Degradable and Recyclable Polymers by Reversible Deactivation Radical Polymerization
Reversible deactivation radical polymerization (RDRP) provides unprecedented control over polymer composition, size, functionality, and topology. Various materials, such as linear polymers, star polymers, branched polymers, graft polymers, polymer networks, and hybrid materials, have been prepared by RDRP. The ability to control polymer topology also enabled precision synthesis of well-defined polymer topologies with degradable functional groups located at specific locations along a polymer chain. This review outlines progress in the synthesis of degradable polymers designed by RDRP, organized by topology and synthetic route. Recent progress in the depolymerization of polymers using RDRP mechanisms is highlighted and critically discussed.
Introduction
Plastics have become an integral part of our world since their discovery more than 100 years ago. They are used in components of electronics, transportation, packaging, biomedical devices, infrastructure, textiles, and other diverse applications due to their low cost, light weight, and tunable mechanical properties. The ubiquitous nature of plastics is reflected in their history and projections of future growth. More than 360 million metric tons of synthetic plastic was produced in 2018.1 The global plastic market size grew to $579.7 billion (U.S.) in 2020 and is expected to continue growing to $750.1 billion (U.S.) by 2028.2 The exponential growth of the plastics industry has, unfortunately, also led to environmental concerns through its reliance on fossil fuels as feedstocks and pollution into the environment. An estimated 6300 Mt of plastic waste was generated worldwide between 1950 and 2015, and the majority was discarded in landfills or the environment.3 To this extent, we must critically assess recent and past progress in the synthesis of degradable and sustainable high-performance polymers to inform future work in this area.4–7
A significant fraction of commercial synthetic polymers are produced via conventional
radical polymerization (RP) of vinyl monomers, such as various olefins, vinyl acetate,
vinyl lactams, styrene, acrylates, and methacrylates (Figure 1).8,9 The widespread use of RP is largely enabled by a broad functional group compatibility
and tolerance to moisture and protic media. Polymers prepared by RP follow a standard
chain growth mechanism consisting of initiation, propagation, transfer, and termination.
Radicals are generated by slow decomposition of a radical initiator, followed by rapid
propagation, and are terminated by biradical termination or transfer. Slow decomposition
of initiator and lack of end-group retainment in the absence of a chain transfer agent
(CTA) provide minimal control over the molecular weight and end-group functionality
of polymers produced by RP.
Figure 1 | Polymerizable monomers by radical polymerization. The monomers denoted with a colored
* were reported to be polymerizable by the respective RDRP method.
The invention of living ionic polymerizations led to advancements in polymer chemistry through optimization of polymerization kinetics. A living polymerization is a chain-growth polymerization which proceeds in the absence of chain-breaking events.10,11 Polymers produced in a living process can have uniform chain length if the rate of initiation is higher than the rate of propagation. This enables the synthesis of low-dispersity polymers with chain-end functionality, as well as with complex topologies such as brushes, stars, and branched polymers.12,13 However, application of living ionic polymerizations are limited by the sensitivity of the cationic and anionic propagating species to specific functional groups and moisture.
A compromise between the robust nature of radical polymerization and precision of
a living polymerization was achieved through reversible deactivation radical polymerizations
(RDRP). RDRP provides control over molecular structures in radical polymerizations
through reversible dissociation and combination with a thermally labile nitroxide
adduct in a nitroxide-mediated polymerization (NMP) (Figure 2a), degenerative chain transfer with a CTA in radical addition-fragmentation transfer
(RAFT) (Figure 2c), and by halogen atom transfer with a transition metal complex in an atom transfer
radical polymerization (ATRP) (Figure 2b).14–18 The fraction of terminated chains is significantly reduced in the presence of a large
amount of dormant species. RDRP achieves uniform growth of polymer chains by maintaining
a fast rate of initiation relative to propagation. These factors enable precise control
over polymer chain length, dispersity, and chain-end functionality for monomers and
reaction conditions suitable for radical polymerization. The use of multifunctional
initiators enables the synthesis of branched polymer topologies, such as stars and
molecular bottlebrushes, by tethering multiple initiators to a core or backbone.8,19–25 Initiators can also be installed on the surface of various inorganic and organic
scaffolds in the preparation of various hybrid composites.8,19,22,24,26,27 Figure 2 | The general mechanism of RDRP mediated by (a) reversible decomposition and combination
in a NMP; (b) reversible halogen atom transfer in an ATRP; (c) degenerative transfer
with a CTA in a RAFT polymerization.
