensitive Gly-Gly-Pro-Nle oligopeptide [42]. When the conjugate attaches to bone, the anabolic agent PGE1 will be released at sites of higher osteoclast activity by cathepsin K mediated cleavage. EP receptors on bone cell surfaces will be activated by released PGE1 resulting in bone formation [407]. Pharmacokinetic and biodistribution studies as well as evaluation of bone formation in ovariectomized rats demonstrated efficacy of the PGE1 conjugates [208,408]. Interestingly, second generation, long-circulating, backbone degradable HPMA copolymer carriers of PGE1 were more effective in promoting bone growth in ovariectomized rats than first generation nondegradable conjugates [409].

Alendronate-targeted HPMA copolymer-TNP470 antiangiogenic conjugates [410,411] as well as alendronate-HPMA copolymer-paclitaxel conjugates [412] were successful in the treatment of animal models of bone cancer [410,411]. Similarly, a poly(oligoethylene glycol)methacrylate – combretastatin A-4 - alendronate conjugate exhibited bone recognition and antiangiogenic activity against HUVEC and U2-OS cells [413].

5.2. Inflammation

Water soluble polymer-drug conjugates are effective in the treatment of persistent and chronic inflammation, such as rheumatoid arthritis, multiple sclerosis, systemic lupus erythematosus, inflammatory bowel disease, and others [414]. Macromolecular therapeutics favorably build up in inflammatory tissues [403] especially in arthritis [415417]. This phenomenon was named “ELVIS” (Extravasation through Leaky Vasculature and the subsequent Inflammatory cell-mediated Sequestration) suggesting the potential of polymer-drug conjugates in the treatment of inflammatory diseases [416,418]. Different biologically active molecules were attached to polymeric carriers to treat rheumatoid arthritis [419] including cathepsin K inhibitors [420], dexamethasone [421], and Janus kinase inhibitor [422].

A comparative study evaluated four DEX containing nanomedicines: liposome, core-crosslinked micelle, slow (drug) releasing HPMA copolymer-DEX conjugate, and fast-releasing HPMA copolymer-DEX conjugate. Following a single i.v. injection, nanomedicines with slower drug release kinetics (micelle and slow releasing HPMA copolymer-DEX conjugate) maintained longer duration of therapeutic activity in a rat model of adjuvant-induced arthritis than those with fast DEX release [423].

HPMA copolymer – tofacinitib conjugate [424] and HPMA copolymer – DEX (P-DEX) conjugate [425] were efficient in ameliorating dextran sulfate sodium induced ulcerative colitis. P-DEX also demonstrated more effective anti-inflammatory effects in the treatment of murine lupus nephritis in mice when compared to free DEX. P-DEX eliminated albuminuria, reduced macrophage recruitment to the kidney and extended the life span of mice [426]. The relationship between structure of P-DEX (molecular weight, DEX content) on one hand and pharmacokinetics and biodistribution on the other hand in mice in an aseptic implant loosening mouse model was recently evaluated [427].

Dextran [428] and PEG conjugates [429,430] were also effective in the treatment of inflammatory diseases on animal models.

6. Conclusions and future prospects

A young drug delivery scientist might be disappointed when reading several negative articles on the status and future prospects of nanomedicines [219,330]. To get a balanced non-biased picture we recommend to read opinions that represent the majority [228,327329,331333]. The facts are well described in recent reviews that summarize the status of the nanomedicine field and show successful current translation efforts, an area of frequent discussions [30,320,321,326]. Obviously, the drug delivery field, as any research area that is well funded, attracts “me-too” scientists. However, the majority of research in the area is based on serious efforts to move the science and translation of the research forward. As shown in several reports [328,329,331] not only results of basic research but also clinical translation has numerous success stories to present. Universities are not structured to translate research data into the clinics. This is the task of the industry. Enormous activities of start-up companies keep the pipeline strong. However, large pharma should not wait for the results of phase II clinical trials before getting involved. There should be a concerted effort to help start-up companies to overcome the “valley-of-the-death”. It would be naïve to anticipate that big pharma would do this voluntary. May be tax credit for such support could initiate the process.

The field of polymer-drug conjugates developed enormously and principles of design are matched with the knowledge of biological and immunological effects and delivery barriers these conjugates are exposed in vivo. Based on the results of early clinical trials second generation backbone degradable polymeric carriers were designed. A plentiful of polymer carriers with different structures are available for the creation of new nanomedicines. The general knowledge of the biological attributes of cancer, including heterogeneity of cancer cells, tumor microenvironment, and metastasis has significantly advanced [11,101,236241]. In addition to systematic research in improving new conjugates and modifying the tumor and its microenvironment, imaging agents are being developed to select patients for clinical evaluation that will be amenable to the treatment with particular macromolecular therapeutics [229,231,232].

In addition, new paradigms in macromolecular therapeutics are emerging. They are an important part of the development and will open new horizons. Two new designs, based on receptor crosslinking, are discussed in this review. The first is the design of a multivalent polymer peptide antagonist (MPPA) to PD-L1. MPPA changes the subcellular fate of PD-L1. Instead of recycling it biases the PD-L1 to the endosomal/lysosomal route for targeted degradation in the lysosomes. The second example are drug-free macromolecular therapeutics (DFMT), a paradigm where apoptosis is initiated by receptor crosslinking without the need for a low molecular weight drug. This concept is already expanding by the design of E-selecting binding polymer-peptide conjugates active in prevention of metastasis and treatment of vascular inflammation, and by constructing the second generation of DFMT, a single platform that triggers two mechanisms of cell death.

We are confident that many new, more efficient macromolecular therapeutics will appear in the literature as a result of systematic basic research and new innovative designs of the future.

Acknowledgments

The research in the authors‘ laboratory was supported in part by the National Institutes of Health (recently NIH grants R42 CA156933, RO1 GM95606, RO1 CA246716), Department of Defense (recently grant W81XWH2019573), TheraTarget, Bastion Biologics, Huntsman Cancer Institute, and the University of Utah Research Foundation.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • [1].Jatzkewitz H, Peptamin (glycyl-L-leucyl-mescaline) bound to blood plasma expander (polyvinylpyrrolidone) as a new depot form of a biologically active primary amine (mescaline), Z. Naturforsch. 10b (1955) 27–31. [Google Scholar]
  • [2].Givetal NI, Ushakov SN, Panarin EF, Popova GO, Experimental studies on penicillin polymer derivatives (in Russian), Antibiotiki 10 (1965) 701–706. [PubMed] [Google Scholar]
  • [3].Shumikina KI, Panarin EF, Ushakov SN, Experimental study of polymer salts of penicillins (in Russian), Antibiotiki 11 (1966) 767–770. [PubMed] [Google Scholar]
  • [4].Panarin EF, Ushakov SN, Synthesis of polymer salts and amidopenicillines (in Russian), Khim. Pharm. Zhur 2 (1968) 28–31. [Google Scholar]
  • [5].De Duve C, De Barsy T, Poole B, Trouet A, Tulkens P, van Hoof F, Lysosomotropic agents, Biochem. Pharmacol. 23 (1974) 2495–2531. [DOI] [PubMed] [Google Scholar]
  • [6].Ringsdorf H, Structure and properties of pharmacologically active polymers, J. Polym. Sci., Polym. Symp 51 (1975) 135–153. [Google Scholar]
  • [7].Matsumura Y, Maeda H, A new concept for macromolecular therapeutics in cancer therapy: mechanism of tumoritropic accumulation of proteins and the antitumor agent SMANCS, Cancer Res. 46 (1986) 6387–6392. [PubMed] [Google Scholar]
  • [8].Kopeček J, Soluble biomedical polymers, Polim. Med. 7 (1977) 191–221. [PubMed] [Google Scholar]
  • [9].Kopeček J, Biomaterials and drug delivery: Past, present, and future, Mol. Pharm. 7 (2010) 922–925. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [10].Yang J, Kopeček J, The light at the end of the tunnel – Second generation HPMA conjugates for cancer treatment, Curr. Opin. Colloid Interface Sci. 31 (2017) 30–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [11].Zhou Y, Kopeček J, Biological rationale for the design of polymeric anti-cancer nanomedicines, J. Drug Target. (2012) doi: 10.3109/1061186X.2012.723213. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [12].Ulbrich K, Holá K, Šubr V, Bakandritos A, Tuček J, Zbořil R, Targeted drug delivery with polymers and magnetic nanoparticles: Covalent and noncovalent approaches, release control, and clinical studies, Chem. Rev. 116 (2016) 5338–5431. [DOI] [PubMed] [Google Scholar]
  • [13].Duncan R, The dawning era of polymer therapeutics, Nat. Rev. Drug Discov 2 (2003) 347–360. [DOI] [PubMed] [Google Scholar]
  • [14].Říhová B, Kovář L, Kovář M, Hovorka O, Cytotoxicity and immunostimulation: double attack on cancer cells with polymer therapeutics, Trends Biotechnol. 27 (2009) 11–17. [DOI] [PubMed] [Google Scholar]
  • [15].van der Meel R, Sulheim E, Shi Y, Kiessling F, Mulder WJM, Lammers T, Smart cancer nanomedicines, Nat. Nanotechnol 14 (2019) 1007–1017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [16].Kopeček J, Kopečková P, HPMA copolymers: Origins, early developments, present, and future, Adv. Drug Deliv. Rev. 62 (2010) 122–149. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [17].Haag R, Kratz F, Polymer therapeutics: Concepts and applications, Angew. Chem. Int. Ed. 45 (2006) 1198–1215. [DOI] [PubMed] [Google Scholar]
  • [18].Yang J, Kopeček J, Design of Smart HPMA Copolymer-Based Nanomedicines, J. Control. Release 240 (2016) 9–23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [19].Duncan R, Vicent MJ, Polymer therapeutics – prospects for 21st century: the end of the beginning, Adv. Drug Deliv. Rev. 65 (2013) 60–70. [DOI] [PubMed] [Google Scholar]
  • [20].Chytil P, Koziolová E, Etrych T, Ulbrich K, HPMA copolymer-drug conjugates with controlled tumor-specific drug release, Macromol. Biosci 18 (2018) doi: 10.1002/mabi.201700209. [DOI] [PubMed] [Google Scholar]
  • [21].Kopeček J, Kopečková P, Design of polymer-drug conjugates In: Drug Delivery in Oncology, Kratz F, Senter P, Steinhagen H, Eds., Wiley-VCH, Weinheim, Germany, 2012, Vol. 2, Chapter 17, pp.485–512. [Google Scholar]
  • [22].Arranja AG, Pathak V, Lammers T, Shi Y, Tumor-targeted nanomedicines for cancer theranostics, Pharmacol. Res. 115 (2017) 87–95. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [23].Kopeček J, Polymer – drug conjugates: Origins, progress to date and future directions, Adv. Drug Deliv. Rev. 65 (2013) 49–59. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [24].Říhová B, Kovář M, Immunogenicity and immunomodulatory properties of HPMA-based polymers, Adv. Drug Deliv. Rev. 62 (2010) 184–191. [DOI] [PubMed] [Google Scholar]
  • [25].Ulbrich K, Šubr V, Structural and chemical aspects of HPMA copolymers as drug carriers, Adv. Drug Deliv. Rev. 62 (2010) 150–166. [DOI] [PubMed] [Google Scholar]
  • [26].Tucker BS, Sumerlin BS, Poly(N-(2-hydroxypropyl)methacrylamide)-based nanotherapeutics, Polym. Chem. 5 (2014) 1566–1572. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [27].Li J, Yu F, Chen Y, Oupický D, Polymeric drugs: Advances in the development of pharmacologically active polymers, J. Control. Release 219 (2015) 369–382. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [28].Yang J, Kopeček J, Polymeric Biomaterials and Nanomedicines, J. Drug Deliv. Sci. Technol 30 (2015) 318–330. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [29].Ljubimova JY, Sun T, Mashouf L, Ljubimov AV, Israel LL, Ljubimov VA, Falahatian V, Holler E, Covalent nano delivery systems for selective imaging and treatment of brain tumors, Adv. Drug Deliv. Rev. 113 (2017) 177–200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [30].Ekladious I, Colson YL, Grinstaff MW, Polymer-drug conjugate therapeutics: advances, insights and prospects, Nature Rev. Drug Disc. 18 (2019) 273–294. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [31].Kopeček J, Kopečková P, Minko T, Lu Z-R, HPMA copolymer – anticancer drug conjugates: Design, activity, and mechanism of action, Eur. J. Pharm. Biopharm 50 (2000) 61–81. [DOI] [PubMed] [Google Scholar]
  • [32].Conner SD, Schmid SL, Regulated portals of entry into the cell, Nature 422 (2002) 37–44. [DOI] [PubMed] [Google Scholar]
  • [33].Chiu H-C, Kopečková P, Deshmane SS, Kopeček J, Lysosomal degradation of poly(α-amino acids), J. Biomed. Mater. Res. 34 (1997) 381–392. [DOI] [PubMed] [Google Scholar]
  • [34].Chiu H-C, Koňák Č, Kopečková P, Kopeček J, Enzymatic degradation of poly(ethylene glycol) modified dextrans, J. Bioact. Comp. Polym. 9 (1994) 388–410. [Google Scholar]
  • [35].Lee B-S, Vert M, Holler E, “Water-soluble aliphatic polyesters: poly(malic acid)s”, in Biopolymers, Doi Y, Steinbüchel A, Eds., Wiley-VCH, Weinheim, 2002, pp. 75–103. [Google Scholar]
  • [36].Portilla-Arias J, Patil R, Hu J, Ding H, Black KL, Garcia-Alvarex M, Munoz-Guerra S, Ljubimova JY, Holler E, Nanoconjugate platforms development based in poly(β,L-malic acid) methyl ester for tumor drug delivery, J. Nanomat 2010, 825363; doi: 10.1155/2010/825363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [37].Kopeček J, Rejmanová P, Enzymatically degradable bonds in synthetic polymers, in: Controlled Drug Delivery, Vol. I, Bruck SD (Ed.), CRC Press, Boca Raton, Florida, 1983, pp. 81–124. [Google Scholar]
  • [38].Rejmanová P, Pohl J, Baudyš M, Kostka V, Kopeček J, Polymers containing enzymatically degradable bonds. 8. Degradation of oligopeptide sequences in N-(2-hydroxypropyl)methacrylamide copolymers by bovine spleen cathepsin B, Makromol. Chem 184 (1983) 2009–2020. [Google Scholar]
  • [39].Rejmanová P, Kopeček J, Duncan R, Lloyd JB, Stability in rat plasma and serum of lysosomally degradable oligopeptide sequences in N-(2-hydroxypropyl)methacrylamide copolymers, Biomaterials 6 (1985) 45–48. [DOI] [PubMed] [Google Scholar]
  • [40].Chytil P, Koziolová E, Etrych T, Ulbrich K, HPMA copolymer-drug conjugates with conntrolled tumor-specific release, Macromol. Biosci. 18(1) (2018): doi: 10.1002/mabi.201700209. [DOI] [PubMed] [Google Scholar]
  • [41].Ulbrich K, Šubr V, Polymeric anticancer drugs with pH-controlled activation, Adv. Drug Deliv. Rev. 56 (2004) 1023–1050. [DOI] [PubMed] [Google Scholar]
  • [42].Pan HZ, Kopečková P, Wang D, Yang J, Miller S, Kopeček J, Water-soluble HPMA copolymer – prostaglandin conjugates containing a cathepsin K sensitive spacer, J. Drug Target. 14 (2006) 425–435. [DOI] [PubMed] [Google Scholar]
  • [43].Shi W, Ogbomo SM, Wagh NK, Zhou Z, Jia Y, Brusnahan SK, Garrison JC, The influence of linker length on the properties of cathepsin S cleavable (177)Lu-labeled HPMA copolymers for pancreatic cancer imaging, Biomaterials 35 (2014) 5760–5770. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [44].Ogbomo SM, Shi W, Wagh NK, Zhou Z, Brusnahan SK, Garrison JC, 177Lu-labeled HPMA copolymers utilizing cathepsin B and S cleavable linkers: Synthesis, characterization and preliminary in vivo investigation in a pancreatic cancer model, Nucl. Med. Biol 40 (2013) 606–617. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [45].Fan W, Shi W, Zhang W, Jia Y, Zhou Z, Brusnahan SK, Garrison JC, Cathepsin S-cleavable, multi-block HPMA copolymers for improved SPECT/CT imaging of pancreatic cancer, Biomaterials 103 (2016) 101–115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [46].Santi DV, Schneider EL, Reid R, Robinson L, Ashley GW, Predictable and tunable half-life extension of therapeutic agents by controlled chemical release from macromolecular conjugates, Proc. Natl. Acad. Sci. U.S.A. 109 (2012) 6211–6216. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [47].Santi DV, Schneider EL, Ashley GW, Macromolecular prodrug that provides the irinotecan (CPT-11) active-metabolite SN-38 with ultralong half-life, low Cmax. and low glucuronide formation, J. Med. Chem. 57 (2014) 2303–2314. [DOI] [PubMed] [Google Scholar]
