Next Article in Journal
The Effects of Vitamin D on Preventing Hyperglycemia and a Novel Approach to Its Treatment
Previous Article in Journal
The Kappa Opioid Receptor: Candidate Pharmacotherapeutic Target for Multiple Sclerosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nanoscale Self-Assemblies from Amphiphilic Block Copolymers as Proficient Templates in Drug Delivery

by
Dhruvi Patel
1,2,
Ketan Kuperkar
2,*,
Shin-ichi Yusa
3 and
Pratap Bahadur
4
1
School of Civil and Environmental Engineering, Cornell University, Ithaca, NY 14850, USA
2
Department of Chemistry, Sardar Vallabhbhai National Institute of Technology (SVNIT), Ichchhanath, Surat 395007, India
3
Department of Applied Chemistry, Graduate School of Engineering, University of Hyogo (UH), 2167 Shosha, Himeji 671-2280, Japan
4
Department of Chemistry, Veer Narmad South Gujarat University (VNSGU), Surat 395007, India
*
Author to whom correspondence should be addressed.
Drugs Drug Candidates 2023, 2(4), 898-922; https://doi.org/10.3390/ddc2040045
Submission received: 10 October 2023 / Revised: 7 November 2023 / Accepted: 8 November 2023 / Published: 22 November 2023
(This article belongs to the Section Preclinical Research)

Abstract

:
This review article emphasizes the current enlargements in the formation and properties of the various nanostructured aggregates resulting from the self-assembly of a variety of block copolymers (BCPs) in an aqueous solution. The development of the different polymerization techniques which produce polymers with a desired predetermined molecular weight and low polydispersity is investigated with regard to their technological and biomedical applications; in particular, their applications as vehicles for drug delivery systems are considered. The solution behavior of amphiphilic BCPs and double-hydrophilic block copolymers (DHBCs), with one or both blocks being responsive to any stimulus, is discussed. Polyion complex micelles (PICMs)/polymersomes obtained from the electrostatic interaction of a polyelectrolyte-neutral BCP with oppositely charged species are also detailed. Lastly, polymerization-induced self-assembly (PISA), which forms nanoscale micellar aggregates with controlled size/shape/surface functionality, and the crystallization-driven self-assembly of semicrystalline BCPs facilitated when one block of the BCP is crystallizable, are also revealed. The scalability of the copolymeric micelles in the drug delivery systems and pharmaceutical formations that are currently being used in clinical trials, research, or preclinical testing is emphasized as these micelles could be used in the future to create novel nanomedicines. The updated literature and the future perspectives of BCP self-assembly are considered.

1. Introduction

As with surface-active agents, which possess dual moieties that behave differently in water and thus adsorb onto interfaces and self-assemble to form nanoscale micelles of different sizes and shapes, block and graft copolymers contain incompatible polymer-size moieties that behave differently in selective solvents and form core–shell micelles and polymersomes. Depending on the polymeric blocks, which have a varied structure and chemical composition and a wide range of polar and non-polar entities linked together, these block copolymers (BCPs) impart unique solid-state and solution properties [1,2].
Self-assembly can be induced in molecularly dissolved polymers with different hydrophilic groups that have one block that is responsive to a stimulus which transforms it and makes it hydrophobic. Furthermore, such copolymers with a hydrophilic polyelectrolyte block may also assemble into ion complex micelles in the presence of an oppositely charged polymer. Also, self-assembly can take place during polymerization. This review focuses on the polymeric nanostructures formed from the self-assembled BCPs that serve as efficient and effective drug carrier candidates and have become an emerging area of research interest from the perspective of drug delivery applications. Unlike polyblends, which are used in the recycling of modified plastics, nanocomposites, and interpenetrating polymer networks and are physical mixtures of two incompatible polymers that undergo a phase separation at the macroscopic scale, these BCPs undergo only a micro-heterogeneous separation due to the covalent bonding that connects the different blocks [3,4]. The varieties of microdomains formed by BCPs in a solid state are highly useful as thermoplastic elastomers and have acquired relevant industrial importance (not discussed in this review) [5]. In addition, the BCPs in a liquid state show unique solution characteristics, which make them versatile candidates in pharmaceutical formulations, catalysis, polyblends, detergency, et cetera [6,7].
The most common BCPs, which have been commercially available for several decades, are ethylene oxide (EO)-propylene oxide (PO)-based BCPs. These have been sold under their BASF trade names as linear triblock copolymers (Pluronics®) and star-block copolymers (Tetronics®). These nano-ionic amphiphilic BCPs are available with varying molecular characteristics and demonstrate superior surface activity and micelle formation and are analogous to poly(ethylene oxide) (PEO, also called polyoxyethylene and poly(ethylene glycol)) condensate-type conventional nonionic surfactants with different nonpolar (lipophilic) parts, such as commercial Triton®, Brij®, Tween®, Soluplus®, Solutol® HS15, Cremophor® EL, Gelucire®, Akypo®, et cetera [8,9].
Recently, polymeric-based nanomaterials have exhibited wonderful properties due to their fabrication ability using various technologies. These have been of immense interest as nanocarriers in the design of pharmaceuticals and in the refining of drug delivery systems; therefore, they offer a suitable tool for the release of hydrophobic drugs and bioactive agents. Polymeric nanocarriers for drug delivery applications include nanogels, nanocapsules, and polymer-coated liposomes (and inorganic nanoparticles); self-assemblies from amphiphilic macromolecules include hydrophobically modified polymers (particularly polysaccharides) and graft/block copolymers. Among these nanocarriers, polymer micelles (PMs) and polymersomes from BCP self-assembly have been of the highest interest due to the ease of fabrication into the desired size/shape/charge, complete characterization, and the tunable properties for drug delivery applications [10,11]. Consequently, in the past few years, BCPs have significantly influenced the drug encapsulation capability and delivery efficacy; so, they are very important in the sustained delivery of drugs due to the core–shell morphology of the aggregates, the excellent thermo-reversible rheological behavior, and the bio-adhesive properties that demonstrate a noteworthy scope [12,13]. There have been several reviews on biomedically important PMs in the literature, but the majority of them focus on their potential applications as nano-vehicles for cancer therapy [14,15,16]. However, less attention is paid to presenting the different aspects of the polymer and colloid chemistry. Therefore, the current review attempts to emphasize the new advances and developments in PMs from a variety of strategies, using different amphiphilic multiphase systems and using blocks from different monomers with varying molecular characteristics and structural features. In giving an account in this direction, the present review, unlike the other reported works, focuses on the self-assembly of amphiphilic BCPs, double-hydrophilic block copolymers (DHBCs) with various stimuli applied, polyion complex micelles (PICMs), polymerization-induced self-assembly (PISA), and crystallization-driven self-assembly (CDSA) in an aqueous solution under one single roof; this will offer an insight into the significance of the copolymeric micellar entities for the desired potential applications.

1.1. Structural Design

Due to the presence of distinct moieties, BCPs display a multiphase system that has been vitally important for a few decades and has aroused considerable research interest among chemists, physicists, biologists, and chemical engineers [17,18]. Figure 1 illustrates the different structures of such smart BCPs, which consist of at least two or more polymeric blocks that are often incompatibly arranged in particular architectures, such as those with a linear, cyclic, or star-like structure, et cetera [19,20].

1.2. Updated Synthesis Route

As the incompatible macromolecular blocks show a variety of microphase domains in solid-state and core–shell micelles and polymersomes in selective solvents (a solvent which is good for one of the blocks), BCPs with the desired structure and molecular characteristics (mol. wt. and % blocks in low polydispersity) can be synthesized using advanced polymerization techniques.
Free radical polymerization is a commonly used synthesis approach. However, owing to the uncontrolled polymerization mechanism, the manufactured polymers are extremely polydispersive, with uncontrolled molecular weight. The living anionic polymerization route first discovered by Szwarc in the 1950s is often followed for a well-established BCP synthesis with a narrow molecular weight distribution or low polydispersity [21]. Thus, several BCPs were prepared, and their self-assembly was investigated in aqueous and non-aqueous solvents. However, this too can be difficult due to the limitation of the fact that only a few monomers can undergo copolymerization to obtain the desired BCPs.
The controlled radical polymerization (CRP) technique has proven to be highly beneficial for the convenient preparation of BCPs from a wide variety of hydrophilic (charged or neutral) and hydrophobic monomers with definite architecture and predetermined molecular weight with very low polydispersity (~1). The commonly known strategies for CRP are atom transfer radical polymerization (ATRP), reversible addition–fragmentation chain transfer polymerization (RAFT) and nitroxide-mediated polymerization (NMP). The RAFT method produces a polymer with a fine molecular weight distribution [22], while ATRP produces the BCPs within an inclusive temperature series [23,24]. NMP uses a nitroxide initiator to produce BCPs with low polydispersity and well-controlled stereochemistry [25,26]. Thus, the literature studies have reviewed multiple synthesis routes, such as tuning polymerization by macroinitiators or ‘living’ free radical polymerization, to produce BCPs [25,27]. Such advances in polymerization synthesis have made BCPs technologically important materials as dispersants and solubilizers for hydrophobic dyes/drugs/pesticides/perfumes; in the fabrication of mesoporous materials; as compatibilizers for two polymers; in the synthesis of nanoparticles and reservoirs for drug delivery systems; and so on. However, the selection of the above-mentioned synthesis technique depends on the structure of the desired copolymer, e.g., diblock, triblock, multiblock, star-shaped, et cetera. The updates on the different controlled polymerization techniques are detailed in several reviews [28].

2. Types of BCPs

2.1. Hydrophilic BCPs

Hydrophilic polymers are easily dissolved in water and cannot form micelles easily due to a water-loving tendency. Double-hydrophilic block copolymers (DHBCs) represent a new class of switchable water-soluble amphiphiles and have been of great interest due to their propensity to undergo self-assembly into varied micellar morphologies when one of the blocks is switched from hydrophobic to hydrophilic because of stimuli such as temperature, pH, or the presence of some additives [29,30]. DHBCs are made of neutral–neutral, neutral–polyelectrolyte, or polyelectrolyte–polyelectrolyte blocks and remain molecularly dissolved in aqueous media. This enables them to provide a new scope of applications in drug carrier systems, gene therapy, crystal growth, colloid synthesis, and desalination membranes [31,32,33].
A few examples of hydrophilic blocks based on several water-soluble polymers, both neutral and charged and from natural sources or synthetically produced, are mentioned below (Figure 2a). (i) The naturally sourced blocks are, for example, hydroxypropyl cellulose, chitosan, alginate, sulfated polysaccharides, and poly(amino acids) (or polypeptides), such as poly(L-glutamic acid), poly(L-aspartic acid), poly(L-lysine), and poly(L-histidine). (ii) The synthetic neutral polymers are poly(ethylene oxide), polyvinylpyrrolidone, polyvinyl alcohol, poly(N-isopropylacrylamide), polyvinylcaprolactam, and various polyelectrolytes (anionic, cationic or zwitterionic), such as polyacrylates, poly(vinylpyridinium chloride), poly(diallyldimethylammonium chloride), poly(styrene sulfonate), et cetera.

2.2. Hydrophobic BCPs

Hydrophobic polymers, being nonpolar, are a bit too rigid to dissolve in the aqueous medium. They form micelles at a very low temperature and have a tendency to undergo phase separation [12,34,35,36]. A few examples of hydrophobic blocks are poly(propylene oxide) (PPO), poly(butylene oxide) (PBO), poly(lactic acid) (PAA), poly(caprolactone) (PCL), poly(butadiene) (PB), poly(styrene) (PS), poly(methylacrylate) (PMA), poly(dimethyl siloxane), poly(vinyl pyridine), et cetera (Figure 2b).
The DHBCs with two distinct hydrophilic blocks remain molecularly dissolved in water. Conversely, if one of the blocks is responsive to any external stimulus, then that block becomes hydrophobic in nature and the DHBC turns to amphiphilic BCP, which would then self-assemble to form core–shell polymer micelles. Such DHBCs can form “stimuli-responsive” or “smart” schizophrenic copolymeric micelles (with a reverse core and shell arrangement), based on which the two blocks in the BCPs have to be hydrophobic (incompatible with water or a given solvent), as demonstrated in Figure 3. Schizophrenic micelles form a water-soluble diblock polymer. Their ability to self-assemble into different structures in response to the applied external stimuli makes them ideal precursors in sensors, actuators, and other thermo-optical devices [33,34,35,36].
Thus, in diluted aqueous solution, AB type DHBC molecules can self-assemble to form two different micelles, one with a core of the hydrophobized A block and a shell of the hydrophilic B block, and the other with a core of the hydrophobized B block and a shell of the hydrophilic A block. The schizophrenic micelles can be switched from conventional to the reverse and vice versa by changing the temperature, solution pH, ionic strength, or solvent composition or in the presence of additives. This type of self-assembly leads to schizophrenic morphological features, as illustrated in Figure 3. Prof. Armes and colleagues, as well as a few others, have reported on extensive investigations of water-soluble diblock copolymers that exhibit so-called “schizophrenic” characteristics [33,34,35]. Mohammad et al. developed the schizophrenic behavior of a water-soluble diblock polymer poly(methacrylic acid-block-N-isopropyl acrylamide) to fabricate a thermo-optical device [36].
Due to their structural fragility, the practical applicability of such micelles is virtually never acknowledged. By utilizing a bifunctional cross-linker and the reactive functional groups of polymer chains, the micellar core and corona are cross-linked to increase their stability. This modulation is represented well in Figure 4.
Additionally, PMs may be produced by electrostatically interacting oppositely charged polymers into a polyelectrolyte block to make it hydrophobic. These are referred to as polyion complex micelles (PICMs) or PIC polymersomes and are discussed later.