There has been significant progress in expanding the monomer scope, environmental impact, and cost efficiency of RDRP methods over the last two decades.4,9,28,29 RDRP of styrene, acrylates, and methacrylates continue to be the state-of-the-art. However, progress has been made in the polymerization of challenging monomers, such as conjugated dienes30–32 and vinyl chloride (Figure 1).33–36 New methods of initiation by photoinduced electron/energy transfer RAFT and photoinduced ATRP enable polymerization under oxygen atmosphere upon exposure to low-intensity visible light.37–39 Catalyst regeneration in the presence of external reducing agents or radical initiators enables well-controlled ATRP with a catalyst concentration as low as 10 ppm.40 Polymers prepared by RDRP are used commercially as high-performance adhesives, sealants, rheology modifiers, surface modifiers, latex binders, chromatographic supports, solid polymer electrolytes, and smart (bio)materials.41–43
The ability to precisely incorporate degradable and reversible bonds into high-performance materials is an attractive opportunity for industry. Polyolefins are often cracked to lower molecular weight waxes and fuels under harsh conditions because they lack the degradable bonds necessary for selective bond scission under mild conditions.44,45 Polyesters, such as poly(lactic acid) and poly (glycolic acid), have beneficial degradable properties, but applications have been limited by their high cost and poor mechanical properties.46,47 There has been recent progress in improving the properties of polyesters which can match the properties of commodity plastics and in the upcycling of plastic waste.48–53 All of these efforts continue to be difficult challenges for our field to overcome.50,54–58
This review focuses on the incorporation of degradable functional groups into high-performance polymers prepared by reversible deactivation radical polymerization. The first portion of this review provides an overview of linear polymers prepared by RDRP of degradable monomers.59 Precise control over end-groups and topologies enables control over both polymer structure and functionality at the junctions of branched polymers. The following sections are organized by polymer topology, starting from linear polymers and then (hyper)branched polymers, star polymers, molecular bottlebrushes, polymer networks, and composites. The final section of this review summarizes recent work on the depolymerization of polymers to monomers by depolymerizations mediated by RDRP mechanisms.
Polymerizations and Copolymerizations of Degradable Monomers
Radical ring-opening polymerization (RROP) has emerged as a powerful method to install
heteroatom-containing functional groups into a vinyl polymer backbone. The resulting
copolymers have a statistical amount of (often) hydrolyzable functional groups along
the backbone which enables polymer degradation to oligomers with low molecular weight
(Scheme 1).
Scheme 1 | Degradable copolymer prepared by copolymerization of a vinyl monomer (black dot) with
a degradable comonomer (red dot) can be cleaved into lower molecular weight oligomers
after scission of the degradable bonds along the backbone.
The amount and position of the degradable bonds along a copolymer backbone are dictated by the reactivity ratios of the monomers and the propensity of the monomer to polymerize via vinyl addition or RROP.