  • [48].Gao SQ, Lu ZR, Petri B, Kopečková P, Kopeček J, Colon-specific 9-aminocamptothecin-HPMA copolymer conjugates containing a 1,6-elimination spacer, J. Control. Release 110 (2006) 323–331. [DOI] [PubMed] [Google Scholar]
  • [49].Gao S, Sun Y, Kopečková P, Peterson CM, Kopeček J, Antitumor efficacy of colon-specific HPMA copolymer – 9-aminocamptothecin conjugate in mice bearing human colon carcinoma xenografts, Macromol. Biosci. 9 (2009) 1135–1142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [50].Daiichi Sankyo, Tokyo, Japan, ADC, DS-8201. [Google Scholar]
  • [51].Schechter I, Berger A, On the size of the active site in proteases. I. Papain, Biochem. Biophys. Res. Commun. 27 (1967) 157–162. [DOI] [PubMed] [Google Scholar]
  • [52].Ulbrich K, Strohalm J, Kopeček J, Poly(ethylene glycol)s containing enzymatically degradable bonds, Makromol. Chem. 187 (1986) 1131–1141 [Google Scholar]
  • [53].Peng Z-H, Sima M, Salama ME, Kopečková P, Kopeček J, Spacer length impacts the efficacy of targeted docetaxel conjugates in prostate specific membrane antigen expressing prostate cancer, J. Drug Target. 21 (2013) 968–980. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [54].Apostolovic B, Deacon SPE, Duncan R, Klok H-A, Hybrid polymer therapeutics incorporating bioresponsive, coiled coil peptide linkers, Biomacromolecules 11 (2010) 1187–1195. [DOI] [PubMed] [Google Scholar]
  • [55].Apostolovic B, Deacon SPE, Duncan R, Klok H-A, Cell uptake and trafficking behavior of non-covalent, coiled-coil based polymer-drug conjugates, Macromol. Rapid Commun. 32 (2011) 11–18. [DOI] [PubMed] [Google Scholar]
  • [56].Pola R, Laga R, Ulbrich K, Sieglová I, Král V, Fábry M, Kabešová M, Kovář M, Pechar M, Polymer therapeutics with a coiled coil motif targeted against murine BCL1 leukemia, Biomacromolecules 14 (2013) 881–889. [DOI] [PubMed] [Google Scholar]
  • [57].Pola R, Král V, Filippov SK, Kaberov L, Etrych T, Sieglová I, Sedláček J, Fábry M, Pechar M, Polymer cancerostatics targeted by recombinant antibody fragments to GD2-positive tumor cells, Biomacromolecules 20 (2019) 412–421. [DOI] [PubMed] [Google Scholar]
  • [58].Noguchi Y, Wu J, Duncan R, Strohalm J, Ulbrich K, Akaike T, Maeda H, Early phase tumor accumulation of macromolecules: A great difference in clearance rate between tumor and normal tissues, Jpn. J. Cancer Res. 89 (1998) 307–314. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [59].Shiah J-G, Dvořák M, Kopečková P, Sun Y, Peterson CM, Kopeček J, Biodistribution and antitumor efficacy of long-circulating N-(2-hydroxypropyl)methacrylamide copolymer-doxorubicin conjugates in nude mice, Eur. J. Cancer 37 (2001) 131–139. [DOI] [PubMed] [Google Scholar]
  • [60].Vasey PA, Kaye SB, Morrison R, Twelves C, Wilson P, Duncan R, Thomson AH, Murray LS, Hilditch TE, Murray T, Burtles S, Fraier D, Frigerio E, Cassidy J, and on behalf of the Cancer Research Campaign Phase I/II Committee, Phase I clinical and pharmacokinetic study of PK1 [N-(2-hydroxypropyl)methacrylamide copolymer doxorubicin]: first member of a new class of chemotherapeutic agents-drug-polymer conjugates, Clin. Cancer Res. 5 (1999) 83–94. [PubMed] [Google Scholar]
  • [61].Thomson AH, Vasey PA, Murray LS, Cassidy J, Fraier D, Frigerio E, Twelves C, Population pharmacokinetics in phase I drug development: a phase I study of PK1 in patients with solid tumors, Brit. J. Cancer 81 (1999) 99–107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [62].Seymour LW, Ferry DR, Kerr DJ, Rea D, Whitlock M, Ponyer R, Boivin C, Hesslewood S, Twelves C, Blackie R, Schätzlein A, Jodrell D, Bissett D, Calvert H, Lind M, Robbins A, Burtless S, Duncan R, Cassidy J, Phase II studies of polymer-doxorubicin (PK1, FCE28068) in the treatment of breast, lung and colorectal carcinoma, Int. J. Oncol. 34 (2009) 1629–1636. [DOI] [PubMed] [Google Scholar]
  • [63].Fox ME, Szoka FC, Fréchet JMJ, Soluble polymer carriers for the treatment of cancer: The importance of molecular architecture, Acc. Chem. Res. 42 (2009) 1141–1151. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [64].Koziolová E, Kostka L, Kotrchová L, Šubr V, Konefal R, Nottelet B, Etrych T, N-(2-hydroxypropyl)methacrylamide-based linear, diblock, and starlike polymer drug carriers: Advanced proces for their simple production, Biomacromolecules 19 (2018) 4003–4013. [DOI] [PubMed] [Google Scholar]
  • [65].Wang CE, Stayton PS, Pun SH, Convertine AJ, Polymer nanostructures synthesized by controlled living polymerization for tumor-targeted drug delivery, J. Control. Release 219 (2015) 345–354. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [66].Etrych T, Šubr V, Strohalm J, Šírová M, Říhová B, Ulbrich K, HPMA copolymer-doxorubicin conjugates: The effects of molecular weight and architecture on biodistribution and in vivo activity, J. Control. Release 164 (2012) 346–354. [DOI] [PubMed] [Google Scholar]
  • [67].Larson N, Ghandehari H, Polymeric conjugates for drug delivery, Chem. Mater. 24 (2012) 840–853. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [68].Wang D, Kopečková P, Minko T, Nanayakkara V, Kopeček J, Synthesis of star-like N-(2-hydroxypropyl)methacrylamide copolymers – potential drug carriers, Biomacromolecules 1 (2000) 313–319. [DOI] [PubMed] [Google Scholar]
  • [69].Nakamura H, Koziolová E, Etrych T, Chytil P, Feng J, Ulbrich K, Maeda H, Comparison between linear and star-like HPMA conjugated pirarubicin (THP) in pharmacokinetics and antitumor activity in tumor bearing mice, Eur. J. Pharm. Biopharm. 90 (2015) 90–96. [DOI] [PubMed] [Google Scholar]
  • [70].Šírová M, Strohalm J, Chytil P, Lidický O, Tomala J, Říhová B, Etrych T, The structure of polymer carriers controls the efficacy of the experimental combination treatment of tumors with HPMA copolymer conjugates carrying doxorubicin and docetaxel, J. Control. Release 246 (2017) 1–11. [DOI] [PubMed] [Google Scholar]
  • [71].Sadekar S, Ray A, Janát-Amsbury M, Peterson CM, Ghandehari H, Comparative biodistribution of PAMAM dendrimers and HPMA copolymers in ovarian-tumor-bearing mice, Biomacromolecules 12 (2011) 88–96. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [72].Mohammadpour R, Safarian S, Buckway B, Ghandehari H, Comparative endocytosis mechanisms and anticancer effect of HPMA copolymer – and PAMAM dendrimer-MTPC conjugates for photodynamic therapy, Macromol. Biosci. 17(4) (2016); doi: 10.1002/mabi.201600333. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [73].Kopeček J, Bažilová H, Poly[N-(2-hydroxypropyl)methacrylamide]. I. Radical polymerization and copolymerization, Eur. Polym. J. 9 (1973) 7–14. [Google Scholar]
  • [74].Bohdanecký M, Bažilová H, Kopeček J, Poly[N-(2-hydroxypropyl)methacrylamide]. II. Hydrodynamic properties of diluted polymer solutions, Eur. Polym. J. 10 (1974) 405–410. [Google Scholar]
  • [75].Šterba O, Paluska E, Jozová O, Špunda J, Nezvalová J, Šprincl L, Kopeček J, Činátl J, New types of synthetic infusion solutions. II. Basic biological properties of poly[N-(2-hydroxypropyl)methacrylamide] (in Czech), Časopis lék. českých 114, 1268–1270 (1975). [PubMed] [Google Scholar]
  • [76].Korčáková L, Paluska E, Hašková V, Kopeček J, A simple test for immunogenicity of colloidal infusion solutions - the draining lymph node activation, Z. Immun. Forsch. 151, 219–223 (1976). [PubMed] [Google Scholar]
  • [77].Šprincl L, Exner J, Šterba O, Kopeček J, New types of synthetic infusion solutions. III. Elimination and retention of poly[N-(2-hydroxypropyl)methacrylamide] in a test organism, J. Biomed. Mater. Res. 10, 953–963 (1976). [DOI] [PubMed] [Google Scholar]
  • [78].Paluska E, Činátl J, Korčáková L, Štěrba O, Kopeček J, Hrubá A, Nezvalová J, Staněk R, Immunosuppressive effect of a synthetic polymer - poly[N-(2-hydroxypropyl)methacrylamide] (Duxon), Folia Biologica 26 (1980) 304–311. [PubMed] [Google Scholar]
  • [79].Paluska E, Hrubá A, Štěrba O, Kopeček J, Effect of a synthetic poly[N-(2-hydroxypropyl)methacrylamide] (Duxon) on haemopoiesis and graft versus host reaction, Folia Biologica 32 (1986) 91–102. [PubMed] [Google Scholar]
  • [80].Šimečková J, Říhová B, Plocová D, Kopeček J, Activity of complement in the presence of N-(2-hydroxypropyl)methacrylamide copolymers, J. Bioact. Comp. Polym. 1 (1986) 20–31. [Google Scholar]
  • [81].Geng J, Li W, Zhang Y, Thottappillil N, Clavadetscher J, Lilienkampf A, Bradley M, Radical polymerization inside living cells, Nature Chem. 11 (2019) 578–586. [DOI] [PubMed] [Google Scholar]
  • [82].Duncan R, Seymour LW, O’Hare KB, Flanagan PA, Wedge S, Hume IC, Ulbrich K, Strohalm J, Šubr V, Sreafico F, Grandi M, Ripamonti M, Farao M, Suarato A, Preclinical evaluation of polymer-bound doxorubicin, J. Control. Release 19 (1992) 331–346. [Google Scholar]
  • [83].Seymour LW, Ferry DR, Anderson D, Hesslewood S, Julyan PJ, Poyner R, Doran J, Young AM, Burtles S, Kerr DJ, Hepatic drug targeting: phase I evaluation of polymer-bound doxorubicin, J. Clin. Oncol. 20 (2002) 1668–1676. [DOI] [PubMed] [Google Scholar]
  • [84].Shoemaker NE, van Kesteren C, Rosing H, Jansen S, Swart M, Lieverst J, Fraier D, Breda M, Pellizzoni C, Spinelli R, Grazia Porro M, Beijnen JH, Schellens JHM, ten Bokkel Huinink WW, A phase I and pharmacokinetic study of MAG-CPT, a water soluble polymer conjugate of camptothecin, Brit. J. Cancer 87 (2002) 608–614. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [85].Meerum Terwogt JM, ten Bokkel Huinink WW, Schellens JH, Schot M, Mandjes IA, Zurlo MG, Rocchetti M, Rosing H, Koopman FJ, Beijnen JH, Phase I clinical and pharmacokinetic study of PNU166945, a novel water-soluble polymer-conjugated prodrug of paclitaxel, Anticancer Drugs 12 (2001) 315–323. [DOI] [PubMed] [Google Scholar]
  • [86].Rademaker-Lakhai JM, Terret C, Howell SB, Baud CM, De Boer RF, Pluim D, Beijnen JH, Schellens JH, Droz JP, A phase I and pharmacological study of the platinum polymer AP5280 given as an intravenous infusion once every 3 weeks in patients with solid tumors, Clin. Cancer Res. 10 (2004) 3386–3395. [DOI] [PubMed] [Google Scholar]
  • [87].Duncan R, Hume IC, Kopečková P, Ulbrich K, Strohalm J, Kopeček J, Anticancer agents coupled to N-(2-hydroxypropyl)methacrylamide copolymers. 3. Evaluation of adriamycin conjugates against mouse leukemia L1210 in vivo, J. Control. Release 9 (1989) 21–32. [Google Scholar]
  • [88].Říhová B, Kopečková P, Strohalm J, Rossmann P, Větvička V, Kopeček J, Antibody directed affinity therapy applied to the immune system: In vivo effectiveness and limited toxicity of daunomycin conjugates to HPMA copolymers and targeting antibody, Clin. Immunol. Immunopathol. 46 (1988) 100–114. [DOI] [PubMed] [Google Scholar]
  • [89].Cassidy J, Duncan R, Morrison GJ, Strohalm J, Plocová D, Kopeček J, Kaye SB, Activity of N-(2-hydroxypropyl)methacrylamide copolymers containing daunomycin against a rat tumour model, Biochem. Pharmacol, 38 (1989) 875–879. [DOI] [PubMed] [Google Scholar]
  • [90].Peterson CM, Lu JM, Sun Y, Peterson CA, Shiah J-G, Straight RC, Kopeček J, Combination chemotherapy and photodynamic therapy with N-(2-hydroxypropyl)methacrylamide copolymer-bound anticancer drugs inhibit human ovarian carcinoma heterotransplanted in nude mice, Cancer Res. 56 (1996) 3980–3985. [PubMed] [Google Scholar]
  • [91].Minko T, Kopečková P, Kopeček J, Efficacy of chemotherapeutic action of HPMA copolymer-bound doxorubicin in a solid tumor model of ovarian carcinoma, Int. J. Cancer 86 (2000) 108–117. [DOI] [PubMed] [Google Scholar]
  • [92].Yang J, Luo K, Pan H, Kopečková P, Kopeček J, Synthesis of biodegradable multiblock copolymers by click coupling of RAFT-generated heterotelechelic polyHPMA conjugates, Reactive Functional Polym. 71 (2011) 294–302. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [93].Luo K, Yang J, Kopečková P, Kopeček J, Biodegradable multiblock N-(2-hydroxypropyl)methacrylamide copolymers via reversible addition-fragmentation chain transfer polymerization and click chemistry, Macromolecules 44 (2011) 2481–2488. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [94].Pan H, Yang J, Kopečková P, Kopeček J, Backbone degradable multiblock N-(2-hydroxypropyl)methacrylamide copolymer conjugates via reversible addition-fragmentation chain transfer polymerization and thiol-ene coupling reaction, Biomacromolecules 12 (2011) 247–252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [95].Yang J, Zhang R, Pan H, Li Y, Fang Y, Zhang L, Kopeček J, Backbone degradable HPMA copolymer conjugates with gemcitabine and paclitaxel: Impact of molecular weight on activity toward human ovarian carcinoma xenografts, Mol. Pharm. 14 (2017) 1384–1394. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [96].Ulbrich K, Koňák Č, Tuzar Z, Kopeček J, Solution properties of drug carriers based on poly[N-(2-hydroxypropyl)methacrylamide] containing biodegradable bonds, Makromol. Chem. 188 (1987) 1261–1272. [Google Scholar]
  • [97].Lammers T, Kühnlein R, Kissel M, Šubr V, Etrych T, Pola R, Pechar M, Ulbrich K, Storm G, Huber P, Peschke P, Effect of physicochemical modification on the biodistribution and tumor accumulation of HPMA copolymers, J. Control. Release 110 (2005) 103–118. [DOI] [PubMed] [Google Scholar]
  • [98].Koňák Č, Kopečková P, Kopeček J, Photoregulated association of N-(2-hydroxypropyl)methacrylamide copolymers with azobenzene-containing side-chains, Macromolecules 25 (1992) 5451–5456. [Google Scholar]
  • [99].Shiah J-G, Koňák Č, Spikes JD, Kopeček J, Solution and photoproperties of N-(2-hydroxypropyl)methacrylamide copolymer – meso-chlorin e6 conjugates, J. Phys. Chem. B 101 (1997) 6803–6809. [DOI] [PubMed] [Google Scholar]
  • [100].Shiah J-G, Koňák Č, Spikes JD, Kopeček J, Influence of pH on aggregation and photoproperties of N-(2-hydroxypropyl)methacrylamide copolymer – meso-chlorin e6 conjugates, Drug Delivery 5 (1998) 119–126. [DOI] [PubMed] [Google Scholar]
  • [101].Al-Abd AM, Aljehani ZK, Gazzaz RW, Fakhri SH, Jabbad AH, Alahdal AM, Torchilin VP, Pharmacokinetic strategies to improve drug penetration and entrapment within solid tumors, J. Control. Release 219 (2015) 269–277. [DOI] [PubMed] [Google Scholar]
  • [102].Yang Y, Pan D, Luo K, Li L, Gu Z, Biodegradable and amphiphilic block copolymer-doxorubicin conjugate as polymeric nanoscale drug delivery vehicle for breast cancer therapy, Biomaterials 34 (2013) 8430–8443. [DOI] [PubMed] [Google Scholar]
  • [103].Ou Y, Chen K, Cai H, Zhang H, Gong Q, Wang J, Chen W, Luo K, Enzyme/pH sensitive polyHPMA-DOX conjugate as a biocompatible and efficient anticancer agent, Biomat. Sci 6 (2018) 1177–1188. [DOI] [PubMed] [Google Scholar]
  • [104].Luo K, Xu F, Peng H, Luo Y, Tian X, Battaglia G, Zhang H, Gong Q, Gu Z, Luo K, Stimuli-responsive polymeric prodrug-based nanomedicine delivering nifuroxazide and doxorubicin against primary breast cancer and pulmonary metastasis, J. Control. Release 318 (2020) 124–135. [DOI] [PubMed] [Google Scholar]
  • [105].Fan W, Zhang W, Jia Y, Brushanan SK, Garrison JC, Investigation into the biological impact of block size on cathepsin S degradable HPMA copolymers, Mol. Pharm. 14 (2017) 1405–1417. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [106].Pechar M, Ulbrich K, Šubr V, Seymour LW, Schacht EH, Poly(ethylene glycol) multiblock copolymer as a carrier of anticancer drug doxorubicin, Bioconjug. Chem 11 (2000) 131–139. [DOI] [PubMed] [Google Scholar]
  • [107].Mužíková G, Pola R, Laga R, Pechar M, Biodegradable multiblock polymers based on N-(2-hydroxypropyl)methacrylamide designed as drug carrier for tumor-targeted drug delivery, Macromol. Chem. Phys. 217 (2017) 1690–1703. [Google Scholar]
  • [108].Yang J, Zhang D R, Radford C, Kopeček J, FRET-trackable biodegradable HPMA copolymer-epirubicin conjugates for ovarian carcinoma therapy, J. Control. Release 218 (2015) 36–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [109]. Nanotechnology Characterization Laboratory, Frederick National Laboratory for Cancer Research, November 2017. awardees.