2.3. Stimuli-Responsive Block Copolymers (SRPs)

“Stimuli-responsive”, “smart”, or “intelligent” polymers are water-soluble smart macromolecules that respond to varied stimuli, such as (i) physical, viz., temperature, solvents, light, magnet, and ultrasound; (ii) chemical, viz., reactants, redox conditions, and pH; and (iii) biological, viz., enzymes and glucose (Figure 5). These are of immense interest in biomedical applications such as drug delivery, biosensors, tissue engineering, and self-heating materials [37,38].
Investigations of SRPs have advanced in the past two decades [39,40,41]. These polymers can undergo phase transitions or exhibit configurational tuning in response to external stimuli such as pH, temperature, light, electric field, chemicals, magnetism, ionic strength, et cetera. These can cause certain physical or chemical changes in these polymers, which can modify their solubility, surface properties, sol–gel transition, and other properties. Noncovalent forces like electrostatic interactions, hydrophobicity, or hydrogen bonding are frequently implicated in the characteristics of solutions. Such responsive polymers are extensively used in separation science, water treatment, water-borne coatings, recyclable catalysis, and oil recovery. In the case of BCPs, one or more blocks is responsive to any of the stimuli and can behave either hydrophilically or hydrophobically. Several studies have presented the use of such stimuli-responsive polymeric micelles for nano-cargos in drug delivery. Additionally, reports have shown the significance of the dual responsiveness or even the multi-responsiveness of BCPs [37,38,39,40,41,42,43].
(i)
Temperature responsiveness
Thermo-responsive BCPs are the most extensively deliberated responsive polymers due to their distinctive temperature-dependent sol–gel transition properties. Furthermore, these can be tuned to display the sol–gel transition at a desired temperature in the presence of various additives or in the presence of a trace amount of another monomer. A number of literature studies have demonstrated the synthesis and self-assembly behavior of temperature-responsive BCPs in an aqueous solution environment for prospective biomedical applications [4,5,44,45,46,47]. A few thermo-responsive homopolymers with their lower critical solution temperature (LCST) values are shown in Figure 6.
Pluronics® and Tetronics® BCPs have shown remarkable thermo-responsive micellization and gelation in aqueous solutions. Several copolymers with varied molecular characteristics as single or mixed surfactants have been extensively studied for their micellization and their rich phase behavior; several distinct liquid crystalline phases with the features of soft and hard gels have been shown; micellization and micellar growth/shape transitions have been investigated by theoretical modelling and a range of experimental methodologies [38,39,40]. The self-assembly of the PEO-containing hydrophilic block and different hydrophobic blocks, such as polybutylene oxide, polystyrene oxide, polybutadiene, polycaprolactone, polydimethylsiloxane, in an aqueous solution have been investigated [2]. The other extensively examined thermo-responsive block copolymer with hydrophilic moiety is the PNIPAM block with different hydrophilic blocks. These studies report on the phase behavior and micelle formation and the micro-/nanogels using a variety of instrumental techniques. The development of PNIPAM-based thermally triggered BCP micelles as cargo for the sustainable release of drugs was scrutinized worldwide and PNIPAM-based BCP micelles as perspective candidates for drug delivery; i.e., PNIPAM-based BCP micelles form a hydrophilic shell below the LCST and a hydrophobic core above the LCST. (PChM-PNIPAM), constructed with poly(cholesteryl 6-methacryloyloxy hexanoate) (PChM), and PNIPAM blocks were designed as the hydrophobic core and hydrophilic shell of the micelles, respectively, with increasing temperature [48,49,50,51].
(ii)
pH responsiveness
pH-responsive BCPs bring about changes in the unique density, chain conformation, solubility, and configuration due to the presence of functional groups in the polymer that can be ionized upon a slight variation in the changing solution pH [45,46,47]. These polymers are enclosed by acidic or basic groups along the chain or at the end of the chain. The charges produced are assumed to induce the electrostatic repulsion among the polymer chains, leading to the alteration in hydrodynamic volume, which is capable of producing the flocculation, chain collapse–extension, and precipitation of the homopolymers. It has proven possible to create random pH-responsive copolymers using traditional free radical polymerization. The pH-responsive BCPs are of significance as their pH-directed self-assembly can act as a suitable vehicle for drug delivery systems. Figure 7 shows the types and examples of a few well-known pH-responsive homopolymers.
Interestingly, block copolymers with one pH-responsive block can self-organize into varied micelle shapes in a solution with subtle changes in pH and are being considered for numerous potential applications. Any alterations in the pH may lead to the protonation or deprotonation of functional groups existing in the polymers, leading to the various morphologies, like spherical micelles, vesicles, microgels, et cetera.
(iii)
Light responsiveness
Light has been recognized as a smart external stimulus for creating responsive drug delivery systems. An enormous assortment of such photo-responsive systems, operating at a particular wavelength to accomplish on-demand drug release, has been reported. In particular, light-responsive BCPs have been of particular interest because of the specific interval and site controls that make them useful for prospective applications in controlled drug delivery systems and in forming self-healing, reversible wettability materials [48,49,50,51]. Also, such BCP-based nanocarriers are considered for the design of non-toxic treatment regimens that deal with spatiotemporal control beyond the release of captured therapeutic cargoes. There are two main types of light-responsive polymers: (i) reversible light-responsive polymers that are usually polymers with photochromic entities with reversible structural change, i.e., structural isomerization, e.g., organic dyes, such as azobenzene and spiropyran and (ii) irreversible light-responsive polymers that are usually polymers with photochromic units and irreversible chemical bond breakage, e.g., o-nitrobenzyl groups, pyrene groups, and coumarin groups (Figure 8). The UV irradiation leads to the cleavage of the photochromic moieties and converts the hydrophobic block to a hydrophilic block. Such apparent structural changes in the light-responsive BCPs may disrupt the self-assembled micelles and release the encapsulated hydrophobic cargo from the micellar core. In order to control the delivery of the hypoxia-activated bio-reductive prodrug tirapazamine for the treatment of metastatic breast cancer using hypoxia-boosted phototherapies, Yuanyuan et al. established near-infrared (NIR) light-decomposable nanomicelles made of PEGylated cypate and mPEG-polylactic acid (mPEG2k-PLA2k) [52,53].
(iv)
Magnetic responsiveness
Among the varied inorganic nanomaterials that are reactive towards the exterior magnetic field, iron oxide (Fe3O4) nanoparticles have gained keen interest in the field of biomedical applications owing to their high surface area to volume, biocompatibility, low toxicity, and easy synthesis [54,55]. The BCPs functionalize the Fe3O4 nanoparticles to encouragingly improve the dispersibility of the nanoparticles in the blood circulation duration [56].
(v)
Multi-responsiveness
Recently, multi-responsive BCPs have also gained considerable attention as they respond to more than one stimulus (temperature, pH, redox, light, and magnetic field) concurrently and thereby exhibit formed micellar morphologies [38,45,57]. These multi-responsive BCPs are made by fusing two or more distinct monomers, each of which has a unique reaction to stimuli. For example, the poly[(2-dimethylaminoethyl) methacrylate] (PDMA) and poly(2-N-morpholinoethyl) methacrylate (PMEMA) polymers are both temperature- and pH-responsive. Multi-stimuli-responsive BCPs are expedient in their ability to release drugs according to their precise biological microenvironment, such as the composition of cells, extracellular matrix (ECM) components, soluble factors, physical forces (e.g., fluid flow and mechanical stress), et cetera [43,58].

3. Physicochemical Features of the Self-Assembly in BCPs

With the possibility to synthesize a variety of BCPs from several different kinds of monomers (neutral or ionic) that can lead to a material with well-defined features and tunable properties, researchers are actively engaged in optimizing self-assembled structures from non-toxic micellar aggregates with a great drug-loading capacity and a decent release profile in the optimized release rate at the targeted organ/tissue/cell. In the following section, we discuss the features of polymer self-assembly with regard to micelles and polymersomes. However, these nanoscale self-assemblies formed by BCPs depend on the structure, chemical nature, and molecular characteristics of the copolymers and the solution conditions.
In the following sections, we describe the self-assembled structures formed from (i) amphiphilic block/graft copolymers, (ii) responsive double-hydrophilic block copolymers, (iii) polyion complex assemblies from polyion-neutral BCPs and oppositely charged substances, and (iv) polymerization-induced self-assembly. The focus is on delivering a brief synopsis of all these aspects and their advances in recent years, along with their possible applications in drug delivery systems.