Incorporation of esters by RROP
RROP of cyclic ketene acetals (CKAs) is the most common method used to statistically incorporate ester functionalities into a vinyl polymer backbone.59 Incorporation of ester functional groups enables degradation by hydrolysis along the backbone, which is impossible for vinyl polymers with only C–C bonds in the backbone. The CKA functionality is typically less reactive than other vinyl comonomers, which leads to compositional drift during copolymerization which enriches vinyl repeat unit content at the beginning of the polymerization and ring-opened repeat units near the end of the reaction.60 Thus, degradation of copolymers prepared by RROP of two monomers with poor compatibility yields ill-defined lower molecular weight oligomers,61 which may influence their biodegradability.62
The 2-methylene-1,3-dioxepane (MDO) CKA was effectively copolymerized with monomers such as vinyl acetates, methacrylates, and pyrrolidones, via RAFT polymerization.63,64 The reactivity of MDO is lower than the comonomers, which leads to slow and incomplete incorporation of MDO into the backbone at high loadings. The ester functional groups along the backbone are resistant to hydrolysis against potassium carbonate, which enabled selective hydrolysis of poly(vinyl chloroacetate)-co-poly(2-methylene-1,3-dioxepane) into poly(vinyl alcohol)-co-poly(2-methylene-1,3-dioxepane)) without degrading the polymer backbone,65 but are degradable in more basic solutions.
The addition of a phenyl functional group to the 7-member MDO CKA scaffold improved
polymerization control due to better resonance stabilization of the opened radical.
The RROP of 5,6-benzo-2-methylene-1,3-dioxepane (BMDO) was nearly quantitative to
the ester, and yields a resonance-stabilized primary radical after ring opening (Schemes 2a and 2b). BMDO was successfully homopolymerized by ATRP, RAFT, and NMP to moderate chain
lengths with decent control.66–68 Similar to other CKAs, copolymerization of BMDO and most methacrylates have a large
difference in reactivity ratios which favor homopolymerization of the methacrylates.69 The copolymerization of pentafluorophenyl methacrylate (PFMA) and BMDO is an exception
due to the electron-withdrawing nature of the pentafluoro group in the side chain.70 The large difference in electronics between electron-rich BMDO and electron-poor
PFMA led to a near-alternating sequence.
Scheme 2 | (a) CKAs polymerized by RDRP include MDO, BMDO, MPDL, and MPDO. (b) RROP of BMDO.
The ATRP and NMP of oligo(ethylene oxide) methacrylate (OEOMA) with a 2-methylene-4-phenyl-1,3-dioxolane (MPDL) CKA with a five-membered ring and a phenyl substituent had better living character than MDO and BMDO.71 The reactivity of MPDL was still less favorable than polymerization of OEOMA. However, only a small fraction of degradable monomer might be necessary to induce a large change in polymer topology. P(OEOMA-co-MPDL) statistical copolymers with low mol fraction of MPDL (FMPDL) of 0.036 had a ∼30% reduction in molecular weight (Mn) after hydrolysis against 5% potassium hydroxide in 24 h.72 Further increasing FMDPL to 0.113 and 0.248 resulted in ∼80% and ∼95% reductions in Mn, respectively.73 P(MPDL-co-OEOMA) copolymers had a similar degradation profile to hydrophobic poly(lactic acid) and poly(caprolactone) (PCL) polyesters over a one-year timeframe under physiological conditions.74 Copolymerization of MPDL and ethyl maleimide (EMA) had an alternating sequence.75 The preference for alternation was also attributed to electronics, where the MDPL was the electron-rich donor which had preference for addition to the electron-deficient EMA acceptor. Copolymerization of 5-Methylene-2-phenyl-1,3-dioxolan-4-one (MPDO) with methyl methacrylate (MMA) or styrene led to copolymers with a higher MPDO content than originally in the feed, due to its captodative structure. However, copolymerization of MPDO predominantly proceeded by 1,2-vinyl addition rather than ring opening.76
Incorporation of thioesters via RROP
Thioester functional groups are an attractive alternative to polyesters due to their
lower reactivity and possibility for enhanced degradability under mild conditions.