  • [110].Pan H, Sima M, Yang J, Kopeček J, Synthesis of long-circulating backbone degradable HPMA copolymer-doxorubicin conjugates and evaluation of molecular weight dependent antitumor efficacy, Macromol. Biosci 13 (2013) 155–160. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [111].Zhang R, Luo K, Yang J, Sima M, Sun Y, Janát-Amsbury MM, Kopeček J, Synthesis and evaluation of a backbone biodegradable multiblock HPMA copolymer nanocarrier for the systemic delivery of paclitaxel, J. Control. Release 166 (2013) 66–74. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [112].Larson N, Yang J, Ray A, Cheney DL, Ghandehari H, Kopeček J, Biodegradable multiblock poly(N-2-hydroxypropyl)methacrylamide gemcitabine and paclitaxel conjugates for ovarian cancer cell combination treatment, Int. J. Pharm. 454 (2013) 435–443. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [113].Zhang R, Yang J, Sima M, Zhou Y, Kopeček J, Sequential combination therapy of ovarian cancer with degradable N-(2-hydroxypropyl)methacrylamide copolymer paclitaxel and gemcitabine conjugates, Proc. Natl. Acad. Sci. U.S.A. 111 (2014) 12181–12186. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [114].Pasut G, Veronese FM, State of the art in PEGylation: The great versatility achieved after forty years of research, J. Control. Release 161 (2012) 461–472. [DOI] [PubMed] [Google Scholar]
  • [115].Turecek PS, Bossard MJ, Schoetens F, Ivens IA, PEGylation of biopharmaceuticals: A review of chemistry and nonclinical safety information of approved drugs, J. Pharm. Sci. 105 (2016) 460–475. [DOI] [PubMed] [Google Scholar]
  • [116].Abuchowski A, van Es T, Palczuk NC, Davis FF, Alteration of immunological properties of bovine serum albumin by covalent attachment of polyethylene glycol, J. Biol. Chem. 252 (1977) 3578–3581. [PubMed] [Google Scholar]
  • [117].Abuchowski A, McCoy JR, Palczuk NC, van Es T, Davis FF, Effect of covalent attachment of polyethylene glycol on immunogenicity and circulating life of bovine liver catalase, J. Biol. Chem. 252 (1977) 3582–3586. [PubMed] [Google Scholar]
  • [118].Hershfield MS, Buckley RH, Greenberg ML, Melton AL, Schiff R, Hatem C, Kurtzberg J, Markert ML, Kobayashi RH, Kobayashi AL, Abuchowski A, Treatment of adenosine deaminase deficiency with polyethylene glycol-modified adenosine deaminase, N. Engl. J. Med. 316 (1987) 589–596. [DOI] [PubMed] [Google Scholar]
  • [119].Abuchowski A, Kazo GM, Verhoest CR Jr., van Es T, Kafkewitz D, Nucci ML, Viau AT, Davis FF, Cancer therapy with chemically modified enzymes. I. Antitumor properties of polyethylene glycol-asparaginase conjugates, Cancer Biochem. Biophys 7 (1984) 175–186. [PubMed] [Google Scholar]
  • [120].Ko JH, Maynard H, A guide to maximizing the therapeutic potential of protein-polymer conjugates by rational design, Chem. Soc. Rev. 47 (2018) 8998–9014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [121].Chiu H-C, Zalipsky S, Kopečkovå P, Kopeček J, Enzymatic activity of chymotrypsin and its poly(ethylene glycol) conjugates toward low and high molecular weight substrates, Bioconjug. Chem. 4 (1993) 290–295. [DOI] [PubMed] [Google Scholar]
  • [122].Hatakeyama H, Akita H, Harashima H, The polyethyleneglycol dilemma: Advantage and disadvantage of liposomes for systemic genes and nucleic acids delivery to tumors, Biol. Pharm. Bull. 36 (2013) 892–899. [DOI] [PubMed] [Google Scholar]
  • [123].Gong Y, Leroux J-C, Gauthier MA, Releasable conjugation of polymers to proteins, Bioconjug. Chem 26 (2015) 1172–1181. [DOI] [PubMed] [Google Scholar]
  • [124].Perez EA, Awada A, O’Shaughnessy J, Rugo HS, Twelves C, Im S-A, Goméz-Pardo P, Schwartzberg LS, Diéras V, Yardley DA, Potter DA, Mailliez A, Moreno-Aspitia A, Ahn J-S, Zhao C, Hoch U, Tagliaferri M, Hannah AL, Cortes J, Etirinotecan pegol (NKTR-102) versus treatment of physician’s choice in women with advanced breast cancer previously treated with an anthracycline, a taxane, and capecitabine (beacon): a randomized, open-label, multicenter, phase 3 trial, Lancet Oncol 16 (2015) 1556–1568. [DOI] [PubMed] [Google Scholar]
  • [125].Backford Vera DR, Fontaine SD, Van Brocklin HF, Heam BR, Reid RR, Ashley GW, Santi DV, PET imaging of the EPR effect in tumor xenografts usinf small 15 nm diameter polyethylene glycol labeled with zirconium-89, Mol. Cancer Ther. 19 (2020) 673–679. [DOI] [PubMed] [Google Scholar]
  • [126].Zhang P, Sun F, Liu S, Jiang S, Anti-PEG antibodies in the clinic: Current issues and beyond PEGylation, J. Control. Release 244(Pt. B) (2016) 184–193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [127].Abu Lila AS, Kiwada H, Ishida T, The accelerated blood clearance (ABC) phenomenon: Clinical challenge and approaches to manage, J. Control. Release 172 (2013) 38–47. [DOI] [PubMed] [Google Scholar]
  • [128].Ishida T, Ichihara M, Wang XY, 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, J. Control. Release 112 (2006) 15–25. [DOI] [PubMed] [Google Scholar]
  • [129].Ishihara T, Maeda T, Sakamoto H, Takasaki N, Shigyo M, Ishida T, Kiwada H, Mizushima Y, Mizushima T, Evasion of nanoparticle accelerated blood clearance phenomenon by coating of nanoparticles with various hydrophilic polymers, Biomacromolecules 11 (2010) 2700–2706 [DOI] [PubMed] [Google Scholar]
  • [130].Armstrong JK, Hempel G, Koling S, Chan LS, Fisher T, Meiselman HJ, Garatty G, Antibody against poly(ethylene glycol) adversely affects PEG-asparaginase therapy in acute lymphoblastic leukemia patients, Cancer 110 (2007) 103–111. [DOI] [PubMed] [Google Scholar]
  • [131].Bendele A, Seely J, Richey C, Sennelo G, Shopp G, Renal tubular vacuolation in animals treated with polyethylene-glycol-conjugated proteins, Toxicol. Sci. 42 (1998) 152–157. [DOI] [PubMed] [Google Scholar]
  • [132].Webster R, Didier E, Harris P, Siegel N, Stadler J, Tilbury L, Smith D, Pegylated proteins: evaluation of their safety in the absence of definitive metabolism studies, Drug Metab. Dispos. 35 (2007) 9–16. [DOI] [PubMed] [Google Scholar]
  • [133].McSweeney MD, Price LSL, Wessler T, Ciociola EC, Herity LB, Piscitelli JA, De Walle AC, Harris TN, Chan AKP, Saw RS, Hu P, Jennette JC, Forrest MG, Cao Y, Montgomery SA, Zamboni WC, Lai SK, Overcoming anti-PEG antibody mediated accelerated blood clearance of PEGylated liposomes by pre-infusion with high molecular weight free PEG, J. Control. Release 311-312 (2019) 138–146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [134].Knop K, Hoogenboom R, Fischer D, Schubert US, Poly(ethylene glycol) in drug delivery: pros and cons as well as potential alternatives, Angew. Chem. Int. Ed. 49 (2010) 6288–6308. [DOI] [PubMed] [Google Scholar]
  • [135].Lu Z-R, Kopečková P, Wu Z, Kopeček J, Functionalized semitelechelic poly[N-(2-hydroxypropyl)methacrylamide] for protein modification, Bioconjug. Chem. 9 (1998) 793–804. [DOI] [PubMed] [Google Scholar]
  • [136].Kamei S, Kopeček J, Prolonged blood circulation in rats of nanospheres surface-modified with semitelechelic poly[N-(2-hydroxypropyl)methacrylamide], Pharm. Res. 12 (1995) 663–668. [DOI] [PubMed] [Google Scholar]
  • [137].Klepáč D, Kostková H, Petrova S, Chytil P, Etrych T, Kereiche S, Raška I, Weitz DA, Filippov SK, Interaction of spin-labeled HPMA-based nanoparticles with human blood plasma proteins - The introduction of protein-corona-free polymer nanomedicine, Nanoscale, 10 (2018) 6194–6204. [DOI] [PubMed] [Google Scholar]
  • [138].Zhang X, Chytil P, Etrych T, Liu W, Rodriguez L, Winter G, Filippov SK, Papadakis CM, Binding of HSA to macromolecular pHPMA based nanoparticles for drug delivery: An investigation using fluorescence methods, Langmuir 34 (2018) 7998–8006. [DOI] [PubMed] [Google Scholar]
  • [139].Janisova L, Gruzinov A, Zaborova OV, Velychkivska N, Vaněk O, Chytil P, Etrych T, Janoušková O, Zhang X, Blanchet C, Papadakis CM, Svergun DI, Filippov SK, Molecular mechanism of the interactions of N-(2-hydroxypropyl)methacrylamide copolymers designed for cancer therapy with blood plasma proteins, Pharmaceutics 12 (2020) 106; doi: 10.3390/pharmaceutics12020106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [140].Kierstead PH, Okochi H, Venditto VJ, Chuong TC, Kivimae S, Fréchet JMJ, Szoka FC, The effect of polymer backbone chemistry on the induction of the accelerated blood clearance in polymer modified liposomes, J. Control. Release 213 (2015) 1–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [141].Zhao J, Qin Z, Wu J, Li L, Jin Q, Ji J, Zwitterionic stealth peptide-protected gold nanoparticles enable long circulation without the accelerated blood clearance phenomenon, Biomater. Sci. 6 (2017) 200–206. [DOI] [PubMed] [Google Scholar]
  • [142].Li Y, Liu R, Shi Y, Zhang Z, Zhang X, Zwitterionic poly(carboxybetaine)-based cationic liposomes for effective delivery od small interfering RNA therapeutics without accelerated blood clearance phenomenon, Theranostics 5 (2015) 583–596. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [143].Hou Y, Lu H, Protein PEPylation: A new paradigm of protein-polymer conjugation, Bioconjug. Chem 30 (2019) 1604–1616. [DOI] [PubMed] [Google Scholar]
  • [144].Lääne A, Aaviksaar A, Haga M, Chytrý V, Kopeček J, Preparation of polymer-modified enzymes of prolonged circulation times. Poly[N-(2-hydroxypropyl)methacrylamide] bound acetylcholinesterase, Makromol. Chem Suppl. 9 (1985) 35–42. [Google Scholar]
  • [145].Oupický D, Ogris M, Howard KA, Dash PR, Ulbrich K, Seymour LW, Importance of lateral and steric stabilization of polyelectrolyte gene delivery vectors for extended systemic circulation, Mol. Ther. 5 (2002) 463–472. [DOI] [PubMed] [Google Scholar]
  • [146].Fisher KD, Seymour LW, HPMA copolymers for masking and retargeting therapeutic viruses, Adv. Drug Deliv. Rev. 62 (2010) 240–245. [DOI] [PubMed] [Google Scholar]
  • [147].Kostka L, Šubr V, Laga R, Chytil P, Ulbrich K, Seymour LW, Etrych T, Nanotherapeutics shielded with a pH responsive polymeric layer, Physiol. Res 64(Suppl. 1) (2015) S29–S40. [DOI] [PubMed] [Google Scholar]
  • [148].Abdellaoui K, Boustta M, Vert M, Morjani H, Manfrat M, Metabolite-derived artificial polymers designed for drug targeting, cell penetration, and bioresorption, Eur. J. Pharm. Sci. 6 (1998) 61–73. [DOI] [PubMed] [Google Scholar]
  • [149].Lee BS, Fujita M, Khazenzon NM, Wawrowsky KA, Wachsmann-Hogiu S, Farkas DL, Black KL, Ljubimova JY, Holler E, Polycefin, a new prototype of a multifunctional nanoconjugate based on poly(β,L-malic acid) for drug delivery, Bioconjug. Chem. 17 (2006) 317–326. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [150].Ljubimova JY, Fujita M, Ljubimov AV, Torchilin VP, Black KL, Holler E, Poly(malic acid) nanoconjugates containing various antibodies and oligonucleotides for multitargeting drug delivery, Nanomedicine (London) 3 (2008) 247–265. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [151].Ding H, Portilla-Arias J, Patil R, Black KL, Ljubimova JY, Holler E, The optimization of polymalic acid peptide copolymers for endosomolytic drug delivery, Biomaterials 32 (2011) 5269–5278. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [152].Ding H, Fox I, Patil R, Galstyan A, Black KL, Ljubimova JY, Holler E, Polymalic acid tritryptophan copolymer interacts with lipid membrane resulting in membrane solubilization, J. Nanomater. 2017:4238697, doi: 10.1155/2017/4238697. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [153].Ding H, Patil R, Portilla-Arias J, Black KL, Ljubimova JY, Holler E, Quantitative analysis of PMLA nanoconjugate components after backbone cleavage, Int. J. Mol. Sci. 16 (2015) 8607–8620. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [154].Ding H, Gangalum PR, Galstyan A, Fox I, Patil R, Hubbard P, Murali R, Ljubimova JY, Holler E, HER2-positive breast cancer targeting and treatment by a peptide-conjugated mini nanodrug, Nanomedicine 13 (2017) 631–639. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [155].Ljubimova JY, Portilla-Arias J, Patil R, Ding H, Inoue S, Markman JL, Rekechenetskiy A, Konda B, Gangalum PR, Chesnokova A, Ljubimov AV, Black KL, Holler E. Toxicity and efficacy evaluation of multiple targeted polymalic acid conjugates for triple-negative breast cancer treatment, J. Drug Target. 21 (2013) 956–967. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [156].Demeule M, Regina A, Che C, Poirier J, Nguyen T, Gabathuler R, Identification and design of new peptides as a drug delivery system for the brain, J. Pharmacol. Exp. Ther. 324 (2008) 1064–1072. [DOI] [PubMed] [Google Scholar]
  • [157].Israel LL, Braubach O, Galstyan A, Chiechi A, Shatalova ES, Grodzinski Z, Ding H, Black KL, Ljubimova JY, Holler E, A combination of tri-leucine and angiopep-2 drives a polyanionic polymalic acid nanodrug platform across the blood brain barrier, ACS Nano 13 (2019) 1253–1271. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [158].Galstyan A, Markman JL, Shatalova ES, Chiechi A, Korman AJ, Patil R, Klymyshyn D, Tourtellotte WG, Israel LL, Braubach O, Ljubimov VA, Mashouf LA, Ramesh A, Grodzinski ZB, Perichet ML, Black KL, Holler E, Sun T, Ding H, Ljubimov AV, Ljubimova JY, Blood-brain barrier permeable nano immunoconjugates induce local immune responses for glioma therapy, Nature Commun. 2019, 10:3850; doi: 10.1038/s41467-019-11719-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [159].Li C, Wallace S, Polymer-drug conjugates: Recent development in clinical oncology, Adv. Drug Deliv. Rev. 60 (2008) 886–898. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [160].Shaffer S, Baker-Lee C, Kennedy J, Lai M, de Vries P, Buhler K, Singer J, In vitro and in vivo metabolism of paclitaxel poliglumex: Identification of metabolites and active proteases, Cancer Chemother. Pharmacol. 59 (2007) 537–548. [DOI] [PubMed] [Google Scholar]