3.1. Polymeric Micelles (PMs)

Unlike classic conventional micelles, the critical micelle concentration (CMC) of BCPs is often very low (<10 mg/L) and not accurately determined. They form kinetically stable nanoscale core–shell architectures that exist in nanoscale sizes (~10–100 nm) and have a fairly narrow size distribution, which enables them to accommodate nonpolar/weakly polar bioactive substances. Furthermore, PMs are less expected to break down and expedite the efficient drug administration over a long period of time. As a consequence, PMs have grabbed most of the research interest as potential cargo for therapeutic applications [20,59,60]. However, the solubilizate site relies on the chemical nature/structure of the solubilizate as well as that of the BCP and its micelle size/shape. As mentioned above, the core signifies the “cargo” for the encapsulation of various therapeutic reagents, where the hydrophilic shell confirms that the micelles remain in a discrete state, thereby minimizing the unwanted drug interactions with cells. Thus, PMs provide promising vectors for the carrying and delivery of drugs and allow the formation of multipurpose roles for the expansion of innovative therapies for incurable diseases [2,61,62].
Additionally, PMs can attain different features; with these features, they can be transformed into varied nanoparticles that remain undissociated in extreme dilutions via cross-linking of the core or shell, as shown in Figure 9. Cross-linking instigates covalent/non-covalent bonding and produces dynamic micellar systems [63,64,65]. Cross-linking raises the solubility by several orders of magnitude without diminishing the drug-loading capability. Furthermore, cross-linking has an effect on the permeability of the shell, which changes the pace at which the drug is released over time.
An overall methodology aims to introduce some functional groups/substituents in the BCPs which ease the cross-linking of the core or shell. Also, the shell permeability can be tuned by changing the degree of cross-linking, which determines the drug loading and the release mechanism of the polymeric micelles. A study by Yilmaz et al. showed a strong influence upon the doxorubicin (DOX) release rate after the cross-linking of the core. It offered efficient trapping of DOX molecules, which later became discharged in a controlled manner. For pH-responsive drug transport, Jie et al. created core-cross-linked PMs which were tunable with BCP with imidazole [66]. For DOX-based drug delivery, Jong et al. produced cross-linked PM based on poly(ethylene oxide)-b-poly(methacrylic acid) (PEO-b-PMA) with ionic cores and divalent metal cations. Because of the carboxylic group protonation in the micelles’ cores, the DOX-loaded polymer micelles displayed observable pH-sensitive action with faster relief of DOX in acidic environments. These micelles also revealed a significant cytotoxicity against human A2780 ovarian cancer cells [67]. PMs can be biocompatible and have the capability to accrue in tumors via enhanced permeability and retention (EPR). Their functionalization through surface modification with specific ligands and the introduction of stimuli-sensitive groups allows specific targeting and release.
The literature reports have shown the formation of various types of PMs that become cross-linked via –SS– linkages that aid the dual purpose: firstly, they avoid impulsive micelle detachment in the blood, and secondly, they allow the release of the entire drug. This is verified in the study of Li and co-workers, where the PMs demonstrated decent stability and a rapid release compared to extensive dilution [68]. Alternatively, the surface functionalization of BCPs can be accomplished using bio-responsive linkers. Such systems are proposed to tune the hydrophilic character of the functional groups into a hydrophobic one, or vice versa [69]. An FDA-approved polymer mPEG-b-PCL, with outstanding biodegradability and biocompatibility in the area of drug delivery, has been extensively used. Chemical modifications, like changing different groups, attaching drugs at the terminal end, and host–guest interaction, have evolved to increase drug loading efficacy and release. The drug loading amount of (DOX) by the transporter was enhanced as a result of the interaction, according to the experimental data from Yan et al.’s development of an amphiphilic poly(caprolactone) of pendant carbamic acid benzyl ester [70]. The BCP of Poly(caprolactone) modified with Phenylboronic acid (PBA), prepared by organocatalytic ring-opening polymerization by Wang et al., has a superior capacity to prevent cancer cell proliferation and the subsequent DOX encapsulation [71]. Runliang et al. prepared DOX-loaded micelles from a PBA-modified polymaleic anhydride-F127 polymer for tumor cell treatment [72].
Studies have confirmed the establishment of core–shell–corona micelles when there is a core-forming hydrophobic A block with an outer shell and corona from two incompatible hydrophilic blocks B and C. Such micellar constructs are referred to as ‘onion’ micelles. By synthesizing Poly(styrene-b-acrylic acid-b-ethylene glycol) (PS-b-PAA-b-PEG), an asymmetric triblock copolymer, Bastakoti et al. created pH-triggered micelles with a core–shell–corona assembly with a PS as a hydrophobic core, PAA as an anionic shell, and PEG as a hydrophilic corona [73]. Schluter et al. described rod-like micelles with two diverse half coronas based on polydendrons. Additionally, PS and PEO macromonomers underwent progressive ring-opening metathesis polymerization to produce cylindrical Janus micelles. Amphiphilic BCPs with hydrophobic blocks with high glass transition temperatures form stable spherical and cylindrical micellar aggregates with “glassy” cores and vesicles with glassy wall interiors in aqueous dispersions. These varied nanoscale micellar geometries are of potential interest in drug delivery applications, where stable and mechanically robust micelles are anticipated. These equilibrium structures, however, do not include the interchange of monomers from micelles like those found in surfactants and BCPs [74,75]. Lyotropic liquid crystals exhibit a liquid crystalline phase in solution, where a liquid crystalline core is formed with a hydrophilic shell. Additionally, studies have shown the development of anionic polymerized crew-cut micelles, which feature enormous insoluble blocks as the core and small soluble blocks as the corona. These PMs can be functionalized through the chemical attachment of a ligand (a bioactive substance, drug, protein, et cetera) at the micelle surface. These polymers can show manifold morphologies, such as spheres, rods, vesicles, lamellae, hybrid micelles, and several other structures, with narrow size distribution, and high stability in water. However, these nanoaggregate morphologies depend on how incompatible the blocks are and what their chemical nature is and on the mol wt. and wt. percent of the blocks. They also depend on the copolymer type, such as AB (di-block), ABA, BAB, and ABC (tri-blocks), and can be roughly predicted by the Israelachvili packing parameter, as described in the published works [2,54]. However, the size of the micelles relies on the preparation method, which signifies the minimum intermicellar exchange in water [12,76,77]. All these varied micellar structures have potential as versatile carriers for drug delivery.

3.2. Polymersomes

Usually, amphiphilic BCPs aggregate to form core–shell micelles of different microstructural features (spheroidal, rod- or wormlike structures). In addition, they exhibit artificial analogues of liposomes (bi-layered vesicles from low molecular weight phospholipids), which are often referred to as polymersomes. [78,79] These soft nanoparticles comprise a spherical aqueous environment enclosed by a bilayer membrane that is composed of a hydrated hydrophilic inner corona and an outer shield made of the hydrophobic middle part. These entities are well known due to their excellent robustness, high stability, longer blood circulation time, structural design, chemical versatility, and ease of surface functionalization, which make them attractive candidates in pharmaceuticals and the medicinal field [80,81]. Drugs, enzymes, proteins, peptides, and fragments of nucleic acids can all be solubilized by polymersomes inside their aqueous core area or membrane region [82,83]. Because of this, polymersomes have been developed into highly intriguing materials that can be used as nano-reservoirs in drug delivery systems (Figure 10).
Discher et al. initially noticed polymersomes in both theoretical and experimental investigations and explained various mechanisms for polymersome preparations [80,81]. Based on BCP chemistry, Letchford et al. presented a novel class of synthetic thin-shelled capsules with a water penetrability that was at least 10 times greater than that of conventional phospholipid bilayers [82]. Also, the polymersomes with thick and tough membranes are able to encapsulate hydrophilic and hydrophobic molecules and deliver higher in vitro and in vivo stability. Robertson et al. have examined the polymersomes which are responsive to pH, light, gas, glucose, enzyme, oxidation, reduction, and temperature, with various examples presented for different drug delivery and biological applications [83]. Pawar et al. developed various methodologies for the fabrication of polymersomes and explained how polymersomes can be applied in multipurpose biomedical research [84]. Mohammadi et al. described the biocompatibility of polymersomes for cancer theranostic with multifunctional nanomedicines [85].

3.3. Ethylene Oxide (EO)-Propylene Oxide (PO)-Based BCP Micelles

BCP has two or more different monomeric blocks in diverse structures and compositions, which give them distinctive solid-state and solution properties and allow them to show their usability in applications involving catalysis, detergency, dispersants, emulsifiers, pharmaceutical vehicles, et cetera. Among the amphiphilic BCPs, Pluronics® (poloxamers) and Tetronics® (poloxamines), the commercially available FDA-approved EO-PO block copolymers, have been extensively examined [86,87,88]. These polymeric surfactants are marketed as triblock and 4-armed branched structures with varying PPO mol wt and %PEO. Pluronics® are poly(ethylene oxide)-poly(propylene oxide)-poly(ethylene oxide) (PEO-PPO-PEO)-type amphiphilic triblock block copolymers, whereas Tetronics® are star-shaped poly(ethylene oxide) (EO)-poly(propylene oxide) (PO)-based amphiphilic block copolymers (BCPs); they have ethylene diamine as the central unit, to which four PO blocks are attached to nitrogen atoms, while the other end of the PO is linked to the EO block. Thus, their self-assembly leads to different sizes/shapes of core–shell micelles and polymersomes, liquid crystalline structures, and thermo-reversible gels, depending on their concentration. Furthermore, the self-assembly is strongly dependent on temperature and the presence of kosmotropic salts, ionic liquids, surfactants, and polar/nonpolar additives, which have been shown to have a profound effect on the formation properties and behavior of these nanoaggregates. Additionally, these copolymers can be functionalized at the end hydroxyl groups with drugs, bioimaging agents, and other functional polymers, which impart useful and stimuli-responsive properties. There have been several excellent reviews that describe the advances in aggregation and rheological aspects using a variety of experimental techniques, such as scattering and thermal techniques in particular, which detail the size/microstructures of the nanoaggregates and thermodynamics of their assemblies. Pluronics® and Tetronics® are chemically altered or functionalized to utilize various synthetic techniques, giving them improved and desirable qualities as vectors for delivery systems that enhance their properties. At the PEO chains of Pluronics® and Tetronics®, the covalent attachment of any small molecule, polymer block, ligand, or stimuli-responsive molecule changes the micellization, surface activity, solubilization power, and rheological behavior, which can be used to fine-tune the desired features in micelles. Medical diagnostics, imaging, and other applications have all taken advantage of these chemical changes in Pluronics® and Tetronics® [36,89,90,91]. The core–shell micelles with an outer hydrophilic and extremely polar and flexible PEO shell provide stability and stealth properties to these micelles and prevent aggregation and interactions with mononuclear phagocytic systems that lead to the removal of micelles from systemic circulation. These micelles show a high drug-loading capacity and possess superior thermodynamic/kinetic properties and stability, which can be easily tuned further using mixed systems of two copolymers in a suitably designed concentration/mixing ratio. Mixed Pluronic (L61/F127) micelles with loaded doxorubicin have already become the first Pluronic®-based micellar formulation to enter clinical trials as SP1049C (Supratek Pharma Inc., Dorval, QC, Canada). The therapeutic efficiency of these polymeric micelles relies on how these overcome biological barriers and deliver and release the drug at the target site [88,89,90,91,92].

3.4. Polyion Complex Micelles (PICMs)

Polyion complex micelles (PICMs) form when aqueous solutions of two oppositely charged polyelectrolytes are mixed. In addition, strong electrostatic interaction increases entropy because of the counter ion release from the macroions; several other interactions, like hydrogen bonding and hydrophobic interactions, could also contribute. Additionally, stoichiometrically mixing two polyelectrolytes with opposing charges leads to the spontaneous formation of PICMs/polymersomes in water (insoluble in other solvents) due to electrostatic attraction. Due to the reciprocal neutralization of the two diametrically opposed charges, PICMs/polymersomes have no charge. Polyelectrolytes with opposing charges neutralize each other, which causes them to drop their colloidal stability and precipitate out of the solution. If polyelectrolytes can be connected to any hydrophilic and nonionic unit, the precipitation of PICMs/polymersomes can be prevented. The complexes can be water-soluble or insoluble (as coacervates or stable colloidal dispersions); their nature in terms of charged groups, molecular weight, flexibility, et cetera depends on the polyelectrolytes used and the stoichiometric mixing, in addition to the solution conditions (pH, temperature, and the presence of salt or other additives). The effects of ionic strength and pH are significant in the formation of PICs. Usually, low ionic strength permits the complex structure to display thermodynamic equilibrium, whereas high ionic strength shrinks it because of the shielding of the polyelectrolyte charges. The stable colloidal dispersions or insoluble coacervates can be examined by turbidimetric or light-scattering measurements. Smart PICMs/polymersomes have single- or multi-responsive properties against external stimuli. Using random copolymers with pendant quaternary ammonium and sulfonate groups, Shukanta et al. created PICMs using the RAFT technique to create anionic random copolymers from MPC and potassium 3-(methacryloyloxy) propane sulfonate and cationic random copolymers from 2-(methacryloyloxy)ethyl phosphorylcholine with methacroylcholine chloride. Thu et al. created PICMs by directly mixing the two opposing charges of PMPC-block-poly(sodium p-styrenesulfonate) and poly(2-(methacryloyloxy)ethylphosphorylcholine)-block-poly(vinylbenzyl trimethylammonium chloride) [93,94,95,96,97,98,99].
PICs have multiple applications in textiles, ink, and paper industries as binders, in coatings, and as flocculants for water purification, to name a few. In 1949, Fuoss et al. first reported that two oppositely charged polymers formed insoluble precipitates, but it was not until 1965 [100] that Michaels et al., after mixing two polyelectrolytes with opposing charges, studied stable nanosized spherical complexes by adjusting the solution pH, temperature, salt content, and other parameters. By successively adsorbing polyelectrolytes from solution onto a surface, layer-b-layer (LbL) films of interacting polyelectrolytes can be created [101,102,103]. The polyelectrolyte complexes formed in solution and multilayers on surfaces are extensively investigated using a variety of polyelectrolytes with several experimental techniques. A thorough theoretical justification describing counterion condensation on polyelectrolyte complexes has been provided [104,105,106,107].
A polyelectrolyte complex can self-assemble into PICMs when it becomes amphiphilic, which is similar to the way that amphiphilic BCPs behave. DHBCs made of a hydrophilic neutral block–polyelectrolyte block are molecularly dissolved in water. Nevertheless, if the solution of an oppositely charged polymer is added to it, self-assembly happens as a consequence of electrostatic attraction, where a core of complexed oppositely charged polymers is formed. The formation of micelles, viz., polyion complex micelles or polyion complex polymersomes and their morphological features, can be finely tuned based on the charge densities of complexing polyelectrolytes, their mixing composition, and their solution conditions, which enable them to be employed in drug and protein delivery [104,105,106,107,108]. The micelles formed with a complexed core and hydrated shells are PICMs (Figure 11).
In an aqueous environment, a variety of synthetic and natural polyelectrolytes can interact with DHBCs to create PICMs or PIC polymersomes, which are extensively characterized by spectral, scattering, thermal, and microscopic methods. [104,105] Two DHBCs with oppositely charged polyelectrolyte blocks and the same or a different hydrophilic neutral block can interact to form interesting shapes. Additionally, these systems may be developed further by employing DHBCs that respond to inputs. Readers interested in learning more about PICMs might refer to several excellent review articles [106,107].