Thioesters were installed by RROP RAFT of macrocyclic dibenzo[c,e]oxepane-5-thione
(DOT) thionolactones with n-butyl acrylate (BA) and t-butyl methacrylate (tBA) comonomers (Scheme 3).77 Characterization of the purified polymers confirmed quantitative ring opening of
the thionolactone. Copolymerization of BA with DOT reached near quantitative consumption
of thionolactone at loadings below 5 mol %. However, lower incorporation of DOT was
observed in copolymerizations with high initial loadings of DOT. Alternating copolymers
of thionolactone and maleimide were prepared by RAFT copolymerization with N-methylmaleimide, N-phenylmaleimide, and N-2,3,4,5,6-pentafluorophenylmaleimide.78 Similar to the polyesters, thionolactone functional groups can be degraded by exposure
to sodium methoxide and cysteine methyl ester.
Scheme 3 | Proposed mechanism for the RROP of DOT.255
Radical ring opening polymerization of macrocycles
Macrocyclic monomers with low ring strain were polymerized and copolymerized by RAFT
radical ring-opening polymerization. Monomers with ester, disulfide, and thioester
functional groups were reported in the literature (Scheme 4a).79 Macrocyclic monomers were designed such that propagation through the allylic sulfide
would readily open the macrocycle through beta-scission to produce a new carbon–carbon
double bond and thiyl radical capable of addition to other monomers (Scheme 4b).79,80 The macrocycles were copolymerized with MMA, N,N-dimethylaminoethyl methacrylate (DMAEMA), 2-hydroxyethyl methacrylate (HEMA), and
2-hydroxypropyl methacrylate (HPMA).79,81 The copolymers were degraded in solutions of sodium methoxide in tetrahydrofuran
(THF),79 and the disulfide-containing polymers were degraded by reduction upon exposure to
a tris(2-carboxyethyl)phosphine reducing agent.81 Scheme 4 | (a) Allylic sulfide macrocyclic monomers polymerized by RAFT copolymerization in the
literature. (b) Mechanism of RROP of allylic sulfide macrocyclic monomers. (c) RROP
by radical cascade ring opening polymerization of an allylic sulfone macrocyclic monomer.
The polymerization of macrocyclic allyl alkylsulfone monomers proceeded through a radical cascade process starting with β-elimination of alkylsulfone, followed by α-scission and liberation of gaseous SO2, providing a 2-propionate radical capable of propagation (Scheme 4c).82 The macrocycles had nearly ideal reactivity with acrylic and acrylamide monomers due to the similar structure of the propagating radical after ring opening and loss of SO2.83 The ideal reactivity between these monomers and vinyl comonomers is a significant improvement over other RROP methods because the degradable bonds should be located roughly the same number of repeat units apart. This is anticipated to yield well-defined oligomers after degradation under the appropriate conditions.
RDRP of degradable vinyl monomers
Some vinyl polymers can be degraded to lower molecular weight oligomers via photoinduced
or catalyst-activated bond scission. Poly(vinyl ketone)s are a unique class of materials
capable of degradation under ultraviolet (UV) irradiation through Norrish Type I and
Norrish Type II reactions upon exposure to UV light, resulting in a degradation of
polymer into shorter oligomers.84,85 Methyl vinyl ketone (MVK) and phenyl vinyl ketone (PVK) were polymerized by RAFT
and subsequently degraded by photolysis (Figure 3).85,86 Multiblock copolymers of poly(vinyl ketones) with nondegradable polymers provided
materials with selectively etchable blocks. The morphology of polystyrene-b-poly(methyl vinyl ketone) thin films changed after etching the poly(methyl vinyl
ketone) block by UV light exposure.86 The modulus of triblock copolymers with a poly(n-butyl acrylate) middle block and PVK outer blocks decreased after selectively etching
the PVK hard blocks.87 Figure 3 | Structure of MVK and PVK.
RAFT copolymerization of methyl α-chloroacrylate (MCA) and MMA provided copolymers
with chlorine functionalities along the backbone.88 The chlorine functionality remained intact through the RAFT polymerization, and resembled
a α-chloroisobutyrate functionality along the backbone. Activation of the chlorines
by ATRP catalysts generated acrylic midchain radicals, which were proposed to degrade
through beta-scission and midchain cleavage of the copolymer along the backbone (Schemes 5a and 5b). The macroinitiator degraded from a starting Mp = 14,900 (Ð = 1.62) to oligomers of Mp = 2000 (Ð = 1.32) upon activation with FeCl2 and the tributylamine cocatalyst.