  • [161].Melancon MP, Wang W, Wang Y, Shao R, Ji X, Gelovani JG, Li C, A novel method for imaging in vivo degradation of poly(L-glutamic acid), a biodegradable drug carrier, Pharm. Res. 24 (2007) 1217–1224. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [162].Singer JW, Shaffer S, Baker R, Bernareggi A, Stromatt S, Nienstedt D, Besman M, Paclitaxel poliglumex (XYPTAC; CT-2103): an intracellularly targeted taxane, Anticancer Drugs 16 (2005) 243–254. [DOI] [PubMed] [Google Scholar]
  • [163].Ross H, Bonomi P, Langer C, O’Brien M, O’Byrne K, Paz-Aares L, Sandler A, Socinski M, Oldham F, Singer J, Effect of gender on outcome in two randomized phase III trials of paclitaxel poliglumex (PPX) in chemonaive patients with advanced NSCLC and poor performance status (PS2), J. Clin. Oncol 24 (2006) Abstract #7039. [Google Scholar]
  • [164].Albain KS, Belani CP, Bonomi P, O’Byrne KJ, Schiller JH, Socinski M, PIONEER: A phase III randomized trial of paclitaxel poliglumex versus paclitaxel in chemotherapy naive women with advanced non-small-cell lung cancer and performance status of 2, Clin. Lung Cancer 7 (2006) 417–419. [DOI] [PubMed] [Google Scholar]
  • [165].Melancom MP, Li C, Multifunctional synthetic poly(L-glutamic acid) – based cancer therapeutics and imaging agents, Mol. Imag 10 (2011) 28–42. [PMC free article] [PubMed] [Google Scholar]
  • [166].Kiew LV, Cheah HY, Voon SH, Gallon E, Movelan J, Ng KH, Alpugan S, Lee HB, Dumoulin F, Vicent MJ, Chung LY, Near-infrared activatable phthalocyanine-poly-L-glutamic acid conjugate: increased cellular uptake and light-dark toxicity ration toward an effective photodynamic cancer therapy, Nanomedicine 13 (2017) 1447–1458. [DOI] [PubMed] [Google Scholar]
  • [167].Conejos-Sánchez I, Cardoso I, Oteo-Vives M, Romero-Sanz E, Paul A, Sauri AR, Morcillo MA, Saraiva MJ, Vicent MJ, Polymer-doxycycline conjugates as fibril disrupters: an approach toward the treatment of a rare amyloidotic disease, J. Control. Release 198 (2015) 80–90. [DOI] [PubMed] [Google Scholar]
  • [168].Duro-Castano A, England RM, Razola D, Romero E, Oteo-Vives M, Morcillo MA, Vicent MJ, Well-defined star-shaped polyglutamates with improved pharmacokinetic profiles as excellent candidates for biomedical applications, Mol. Pharm. 12 (2015) 3639–3649. [DOI] [PubMed] [Google Scholar]
  • [169].Córdoba-David G, Duro-Castano A, Castelo-Branco RC, Gonzáles-Guerrero C, Cannata P, Sanz AB, Vicent MJ, Ortiz A, Ramos AM, Effective nephroprotection against acute kidney injury with a star-shaped polyglutamate-curcuminoid conjugate, Sci. Rep 10 (2020) 2056; doi: 10.1038/s41598-020-58974-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [170].Davis ME, Design and development of IT-101, a cyclodextrin-containing polymer conjugate of camptothecin, Adv. Drug Deliv. Rev. 61 (2009) 1189–1192. [DOI] [PubMed] [Google Scholar]
  • [171].Davis ME, Pun SH, Bellocq NC, Reinecke TM, Popielarski SR, Mishra S, Heidel JD, Self-assembling nucleic acid delivery vehicles via linear, water-soluble, cyclodextrin-containing polymers, Curr. Med. Chem. 11 (2004) 179–197. [DOI] [PubMed] [Google Scholar]
  • [172].Cheng J, Khin KT, Jensen GS, Liu A, Davis ME, Synthesis of linear, β-cyclodextrin-based polymers and their camptothecin conjugates, Bioconjug. Chem 14 (2003) 1007–1017. [DOI] [PubMed] [Google Scholar]
  • [173].Schluep T, Hwang J, Cheng J, Heidel JD, Bartlett DW, Hollister B, Davis ME, Preclinical efficacy of the camptothecin-polymer conjugate IT-101 in multiple cancer models, Clin. Cancer Res. 12 (2006) 1606–1614. [DOI] [PubMed] [Google Scholar]
  • [174].Schluep T, Cheng J, Khin KT, Davis ME, Pharmacokinetics and biodistribution of the camptothecin-polymer conjugate IT-101 in rats and tumor bearing mice, Cancer Chemother. Pharmacol. 57 (2006) 654–662. [DOI] [PubMed] [Google Scholar]
  • [175].Weiss GJ, Chao J, Neidhart JD, Ramanathan RK, Bassett D, Neidhart JA, Choi CHJ, Chow W, Chung V, Forman SJ, Garmey E, Hwang J, Kalinoski DL, Koczywas M, Longmate J, Melton RJ, Morgan R, Oliver J, Peterkin JJ, Ryan JL, Schluep T, Synold TW, Twardowski P, Davis ME, Yen Y, First-in-human phase 1/2a trial of CRLX101, a cyclodextrin containing polymer-camptothecin nanopharmacutical in patients with advanced solid tumor malignancies, Invest. New Drugs 31 (2013) 986–1000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [176].Tian B, Hua S, Liu J, Cyclodextrin-based delivery system for chemotherapeutic anticancer drugs: A review, Carbohydrate Polym 232 (2020) 115805. [DOI] [PubMed] [Google Scholar]
  • [177].Clark AJ, Wiley DT, Zuckerman JE, Webster P, Chao J, Lin J, Yen Y, Davis ME, CRLX101 nanoparticles localize in human tumors and not in adjacent, nonneoplastic tissue after intravenous dosing, Proc. Natl. Acad. Sci. U.S.A. 113 (2016) 3850–3854. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [178].Chen Y-F, Wang Y-H, Lee C-S, Changou CA, Davis ME, Yen Y, Host immune response to anti-cancer camptothecin conjugated cyclodextrin-based polymers, J. Biomed. Sci. 26 (2019) 85; doi:10/1186/s12929-019-0583-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [179].Eliasof S, Lazarus D, Peters CG, Chase RI, Cole RO, Hwang J, Schluep T, Chao J, Li J, Yen Y, Han H, Wiley DT, Zuckerman JE, Davis ME, Correlating preclinical animal studies and human clinical trials of a multifunctional, polymeric nanoparticle, Proc. Natl. Acad. Sci. U.S.A. 110 (2013) 15127–15132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [180].Monnery BD, Jerca VV, Sedláček O, Verbraeken B, Cavill R, Hoogenboom R, Defined high molar mass poly(2-oxazoline)s, Angew. Chem. Int. Ed. 57 (2018) 15400–15404. [DOI] [PubMed] [Google Scholar]
  • [181].Sedláček O, Hoogenboom R, Drug delivery systems based on poly(2-oxazoline)s and poly(2-oxazine)s, Adv. Therap 3 (2020) 1900168. [Google Scholar]
  • [182].Viegas TX, Bentley MD, Harris JM, Fang Z, Yoon K, Dizman B, Weimer R, Mero A, Pasut G, Veronese FM, Polyoxazoline: chemistry, properties, and applications in drug delivery, Bioconjug. Chem 22 (2011) 976–986. [DOI] [PubMed] [Google Scholar]
  • [183].Sedláček O, Moonery BD, Mattová J, Kučka J, Pánek O, Janoušková O, Hocherl A, Verbraeken B, Vergaelen M, Zadinová M, Poly(2-ethyl-2-oxazoline) conjugates with doxorubicin for cancer therapy: In vitro and in vivo evaluation and direct comparison to poly[N-(2-hydroxypropyl)methacrylamide] analogues, Biomaterials 146 (2017) 1–12. [DOI] [PubMed] [Google Scholar]
  • [184].Moreadith RW, Viegas TX, Bentley MD, Harris JM, Fang Z, Yoon K, Dizman B, Weimer R, Rae BP, Ki X, Rader C, Standert D, Olanow W, Clinical development of a poly(2-oxazoline) (POZ) polymer therapeutics for the treatment of Parkinson’s disease – Proof of concept of POZ as a versatile polymer platform for drug development in multiple therapeutic indications, Eur. Polym. J. 88 (2017) 524–552. [Google Scholar]
  • [185].Harris JM, Bentley MD, Moreadith RW, Viegas TX, Fang Z, Yoon K, Weimer R, Dizman B, Nordstierna L, Tuning drug release from polyoxazoline-drug conjugates, Eur. Polym. J. 120 (2019) 109241. [Google Scholar]
  • [186].Park J-R, van Guyse JFR, Podevyn A, Bolle ECL, Bock N, Linde E, Celina M, Hoogenboom R, Dargaville TR, Influence of side-chain length on long-term release kinetics from poly(2-oxazoline)-drug conjugates, Eur. Polym. J. 120 (2019) 109217. [Google Scholar]
  • [187].Shliakhtsin SV, Trukhachova TV, Isakau HA, Istomin YP, Pharmacokinetics and biodistribution of Photolon (Fotolon) in intact and tumor-bearing rats, Photodiagnosis Photodyn. Ther. 6 (2009) 97–104. [DOI] [PubMed] [Google Scholar]
  • [188].Huang G, Huang H, Application of dextran as nanoscale drug carrier, Nanomedicine (Lond.) 13 (2018) 3149–3158. [DOI] [PubMed] [Google Scholar]
  • [189].Huang G, Huang H, Hyaluronic acid-based biopharmaceutical delivery and tumor-targeted drug delivery system, J. Control. Release 278 (2018) 122–126. [DOI] [PubMed] [Google Scholar]
  • [190].Serrano-Sevilla I, Artiga A, Mitchell SG, De Matteis L, de la Funte JM, Natural polysaccharides for siRNA delivery: Nanocarriers based on chitosan, hyaluronic acid, and their derivatives, Molecules 24 (2019); doi: 10.3390/molecules24142570. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [191].Choi Y, Lim S, Yoon HY, Kim BS, Kwon IC, Kim K, Tunor-targeting glycol chitosan nanocarriers: overcoming the challenges posed by chemotherapeutics, Expert Opin. Drug Deliv. 16 (2019) 835–846. [DOI] [PubMed] [Google Scholar]
  • [192].Secker C, Brosnan SM, Luxenhofer R, Schlaad H, Poly(α-peptoid)s revisited: Synthesis, properties, and use as biomaterial, Macromol. Biosci. 15 (2015) 881–891. [DOI] [PubMed] [Google Scholar]
  • [193].Schroeder T, Niemeier N, Afonin S, Ulrich AS, Krug HF, Brase S, Peptoid amino- and guanidinium-carrier systems: Targeted drug delivery into the cell cytosol of the nucleus, J. Med. Chem. 51 (2008) 376–379. [DOI] [PubMed] [Google Scholar]
  • [194].Yurkovetskiy AV, Fram RJ, XMT-1001, a novel polymeric camptothecin pro-drug in clinical development for patients with advanced cancer, Adv. Drug Deliv. Rev. 61 (2009) 1193–1202. [DOI] [PubMed] [Google Scholar]
  • [195].Yurkovetskiy AV, Fram RJ, XMT-1001, a novel biodegradable polyacetal polymer conjugate of camptothecin in clinical development, Curr. Bioactive Comp 7 (2011); doi: 10.2174/157340711795163839. [DOI] [Google Scholar]
  • [196].Tolcher AW, Ulahannan SV, Papadopoulos KP, Edenfield WJ, Matulonis UA, Burns TF, Mosher R, Huebner D, Hailman E, Burris III HA, Moore KN, Hamilton EP, Phase 1 dose escalation study of XMT-1536, a novel NaPi2b-targeting antibody-drug conjugate (ADC), in patients with solid tumors likely to express NaPi2b, American Society for Clinical Oncology Meeting, June 2019, Abstract #255157; poster # 3010. [Google Scholar]
  • [197].Omelyanenko V, Gentry C, Kopečková P, Kopeček J, HPMA copolymer - anticancer drug - OV-TL16 antibody conjugates. 2. Processing in epithelial ovarian carcinoma cells in vitro, Int. J. Cancer 75 (1998) 600–608. [DOI] [PubMed] [Google Scholar]
  • [198].Liu J, Bauer H, Callahan J, Kopečková P, Pan H, Kopeček J, Endocytic uptake of a large array of HPMA copolymers: Elucidation into the dependence on the physicochemical characteristics, J. Control. Release 143 (2010) 71–79. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [199].Putnam D, Shiah J-G, Kopeček J, Intracellularly Biorecognizable derivatives of 5-fluorouracil: Implications of targetable delivery in the human condition, Biochem. Pharmacol. 52 (1996) 957–962. [DOI] [PubMed] [Google Scholar]
  • [200].Kasuya Y, Lu Z-R, Kopečková P, Tabibi SE, Kopeček J, Influence of the structure of drug moieties on the in vitro efficacy of HPMA copolymer-geldanamycin derivative conjugate, Pharm. Res. 19 (2002) 115–123. [DOI] [PubMed] [Google Scholar]
  • [201].Callahan J, Kopečková P, Kopeček J, Intracellular trafficking and subcellular distribution of a large array of HPMA copolymers, Biomacromolecules 10 (2009) 1704–1714. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [202].Armiñan A, Palomino-Schätzlein M, Deladriere C, Arroyo-Crespo JJ, Vicente-Ruiz S, Vicent MJ, Pineda-Lucena A, Metabolomics facilitates the discrimination of the specific anti-cancer effects of free- and polymer-conjugated doxorubicin in breast cancer models, Biomaterials 162 (2018) 144–153. [DOI] [PubMed] [Google Scholar]
  • [203].Wang Y, Xie Y, Oupický D, Potential of CXCR4/CXCL12 chemokine axis in cancer drug delivery, Curr. Pharmacol. Rep 2 (2016) 1–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [204].Nishiyama N, Nori A, Malugin A, Kasuya Y, Kopečková P, Kopeček J, Free and N-(2-hydroxypropyl)methacrylamide (HPMA) copolymer-bound geldanamycin derivative induce different stress responses in A2780 human ovarian carcinoma cells, Cancer Res. 63 (2003) 7876–7882. [PubMed] [Google Scholar]
  • [205].Radford DC, Yang J, Doan M, Li L, Dixon AS, Owen SC, Kopeček J, Multivalent HER2-binding polymer conjugates facilitate rapid endocytosis and enhance intracellular drug delivery, J. Control. Release 319 (2020) 285–299. [DOI] [PubMed] [Google Scholar]
  • [206].Hennig R, Veser A, Kirchof S, Goepferich A, Branched polymer-drug conjugates for multivalent blockade of angiotensin II receptors, Mol. Pharm. 12 (2015) 3292–3302. [DOI] [PubMed] [Google Scholar]
  • [207].Yu F, Li J, Xie Y, Sleightholm RL, Oupický D, Polymeric chloroquine as an inhibitor of cancer cell migration and experimental lung metastasis, J. Control. Release 244(pt.B) (2016) 347–356 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [208].Miller SC, Pan H, Wang D, Bowman BM, Kopečková P, Kopeček J, Feasibility of using a bone-targeted, macromolecular delivery system coupled with prostaglandin E1 to promote bone formation in aged, estrogen-deficient rats, Pharm. Res. 25 (2008) 2889–2895. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [209].Nicoletti S, Seifert K, Gilbert IH, N-(2-Hydroxypropyl)methacrylamide-amphotericin B (HPMA-AmB) copolymer conjugates as antileishmanial agents, Int. J. Antimicrob. Agents 33 (2009) 441–448. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [210].Ward SM, Skinner M, Saha B, Emrick T, Polymer-temozolomide conjugates as therapeutics for treating glioblastoma, Mol. Pharm 15 (2018) 5263–5276. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [211].Zaiden M, Feinshtein V, David A, Inhibition of CD44v3 and CD44v6 function blocks tumor invasion and metastatic colonization, J. Control. Release 257 (2017) 10–20. [DOI] [PubMed] [Google Scholar]