3.5. Polymerization-Induced Self-Assembly (PISA)

Polymerization-induced self-assembly (PISA) is a versatile approach that associates polymerization and self-assembly in a more concentrated solution for the rational production of concentrated BCP assemblies in a high yield using controlled/living polymerization techniques. It is a straightforward one-pot process with convenient scalability and economic viability. [108,109,110] The synthesis of linear as well as nonlinear (e.g., star-shaped, or branched) BCPs can be performed by PISA. PISA mediated by RAFT using suitable chain transfer agents and a water-miscible monomer has become a powerful technique that provides BCP micelles with a defined morphology, a controlled size, and surface functionality [111,112,113]. In contrast to the conventional approach, PISA effectively produces a higher concentration of nano-assemblies without adding a non-solvent to induce micellization. The first polymeric block and the second monomer that are polymerized in PISA must both be soluble in the same solvent. When the second block is sufficiently lengthy, its solubility is reduced, allowing the resultant BCP to self-assemble in place and produce nanostructures. PISA involves the aqueous dispersion polymerization of a water-miscible monomer till a critical degree of polymerization (DP) transforms it into a water insoluble polymer, which then self-assembles into a core–shell micelle of different morphologies, depending on the molecular weight and nature/composition of the blocks (Figure 12). Here, the macroscopic precipitation is avoided by appropriately controlling the colloid stability through steric stabilization.
PISA can be achieved with comparatively high solid contents (up to 50% wt) and can undergo very high monomer conversions within a short reaction time. Armes and co-workers have reviewed the polymer synthesis by PISA (RAFT) in polar or non-polar media (including supercritical CO2 and ionic liquids) and non-aqueous media [114,115,116]. Thus, PISA enables the in situ production of self-assembled nano-objects with various morphologies ranging from sphere to rod to polymersome; finally, a precipitate is formed by modifying the solution conditions. It has become very important for nanocarriers in drug delivery and can achieve better cellular uptake, tracking, and enhanced bioavailability. The synthesis and in situ formation of nano-assemblies of diblock copolymers (PMPC-b-PPFPMA) can be achieved via PISA using RAFT in ethanol as a selective solvent.

3.6. Crystallization-Driven Self-Assembly (CDSA)

Crystallization-driven self-assembly (CDSA) represents a handy methodology for the production of 1D and 2D anisotropic nanostructures with a high level of dimensional control and a tunable core–shell micelle in solution. This approach has shown that semicrystalline BCPs are promising for the development of materials with complex architectures [117,118,119]. This propensity is caused by the core-forming block crystallizing after being heated over the glass transition temperature and cooled in a weak solvent. As a result, polymers that are easily accessible by coupled ring-opening and RAFT polymerizations can be used to create nanostructures with anisotropic morphologies and excellent solution stability. The self-assembly process is facilitated when one block of the BCP is crystallizable. For a copolymer with a crystallizable insoluble block, the crystal packing forces play an important role in determining the morphologies of the different core–shell micelles formed. A few recent reviews on such self-assembly of semicrystalline block copolymers have been published [120,121,122,123,124,125].

4. Applications of BCPs

The applications of BCP-derived nanopatterns in real-life technologies are of great significance in several areas. BCPs are extensively used in the advancing field of nanopatterning, including next-generation lithography for semiconductors. The capabilities constituting the BCP toolbox have grown over the ages and have delivered numerous extraordinary achievements. However, a few challenges remain to be overcome in ensuring its integrity in the development process. The nanoscale self-assembly of BCPs like micelles and vesicles has received great attention in the field of drug delivery applications due to its effective control of morphology, surface chemistry, and responsiveness [126,127,128,129,130,131,132,133,134,135,136,137,138]. Organic dyes are only partly water soluble, which is a major challenge for dye and paint industries. BCPs tend to solubilize such hydrophobic moieties in their micellar core in a manner which depends on the hydrophobicity of the BCPs. Several studies have reported the ease in which dye solubilization can be influenced in the presence of external stimuli or additives [139,140,141,142]. Due to their ideal drug loading and release features, lengthy shelf life, and low toxicity, smart BCPs are particularly successful in controlled drug delivery applications, as well as for treatment and diagnostics. In particular, polymersomes serve as successful carriers for drug delivery systems [143,144,145,146]. Consequently, due to their high functionality, these BCPs might be used in a variety of medical devices, including nanoreactors, “semi-artificial” enzymes, and biodevices with blood-compatible surface treatments [147,148,149,150].
Polymeric micelles have emerged as promising nanocarriers due to their ability to control the drug release profile and the release at the target site; they have enhanced permeability and retention and enhance the solubility of hydrophobic bioactive substances, particularly anticancer drugs. They can build up in the tumor microenvironment because of the increased permeability and retention impact of their nanoscale size. These micelles from biocompatible and nontoxic amphiphilic polymers are anticipated to be a fruitful treatment option for drug and gene delivery (DNA/siRNA) in various cancer therapeutic strategies, like the crossing of the blood–brain barrier, photothermal therapy, gene therapy, and immunotherapy. The majority of polymeric micelles are undergoing clinical trials for various cancer therapies. Here is a description of both the commercially available polymeric micelles and the micelles presently undergoing clinical trials. The paclitaxel (PTX)-loaded polymeric micelles (NK105) have improved clinical efficacy against ovarian, lung, neck, breast, and brain cancers. Clinical testing for breast, lung, and pancreatic cancer has been approved for the FDA-approved “Abraxane®” albumin nanoparticle. Long-circulating polymeric micelles with an amide linkage and a hydrophobic PEG-b-poly (aspartic acid) copolymer backbone make up NK105. PEG-PLA copolymer-based micelles were authorized in 2007 in the Republic of Korea. Genexol®-PM, a PM formulation containing paclitaxel, is offered by Samyang Co. in Seoul, Korea. Genexol®-PM has shown increased tumor tissue concentration and anticancer activity when compared to paclitaxel. Genexol®-PM has been approved for use in the treatment of breast, lung, and ovarian cancer. It is commercially available in India under the name Nanoxel®, by Dabur Pharma Ltd., New Delhi, India. Doxorubicin (DOX)-containing micelles (NK911) have included the irreversible binding of DOX with a polymer, such as N-(2-hydroxypropyl) methacrylamide (HPMA), and an enzyme-sensitive spacer of glycyl-phenylalanyl-leucyl-glycine reached Phase-I clinical trials. Myocet and Doxil/Cealyx (PEGylated) were approved by the USFDA for ovarian and breast cancer. PEG-b-poly (α, β-aspartic acid) was verified by preclinical studies in cancer mouse models under the name NK911. SN-38-loaded polymeric micelles (NK012), which contain a micellar formulation comprising 7-ethyl-10-hydroxy-CPT (SN-38) attached through esterification via PEG-PLGA, have a greater half-life and superior bioactivity level at the tumor level in mouse models of brain cancer and renal, colon, pancreatic, and gastric cancer. In clinical studies for patients with esophageal and gastroesophageal cancer, the Kabanov research group and Supratek pharma developed SP1049C, Pluronic®-based DOX-loaded polymeric micelles. Pluronic® L16 and Pluronic® F127 (1:8 ratio) were combined to create SP1049C. For the formulation of platinum-based polymeric micelles used to treat diseases such as myeloma and lymphoma and testicular, ovarian, bladder, and lung cancer, cisplatin and oxaliplatin were widely used. In late 2004, the USFDA certified oxaliplatin (oxalato(trans-l-1,2-4 diaminocyclohexane) platinum (II)) as a main therapy for colorectal cancer when used in combination with 5-FU. Modifications were made to a formulation of oxaliplatin and platinum (II) (PtDACH), and a PEG-b-poly(glutamic acid) copolymer developed self-assembled polymeric micelles. Finally, clinical studies have begun for both formulations [151,152,153,154,155,156,157,158,159,160].
Pharmaceutical formulations are generally very complex and not completely understood. An enormous number of BCP-based drug delivery systems are widely explored to understand the drug release from the dosage, with the aim of developing the formulations that ensure safety and efficacy for patients [161,162,163,164,165]. Several different strategies have been engaged to improve the solubility and bioavailability of hydrophobic drugs. The nanoscale self-assemblies (polymer micelles and polymersomes) have been of great interest in recent years as these offer several advantages. The use of cosolvents, hydrotropes, micronization, and supramolecular complexes and the altering of the crystal structure have been adopted for the initial processes. Hydrogels, solid lipid nanoparticles, self-emulsifying dispersions, nano- and microemulsions, and liposomes have been developed over time. Several nanocarriers, such as inorganic metal/metal oxide nanoparticles, particularly AuNPs and magnetic NPs, and mesoporous silica; nanocarbons such as CNTs, graphene, and quantum dots; and organic nanocarriers such as dendrimers and NPs derived from polysaccharides, proteins, and biodegradable polymers have been found to be quite promising for drug encapsulation. Here, we provide the references of the review articles covering all such features [164,165,166].
The application of BCPs in the sensing field is also gaining momentum significantly. In an effort to create sensors with high sensitivity and selectivity, novel transduction methods and manufacturing techniques have been disclosed [167]. Metal oxides, non-oxide inorganics, and carbon-based ordered porous materials exhibit aligned structures and uniform cavities with micro- to meso- to macropore sizes, which lead to a very high surface area. Such features enable their utilization in energy conversion and storage, catalysis, gas capture, and water purification. Apart from the above-mentioned applications of these BCPs, their extensive use in separation and surface coating is anticipated [139,140,141,142,143].

5. Conclusions and Future Perspectives

This review presents a detailed description of different nano-assemblies from block copolymers obtained in tailor-made structures using a variety of hydrophilic/hydrophobic blocks, both charged and uncharged, and it highlights the evolution in the expansion of multifunctional nanoaggregates as vectors for delivery systems. Furthermore, it offers an insight into the nanoscale micellar aggregates formed from amphiphilic block copolymers (BCPs), stimuli-responsive double-hydrophilic block copolymers (DHBCs), and polyion complexes (PICs). Advanced polymerization synthesis techniques have been discussed; these techniques enable us to craft the novel stimuli-responsive BCPs. The formation and characterizations of BCP micelles are explained. However, the drug transport throughout the body and to its targeted delivery site is among the modern challenges. The immediate challenge in BCPs is to design and formulate an effective micellar system that enhances their specific delivery to the targeted site along with an efficacious penetration ability that fights against the infected cells. Also, there are pressing challenges ahead, such as the expense of the regulatory approval for the commercialization of these aspirant BCP-based delivery systems. Accordingly, from the point of view of the industrial pharmaceutical trade, the customized BCPs are novelties. Thus, a judicious design of non-toxic probes in drug delivery systems will improve bioavailability, patient compliance, and therapeutic outcomes. Thus, this review will surely inspire the readers to consider all the BCP features, whose modes of operation can be tailor-made as per the desired needs and applications.

Author Contributions

D.P. is responsible for writing the review paper; K.K. is responsible for conceptualization and for editing this review writing; S.-i.Y. and P.B. directed this review. All authors have read and agreed to the published version of the manuscript.

Funding

This work did not receive any specific grant from any funding agencies in the public, commercial, or not-for-profit sectors.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Acknowledgments

The authors acknowledge Sardar Vallabhbhai National Institute of Technology (SVNIT), Gujarat, India for providing the facilities.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence this review.