  • [212].Peng Z-H, Kopeček J, HPMA copolymer CXCR4 Antagonist conjugates substantially inhibited the migration of prostate cancer cells, ACS Macro Letters 3 (2014) 1240–1243. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [213].Nel A, Ruoslahti E, Meng H, New insight into “permeability” as in the enhanced permeability and retention effect of cancer nanotherapeutics, ACS Nano 11 (2017) 9567–9569. [DOI] [PubMed] [Google Scholar]
  • [214].Harrington KJ, Mohammadtaghi S, Uster PS, Glass D, Peters AM, Vile RG, Stewart JSW, Effective targeting of solid tumors in patients with locally advanced cancers by radiolabeled pegylated liposomes, Clin. Cancer Res. 7 (2001) 243–254. [PubMed] [Google Scholar]
  • [215].Ramanathan RK, Korn RL, Raghunand N, Sachdev JC, Newbold RG, Jameson G, Fetterly GJ, Prey J, Klinz SG, Kim J, Cain J, Hendricks BS, Drummond DC, Bayever E, Fitzgerald JB, Correlation between ferumoxytol uptake in tumor lesions by MRI and response to nanoliposomal irinotecan in patients with advanced solid tumors: A pilot study, Clin. Cancer Res. 23 (2017) 3638–3648. [DOI] [PubMed] [Google Scholar]
  • [216].Lee H, Shields AF, Siegel BA, Miller KD, Krop I, Ma CX, LoRusso PM, Munster PM, Campbell K, Gaddy DF, Leonard SC, Geretti E, Blocker SJ, Kirpotin DB, Moyo V, Wickham TJ, Hendricks BS, 64Cu-MM-302 positron emission tomography quantifies variability of enhanced permeability of nanoparticles in relation to treatment response in patients with metastatic breast cancer, Clin. Cancer Res. 23 (2017) 4190–4202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [217].Hare JL, Lammers T, Ashford MB, Puri S, Storm G, Barry ST, Challenges and strategies in anti-cancer nanomedicine development: An industry perspective, Adv. Drug Deliv. Rev. 108 (2017) 25–38. [DOI] [PubMed] [Google Scholar]
  • [218].Lammers T, Macro-nanomedicine: Targeting the big picture, J. Control. Release 294 (2019) 372–375. [DOI] [PubMed] [Google Scholar]
  • [219].Park K, The drug delivery field at the inflection point: Time to fight its way out of the egg, J. Control. Release 267 (2017) 2–14. [DOI] [PubMed] [Google Scholar]
  • [220].Danhier F, To exploit the tumor microenvironment: Since the EPR effect fails in the clinic, what is the future of nanomedicine? J. Control. Release 244 (2016) 108–121. [DOI] [PubMed] [Google Scholar]
  • [221].Chauhan VP, Stylianopoulos T, Martin JD, Popović Z, Chen O, Kamoun WS, Bawendi MG, Fukumura D, Jain RK, Normalization of tumor blood vessels improves the delivery of nanomedicines in a size-dependent manner, Nature Nanotechnol. 7 (2012) 383–388. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [222].Golombek SK, May J-N, Theek B, Appold L, Drude N, Kiessling F, Lammers T, Tumor targeting via EPR: Strategies to enhance patient responses, Adv. Drug Deliv. Rev. 130 (2018) 17–38. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [223].Studenovský M, Sivák L, Sedláček O, Konefal R, Horková V, Etrych T, Kovář M, Říhová B, Šírová M, Polymeric nitric oxide donors potentiate the treatment of experimental solid tumors by increasing drug accumulation in the tumor tissue, J. Control. Release 269 (2018) 214–224. [DOI] [PubMed] [Google Scholar]
  • [224].Fang J, Long L, Maeda H, Enhancement of tumor-targeted delivery of bacteria with nitroglycerin involving augmentation of the EPR effect, Methods Mol. Biol. 1409 (2016) 9–23. [DOI] [PubMed] [Google Scholar]
  • [225].Nakamura H, Jun F, Maeda H, Development of next-generation macromolecular drugs based on the EPR effect: challenges and pitfalls, Expert Opin. Drug Deliv 12 (2015) 53–64. [DOI] [PubMed] [Google Scholar]
  • [226].Fang J, Islam R, Islam W, Yin H, Šubr V, Etrych T, Ulbrich K, Maeda H, Augmentation of EPR effect and efficacy of anticancer nanomedicine by carbon monoxide generating agents, Pharmaceutics 11 (2019) 343; doi: 10.3390/pharmaceutics11070343. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [227].Islam W, Fang J, Imamura T, Etrych T, Šubr V, Ulbrich K, Maeda H, Augmentation of the enhanced permeability and retention effect with nitricoxide-generating agents improves the therapeutic effects of nanomedicines, Mol. Cancer Ther. 17 (2018) 2643–2653. [DOI] [PubMed] [Google Scholar]
  • [228].Allen C, Why I’m holding onto hope for nano in oncology, Mol. Pharm. 13 (2016) 2603–2604. [DOI] [PubMed] [Google Scholar]
  • [229].Miller MA, Arlauckas S, Weissleder R, Prediction of anti-cancer nanotherapy efficacy by imaging, Nanotheranostics 1 (2017) 296–312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [230].van der Meel R, Lammers T, Hennink WE, Cancer nanomedicines: oversold or underappreciated? Expert Opin. Drug Deliv 14 (2017) 1–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [231].Wang Z, Xue X, Lu H, He Y, Lu Z, Chen Z, Yuan Y, Tang N, Dreyer CA, Quigley L, Curro N, Lam KS, Walton JH, Lin T-Y, Louie AY, Gilbert DA, Liu K, Ferrara KW, Li Y, Two-way magnetic resonance tuning and enhanced subtraction imaging for non-invasive and quantitative biological imaging, Nature Nanotechnol. (2020) doi: 10.1038/s41565-020-0678-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [232].Lu ZR, Molecular imaging of HPMA copolymers: visualizing drug delivery in cell, mouse and man, Adv. Drug Deliv. Rev. 62 (2010) 246–257. [DOI] [PubMed] [Google Scholar]
  • [233].Buckway B, Wang Y, Ray A, Ghandehari H, In vitro evaluation of HPMA-copolymers targeted to HER2 expressing pancreatic tumor cells for image guided drug delivery, Macromol. Biosci. 14 (2014) 92–99. [DOI] [PubMed] [Google Scholar]
  • [234].Lee H, Gaddy D, Ventura M, Barnards N, de Souza R, Kirpotin D, Wickham T, Fitzgerald J, Zheng J, Hendricks BS, Companion diagnostics 64Cu-liposome positron emission tomography enables characterization of drug delivery to tumors and predicts response to cancer nanomedicines, Theranostics 8 (2018) 2300–2312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [235].Koziolová E, Goel S, Chytil P, Janoušková O, Barnhart TE, Cai W, Etrych T, A tumor-targeted polymer theranostic platform for positron emission tomography and fluorescence imaging, Nanoscale 9 (2017) 10906–10918. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [236].Whatcott C, Han H, Posner RG, Von Hoff DD, Tumor-stromal interactions in pancreatic cancer, Crit. Rev. Oncog 18 (2013) 135–151. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [237].Provenzano PP, Cueves C, Chang AE, Goel VK, Von Hoff DD, Hingorani SR, Enzymatic targeting of the stroma ablates physical barriers to treatment of pancreatic ductal adenocarcinoma, Cancer Cell 21 (2012) 418–421. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [238].Kang Y, Dissecting tumor-stromal interactions in breast cancer bone metastasis, Endocrinol. Metab 31 (2016) 206–212. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [239].Stylianopoulos T, Martin JD, Chauham VP, Jain SR, Diop-Frimpong B, Bardeesy N, Smith BL, Ferrone CR, Hornicek FJ, Boucher Y, Munn LL, Jain RK, Causes, consequences, and remedies for growth-induced solid stress in murine and human tumors, Proc. Natl. Acad. Sci. U.S.A. 109 (2012) 15101–15108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [240].Peng Z-H, Kopeček J, Enhancing accumulation and penetration of HPMA copolymer doxorubicin conjugates in 2D and 3D prostate cancer cells via iRGD conjugation with an MMP-2 cleavable spacer, J. Am. Chem. Soc. 137 (2015) 6726–6729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [241].Corti A, Pastorino F, Curnis F, Arap W, Ponzoni M, Pasqualini R, Trageted drug delivery and penetration into solid tumor, Med. Res. Rev. 32 (2012) 1078–1091. [DOI] [PubMed] [Google Scholar]
  • [242].Seymour LW, Ulbrich K, Wedge SR, Hume IC, Strohalm J, Duncan R, N-(2-hydroxypropyl)methacrylamide copolymers targeted to the hepatocyte galactose-receptor: Pharmacokinetics in BDA2 mice, Br. J. Cancer 63 (1991) 859–866. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [243].Swain SM, Whaley FS, Ever MS, Congestive heart failure in patients treated with doxorubicin: A retrospective analysis of three trials, Cancer 97 (2003) 2869–2879. [DOI] [PubMed] [Google Scholar]
  • [244].Fowers KD, Kopeček J, Targeting of multidrug-resistant human ovarian carcinoma cells with anti-P-glycoprotein antibody conjugates, Macromol. Biosci 12 (2012) 502–514. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [245].Chu T-W, Yang J, Kopeček J, Anti-CD20 multivalent HPMA copolymer-Fab’ conjugates for the direct induction of apoptosis, Biomaterials 33 (2012) 7174–7181. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [246].Lu Z-R, Kopečková P, Kopeček J, Polymerizable Fab’ antibody fragments for targeting of anticancer drugs, Nature Biotechnol. 17 (1999) 1101–1104. [DOI] [PubMed] [Google Scholar]
  • [247].Lu Z-R, Shiah J-G, Kopečková P, Kopeček J, Polymerizable Fab’ antibody fragment targeted photodynamic cancer therapy in nude mice, STP Pharma Sci 13 (2003) 69–75. [Google Scholar]
  • [248].Journo-Gershfeld G, Kapp D, Shamay Y, Kopeček J, David A, Hyalorunan oligomers-HPMA copolymer conjugates for targeting paclitaxel to CD44-overexpressing ovarian carcinoma, Pharm. Res. 29 (2012) 1121–1133. [DOI] [PubMed] [Google Scholar]
  • [249].David A, Kopečková P, Kopeček J, Rubinstein A, The role of galactose, lactose and their three-dimensional arrangement in the biorecognition of HPMA copolymers by human colon adenocarcinoma cells, Pharm. Res. 19 (2002) 1114–1122. [DOI] [PubMed] [Google Scholar]
  • [250].Luo Y, Bernshaw NJ, Lu Z-R, Kopeček J, Prestwich GD, Targeted delivery of HPMA copolymer – hyaluronan – doxorubicin bioconjugates, Pharm. Res. 19 (2002) 396–402. [DOI] [PubMed] [Google Scholar]
  • [251].David A, Kopečková P, Minko T, Rubinstein A, Kopeček J, Design of a multivalent galactoside ligand for selective targeting of HPMA copolymer-doxorubicin conjugates to human colon cancer cells, Eur. J. Cancer 40 (2004) 148–157. [DOI] [PubMed] [Google Scholar]
  • [252].Wróblewski S, Berenson M, Kopečková P, Kopeček J, Biorecognition of HPMA copolymer-lectin conjugates as an indicator of differentiation of cell-surface glycoproteins in development, maturation and diseases of human and rodent gastrointestinal tissues, J. Biomed. Mater. Res. 51 (2000) 329–342. [DOI] [PubMed] [Google Scholar]
  • [253].Shamay Y, Paulin D, Ashkenazy G, David A, E-selectin binding peptide-polymer-drug conjugates and their selective cytotoxicity against vascular endothelial cells, Biomaterials 30 (2009) 6460–6468. [DOI] [PubMed] [Google Scholar]
  • [254].Crownover E, Duvall CL, Convertine A, Hoffman AS, Stayton PS, RAFT-synthesized graft copolymers that enhance pH-dependent membrane destabilization and protein circulation times, J. Control. Release 155 (2011) 167–174. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [255].Tang A, Kopečková P, Kopeček J, Binding and cytotoxicity of HPMA copolymer conjugates to lymphocytes mediated by receptor-binding epitopes, Pharm. Res. 20 (2003) 360–367. [DOI] [PubMed] [Google Scholar]
  • [256].Fletcher NL, Houston ZH, Simpson JD, Veedu RN, Thurecht KJ, Designed multifunctional polymeric nanomedicines: Long-term biodistribution and tumor accumulation of aptamer-targeted nanomaterials, Chem. Commun. 54 (2018) 11538–11541. [DOI] [PubMed] [Google Scholar]
  • [257].Biffi S, Voltan R, Bortot B, Zauli G, Secchiero P, Actively targeted nanocarriers for drug delivery to cancer cells, Expert Opin. Drug Deliv 16 (2019) 481–496. [DOI] [PubMed] [Google Scholar]
  • [258].Maso K, Grigoletto A, Vicent MJ, Pasut G, Molecular platforms for targeted drug delivery, Int. Rev. Cell Mol. Biol. 346 (2019) 1–50. [DOI] [PubMed] [Google Scholar]
  • [259].Lammers T, Kiessling F, Hennink W, Storm G, Drug targeting to tumors: Principles, pitfalls and (pre-)clinical progress, J. Control. Release 161 (2012) 175–187 [DOI] [PubMed] [Google Scholar]
  • [260].Allen TM, Ligand-targeted therapeutics in anticancer therapy, Nat. Rev. Cancer 2 (2002) 750–763. [DOI] [PubMed] [Google Scholar]
  • [261].Srinivasarao M, Low PS, Ligand-targeted drug delivery, Chem. Rev. 117 (2017) 12133–12164 [DOI] [PubMed] [Google Scholar]
  • [262].Ding H, Prodinger WM, Kopeček J, Identification of CD21-binding peptides with phage display and investigation of binding properties of HPMA copolymer – peptide conjugates, Bioconjug. Chem 17 (2006) 514–523. [DOI] [PubMed] [Google Scholar]
  • [263].Ding H, Prodinger WM, Kopeček J, Two-step fluorescence screening of CD21-binding peptides with one-bead one-compound library and investigation of binding of HPMA copolymer – peptide conjugates, Biomacromolecules 7 (2006) 3037–3046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [264].Sefah K, Shangguan D, Xiong X, O’Donoghue MB, Tan W, Development of DNA aptamers using Cell-SELEX, Nature Protocols 5 (2010) 1169–1185, doi: 10.1038/nprot.2010.66 (2010). [DOI] [PubMed] [Google Scholar]
  • [265].Moody PR, Sayers EJ, Magnusson JP, Alexander C, Borri P, Watson P, Jones AT, Receptor crosslinking: A general method to trigger internalization and lysosomal targeting of therapeutic receptor:ligand complexes, Mol. Ther. 23 (2015) 1888–1898. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [266].Tjandra KC, Thordarson P, Multivalency in drug delivery- when is it too much of a good thing? Bioconjug. Chem 30 (2019) 503–514. [DOI] [PubMed] [Google Scholar]
  • [267].Chu T-W, Zhang R, Yang J, Chao MP, Shami PJ, Kopeček J, A two-step pretargeted nanotherapy for CD20 crosslinking may achieve superior anti-lymphoma efficacy to rituximab, Theranostics 5 (2015) 834–846. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [268].Li L, Yang J, Soodvilai S, Wang J, Opanasopit P, Kopeček J, Drug-free albumin-triggered sensitization of cancer cells to anticancer drugs, J. Control. Release 293 (2019) 84–93 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [269].Szymanska M, Fosdahl AM, Nikolaysen F, Pedersen MW, Grandal MM, Stang E, Bertelsen V, A combination of two antibodies recognizing non-overlapping epitopes of HER2 induces kinase activity-dependent internalization of HER2, J. Cell. Mol. Med. 20 (2016) 1999–2010. doi: 10.1111/jcmm.12899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [270].Li L, Li Y, Yang C-H, Radford DC, Wang J, Janát-Amsbury M, Kopeček J, Yang J, Inhibition of immunosuppresive tumors by polymer-assisted inductions of immunogenic cell death and multivalent PD-L1 crosslinking. Adv. Funct. Mater (2020) doi: 10.1002/admf.201908961. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [271].Thomas A, Teicher BA, Hassan R, Antibody-drug conjugates for cancer therapy, Lancet Oncol 17 (2016) e254–e262. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [272].Zhang L, Fang Y, Kopeček J, Yang J, A New construct of antibody-drug conjugates for treatment of non-Hodgkin’s lymphoma, Eur. J. Pharm. Sci. 103, (2017) 36–46. [DOI] [PubMed] [Google Scholar]
  • [273].Lidický O, Janoušková O, Strohalm J, Alam M, Klener P, Etrych T, Anti-lymphoma efficacy comparison of anti-CD20 monoclonal antibody-targeted and non-targeted star-shaped polymer-prodrug conjugates, Molecules 20 (2015) 19849–19864. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [274].Tijerina M, Kopečková P, Kopeček J, Correlation of subcellular compartmentalization of HPMA copolymer-Mce6 conjugates with chemotherapeutic activity in human ovarian carcinoma cells, Pharm. Res. 20 (2003) 728–737. [DOI] [PubMed] [Google Scholar]
  • [275].Duvall CJ, Convertine AJ, Benoit DSW, Hoffman AS, Stayton PS, Intracellular delivery of proapoptotic peptide via conjugation to a RAFT synthesized endosomolytic polymer, Mol. Pharm. 7 (2010) 468–476. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [276].He W, Xing X, Wang X, Wu D, Wu W, Guo J, Mitragotri S, Nanocarrier-mediated cytosolic delivery of biopharmaceuticals, Adv. Funct. Mat (2020) 1910566; doi: 10.1002/adfm.201910566. [DOI] [Google Scholar]
  • [277].Schmidt HB, Görlich D, Transport selectivity of nuclear pores, phase separation, and mebraneless organelles, Trends Biochem. Sci. 41 (2016) 46–61. [DOI] [PubMed] [Google Scholar]
  • [278].Zhong J, Li L, Zhu X, Guan S, Yang Q, Zhou Z, Zhang Z, Huang Y, A smart polymeric platform for multistage nucleus-targeted anticancer drug delivery, Biomaterials 65 (2015) 43–55. [DOI] [PubMed] [Google Scholar]
  • [279].Li L, Sun W, Zhang Z, Huang Y, Time-staggered delivery of docetaxel and H1-S6A, F8A peptide for sequential dual-strike chemotherapy through tumor priming and nuclear targeting, J. Control. Release 232 (2016) 62–74. [DOI] [PubMed] [Google Scholar]
  • [280].Cuchelkar V, Kopečková P, Kopeček J, Novel HPMA copolymer-bound constructs for combined tumor and mitochondrial targeting, Mol. Pharm. 5 (2008) 776–786. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [281].Rebuffat A, Bernasconi A, Ceppi M, Wehrli H, Brenz Verca S, Ibrahim M, Frey BM, Frey FJ, Rusconi S, Selective enhancement of gene transfer by steroid-mediated gene delivery, Nature Biotechnol. 19 (2001) 1155–1161. [DOI] [PubMed] [Google Scholar]
  • [282].Cuchelkar V, Strategies for enhancing the photodynamic effect of N-(2-hydroxypropyl)-methacrylamide copolymer bound mesochlorin e6, Ph.D. Dissertation, Department of Biomedical Engineering, University of Utah, 2008.
  • [283].Crawford L, Rosch J, Putnam D, Concepts, technologies, and practices for drug delivery past the blood-brain barrier to the central nervous system, J. Control. Release 240 (2016) 251–266 [DOI] [PubMed] [Google Scholar]
  • [284].Armstrong BK, Smith Q, Rapoport SI, Strohalm J, Kopeček J, Duncan R, Osmotic opening of the blood-brain barrier permeability to N-(2-hydroxypropyl)methacrylamide copolymers. Effect of polymer Mw, charge and hydrophobicity, J. Control. Release 10 (1989) 27–35. [Google Scholar]
  • [285].Lammers T, Koczera P, Fokong S, Gremse F, Ehling J, Vogt M, Pich A, Storm G, van Zandvoort M, Kiessling F, Theranostic USPIO loaded microbubbles for mediating and monitoring bloob-brain barrier permeation, Adv. Funct. Mater. 25 (2015) 36–43. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [286].Kim YS, Sung DK, Kim H, Kong WH, Kim YE, Hahn SK, Nose-to-brain delivery of hyaluronate – FG loop peptide conjugate for non-invasive hypoxic-ischemic encephalopathy therapy, J. Control. Release 307 (2019) 76–89. [DOI] [PubMed] [Google Scholar]
  • [287].Johnsen KB, Burkhart A, Thomsen LB, Andersen T TL Moos, Targeting the transferrin receptor for brain drug delivery, Prog. Neurobiol. 181 (2019) 101665; doi: 10.1016/j.pneurobio.2019.101655. [DOI] [PubMed] [Google Scholar]
  • [288].Pardridge WM, Blood-brain barrier drug delivery of IgG fusion proteins with a transferrin receptor monoclonal antibody, Expert Opin. Drug Deliv 12 (2015) 207–222. [DOI] [PubMed] [Google Scholar]
  • [289].Demeule M, Curri JC, Bertrand Y, Ché C, Nguyen T, Régina A, Gabathuler R, Castaigne JP, Béliveau R, Involvement of the low-density lipoprotein receptor-related protein in the transcytosis of the brain delivery vector angiopep-2, J. Neurochem. 106 (2008) 1534–1544. [DOI] [PubMed] [Google Scholar]
  • [290].Huile G, Shuaiqi P, Zhi Y, Shijie C, Chen C, Xinguo J, Shun S, Zhiqing P, Yu H, A cascade targeting strategy for brain neuroglial cells employing nanoparticles modified with angiopep-2 peptide and EGFP-EGF1 protein, Biomaterials 32 (2011) 8669–8675. [DOI] [PubMed] [Google Scholar]
  • [291].Li Y, Zheng X, Gong M, Zhang J, Delivery of a peptide-drug conjugate targeting the blood brain barrier improved the efficacy of paclitaxel against glioma, Oncotarget 7(48) (2016) 79401–79407. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [292].Thomas FC, Taskar K, Rudraraju V, Goda S, Thorsheim HR, Gaasch JA, Mittapalli RK, Palmieri D, Steeg PS, Lockman PR, Smith QR, Uptake of ANG1005, a novel paclitaxel derivative, through the blood-brain barrier into brain and experimental brain metastases of breast cancer, Pharm. Res. 26 (2009) 2486–2494. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [293].Sun T,* Yang J,* Galstyan A, Braubach O, Wang J, Penichet ML, Li L, Shatalova E, Grodzinski ZB, Klymyshyn D, Ding H, Black KL, Holler E, Kopeček J, Ljubimova JY, BBB cCrossing, c-Myc targeting nanodrug for treating primary central nervous system lymphoma in combination with immune checkpoint inhibition, J. Control. Release, submitted. [Google Scholar]
  • [294].Kumthekar P, Tang SC, Brenner AJ, Kesari S, E Piccioni D, Anders CK, Carrillo JA, Chalasani P, Kabos P, Puhalla SL, Tkaczuk KHR, Garcia A, Ahluwalia MS, Wefel JS, Lakhani N, Ibrahim NK, ANG1005, a brain penetrating peptide-drug conjugate, shows activity in patients with breast cancer with leptomeningeal carcinomatosis and recurrent brain metastases, Clin. Cancer Res. (2020); doi: 10.1158/1078-0432.CCR-19-3258. [DOI] [PubMed] [Google Scholar]
  • [295].Minko T, HPMA copolymers for modulating cellular signaling and overcoming multidrug resistance, Adv. Drug Deliv. Rev. 62 (2010) 192–202 [DOI] [PubMed] [Google Scholar]
  • [296].Aller SG, Yu J, Ward A, Chittaboina S, Zhuo R, Harell PM, Trinh YT, Zhang Q, Urbatsch IL, Chang G, Structure of P-glycoprotein reveals a molecular basis for poly-specific drug binding, Science 323 (2009) 1718–1722. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [297].Minko T, Kopečková P, Pozharov V, Kopeček J, HPMA copolymer bound adriamycin overcomes MDR1 gene encoded resistance in a human ovarian carcinoma cell line, J. Control. Release 54 (1998) 223–233. [DOI] [PubMed] [Google Scholar]
  • [298].Minko T, Kopečková P, Kopeček J, Chronic exposure to HPMA copolymer-bound adriamycin does not induce multidrug resistance in a human ovarian carcinoma cell line, J. Control. Release 59 (1999) 133–148. [DOI] [PubMed] [Google Scholar]
  • [299].Tijerina M, Fowers K, Kopečková P, Kopeček J, Chronic exposure of human ovarian carcinoma cells to free or HPMA copolymer-bound mesochlorin e6 does not induce P-glycoprotein mediated multidrug resistance, Biomaterials 21 (2000) 2203–2210. [DOI] [PubMed] [Google Scholar]
  • [300].Sivák L, Šubr V, Tomala J, Říhová B, Strohalm J, Etrych T, Kovář M, Overcoming drug resistance via simultaneous delivery of cytostatic drug and P-glycoprotein inhibitor to cancer cells by HPMA copolymer conjugate, Biomaterials 115 (2017) 65–80. [DOI] [PubMed] [Google Scholar]
  • [301].Krinick NL, Sun Y, Joyner D, Spikes JD, Straight RC, Kopeček J, A polymeric drug delivery system for the simultaneous delivery of drugs activatable by enzymes and/or light, J. Biomat. Sci. Polym. Ed 5 (1994) 303–324. [DOI] [PubMed] [Google Scholar]
  • [302].Markovsky E, Baabur-Cohen H, Satchi-Fainaro R, Anticancer polymeric nanomedicines bearing synergistic drug combination is superior to a mixture of individually conjugated drugs, J. Control. Release 187 (2014) 145–157. [DOI] [PubMed] [Google Scholar]
  • [303].Baabur-Cohen H, Vossen LI, Krüger HR, Eldar-Boock A, Jeini E, Landa-Rouben N, Tiram G, Wedepohl S, Markovsky E, Leor J, Calderón M, Satchi-Fainaro R, In vivo comparative study of distinct polymeric architectures bearing a combination of paclitaxel and doxorubicin at a synergistic ratio, J. Control. Release 257 (2017) 118–131. [DOI] [PubMed] [Google Scholar]
  • [304].Larson N, Roberts S, Ray A, Buckway B, Cheney DL, Ghandehari H, In vitro synergistic action of geldanamycin- and docetaxel-containing HPMA copolymer-RGDfK conjugates against ovarian cancer, Macromol. Biosci 14 (2014) 1735–1747. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [305].Zhou Y, Yang J, Kopeček J, Selective inhibitory effect of HPMA copolymer – cyclopamine conjugate on prostate cancer stem cells, Biomaterials 33 (2012) 1863–1872. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [306].Zhou Y, Yang J, Rhim J, Kopeček J, HPMA copolymer-based combination therapy toxic to both prostate cancer stem/progenitor cells and differentiated cells induces durable anti- tumor effect, J. Control. Release 172 (2013) 946–953. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [307].Zhou Y, Yang J, Zhang R, Kopeček J, Combination therapy of prostate cancer with HPMA copolymer conjugates containing PI3K/mTOR inhibitor and docetaxel, Eur. J. Pharm. Biopharm 89 (2014) 107–115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [308].Peterson CM, Lu JM, Sun Y, Peterson CA, Shiah J-G, Straight RC, Kopeček J, Combination chemotherapy and photodynamic therapy with N-(2-hydroxypropyl)methacrylamide copolymer-bound anticancer drugs inhibit human ovarian carcinoma heterotransplanted in nude mice, Cancer Res. 56 (1996) 3980–3985. [PubMed] [Google Scholar]
  • [309].Shiah JG, Sun Y, Kopečková P, Peterson CM, Straight RC, Kopeček J, Combination chemotherapy and photodynamic therapy of targetable N-(2-hydroxypropyl)methacrylamide copolymer – doxorubicin/mesochlorin e6 – OV-TL16 antibody immunoconjugates, J. Control. Release 74 (2001) 249–253. [DOI] [PubMed] [Google Scholar]
  • [310].Larson N, Gormley A, Frazier N, Ghandehari H, Synergistic enhancement of cancer therapy using a combination of heat shock protein targeted HPMA copolymer-drug conjugates and gold nanorod induced hyperthermia, J. Control. Release 170 (2013) 41–50. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [311].Frazier N, Payne A, Dillon C, Subrahmanyam N, Ghandehari H, Enhanced efficacy of combination heat shock targeted polymer therapeutics with high intensity focused ultrasound, Nanomedicine 13 (2017) 1235–1243. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [312].Dobrovolskaia MA, Pre-clinical immunotoxicity studies of nanotechnology-formulated drugs: Challenges, considerations and strategy, J. Control. Release 220 (2015) 571–583. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [313].Říhová B, Kubáčková M, Clinical implications of N-(2-hydroxypropyl)methacrylamide copolymers, Curr. Pharm. Biotechnol. 4 (2003) 311–322. [DOI] [PubMed] [Google Scholar]
  • [314].Říhová B, Clinical experience with anthracycline antibiotics-HPMA copolymer-human immunoglobulin conjugates. Adv. Drug Deliv. Rev. 61 (2009) 1149–1158. [DOI] [PubMed] [Google Scholar]
  • [315].Říhová B, Strohalm J, Kubáčková K, Jelínková M, Rozprimová L, Šírová M, Plocová D, Mrkvan T, Kovář M, Pokorná J, Etrych T, Ulbrich K, Drug-HPMA-HuIg conjugates effective against human solid cancer, Adv. Exp. Med. Biol. 519 (2003) 125–143. [DOI] [PubMed] [Google Scholar]
  • [316].Říhová B, Strohalm J, Kubáčková M, Jelínková M, Hovorka O, Kovář M, Plocová D, Šírová M, Štastný M, Rozprimová L, Ulbrich K, Acquired and specific immunological mechanisms co-resposible for efficacy of polymer-bound drugs, J. Control. Release 78 (2002) 97–104. [DOI] [PubMed] [Google Scholar]
  • [317].Říhová B, Kopeček J, Ulbrich K, Chytrý V, Immunogenicity of N-(2-hydroxypropyl)-methacrylamide copolymers, Makromol. Chem. Suppl. 9 (1985) 13–24. [Google Scholar]
  • [318].Říhová B, Kopeček J, Ulbrich K, Pospíšil J, Mančal P, Effect of the chemical structure of N-(2-hydroxypropyl)methacrylamide copolymers on their ability to induce antibody formation in inbred strains of mice, Biomaterials 5 (1984) 143–148. [DOI] [PubMed] [Google Scholar]
  • [319].Kverka M, Hartley JM, Chu T-W, Yang J, Heidchen R, Kopeček J, Immunogenicity of coiled-coil based drug-free macromolecular therapeutics, Biomaterials 35 (2014) 5886–5896. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [320].Anselmo AC, Mitragotri S, Nanoparticles in the clinic, Bioeng. Transl. Med 1 (2016) 10–29. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [321].Anselmo AC, Mitragotri S, Nanoparticles in the clinic: An update, Bioeng. Transl. Med 4 (2019) e10143; doi:10/1002/btm2.10143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [322].Lu ZR, Qiao P, Drug delivery in cancer therapy, quo vadis? Mol. Pharm. 15 (2018) 3603–3616. [DOI] [PubMed] [Google Scholar]