References

  1. Riess, G.; Hurtrej, G.; Bahadur, P. Block Copolymers, Encyclopedia of Polymer Science Engineering. Wiley 1985, 2, 324–434. [Google Scholar]
  2. Riess, G. Micellization of block copolymers. Prog. Polym. Sci. 2003, 28, 1107–1170. [Google Scholar] [CrossRef]
  3. Perin, F.; Motta, A.; Maniglio, D. Amphiphilic copolymers in biomedical applications: Synthesis routes and property control. Mater. Sci. Eng. C 2021, 123, 111952–111967. [Google Scholar] [CrossRef] [PubMed]
  4. Li, M.; Coenjarts, C.; Ober, C. Patternable block copolymers. Adv. Polym. Sci. 2005, 190, 183–226. [Google Scholar]
  5. Li, Z.; Lin, Z. Self-assembly of block copolymers for biological applications. Polym. Int. 2021, 71, 366–370. [Google Scholar] [CrossRef]
  6. Hadjichristidis, N.; Pispas, S.; Floudas, G. Block Copolymers: Synthetic Strategies, Physical Properties, and Applications; John Wiley and Sons: Hoboken, NJ, USA, 2003. [Google Scholar]
  7. Lazzari, M.; Tornerio, M. A global view on block copolymers. Polymers 2020, 12, 869. [Google Scholar] [CrossRef] [PubMed]
  8. Alexandridis, P.; Lindman, B. Amphiphilic Block Copolymers: Self-Assembly and Applications; Elsevier: Amsterdam, The Netherlands, 2000. [Google Scholar]
  9. Lu, Y.; Lin, J.; Wang, L.; Zhang, L.; Cai, C. Self-Assembly of Copolymer Micelles: Higher-Level Assembly for Constructing Hierarchical Structure. Chem. Rev. 2020, 120, 4111–4140. [Google Scholar] [CrossRef]
  10. Hu, X.; Xiong, S. Fabrication of Nanodevices Through Block Copolymer Self-Assembly. Front. Nanotechnol. 2022, 4, 762996–763012. [Google Scholar] [CrossRef]
  11. Xi, S.; Zhu, Y.; Lu, J.; Chapmana, W. Block copolymer self-assembly: Melt and solution by molecular density functional theory. J. Chem. Phys. 2022, 156, 054902–054910. [Google Scholar] [CrossRef]
  12. Kuperkar, K.; Patel, D.; Atanase, L.; Bahadur, P. Amphiphilic Block Copolymers: Their Structures, and Self-Assembly to Polymeric Micelles and Polymersomes as Drug Delivery Vehicles. Polymers 2022, 14, 4072. [Google Scholar] [CrossRef]
  13. Cabral, H.; Miyata, K.; Osada, K.; Kataoka, K. Block Copolymer Micelles in Nanomedicine Applications. Chem. Rev. 2018, 118, 6844–6892. [Google Scholar] [CrossRef]
  14. Genevieve, G.; Marie-Helene, D.; Vinayak, P.; Sant, N.; Dusica, M.; Jean-Christophe, L. Block copolymer micelles: Preparation, characterization and application in drug delivery. J. Control. Release 2005, 109, 169–188. [Google Scholar]
  15. Kazunori, K.; Atsushi, H.; Yukio, N. Block copolymer micelles for drug delivery: Design, characterization and biological significance. Adv. Drug Deliv. Rev. 2012, 64, 37–48. [Google Scholar]
  16. Xiong, X.; Binkhathlan, Z.; Molavi, O.; Lavasanifar, A. Amphiphilic block co-polymers: Preparation and application in nanodrug and gene delivery. Acta Biomater. 2012, 8, 2017–2033. [Google Scholar] [CrossRef] [PubMed]
  17. Arotçaréna, M.; Heise, B.; Ishaya, S.; Laschewsky, A. Switching the inside and the outside of aggregates of water-soluble block copolymers with double thermoresponsivity. J. Am. Chem. Soc. 2002, 124, 3787–3793. [Google Scholar] [CrossRef] [PubMed]
  18. Bielawski, C.; Grubbs, R. Living ring-opening metathesis polymerization. Prog. Polym. Sci. 2007, 32, 1–29. [Google Scholar] [CrossRef]
  19. Braunecker, W.; Matyjaszewski, K. Controlled/living radical polymerization: Features, developments and perspectives. Prog. Polym. Sci. 2007, 32, 93–146. [Google Scholar] [CrossRef]
  20. Kataoka, K.; Harada, A.; Nagasaki, Y. Block copolymer micelles for drug delivery: Design, characterization and biological significance. Adv. Drug Deliv. Rev. 2001, 47, 113–131. [Google Scholar] [CrossRef]
  21. Zhao, D.; Feng, J.; Huo, Q.; Chmelka, B.F.; Stucky, G.D. Triblock copolymer syntheses of mesoporous silica with periodic 50 to 300 angstrom pores. Science 1998, 279, 548–552. [Google Scholar] [CrossRef]
  22. Hadjichristidis, N.; Iatrou, H.; Pitsikalis, M.; Mays, J. Macromolecular architectures by living and controlled/living polymerizations. Prog. Polym. Sci. 2006, 31, 1068–1132. [Google Scholar] [CrossRef]
  23. Matyjaszewski, K.; Sumerlin, B.; Tsarevsky, N. Progress in Controlled Radical Polymerization: Mechanisms and Techniques. Am. Chem. Soc. 2012, 1100, 317–331. [Google Scholar]
  24. Matyjaszewski, K. Advanced Materials by Atom Transfer Radical Polymerization. Adv. Mater. 2018, 30, 1706441–1706473. [Google Scholar] [CrossRef] [PubMed]
  25. Mertoglu, M.; Garnier, S.; Laschewsky, A.; Skrabania, K.; Storsberg, J. Stimuli responsive amphiphilic block copolymers for aqueous media synthesised via reversible addition fragmentation chain transfer polymerisation (RAFT). Polymer 2005, 46, 7726–7740. [Google Scholar] [CrossRef]
  26. Siegwart, D.; Oh, J.; Matyjaszewski, K. ATRP in the design of functional materials for biomedical applications. Prog. Polym. Sci. 2012, 37, 18–37. [Google Scholar] [CrossRef] [PubMed]
  27. Szwarc, M. ‘Living’ polymers. Nature 1956, 178, 1168–1169. [Google Scholar] [CrossRef]
  28. Moad, G.; Rizzardo, E.; Thang, S. Radical addition-fragmentation chemistry in polymer synthesis. Polymer 2008, 49, 1079–1131. [Google Scholar] [CrossRef]
  29. Nabiyan, A.; Max, J.B.; Schacher, F.H. Double hydrophilic copolymers-synthetic approaches, architectural variety, and current application fields. Chem. Soc. Rev. 2022, 51, 995–1044. [Google Scholar] [CrossRef] [PubMed]
  30. Nadal, C.; Gineste, S.; Coutelier, O.; Marty, J.D.; Destarac, M. A deeper insight into the dual temperature- and pH-responsiveness of poly(vinylamine)-b-poly(N-isopropylacrylamide) double hydrophilic block copolymers. Colloids Surf. A Physicochem. Eng. Asp. 2022, 641, 128502–128513. [Google Scholar] [CrossRef]
  31. Schmidt, B.V.K.J. Double Hydrophilic Block Copolymer Self-Assembly in Aqueous Solution. Macromol. Chem. Phys. 2018, 219, 1700494–1700508. [Google Scholar] [CrossRef]
  32. Butun, V.; Armes, S.; Billingham, N.; Tuzar, Z.; Rankin, A.; Eastoe, J.; Heenan, R. Synthesis and aqueous solution properties of a well-defined thermo-responsive schizophrenic diblock copolymer. Macromolecules 2001, 34, 1503. [Google Scholar]
  33. Liu, S.; Billingham, N.C.; Armes, S.P. A Schizophrenic Water-Soluble Diblock Copolymer. Angew. Chem. Int. Ed. 2001, 40, 2328–2332. [Google Scholar] [CrossRef]
  34. Liu, S.; Armes, S.P. Micelle Formation and Inversion Kinetics of a Schizophrenic Diblock Copolymer. Langmuir 2002, 19, 4432. [Google Scholar] [CrossRef]
  35. Papadakis, C.M.; Müller-Buschbaum, P.; Laschewsky, A. Switch it inside-out: “Schizophrenic” behavior of all thermoresponsive UCST-LCST diblock copolymers. Langmuir 2019, 35, 9660–9676. [Google Scholar] [CrossRef]
  36. Alam, M.; Keiko, H.; Yusa, S.; Nakashima, K. Schizophrenic micelle of a water-soluble diblock polymer and its application to a thermo-optical device. Colloid Polym. Sci. 2014, 292, 1611–1617. [Google Scholar] [CrossRef]
  37. Guragain, S.; Bastakoti, B.; Malgras, V.; Nakashima, K.; Yamauchi, Y. Multi-Stimuli-Responsive Polymeric Materials. Chem.-A Eur. J. 2021, 21, 13164–13174. [Google Scholar] [CrossRef] [PubMed]
  38. Mura, S.; Nicolas, J.; Couvreur, P. Stimuli-responsive nanocarriers for drug delivery. Nat. Mater. 2013, 12, 991–1003. [Google Scholar] [CrossRef] [PubMed]
  39. Stuart, M.A.C.; Huck, W.T.S.; Genzer, J.; Müller, M.; Ober, C.; Stamm, M.; Sukhorukov, G.B.; Szleifer, T.; Urban, M.; Winnik, F.; et al. Emerging applications of stimuli-responsive polymer materials. Nat. Mater. 2010, 9, 101–113. [Google Scholar] [CrossRef]
  40. Mizoue, Y.; Takahashi, R.; Sakurai, K.; Yusa, S. A Thermo-Responsive Polymer Micelle with a Liquid Crystalline Core. Polymers 2023, 15, 770. [Google Scholar] [CrossRef]
  41. Bhattacharya, D.; Behera, B.; Sahu, S.; Ananthakrishnan, R.; Maiti, T.; Pramanik, P. Design of Dual Stimuli Responsive Polymer Modified Magnetic Nanoparticles for Targeted Anti-Cancer Drug Delivery and Enhanced MR Imaging. New J. Chem. 2016, 40, 545–557. [Google Scholar] [CrossRef]
  42. Appold, M.; Mari, C.; Lederle, C.; Elbert, J.; Schmidt, C.; Ott, I.; Stühn, B.; Gasser, G.; Gallei, M. Multi-stimuli responsive block copolymers as a smart release platform for a polypyridyl ruthenium complex. Polym. Chem. 2016, 8, 890–900. [Google Scholar] [CrossRef]
  43. Fleige, E.; Quadir, M.; Haag, R. Stimuli-responsive polymeric nanocarriers for the controlled transport of active compounds: Concepts and applications. Adv. Drug. Deliv. Rev. 2012, 64, 866–884. [Google Scholar] [CrossRef]
  44. You, Y.; Oupicky, D. Synthesis of Temperature-Responsive Heterobifunctional Block Copolymers of Poly(ethylene glycol) and Poly(N-isopropylacrylamide). Biomacromolecules 2007, 8, 98–105. [Google Scholar] [CrossRef] [PubMed]
  45. Mukerabigwi, J.; Yin, W.; Zha, Z.; Ke, W.; Wang, Y.; Chen, W.; Japir, A.; Wang, Y.; Ge, Z. Polymersome nanoreactors with tumor pH-triggered selective membrane permeability for prodrug delivery, activation, and combined oxidation-chemotherapy. J. Control. Release 2019, 303, 209–222. [Google Scholar] [CrossRef] [PubMed]
  46. Shao, B.; Huang, X.; Xu, F.; Pan, J.; Wang, Y.; Zhou, S. A pH-responsive polymersome depleting regulatory T cells and blocking A2A receptor for cancer immunotherapy. Nano Res. 2022, 15, 2324–2334. [Google Scholar] [CrossRef]
  47. Zhuang, J.; Gordon, M.; Ventura, J.; Li, L.; Thayumanavan, S. Multi-stimuli responsive macromolecules and their assemblies. Chem. Soc. Rev. 2013, 42, 7421–7436. [Google Scholar] [CrossRef] [PubMed]
  48. Tan, S.; Zhao, D.; Yuan, D.; Wang, H.; Tu, K.; Wang, L. Influence of indomethacin-loading on the micellization and drug release of thermosensitive dextran-graft-poly(N-isopropylacrylamide). Reactive Func. Polym. 2011, 71, 820–827. [Google Scholar] [CrossRef]
  49. Wei, H.; Cheng, S.X.; Zhang, X.Z.; Zhuo, R.X. Thermo-sensitive polymeric micelles based on poly (N-isopropylacrylamide) as drug carriers. Progr. Polym. Sci. 2009, 34, 893–910. [Google Scholar] [CrossRef]
  50. He, M.; Zhang, Z.; Jiao, Z.; Yan, M.; Miao, P.; Wei, Z.; Leng, X.; Li, Y.; Fan, J.; Sun, W.; et al. Redox-responsive phenyl-functionalized polylactide micelles for enhancing Ru complexes delivery and phototherapy. Chin. Chem. Lett. 2023, 34, 1075741–1075746. [Google Scholar] [CrossRef]