  • [323].Dozono H, Yanazume S, Nakamura H, Etrych T, Chytil P, Ulbrich K, Fang J, Arimura T, Douchi T, Kobayashi H, Ikoma M, Maeda H, HPMA copolymer conjugated pirarubicin in multimodal treatment of a patient with stage IV prostate cancer and extensive lung and bone metastases, Target Oncol 11 (2016) 101–106. [DOI] [PubMed] [Google Scholar]
  • [324].Nowotnik DP, Cvitkovic E, ProLindacTM (AP5346): A review of the development of an HPMA DACH platinum polymer therapeutic, Adv. Drug Deliv. Rev. 61 (2009) 1214–1219. [DOI] [PubMed] [Google Scholar]
  • [325].Bobo D, Robinson KJ, Islam J, Thurecht KJ, Corrie SR, Nanoparticle-based nanomedicines: A review of FDA-approved materials and clinical trials to date, Pharm. Res. 33 (2016) 2373–2387. [DOI] [PubMed] [Google Scholar]
  • [326].Mi P, Cabral H, Kataoka K, Ligand-installed nanocarriers toward precision therapy, Adv. Mater. 32 (2019) 1902604; doi: 10.1002/adma.201902604. [DOI] [PubMed] [Google Scholar]
  • [327].Lammers T, Ferrari M, The success of nanomedicine, Nano Today 31 (2020) 100853. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [328].Couvreur P, Nanomedicine: From where we are coming and where we are going? J. Control. Release 311–312 (2019) 319–321. [DOI] [PubMed] [Google Scholar]
  • [329].Martins JP, des Neves J, de la Fuente M, Cella C, Florindo H, Gündai-Türelli N, Popat A, Santos JL, Sousa F, Schmid R, Wolfram J, Sarmento B, Santos HA, The solid progress of nanomedicine, Drug Deliv. Transl. Res 10 (2020) 726–729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [330].Park K, The beginning of the end of the nanomedicine hype, J. Control. Release 305 (2019) 221. [DOI] [PubMed] [Google Scholar]
  • [331].Grodzinski P, NCI Centers of Cancer Nanotechnology Excellence (CCNEs) – a full story to set the record straight, J. Control. Release 309 (2019) 341–342. [DOI] [PubMed] [Google Scholar]
  • [332].Hrkach J, Langer R, From micro to nano: evolution and impact of drug delivery in treating disease, Drug Deliv. Transl. Res 10 (2020) 567–570. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [333].Metselaar JM, Lammers T, Challenges in nanomedicine clinical translation, Drug Deliv. Transl. Res 10 (2020) 721–725. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [334].Gartner hype-cycle; gartner.com
  • [335].Leroux J-P, The novelty bubble, J Control. Release 278 (2018) 140–141. [DOI] [PubMed] [Google Scholar]
  • [336].Micha-Screttas M, Ringsdorf H, Polymer science and polymer therapeutics: Macromolecules, dendrimers, and nanomedicine, Curr. Topics Med. Chem 8 (2008) 1161–1164. [DOI] [PubMed] [Google Scholar]
  • [337].Galluzzi L, Buqué A, Kepp O, Zitvogel L, Kroemer G, Immunological effects of conventional chemotherapy and targeted anticancer agents, Cancer Cell 28 (2015) 690–714. [DOI] [PubMed] [Google Scholar]
  • [338].Pfirschke C, Engblom C, Rickelt S, Cortez-Retamozo V, Garris C, Pucci F, Yamazaki T, Poirier-Colame V, Newton A, Redouane Y, Lin YJ, Wojtkiewicz G, Iwamoto Y, Mino-Kenudson M, Huynh TG, Hynes RO, Freedman GJ, Kroemer G, Zitvogel L, Weissleder R, Pittet MJ, Immunogenic chemotherapy sensitizes tumors to checkpoint blockade therapy, Immunity 44 (2016) 343–354. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [339].Kashima H, Momose F, Umehara H, Miyoshi N, Ogo N, Muraoka D, Shiku H, Harada N, Asai A, Epirubicin, identified using a novel luciferase reporter assay for Foxp3 inhibitors, inhibits regulatory T cell activity, PLoS ONE 11(6) (2016) e0156643; doi: 10.1371/journal.pone.0156643. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [340].Peng J, Hamanishi J, Matsumura N, Abiko K, Murat K, Baba T, Yamaguchi K, Horikawa N, Hosoe Y, Murphy SK, Konishi I, Mandai M, Chemotherapy induces programmed cell death-ligand 1 overexpression via the nuclear factor-κB to foster an immunosuppressive tumor microenvironment in ovarian cancer, Cancer Res. 75 (2015) 5034–5045. [DOI] [PubMed] [Google Scholar]
  • [341].Acúrcio RC, Scomparin A, Conniot J, Salvador JAR, Satchi-Fainaro R, Florindo HF, Guedes RC, Structure-function analysis of immune checkpoint receptors to guide emerging anticancer immunotherapy, J. Med. Chem. 61 (2018) 10957–10975. [DOI] [PubMed] [Google Scholar]
  • [342].Chen L, Han X, Anti-PD-1/PD-L1 therapy of human cancer: Past, present, and future, J. Clin. Inv 125 (2015) 3384–3391. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [343].Salmaninejad A, Valilou SF, Shabgah AG, Aslani S, Alimardani M, Padar A, Sahabkar A, PD-1/PD-L1 pathway: Basic biology and role in cancer immunotherapy, J. Cell Physiol. 234 (2019) 16824–16837. [DOI] [PubMed] [Google Scholar]
  • [344].Shaabani S, Huizinga HPS, Butera R, Kouchi A, Guzik K, Magiera-Mularz K, Holak TA, Dömling A, A patent review on PD-1/PD-L1 antagonists: small molecules, peptides, and macrocycles (2015–2018), Exp. Opinion Ther. Patents 28 (2018) 665–678. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [345]. https://www.pfizer.com/science/oncology-cancer/immuno-oncology/the-science-of-io.
  • [346].Sun Q, Barz M, De Geest BG, Diken M, Hennink WE, Kiessling F, Lammers T, Shi Y, Nanomedicine and macroscale materials in immune-oncology, Chem. Soc. Rev. 48 (2019) 351–381. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [347].Haanen JBAG, Converting cold into hot tumors by combining immunotherapies, Cell 170 (2017) 1055–1056. [DOI] [PubMed] [Google Scholar]
  • [348].Obeid M, Tesniere A, Ghiringhelli F, Fimia GM, Apetoh L, Perfettini J-L, Castedo M, Mignot G, Panaretakis T, Casares N, Métivier D, Larochette N, van Endert P, Ciccosanti F, Piacentini M, Zitvogel L, Kroemer G, Calreticulin exposure dictates the immunogenicity of cell death, Nature Med 13 (2007) 54–61. [DOI] [PubMed] [Google Scholar]
  • [349].Galluzzi L, Buqué A, Kepp O, Zitvogel L, Kroemer G Immunogenic cell death in cancer and infectious disease, Nature Rev. Immunol 17 (2017) 97–111. [DOI] [PubMed] [Google Scholar]
  • [350].Šírová M, Strohalm J, Šubr V, Plocová D, Rossmann P, Mrkvan T, Ulbrich K, Říhová B, Treatment with HPMA copolymer-based doxorubicin conjugate containing human immunoglobulin induces long-lasting systemic anti-tumour immunity in mice, Cancer Immunol. Immunother. 56 (2007) 35–47. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [351].Kuai R, Yuan W, Son S, Nam J, Xu Y, Fan Y, Schwendeman A, Moon JJ, Elimination of established tumors with nanodisc-based combination chemoimmunotherapy, Sci. Adv 4 (2018) eaa01736. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [352].Soliman HH, nab-Paclitaxel as a potential partner with checkpoint inhibitors in solid tumors, OncoTargets Ther. 10 (2017) 101–112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [353].Fan Y, Kuai R, Xu Y, Ochyl LJ, Irvine DJ, Moon JJ, Immunogenic cell death amplified by co-localized adjuvant delivery for cancer immunotherapy, Nano Letters 17 (2017) 7387–7393. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [354].Rios-Doria J, Durham N, Wetzel L, Rohstein R, Chesebrough J, Holoweckyj N, Zhao W, Leow CC, Hollingsworth R, Doxil synergizes with cancer immunotherapies to enhance antitumor responses in syngeneic mouse models, Neoplasia 17 (2015) 661–670. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [355].He X, Xu C, Immune checkpoint signaling and cancer immunotherapy, Cell Res. (2020) doi: 10.1038/s41422-020-0343-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [356].Pardoll DM, The blockade of immune checkpoints in cancer immunotherapy, Nature Rev. Cancer 12 (2012) 252–264. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [357].Wang H, Yao H, Li C, Shi H, Lan J, Zhang Y, Liang L, Fang JY, Xu J, HIP1R targets PD-L1 to lysosomal degradation to alter T cell-mediated cytotoxicity, Nature Chem. Biol. 15 (2019) 42–50. [DOI] [PubMed] [Google Scholar]
  • [358].Yao H, Lan J, Shi H, Brosseau JP, Wang H, Li H, Fang C, Zhang Y, Liang L, Zhou X, Wang C, Xue Y, Y Cui J Xu, Inhibiting PD-L1 palmitoylation enhances T-cell immune responses against tumors, Nature Biomed. Eng. 3 (2019) 306–317. [DOI] [PubMed] [Google Scholar]
  • [359].Burr ML, Sparbier CE, Chan YC, Williamson JC, Woods K, Beavis PA, Lam EYN, Henderson MA, Bell CC, Stolzenburg S, Gilan O, Bloor S, Noori T, Morgens DW, Bassik MC, Neeson PJ, Behren A, Darcy PK, Dawson SJ, Voskoboinik I, Trapani JA, Cebon J, Lehner PJ, Dawson MA, CMTM6 maintains the expression of PD-L1 and regulates anti-tumor immunity, Nature 549 (2017) 101–105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [360].Li C, Zhang N, Zhou J, Ding C, Jin Y, Cui X, Pu K, Zhu Y, Peptide blocking of PD-1/PD-L1 interaction for cancer immunotherapy, Cancer Immunol. Res 6 (2018) 178–188. [DOI] [PubMed] [Google Scholar]
  • [361].Chang HN, Liu BY, Qi YK, Zhou Y, Chen YP, Pan KM, Li WW, Zhou XM, Ma WW, Fu CY, Qi YM, Liu L, Gao YF, Blocking of the PD-1/PD-L1 interaction by a D-peptide antagonist for cancer immunotherapy, Ang. Chem. Int. Ed 54 (2015) 11760–11764. [DOI] [PubMed] [Google Scholar]
  • [362].Chu T-W, Kopeček J, Drug-free macromolecular therapeutics – A new paradigm in polymeric nanomedicines, Biomaterials Sci 3 (2015) 908–922. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [363].Yang J, Li L, Kopeček J, Biorecognition - a key to drug-free macromolecular therapeutics, Biomaterials 190–191 (2019) 11–23. [DOI] [PubMed] [Google Scholar]
  • [364].Yang J, Xu C, Wang C, Kopeček J, Refolding hydrogels self-assembled from HPMA graft copolymers by antiparallel coiled-coil formation, Biomacromolecules 7 (2006) 1187–1195. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [365].Janas E, Priest R, Wilde JI, White JH, Malhotra R, Rituxan (anti-CD20 antibody)-induced translocation of CD20 into lipid rafts is crucial for calcium influx and apoptosis, Clin. Exp. Immunol. 139 (2005) 439–446. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [366].Seyfizadeh N, Seyfizadeh N, Hasenkamp J, Huerta-Yepez SA, Molecular perspective on rituximab: a monoclonal antibody for B cell non-Hodgkin lymphoma and other affections, Crit. Rev. Oncol. Hemat 97 (2016) 275–290. [DOI] [PubMed] [Google Scholar]
  • [367].Okroj M, Österborg A, Blom AM, Effector mechanisms of anti-CD20 monoclonal antibodies in B cell malignancies, Cancer Treat. Rev 39 (2013) 632–639. [DOI] [PubMed] [Google Scholar]
  • [368].Wu K, Liu J, Johnson RN, Yang J, Kopeček J, Drug-free macromolecular therapeutics: Induction of apoptosis by coiled-coil-mediated cross-linking of antigens on the cell surface, Angew. Chem. Int. Ed. 49 (2010) 1451–1455. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [369].Wu K, Yang J, Liu J, Kopeček J, Coiled-coil based drug-free macromolecular therapeutics: In vivo efficacy, J. Control. Release 157 (2012) 126–131. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [370].Zhang L, Fang Y, Li L, Yang J, Radford DC, Kopeček J, Human serum albumin-based drug-free macromolecular therapeutics: Apoptosis induction by coiled-coil-mediated cross-linking of CD20 antigens on lymphoma B cell surface, Macromol. Biosci 18 (2018) 1800224; doi: 10.1002/mabi.201800224. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [371].Li L, Yang J, Wang J, Kopeček J, Drug-free macromolecular therapeutics induce apoptosis via calcium influx and mitochondrial signaling pathway, Macromol. Biosci 18(1) (2018); doi: 10.1002/mabi.201700196. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [372].Hartley JM, Chu T-W, Peterson EM, Zhang R, Yang J, Harris J, Kopeček J, Super-resolution imaging and quantitative analysis of membrane protein/lipid raft clustering mediated by cell surface self-assembly of hybrid nanoconjugates, ChemBioChem 16 (2015) 1725–1729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [373].Hartley JM, Zhang R, Gudheti M, Yang J, Kopeček J, Tracking and quantifying polymer therapeutic distribution on a cellular level using 3D dSTORM, J. Control. Release 231 (2016) 50–59. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [374].Li L, Yang J, Wang J, Kopeček J, Amplification of CD20 crosslinking in rituximab resistant B-lymphoma cells enhances apoptosis induction by drug-free macromolecular therapeutics, ACS Nano 12 (2018) 3658–3670. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [375].Wang J, Li L, Yang J, Clair P, Glenn M, Stephens DM, Radford DC, Kosak KM, Deininger MW, Shami PJ, Kopeček J, Drug-free macromolecular therapeutics induce apoptosis in cells isolated from clinical patients with B cell malignancies with enhanced apoptosis induction by pretreatment with gemcitabine, Nanomedicine: NBM 16 (2019) 217–225; doi: 10.1016/j.nano.2018.12.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [376].Chu T-W, Kosak KM, Shami PJ, Kopeček J, Drug-free macromolecular therapeutics induce apoptosis of patient chronic lymphocytic leukemia cells, Drug Deliv. Transl. Res 4 (2014) 389–394. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [377].Gabellier L, Carton G, Obinutuzumab for relapsed or refractory indolent non-Hodgkin’s lymphomas, Ther. Adv. Hematol. 7 (2016) 85–93. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [378].Alduaij W, Ivanov A, Honeychurch J, Cheadle EJ, Potluri S, Lim SH, Shimada K, Chan CH, Tutt A, Beers SA, Glennie M, Novel type II anti-CD20 monoclonal antibody (GA101) evokes homotypic adhesion and actin-dependent, lysosome-mediated cell death in B-cell malignancies, Blood 117 (2010) 4519–4529. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [379].Niederfellner G, Lammens A, Mundigl O, Georges GJ, Schaefer W, Schwaiger M, Franke A, Wiechmann K, Jenewein S, Slootstra JW, Timmerman P, Brännström A, Lindstrom F, Mössner E, Umana P, Hopfner KP, Klein C, Epitope characterization and crystal structure of GA101 provide insights into the molecular basis for Type I/II distinction of CD20 antibodies, Blood 118 (2011) 358–367. [DOI] [PubMed] [Google Scholar]