  51. Li, Y.; Tong, A.; Niu, P.; Guo, W.; Jin, Y.; Hu, Y.; Tao, P.; Miao, W. Light-Decomposable Polymeric Micelles with Hypoxia-Enhanced Phototherapeutic Efficacy for Combating Metastatic Breast Cancer. Pharmaceutics 2022, 14, 253. [Google Scholar] [CrossRef] [PubMed]
  52. Tiwari, S.; Singh, K. Amphiphilic star block copolymer micelles in saline as effective vehicle for quercetin solubilization. J. Mol. Liq. 2021, 345, 118259–118265. [Google Scholar] [CrossRef]
  53. Kuperkar, K.; Tiwari, S.; Bahadur, P. Self-assembled block copolymer nanoaggregates for drug delivery applications. In Applications of Polymers in Drug Delivery, 2nd ed.; Elsevier: Amsterdam, The Netherlands, 2021; pp. 423–447. ISBN 978-0-12-819659-5. [Google Scholar]
  54. Nakashima, K.; Bahadur, P. Aggregation of water-soluble block copolymers in aqueous solutions: Recent trends. Adv. Colloid Interface Sci. 2006, 123, 75–96. [Google Scholar] [CrossRef] [PubMed]
  55. Hu, J.; Liu, S. Responsive Polymers for Detection and Sensing Applications: Current Status and Future Developments. Macromolecules 2010, 43, 8315–8330. [Google Scholar] [CrossRef]
  56. Stubenrauch, K.; Voets, I.; Fritz-Popovski, G.; Trimmel, G. pH and Ionic Strength Responsive Polyelectrolyte Block copolymer micelles prepared by ring opening metathesis polymerization. J. Polym. Sci. A 2008, 47, 1178–1191. [Google Scholar] [CrossRef]
  57. Qian, S.; Li, S.; Xiong, W.; Khan, H.; Huang, J.; Zhang, W. A new visible light and temperature responsive diblock copolymer. Polym. Chem. 2019, 10, 5001–5009. [Google Scholar] [CrossRef]
  58. Blanazs, A.; Armes, S.; Ryan, A. Self-assembled block copolymer aggregates: From micelles to vesicles and their biological applications. Macromol. Rapid Commun. 2009, 30, 267–277. [Google Scholar] [CrossRef] [PubMed]
  59. Mai, Y.; Eisenberg, A. Self-assembly of block copolymers. Chem. Soc. Rev. 2012, 41, 5969–5985. [Google Scholar] [CrossRef] [PubMed]
  60. Procházka, K.; Limpouchová, Z.; Štěpánek, M.; Šindelka, K.; Lísal, M. DPD Modelling of the Self-and Co-Assembly of Polymers and Polyelectrolytes in Aqueous Media: Impact on Polymer Science. Polymers 2022, 14, 404. [Google Scholar] [CrossRef]
  61. Tuzar, Z.; Kratochvil, P. Block and graft copolymer micelles in solution. Adv. Colloid Interface Sci. 1976, 6, 201–232. [Google Scholar] [CrossRef]
  62. Aqeel, R.; Srivastava, N.; Kushwaha, P. Micelles in Cancer Therapy: An Update on Preclinical and Clinical Status. Recent Pat. Nanotechnol. 2022, 16, 283–294. [Google Scholar]
  63. Wang, H.; Fliedel, C.; Manoury, E.; Poli, R. Core-crosslinked micelles with a poly-anionic poly(styrene sulfonate)-based outer shell made by RAFT polymerization. Polymer 2022, 243, 124640–124649. [Google Scholar] [CrossRef]
  64. Xiong, D.; Yao, N.; Gu, H.; Wang, J.; Zhang, L. Stimuli-responsive shell cross-linked micelles from amphiphilic four-arm star copolymers as potential nanocarriers for “pH/redox-triggered” anticancer drug release. Polymer 2017, 114, 161–172. [Google Scholar] [CrossRef]
  65. Bai, J.; Wang, J.; Feng, Y.; Yao, Y.; Zhao, X. Stability-tunable core-crosslinked polymeric micelles based on an imidazole-bearing block polymer for pH-responsive drug delivery. Colloids Surf. A 2022, 639, 128353–128361. [Google Scholar] [CrossRef]
  66. Kim, J.O.; Kabanov, A.V.; Bronich, T.K. Polymer micelles with cross-linked polyanion core for delivery of a cationic drug doxorubicin. J. Control. Release 2009, 138, 197–204. [Google Scholar] [CrossRef] [PubMed]
  67. Li, L.; Yang, W.; Xu, D. Stimuli-responsive nanoscale drug delivery systems for cancer therapy. J. Drug Target. 2018, 27, 423–433. [Google Scholar] [CrossRef] [PubMed]
  68. Li, K.; Chen, F.; Wang, Y.; Stenzel, M.; Chapman, R. Polyion Complex Micelles for Protein Delivery Benefit from Flexible Hydrophobic Spacers in the Binding Group. Macromol. Rapid Commun. 2022, 41, 2000208–2000214. [Google Scholar] [CrossRef] [PubMed]
  69. Yin, W.; Wang, Y.; Xiao, Y.; Mao, A.; Lang, M. Phenylboronic acid conjugated mPEG-b-PCL micelles as DOX carriers for enhanced drug encapsulation and controlled drug release. Eur. Polym. J. 2022, 173, 111235. [Google Scholar] [CrossRef]
  70. Feng, R.; Li, Z.; Fangfang, T.; Min, W.; Shiyu, C.; Zhimei, S.; Hongmei, L. Phenylboronic acid-modified polymeric anhydride-F127 micelles for pH-activated targeting delivery of doxorubicin. Colloids Surf. B 2022, 216, 112559–112568. [Google Scholar] [CrossRef] [PubMed]
  71. Wang, S.; Tan, X.; Zhou, Q.; Geng, P.; Wang, J.; Zou, P.; Deng, A.; Hu, J. Co-delivery of doxorubicin and SIS3 by folate-targeted polymeric micelles for overcoming tumor multidrug resistance. Drug Deliv. Transl. Res. 2022, 12, 167–179. [Google Scholar] [CrossRef]
  72. Bastakoti, B.; Liao, S.; Inoue, M.; Yusa, S.; Imura, M.; Nakashima, K.; Wu, K.; Yamauchi, Y. pH-responsive polymeric micelles with core-shell-corona architectures as intracellular anti-cancer drug carriers. Sci. Technol. Adv. Mater. 2013, 14, 044402–044407. [Google Scholar] [CrossRef]
  73. Nicolai, T.; Colombani, O.; Chassenieux, C. Dynamic polymeric micelles versus frozen nanoparticles formed by block copolymers. Soft Matter. 2010, 6, 3111–3118. [Google Scholar] [CrossRef]
  74. Gohy, J.; Zhao, Y. Photo-responsive block copolymer micelles: Design and behavior. Chem. Soc. Rev. 2013, 42, 7117–7129. [Google Scholar] [CrossRef] [PubMed]
  75. Gao, Z.; Varshney, S.; Wong, S.; Eisenberg, A. Block Copolymer “Crew-Cut” Micelles in Water. Macromolecules 1994, 27, 7923–7927. [Google Scholar] [CrossRef]
  76. Zhang, Q.; Zhu, S. Ionic liquids: Versatile media for preparation of vesicles from polymerization-induced self-assembly. ACS Macro Lett. 2015, 4, 755–758. [Google Scholar] [CrossRef] [PubMed]
  77. Becerra, E.; Quinchia, J.; Castro, C.; Orozco, J. Light-Triggered Polymersome-based anticancer therapeutics delivery. Nanomaterials 2022, 12, 836. [Google Scholar] [CrossRef] [PubMed]
  78. Chen, W.; Meng, F.; Cheng, R.; Zhong, Z. pH-Sensitive degradable polymersomes for triggered release of anticancer drugs: A comparative study with micelles. J. Control. Release 2010, 142, 40–46. [Google Scholar] [CrossRef] [PubMed]
  79. Discher, D.E.; Eisenberg, A. Polymer vesicles. Science 2002, 297, 967–973. [Google Scholar] [CrossRef] [PubMed]
  80. Discher, D.E.; Ahmed, F. Polymersomes. Annu. Rev. Biomed. Eng. 2006, 8, 323–341. [Google Scholar] [CrossRef]
  81. Letchford, K.; Burt, H. A review of the formation and classification of amphiphilic block copolymer nanoparticulate structures: Micelles, nanospheres, nanocapsules and polymersomes. Eur. J. Pharm. Biopharm. 2007, 65, 259–269. [Google Scholar] [CrossRef]
  82. Robertson, J.; Yealland, G.; Avila-Olias, M.; Chierico, L.; Bandmann, O.; Renshaw, S.; Battaglia, G. pH-sensitive tubular polymersomes: Formation and applications in cellular delivery. ACS Nano 2014, 8, 4650–4661. [Google Scholar] [CrossRef]
  83. Pawar, P.; Gohil, S.; Jain, J.; Kumar, N. Functionalized polymersomes for biomedical applications. Polym. Chem. 2013, 4, 3160–3176. [Google Scholar] [CrossRef]
  84. Mohammadi, M.; Ramezani, M.; Abnous, K.; Alibolandi, M. Biocompatible polymersomes-based cancer theranostics: Towards multifunctional nanomedicine. Int. J. Pharm. 2017, 519, 287–303. [Google Scholar] [CrossRef]
  85. Raffa, P.; Wever, D.; Picchioni, F.; Broekhuis, A. Polymeric Surfactants: Synthesis, Properties, and Links to Applications. Chem. Rev. 2015, 115, 8504–8563. [Google Scholar] [CrossRef] [PubMed]
  86. Pitto-Barry, A.; Barry, N.P.E. Pluronic® block-copolymers in medicine: From chemical and biological versatility to rationalisation and clinical advances. Polym. Chem. 2014, 5, 3291–3297. [Google Scholar] [CrossRef]
  87. Singla, P.; Garg, S.; McClements, J.; Jamieson, O.; Peeters, M.; Mahajan, R. Advances in the therapeutic delivery and applications of functionalized Pluronics: A critical review. Adv. Colloid Interface Sci. 2022, 299, 102563–102594. [Google Scholar] [CrossRef] [PubMed]
  88. Chiappetta, D.A.; Sosnik, A. Poly(ethylene oxide)-poly(propylene oxide) block copolymer micelles as drug delivery agents: Improved hydrosolubility, stability and bioavailability of drugs. Eur. J. Pharm. Biopharm. 2007, 66, 303–317. [Google Scholar] [CrossRef] [PubMed]
  89. Zlotnikov, I.D.; Ezhov, A.A.; Ferberg, A.S.; Krylov, S.S.; Semenova, M.N.; Semenov, V.V.; Kudryashova, E.V. Polymeric Micelles Formulation of Combretastatin Derivatives with Enhanced Solubility, Cytostatic Activity and Selectivity against Cancer Cells. Pharmaceutics 2023, 15, 1613. [Google Scholar] [CrossRef] [PubMed]
  90. Jundi, A.E.; Buwalda, S.J.; Bakkour, Y.; Garric, X.; Nottelet, B. Double hydrophilic block copolymers self-assemblies in biomedical applications. Adv. Colloid Int. Sci. 2020, 283, 102213–102279. [Google Scholar] [CrossRef] [PubMed]
  91. Giaouzi, D.; Pispas, S. PNIPAM-b-PDMAEA double stimuli responsive copolymers: Effects of composition, end groups and chemical modification on solution self-assembly. Eur. Polym. J. 2020, 135, 109867. [Google Scholar] [CrossRef]
  92. Bhowmik, S.; Pham, T.; Takahashi, R.; Kim, D.; Matsuoka, H.; Ishihara, K.; Yusa, S. Preparation of Water-Soluble Polyion Complex (PIC) Micelles with Random Copolymers Containing Pendant Quaternary Ammonium and Sulfonate Groups. Langmuir 2023, 39, 8120–8129. [Google Scholar] [CrossRef]
  93. Pham, T.; Pham, T.; Yusa, S. Polyion complex (PIC) micelles formed from oppositely charged styrene-based polyelectrolytes via electrostatic, hydrophobic, and π–π interactions. Polym. J. 2022, 54, 1091–1101. [Google Scholar] [CrossRef]
  94. Song, Y.; Tian, Q.; Huang, Z.; Fan, D.; She, Z.; Liu, X.; Cheng, X.; Yu, B.; Deng, Y. Self-assembled micelles of novel amphiphilic copolymer cholesterol-coupled F68 containing cabazitaxel as a drug delivery system. Int. J. Nanomed. 2014, 9, 2307–2317. [Google Scholar]
  95. Kim, D.; Vitol, E.; Liu, J.; Balasubramanian, S.; Gosztola, D.; Cohen, E.; Novosad, V.; Rozhkova, E. Stimuli-Responsive Magnetic Nanomicelles as Multifunctional Heat and Cargo Delivery Vehicles. Langmuir 2013, 29, 7425–7432. [Google Scholar] [CrossRef] [PubMed]
  96. Medeiros, S.; Santos, S.; Fessi, H.; Elaissari, A. Stimuli-responsive magnetic particles for biomedical applications. Int. J. Pharm. 2011, 403, 139–161. [Google Scholar] [CrossRef] [PubMed]
  97. Zhang, Y.; Chen, S.; Pang, M.; Zhang, W. Synthesis and micellization of a multi-stimuli responsive block copolymer based on spiropyran. Polym. Chem. 2016, 7, 6880–6885. [Google Scholar] [CrossRef]
  98. Corten, C.; Kretschmer, K.; Kuckling, D. Novel multi-responsive P2VP-block-PNIPAAm block copolymers via nitroxide-mediated radical polymerization. Beilstein J. Org. Chem. 2010, 6, 756–765. [Google Scholar] [CrossRef] [PubMed]
  99. Fuoss, R.; Sadek, H. Mutual interaction of polyelectrolytes. Science 1949, 110, 552–554. [Google Scholar] [CrossRef] [PubMed]
  100. Nakai, K.; Nishiuchi, M.; Inoue, M.; Ishihara, K.; Sanada, Y.; Sakurai, K.; Yusa, S. Preparation and characterization of polyion complex micelles with phosphobetaine shells. Langmuir 2013, 29, 9651–9661. [Google Scholar] [CrossRef] [PubMed]
  101. Damsongsang, P.; Yusa, S.; Hoven, V. Zwitterionic nano-objects having functionalizable hydrophobic core: Formation via polymerization-induced self-assembly and their morphology. Eur. Polym. J. 2022, 179, 111536. [Google Scholar] [CrossRef]
  102. Ohno, S.; Ishihara, K.; Yusa, S. Formation of Polyion Complex (PIC) Micelles and Vesicles with Anionic pH-Responsive Unimer Micelles and Cationic Diblock Copolymers in Water. Langmuir 2016, 32, 3945–3953. [Google Scholar] [CrossRef]
  103. Yusa, S.; Yokoyama, Y.; Morishima, Y. Synthesis of oppositely charged block copolymers of polyethylene glycol via reversible addition-fragmentation chain transfer radical polymerization and characterization of their polyion complex micelles in water. Macromolecules 2009, 42, 376–383. [Google Scholar] [CrossRef]
  104. Chen, F.; Stenzel, M. Polyion Complex Micelles for Protein Delivery. Aust. J. Chem. 2018, 71, 768–780. [Google Scholar] [CrossRef]
  105. Harada, A.; Kataoka, K. Polyion complex micelle formation from double-hydrophilic block copolymers composed of charged and non-charged segments in aqueous media. Polym. J. 2018, 50, 95–100. [Google Scholar] [CrossRef]
  106. Nakamura, N.; Mochida, Y.; Toh, K.; Anraku, Y.; Cabral, H. Effect of mixing ratio of oppositely charged block copolymers on polyion complex micelles for in vivo application. Polymers 2021, 13, 5. [Google Scholar] [CrossRef] [PubMed]
  107. Charleux, B.; Delaittre, G.; Rieger, J.; D’Agosto, F. Polymerization-Induced Self-Assembly: From Soluble Macromolecules to Block Copolymer Nano-Objects in One Step. Macromolecules 2012, 45, 6753–6765. [Google Scholar] [CrossRef]
  108. Derry, M.; Fielding, L.; Armes, S. Polymerization-induced self-assembly of block copolymer nanoparticles via RAFT non-aqueous dispersion polymerization. Prog. Polym. Sci. 2016, 52, 1–18. [Google Scholar] [CrossRef]
  109. Penfold, N.; Yeow, J.; Boyer, C.; Armes, S. Emerging Trends in Polymerization-Induced Self-Assembly. ACS Macro Lett. 2019, 8, 1029–1054. [Google Scholar] [CrossRef] [PubMed]
  110. Warren, N.; Armes, S. Polymerization-Induced Self-Assembly of Block Copolymer Nano-objects via RAFT Aqueous Dispersion Polymerization. J. Am. Chem. Soc. 2014, 136, 10174–10185. [Google Scholar] [CrossRef] [PubMed]
  111. Pei, Y.; Lowe, A. Polymerization-induced self-assembly: Ethanolic RAFT dispersion polymerization of 2-phenylethyl methacrylate. Polym. Chem. 2014, 5, 2342–2351. [Google Scholar] [CrossRef]
  112. Zhang, W.; Kadirkhanov, J.; Wang, C.; Ding, S.; Hong, C.; Wang, F.; You, Y. Polymerization-induced self-assembly for the fabrication of polymeric nano-objects with enhanced structural stability by cross-linking. Polym. Chem. 2020, 11, 3654–3672. [Google Scholar] [CrossRef]
  113. Cao, J.; Tan, Y.; Chen, Y.; Zhang, L.; Tan, J. Expanding the Scope of Polymerization-Induced Self-Assembly: Recent Advances and New Horizons. Macromol. Rapid Comm. 2021, 42, 2100498–2100510. [Google Scholar] [CrossRef]
  114. D’Agosto, F.; Rieger, J.; Lansalot, M. RAFT-Mediated Polymerization-Induced Self-Assembly. Angew. Chem. 2020, 59, 8368–8392. [Google Scholar] [CrossRef] [PubMed]
  115. Liu, C.; Hong, C.; Pan, C. Polymerization techniques in polymerization-induced self-assembly (PISA). Polym. Chem. 2020, 11, 3673–3689. [Google Scholar] [CrossRef]
  116. Gonzalez, J.; Schmarsow, R.; Rojo, U.; Puig, J.; Schroeder, W.; Zucchi, I. Block Copolymer Micelles Generated by Crystallization-Driven Self-Assembly in Polymer Matrices. Sci. Rev. 2020, 1, 47–64. [Google Scholar]
  117. Inam, M.; Cambridge, G.; Pitto-Barry, A.; Laker, Z.L.; Wilson, N.R.; Mathers, R.T.; Dove, A.P.; O’Reilly, R.K. 1D vs. 2D shape selectivity in the crystallization-driven self-assembly of polylactide block copolymers. Chem. Sci. 2017, 8, 4223–4230. [Google Scholar] [CrossRef] [PubMed]
  118. Gädt, T.; Ieong, N.; Cambridge, G.; Winnik, M.; Manners, I. Complex and hierarchical micelle architectures from diblock copolymers using living, crystallization-driven polymerizations. Nat. Mater. 2009, 8, 144–150. [Google Scholar] [CrossRef] [PubMed]
  119. Ganda, S.; Stenzel, M. Concepts, fabrication methods and applications of living crystallization-driven self-assembly of block copolymers. Prog. Polym. Sci. 2020, 101, 101195–101205. [Google Scholar] [CrossRef]
  120. He, W.; Xu, J. Crystallization assisted self-assembly of semicrystalline block copolymers. Prog. Polym. Sci. 2012, 37, 1350–1400. [Google Scholar] [CrossRef]
  121. Finnegan, J.; Pilkington, E.; Alt, K.; Rahim, A.; Kent, S.; Davis, T.; Kempe, K. Stealth Nanorods via the Aqueous Living Crystallisation-Driven Self-Assembly of Poly(2-oxazoline)s. Chem. Sci. 2021, 12, 7350–7360. [Google Scholar] [CrossRef]
  122. Gilroy, J.; Gädt, T.; Whittell, G.; Chabanne, L.; Mitchels, J.; Richardson, R.; Winnik, M.; Manners, I. Monodisperse cylindrical micelles by crystallization-driven living self-assembly. Nat. Chem. 2010, 2, 566–570. [Google Scholar] [CrossRef]
  123. MacFarlane, L.; Zhao, C.; Cai, J.; Qiu, H.; Manners, I. Emerging applications for living crystallization-driven self-assembly. Chem. Sci. 2021, 12, 4661–4682. [Google Scholar] [CrossRef]
  124. Sha, Y.; Rahman, A.; Zhu, T.; Cha, Y.; McAlister, C.; Tang, C. ROMPI-CDSA: Ring-opening metathesis polymerization-induced crystallization-driven self-assembly of metallo-block copolymers. Chem. Sci. 2019, 10, 9782–9787. [Google Scholar] [CrossRef] [PubMed]
  125. López-Lorente, A.; Mizaikoff, B. Recent advances on the characterization of nanoparticles using infrared spectroscopy. Trends Analyt. Chem. 2016, 84, 97–106. [Google Scholar] [CrossRef]
  126. Modena, M.; Rühle, B.; Burg, T.; Wuttke, S. Nanoparticle Characterization: What to Measure. Adv. Mater. 2019, 31, 1901556–1901580. [Google Scholar] [CrossRef] [PubMed]
  127. Mourdikoudis, S.; Pallares, R.; Thanh, N. Characterization Techniques for Nanoparticles: Comparison and Complementarity upon Studying Nanoparticle Properties. Nanoscale 2018, 10, 12871–12934. [Google Scholar] [CrossRef] [PubMed]
  128. Xu, L.; Zhang, Z.; Wang, F.; Xie, D.; Yang, S.; Wang, T.; Feng, L.; Chu, C. Synthesis, characterization, and self-assembly of linear poly(ethyleneoxide)-block–poly(propylene oxide)-block–poly(e-caprolactone) (PEO–PPO–PCL) copolymers. J. Colloid Interf. Sci. 2013, 393, 174–181. [Google Scholar] [CrossRef] [PubMed]
  129. Jung, Y.; Park, W.; Park, H.; Lee, D.; Na, K. Thermo-sensitive injectable hydrogel based on the physical mixing of hyaluronic acid and Pluronic F-127 for sustained NSAID delivery. Carbohydr. Polym. 2017, 156, 403–408. [Google Scholar] [CrossRef] [PubMed]
  130. Patel, D.; Ray, D.; Aswal, V.; Kuperkar, K.; Bahadur, P. Temperature stimulated self-association and micellar transition for star shaped normal and reverse EO-PO block copolymers and their mixed systems as potential use for anticancer drug solubilization. Soft Matter. 2022, 18, 4543–4553. [Google Scholar] [CrossRef]
  131. Patel, D.; Jana, R.; Lin, M.; Kuperkar, K.; Seth, D.; Chen, L.; Bahadur, P. Revisiting the salt-triggered self-assembly in very hydrophilic triblock copolymer Pluronic® F88 using multitechnique approach. Colloid Polym. Sci. 2021, 299, 229–239. [Google Scholar] [CrossRef]
  132. Lunagariya, J.; Sivakumar, N.; Asif, M.; Dhar, A.; Vekariya, R. Dependency of Anion and Chain Length of Imidazolium Based Ionic Liquid on Micellization of the Block Copolymer F127 in Aqueous Solution: An Experimental Deep Insight. Polymers 2017, 9, 285. [Google Scholar] [CrossRef]
  133. Patel, D.; Ray, D.; Kuperkar, K.; Pal, H.; Aswal, V.; Bahadur, P. Solubilization, micellar transition and biocidal assay of loaded multifunctional antioxidants in Tetronic® 1304 micelles. Polym. Int. 2020, 69, 1097–1104. [Google Scholar] [CrossRef]
  134. Patel, D.; Ray, D.; Kuperkar, K.; Aswal, V.; Bahadur, P. Parabens induced spherical micelle to polymersome transition in thermo-responsive amphiphilic linear and star-shaped EO-PO block copolymers. J. Mol. Liq. 2020, 316, 113897–113908. [Google Scholar] [CrossRef]
  135. Singla, P.; Singh, O.; Chabba, S.; Aswal, V.; Mahajan, R. Sodium deoxycholate mediated enhanced solubilization and stability of hydrophobic drug Clozapine in pluronic micelles. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2018, 15, 143–154. [Google Scholar] [CrossRef] [PubMed]
  136. Kim, H.; Park, S.; Hinsberg, W. Block copolymer-based nanostructures: Materials, processes, and applications to electronics. Chem. Rev. 2010, 110, 146–177. [Google Scholar] [CrossRef] [PubMed]
  137. Kim, J.; Lee, J. Preparation and Properties of Poly(l-lactide)-block-poly(trimethylenecarbonate) as Biodegradable Thermoplastic Elastomer. Polym. J. 2002, 34, 203–208. [Google Scholar] [CrossRef]
  138. Chowdhry, B.; Snowdena, M.; Leharne, S. A scanning calorimetric investigation of phase transitions in a PPO-PEO-PPO block copolymer. Eur. Polym. J. 1999, 35, 273–278. [Google Scholar] [CrossRef]
  139. Sarolia, J.; Kumar, D.; Shah, S.; Bahadur, P.; Tiwari, S. Thermodynamics of pluronic 103 micellization in mannitol solution: Analyses based on isothermal titration calorimetry. Colloid Surf. A 2022, 648, 129240–129244. [Google Scholar] [CrossRef]