  • [380].Li L, Wang J, Li Y, Radford DC, Yang J, Kopeček J, Broadening and enhancing functions of antibodies by self-assembling multimerization at cell surface, ACS Nano 13 (2019) 11422–11432; doi. 10.1021/acsnano.9b04868. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [381].Niwa T, Kasuya Y, Suzuki Y, Ichikawa K, Yoshida H, Kurimoto A, Tanaka K, Morita K, Novel immunoliposome technology for enhancing the activity of the agonistic antibody against the tumor necrosis factor receptor superfamily, Mol. Pharm. 15 (2018) 3729–3740. [DOI] [PubMed] [Google Scholar]
  • [382].Huet HA, Growney JD, Johnson JA, Li J, Bilic S, Ostrom L, Zafari M, Kowal C, Yang G, Royo A, Jensen M, Dombrecht B, Meerschaert KRA, Kolkman JA, Cromie KD, Mosher R, Gao H, Schuller A, Isaaks R, RC Sellers W, Ettenberg SA, Multivalent nanobodies targeting death receptor 5 elicit superior tumor cell killing through efficient caspase induction, mAbs 6 (2014) 1560–1570. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [383].Valldorf B, Fittler H, Deweid L, Ebenig A, Dickgiesser S, Sellmann C, Becker J, Zielonka S, Empting M, Avrutina O, Kolmar H, An apoptosis-inducing peptide heptad that efficiently clusters death receptor 5, Ang. Chem. Int. Ed 55 (2016) 5085–5089. [DOI] [PubMed] [Google Scholar]
  • [384].Liu F, Si Y, Liu G, Li S, Zhang J, Ma Y, The tetravalent anti-DR5 antibody without cross-linking direct induces apoptosis of cancer cells, Biomed. Pharmacother. 70 (2015) 41–45. [DOI] [PubMed] [Google Scholar]
  • [385].Gieffers C, Kluge M, Merz C, Sykora J, Thiemann M, Schaal R, Fischer C, Branschädel M, Abhari BA, Hohenberger P, Fulda S, Fricke H, Hill O, APG350 induces superior clustering of TRAIL receptors and shows therapeutic antitumor efficacy independent of crosslinking via Fcγ receptors, Mol. Cancer Ther. 12 (2013) 2735–2747. [DOI] [PubMed] [Google Scholar]
  • [386].Pavet V, Beyrath J, Pardin C, Morizot A, Lechner M-C, Briand J-P, Wenlannd M, Maison W, Fournel S, Micheau O, Guichard G, Gronemeyer H, Multivalent DR5 peptides activate the TRAIL death pathway and exert tumoricidial activity, Cancer Res. 70 (2010) 1101–1110. [DOI] [PubMed] [Google Scholar]
  • [387].Steff A-M, Fortin M, Philippoussis F, Lesage S, Arguin C, Johnson P, Hugo P, A cell death pathway induced by antibody-mediated crosslinking of CD45 on lymphocytes, Crit. Rev. Immunol 23 (2003) 421–440. [DOI] [PubMed] [Google Scholar]
  • [388].Gu Z, Yamashiro J, Kono E, Reiter RE, Anti-prostate stem cell antigen monoclonal antibody 1G8 induces cell death in vitro and inhibits tumor growth in vivo via a Fc-independent mechanism, Cancer Res. 65 (2005) 9495–9500. [DOI] [PubMed] [Google Scholar]
  • [389].Wirth T, Soeth E, Czubayko F, Juhl H, Inhibition of endogenous carcinoembryonic antigen (CEA) increases the apoptotic rate of colon cancer cells and inhibits metastatic tumor growth, Clin. Exp. Metastasis 19 (2002) 155–160. [DOI] [PubMed] [Google Scholar]
  • [390].Brühl H, Cihak J, Talke Y, Gomez MR, Hermann F, Goebel N, Renner K, Plachý J, Stangassinger M, Aschermann S, Nimmerjahn F, Mack M, B-cell inhibition by crosslinking CD79b is superior to B-cell depletion with anti-CD20 antibodies in treating collagen induced arthritis, Eur. J. Immunol. 45 (2015) 705–715. [DOI] [PubMed] [Google Scholar]
  • [391].Chang C-H, Wang Y, Gupta P, Goldenberg DM, Extensive crosslinking of CD22 by epratuzumab triggers BCR signaling and caspase-dependent apoptosis in human lymphoma cells, MAbs 7 (2015) 199–211. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [392].Wang J, Li Y, Li L, Yang J, Kopeček J, Exploration and evaluation of therapeutic efficacy of drug-free macromolecular therapeutics in collagen-induced rheumatoid arthritis mouse model, Macromol. Biosci (2020) 1900445; doi: 10.1002/mabi.201900445. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [393].Yazici Y, Rheumatoid arthritis: When should we use rituximab to treat RA? Nat. Rev. Rheumatol. 7 (2011) 379–380. [DOI] [PubMed] [Google Scholar]
  • [394].Chan VS-F, Tsang HH-L, Tam RC-Y, Lu L, Lau C-S, B-cell-targeted therapies in systemic lupus erythematosus, Cell. Mol. Immunol. 10 (2013) 133–142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [395].Hauser SL, Waubant E, Arnold DL, Vollmer T, Antel J, Fox RJ, Bar-Or A, Panzara M, Sarkar N, Agarwal S, Langer-Gould A, Smith CH; HERMES Trial Group, B-cell depletion with rituximab in relapsing-remitting multiple sclerosis, N. Engl. J. Med. 358 (2008) 676–688. [DOI] [PubMed] [Google Scholar]
  • [396].Ramanath V, Nistala R, Chaudhary K, Update on the role of rituximab in kidney diseases and transplant, Expert Opin. Biol. Ther 12 (2012) 223–233. [DOI] [PubMed] [Google Scholar]
  • [397].Chung BL, Toth MJ, Kamaly N, Sei YJ, Becraft J, Mulder WJM, Fayad ZA, Farokhzad OC, Kim YT, Langer R, Nanomedicines for endothelial disorders, Nano Today 10 (2015) 759–776 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [398].Martens CL, Cwirla SE, Lee RY,-W, Whitehorn E, Chen EY-F, Bakker A, Martin EL, Wagstrom C, Gopalan P, Smith CW, Tate E, Koller KJ, Schatz PJ, Dower WJ, Barrett RW, Peptides which bind E-selectin and block neutrophil adhesion, J. Biol. Chem. 270 (1995) 21129–21136. [DOI] [PubMed] [Google Scholar]
  • [399].Fukuda MN, Ohyama C, Lowitz K, Matsuo O, Pasqualini R, Ruoslahti E, Fukuda M, A peptide mimic of E-selectin ligand inhibits sialyl Lewis X-dependent lung colonization of tumor cells, Cancer Res. 60 (2000) 450–456. [PubMed] [Google Scholar]
  • [400].Shamay Y, Raviv L, Golan M, Voronov E, Apte RN, David A, <Inhibition of primary and metastatic tumors in mice by E-selectin-targeted polymer-drug conjugates, J. Control. Release 217 (2015) 102–112. [DOI] [PubMed] [Google Scholar]
  • [401].Tsoref O, Tyomkin D, Amit U, Landa N, Cohen-Rosenboim O, Kain D, Golan M, David A, Leor J, E-selectin-targeted copolymer reduces atherosclerotic lesions, adverse cardiac remodeling, and dysfunction, J. Control. Release 288 (2018) 136–147. [DOI] [PubMed] [Google Scholar]
  • [402].Zaiden M, Rütter M, Shpirt L, Ventura Y, Feinshtein V, David A, CD44-targeted polymer-paclitaxel conjugates to control the spread and growth of metastatic tumors, Mol. Pharm. 15 (2018) 3690–3699. [DOI] [PubMed] [Google Scholar]
  • [403].Liu X-M, Miller SC, Wang D, Beyond oncology – application of HPMA copolymers in non-cancerous diseases, Adv. Drug Deliv. Rev. 62 (2010) 258–271. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [404].Low SA, Kopeček J, Targeting polymer therapeutics to bone, Adv. Drug Deliv. Rev. 64 (2012) 1189–1204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [405].Wang D, Miller S, Kopečková P, Kopeček J, Bone-targeting macromolecular therapeutics, Adv. Drug Deliv. Rev. 57 (2005) 1049–1076. [DOI] [PubMed] [Google Scholar]
  • [406].Wang D, Miller SC, Shlyakhtenko LS, Portillo AM, Liu X-M, Papangkorn K, Kopečková P, Lyubchenko Y, Higuchi WI, Kopeček J, Osteotropic peptide that differentiates functional domains of the skeleton, Bioconjug. Chem. 18 (2007) 1375–1378. [DOI] [PubMed] [Google Scholar]
  • [407].Pan H, Liu J, Dong Y, Sima M, Kopečková P, Brandi ML, Kopeček J, Release of prostaglandin E1 from N-(2-hydroxypropyl)methacrylamide copolymer conjugates by bone cells, Macromol. Biosci 8 (2008) 599–605. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [408].Wang D, Sima M, Mosley RL, Davda JP, Tietze N, Miller SC, Gwilt PR, Kopečková P, Kopeček J, Pharmacokinetic and biodistribution studies of bone-targeting drug delivery system based on N-(2-hydroxypropyl)methacrylamide) copolymers, Mol. Pharm. 3 (2006) 717–725. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [409].Pan H, Sima M, Miller SC, Kopečková P, Yang J, Kopeček J, Efficiency of high molecular weight backbone degradable HPMA copolymer-prostaglandin E1 conjugate in promotion of bone formation in overiectomized rats, Biomaterials 34 (2013) 6528–6538. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [410].Segal E, Pan H, Ofek P, Ugadawa T, Kopečková P, Kopeček J, Satchi-Fainaro R, Targeting angiogenesis-dependent calcified neoplasms using combined polymer therapeutics, PLoS ONE 4(4) e5233 (2009); doi: 10.1371/journal.pone.0005233 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [411].Segal E, Pan H, Benayoun L, Kopečková P, Shaked Y, Kopeček J, Satchi-Fainaro R, Enhanced antitumor activity and safety profile of targeted nano-scaled HPMA copolymer – alendronate – TNP470 conjugate in the treatment of bone malignancies, Biomaterials 32 (2011) 4450–4463. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [412].Miller K, Eldar-Boock A, Polyak D, Segal E, Benayoun L, Shaked Y, Satchi-Fainaro R, Antiangiogenic antitumor activity of HPMA copolymer-paclitaxel-alendronate conjugate on breast cancer bone metastasis mouse model, Mol. Pharm. 8 (2011) 1052–1062. [DOI] [PubMed] [Google Scholar]
  • [413].Karacivi M, Bolu BS, Sanyal R, Targeting to the bone: Alendronate-directed combretastatin A-4 bearing antiangiogenic polymer-drug conjugate, Mol. Pharm. 14 (2017) 1373–1383. [DOI] [PubMed] [Google Scholar]
  • [414].Koziolová E, Venclíková K, Etrych T, Polymer-drug conjugates in inflammation treatment, Physiol. Res 67(Suppl. 2) (2018) S281–S292. [DOI] [PubMed] [Google Scholar]
  • [415].Wang D, Miller SC, Sima M, Parker D, Buswell H, Goodrich KC, Kopečková P, Kopeček J, The arthrotropism of macromolecules in adjuvant-induced arthritis rat model: a preliminary study, Pharm. Res. 21 (2004) 1741–1749. [DOI] [PubMed] [Google Scholar]
  • [416].Yuan F, D Quan L, Cui L, Goldring SR, Wang D, Development of macromolecular prodrug for rheumatoid arthritis, Adv. Drug Deliv. Rev. 64 (2012) 1205–1219. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [417].Liu XM, Quan LD, Tian J, Alnouti Y, Fu K, Thiele GM, Wang D, Synthesis and evaluation of a well-defined HPMA copolymer-dexamethasone conjugate for effective treatment of rheumatoid arthritis, Pharm. Res. 25 (2008) 2910–2919. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [418].Wang D, Goldring SR, The bone, the joints and the balm of gilead, Mol. Pharm. 8 (2011) 991–993. [DOI] [PubMed] [Google Scholar]
  • [419].Gorantla S, Singhvi G, Rapalli VK, Wasghule T, Dubey SK, Saha RN, Targeted drug-delivery systems in the treatment of rheumatoid arthritis: Recent advancement and clinical status, Ther. Deliv 11 (2020) 269–284. [DOI] [PubMed] [Google Scholar]
  • [420].Wang D, Li W, Pechar M, Kopečková P, Brömme D, Kopeček J, Cathepsin K inhibitor – polymer conjugates: Potential drugs for the treatment of osteoporosis and rheumatoid arthritis, Int. J. Pharm. 277 (2004) 73–79. [DOI] [PubMed] [Google Scholar]
  • [421].Quan L, Zhang Y, Dusad A, Ren K, Purdue PE, Goldring SR, Wang D, The evaluation of the therapeutic efficacy and side effects of a macromolecular dexamethasone prodrug in the collagen-induced arthritis mouse model, Pharm. Res. 33 (2016) 186–193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [422].Wei X, Wu J, Zhao G, Galdamez J, Lele SM, Wang X, Liu Y, Soni DM, Purdue PE, Mikuls TR, Goldring SR, Wang D, Development of a Janus kinase inhibitor prodrug for the treatment of rheumatoid arthritis, Mol. Pharm. 15 (2018) 3456–3467. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [423].Quan L, Zhang Y, Crielaard BJ, Dusad A, Lele SM, Rijcken CJ, Metsalaar JM, Kostková H, Etrych T, Ulbrich K, Kiessling F, Mikuls TR, Hennink WE, Storm G, Lammers T, Wang D, Nanomedicines for inflammatory arthritis: Head-to-head comparison of glucocorticoid-containing polymers, micelles, and liposomes, ACS Nano 8 (2014) 458–466. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [424].Zhao G, Wei X, Eichele DE, Lele SM, Yang L, Zhang F, Wang D, A macromolecular Janus kinase (JAK) inhibitor prodrug effectively ameliorates dextran sulfate sodium – induced ulcerative colitis in mice, Pharm. Res. 36(4) (2019) 64; doi: 10.1007/s11095-019-2587-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [425].Ren K, Yuan H, Zhang Y, Wei X, Wang D, Macromolecular glucocorticoid prodrug improves the treatment of dextran sulfate sodium-induced mice ulcerative colitis, Clin. Immunol. 160 (2015) 71–81. [DOI] [PubMed] [Google Scholar]
  • [426].Yuan F, Tabor DE, Nelson RK, Yuan H, Zhang Y, Nuxoll J, Bynoté KK, Lele SM, Wang D, Gould KA, A dexamethasone prodrug reduces the renal macrophage response and provides enhanced resolution of established murine lupus nephritis, PLOS One 8(11) (2013) e82483; doi: 10.1371/journal.pone.0081482. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [427].Wei X, Zhao G, Chhonker YS, Averill C, Galdamez J, Purdue PE, Wang X, Fehringer EV, Garvin KL, Goldring SR, Alnouti Y, Wang D, Pharmacokinetics and biodistribution studies of HPMA copolymer conjugates in an aseptic implant loosening mouse model, Mol. Pharm. 14 (2017) 1418–1428. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [428].Lee S, Stubelius A, Hamelmann N, Tran V, Almutari A, Inflmmation-responsive drug-cenjugated dextran nanoparticles ebhance anti-inflammatory drug efficacy, ACS Appl. Mater. Interfaces 10 (2018) 40378–40387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [429].Liu X-M, Quan L-D, Tan J, Laquer FC, Ciborowski R, Wang D, Synthesis of click PEG-dexamethasone conjugates for the treatment of rheumatoid arthritis, Biomacromolecules 11 (2010) 2621–2628. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [430].Luo T, Loira-Pastoriza C, Patil HP, Ucakar B, Muccioli GG, 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, J. Control. Release 239 (2016) 62–71. [DOI] [PubMed] [Google Scholar]

RESOURCES