  140. Santiago, A.; Vargas, J.; Jorge, V.; Cruz-Morales, J.; Mikhail, T.; Gavino, R.; Malkanduev, Y.; Sivov, N. Synthesis of New Polymer Ionomers via Ring-Opening Metathesis Polymerization. Open J. Organic Polym. Mat. 2014, 4, 84–91. [Google Scholar] [CrossRef]
  141. Patel, D.; Vaswani, P.; Sengupta, S.; Ray, D.; Bhatia, D.; Choudhury, S.D.; Aswal, V.K.; Kuperkar, K.; Bahadur, P. Thermoresponsive phase behavior and nanoscale self-assembly generation in normal and reverse Pluronics®. Colloid Polym. Sci. 2023, 301, 75–92. [Google Scholar] [CrossRef]
  142. Akhlaghi, S.; Ribeiro, I.; Boyd, B.; Loh, W. Impact of preparation method and variables on the internal structure, morphology, and presence of liposomes in phytantriol-Pluronic® F127cubosomes. Colloids Surf. B 2016, 145, 845–853. [Google Scholar] [CrossRef]
  143. Rodrigues, E.; Morales, M.; Medeiros, S.; Suguihiro, M.; Baggio-Saitovitch, E. Pluronics coated sterically stabilized magnetite nanoparticles for hyperthermia applications. J. Magn. Magn. Mater. 2016, 416, 434–440. [Google Scholar] [CrossRef]
  144. Kanga, E.; Sharker, S.; Inc, I.; Park, S. Pluronic mimicking fluorescent carbon nanoparticles conjugated with doxorubicin via acid-cleavable linkage for tumor-targeted drug delivery and bioimaging. J. Ind. Eng. Chem. 2016, 43, 150–157. [Google Scholar] [CrossRef]
  145. Bhattacharjee, A.; Kumar, K.; Arora, A.; Katti, D. Fabrication and characterization of Pluronic modified poly(hydroxybutyrate) fibers for potential wound dressing applications. Mat. Sci. Eng. C 2016, 63, 266–273. [Google Scholar] [CrossRef] [PubMed]
  146. Pellosi, D.; Calori, I.; Paula, L.; Hioka, N.; Quaglia, F.; Tedesco, A. Multifunctional theranostic Pluronic mixed micelles improve targeted photoactivity of Verteporfin in cancer cells. Mat. Sci. Eng. C 2017, 71, 1–9. [Google Scholar] [CrossRef] [PubMed]
  147. Yap, L.; Yang, M. Evaluation of hydrogel composing of Pluronic F127 and carboxymethyl hexanoyl chitosan as injectable scaffold for tissue engineering applications. Colloids Surf. B 2016, 146, 204–211. [Google Scholar] [CrossRef] [PubMed]
  148. Negut, I.; Bita, B. Polymeric Micellar Systems—A Special Emphasis on “Smart” Drug Delivery. Pharmaceutics 2023, 15, 976. [Google Scholar] [CrossRef] [PubMed]
  149. Talelli, M.; Barz, M.; Rijcken, C.; Kiessling, F.; Hennink, W.; Lammers, T. Core-crosslinked polymeric micelles: Principles, preparation, biomedical applications and clinical translation. Nano Today 2015, 10, 93–117. [Google Scholar] [CrossRef] [PubMed]
  150. Colfen, H. Double-Hydrophilic Block Copolymers: Synthesis and Application as Novel Surfactants and Crystal Growth Modifiers. Macromol. Rapid Commun. 2001, 4, 220–252. [Google Scholar]
  151. Zou, Y.; Zhou, X.; Ma, J.; Yang, X.; Deng, Y. Recent advances in amphiphilic block copolymer templated mesoporous metal-based materials: Assembly engineering and applications. Chem. Soc. Rev. 2020, 49, 1173–1208. [Google Scholar] [CrossRef]
  152. Lu, H.; Jiang, K.; Liang, X.; Liu, H.; Li, Y. Small molecule-mediated self-assembly behaviors of Pluronic block copolymers in aqueous solution: Impact of hydrogen bonding on the morphological transition of Pluronic micelles. Soft Matter. 2020, 16, 142–152. [Google Scholar] [CrossRef]
  153. Dhapte, V.; Mehta, P. Advances in hydrotropic solutions: An updated review. Polytech. J. 2015, 1, 424–435. [Google Scholar] [CrossRef]
  154. Hodgdon, T.; Kaler, E. Hydrotropic solutions. Curr. Opin. Colloid Interface Sci. 2007, 12, 121–128. [Google Scholar] [CrossRef]
  155. Vemula, V.; Lagishetty, V.; Lingala, S. Solubility enhancement techniques. Int. J. Pharm. Sci. Rev. Res. 2010, 5, 41–51. [Google Scholar]
  156. Klermund, L.; Castiglione, K. Polymersomes as nanoreactors for preparative biocatalytic applications: Current challenges and future perspectives. Bioprocess Biosyst. Eng. 2018, 41, 1233–1246. [Google Scholar] [CrossRef] [PubMed]
  157. Ruttala, H.; Ramasamy, T.; Madeshwaran, T.; Hiep, T.; Kandasamy, U.; Oh, K.; Choi, H.; Yong, C.; Kim, J. Emerging potential of stimulus-responsive nanosized anticancer drug delivery systems for systemic applications. Arch. Pharm. Res. 2017, 41, 111–129. [Google Scholar] [CrossRef] [PubMed]
  158. Aliabadi, H.; Lavasanifar, A. Polymeric micelles for drug delivery. Expert Opin. Drug Deliv. 2006, 3, 139–162. [Google Scholar] [CrossRef] [PubMed]
  159. Manimaran, V.; Nivetha, R.P.; Tamilanban, T.; Narayanan, J.; Vetriselvan, S.; Fuloria, N.K.; Chinni, S.V.; Sekar, M.; Fuloria, S.; Wong, L.S.; et al. Nanogels as novel drug nanocarriers for CNS drug delivery. Front. Mol. Biosci. 2023, 10, 1232109. [Google Scholar] [CrossRef] [PubMed]
  160. Hlavatovičová, E.; Fernandez-Alvarez, R.; Byś, K.; Kereïche, S.; Mandal, T.; Atanase, L.; Štěpánek, M.; Uchman, M. Stimuli-Responsive Triblock Terpolymer Conversion into Multi-Stimuli-Responsive Micelles with Dynamic Covalent Bonds for Drug Delivery through a Quick and Controllable Post-Polymerization Reaction. Pharmaceutics 2023, 15, 288. [Google Scholar] [CrossRef]
  161. Zlotnikov, I.; Streltsov, D.; Ezhov, A.; Kudryashova, E. Smart pH- and Temperature-Sensitive Micelles Based on Chitosan Grafted with Fatty Acids to Increase the Efficiency and Selectivity of Doxorubicin and Its Adjuvant Regarding the Tumor Cells. Pharmaceutics 2023, 15, 1135. [Google Scholar] [CrossRef]
  162. Szewczyk-Łagodzińska, M.; Plichta, A.; Dębowski, M.; Kowalczyk, S.; Iuliano, A.; Florjańczyk, Z. Recent Advances in the Application of ATRP in the Synthesis of Drug Delivery Systems. Polymers 2023, 15, 1234. [Google Scholar] [CrossRef]
  163. Movassaghian, S.; Merkel, O.M.; Torchilin, V.P. Applications of polymer micelles for imaging and drug delivery. Wiley Interdiscip. Rev. Nanomed. 2015, 7, 691–707. [Google Scholar] [CrossRef]
  164. Varela-Moreira, A.; Yang, S.; Marcel, H.F.; Twan, L.; Wim, E.H.; Raymond, M.S. Clinical application of polymeric micelles for the treatment of cancer. Mater. Chem. Front. 2017, 1, 1485–1501. [Google Scholar] [CrossRef]
  165. Ghezzi, M.; Pescina, S.; Padula, C.; Santi, P.; Favero, E.D.; Cantù, L.; Nicoli, S. Polymeric micelles in drug delivery: An insight of the techniques for their characterization and assessment in biorelevant conditions. J. Control. Release. 2021, 332, 312–336. [Google Scholar] [CrossRef] [PubMed]
  166. Kaur, J.; Monica, G.; Niraj, K.J.; John, D.; Vandana, P.; Kamal, D.; Sachin, K.S. Recent advances in developing polymeric micelles for treating cancer: Breakthroughs and bottlenecks in their clinical translation. Drug Discov. Today 2022, 27, 1495–1512. [Google Scholar] [CrossRef] [PubMed]
  167. Junnuthula, V.; Kolimi, P.; Nyavanandi, D.; Sampathi, S.; Vora, L.K.; Dyawanapelly, S. Polymeric Micelles for Breast Cancer Therapy: Recent Updates, Clinical Translation and Regulatory Considerations. Pharmaceutics 2022, 4, 1860. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of some block copolymer structures.
Figure 1. Schematic representation of some block copolymer structures.
Ddc 02 00045 g001
Figure 2. Examples of some (a) water–soluble homopolymers from natural or synthetic origin (charged or uncharged), and (b) hydrophobic polymers that may undergo polymerization to form BCPs.
Figure 2. Examples of some (a) water–soluble homopolymers from natural or synthetic origin (charged or uncharged), and (b) hydrophobic polymers that may undergo polymerization to form BCPs.
Ddc 02 00045 g002
Figure 3. Formation of schizophrenic micelles in AB type DHBC.
Figure 3. Formation of schizophrenic micelles in AB type DHBC.
Ddc 02 00045 g003
Figure 4. Self-assembly of core–shell cross-linking using DHBC.
Figure 4. Self-assembly of core–shell cross-linking using DHBC.
Ddc 02 00045 g004
Figure 5. Some common examples of stimuli-responsive blocks in BCPs.
Figure 5. Some common examples of stimuli-responsive blocks in BCPs.
Ddc 02 00045 g005
Figure 6. Some thermo–responsive polymers with their respective LCST values.
Figure 6. Some thermo–responsive polymers with their respective LCST values.
Ddc 02 00045 g006
Figure 7. Some examples of pH-responsive polymers.
Figure 7. Some examples of pH-responsive polymers.
Ddc 02 00045 g007
Figure 8. Light-responsive polymers with reversible photochromic and irreversible photo-cleavable moieties.
Figure 8. Light-responsive polymers with reversible photochromic and irreversible photo-cleavable moieties.
Ddc 02 00045 g008
Figure 9. Alteration in the PMs via cross-linking of the core, shell, and surface functionalization. Also, other possible PMs structures, viz., core–shell–corona, Janus micelles, frozen micelles, and crew-cut micelles, are schematically shown.
Figure 9. Alteration in the PMs via cross-linking of the core, shell, and surface functionalization. Also, other possible PMs structures, viz., core–shell–corona, Janus micelles, frozen micelles, and crew-cut micelles, are schematically shown.
Ddc 02 00045 g009
Figure 10. Hydrophobic/hydrophilic molecules entrapped in polymersomes.
Figure 10. Hydrophobic/hydrophilic molecules entrapped in polymersomes.
Ddc 02 00045 g010
Figure 11. Polyion complex (PIC) micelles and polymersomes.
Figure 11. Polyion complex (PIC) micelles and polymersomes.
Ddc 02 00045 g011
Figure 12. Schematic representation of PISA in solution.
Figure 12. Schematic representation of PISA in solution.
Ddc 02 00045 g012
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Patel, D.; Kuperkar, K.; Yusa, S.-i.; Bahadur, P. Nanoscale Self-Assemblies from Amphiphilic Block Copolymers as Proficient Templates in Drug Delivery. Drugs Drug Candidates 2023, 2, 898-922. https://doi.org/10.3390/ddc2040045

AMA Style

Patel D, Kuperkar K, Yusa S-i, Bahadur P. Nanoscale Self-Assemblies from Amphiphilic Block Copolymers as Proficient Templates in Drug Delivery. Drugs and Drug Candidates. 2023; 2(4):898-922. https://doi.org/10.3390/ddc2040045

Chicago/Turabian Style

Patel, Dhruvi, Ketan Kuperkar, Shin-ichi Yusa, and Pratap Bahadur. 2023. "Nanoscale Self-Assemblies from Amphiphilic Block Copolymers as Proficient Templates in Drug Delivery" Drugs and Drug Candidates 2, no. 4: 898-922. https://doi.org/10.3390/ddc2040045

Article Metrics

Back to TopTop