Scheme 5 | (a) RAFT polymerization of MMA and MCA. (b) Degradation of a P(MMA-co-MCA) copolymer
by activation of chlorine bonds along the backbone with an ATRP catalyst, leading
to beta scission into lower molecular weight oligomers.
Degradable Polymers from the Backbone via the Use of Bifunctional Initiators
Linear polymers with one degradable bond at the center of a linear polymer could be
prepared by RDRP using bifunctional initiators.89 The use of a bifunctional initiator enables growth of a polymer on both sides of
the degradable bond. Thus, degradation of the central linkage via chemical or external
stimuli cleaves the high molecular weight polymer into two linear polymers with half
the molecular weight of the precursor (Scheme 6). A diverse library of degradable bonds were installed in the middle of two ATRP
initiators (Figure 4).
Scheme 6 | Scheme of an RDRP using a red bifunctional initiator installs the degradable bond
at the middle of the polymer chain. Exposure to chemical or external stimuli cleaves
the polymer into two polymers with half the molecular weight of the precursor. The
degraded halves can reform the original polymer if the reaction is reversible.
Redox degradable polyacrylates and PSs were prepared by ATRP with bifunctional 2-bromopropionic
acid and 2-bromoisobutyrate diesters of bis(2hydroxyethyl) disulfide initiators.90,91 The ATRP reaction can be performed without reduction of the disulfide bond, providing
high molecular weight polymers with one central disulfide bond. The disulfide linkage
could be reversibly reduced by dithiothreitol (DTT) or tributylphosphine to yield
linear polymers with half the molecular weight of the original material. The degraded
mercapto-functional oligomers could be recoupled after oxidation in the presence of
weak oxidizing agents, such as FeCl3 or iodine, to reform the high molecular weight polymer linked by the disulfide bonds.
Figure 4 | Bifunctional ATRP initiators with degraded products and conditions required to cleave
the degradable bond.
The method of disulfide reduction and oxidation can affect the efficiency of the self-healing
reaction. Reductive cleavage/nucleophilic substitution of a poly(methyl acrylate)
(PMA) with a central disulfide bond and terminal bromine functionality with diethylamine
base and DTT reduces the disulfide bond and installs an additional thiol at the chain-end
by substitution of the bromine. Thus, oxidation of telechelic α,ω-bis(thiol)-functionalized
PMA with I2 leads to step-growth polymerization by bridging disulfide bonds (Figures 5a and 5b).92 Reduction of the disulfide bond over Bu3SnH and Azobisisobutyronitrile (AIBN) saturates the bromine chain-end while reducing
the disulfide bond, yielding two saturated chains with terminal thiol functionality.
Oxidation of the polymer produced a new material with comparable molecular weight
to the precursor.
Figure 5 | (a) Reaction scheme and gel permeation chromatography (GPC) traces illustrating the
reduction of a polymer with a central disulfide bond with diethylamine as base and
DTT as reducing agent, followed by oxidation to a higher molecular weight polymer
containing multiple disulfide bonds. (b) Reaction scheme and GPC illustrating the
reduction and hydrogenolysis of the same polymer by Bu3SnH, followed by subsequent oxidation to disulfide polymer, then reduction with DTT
in dimethyl sulfoxide. (Mn determined by GPC; values in parentheses determined by 1H NMR integration). Reproduced with permission from ref 92. Copyright 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Furfuryl-maleimide Diels–Alder (DA) functionalities enabled linear polymer degradation after thermal and mechanical treatment. Maleimide functionalized PMA was prepared by ATRP of methyl acrylate with a maleimide functionalized α-bromoisobutyrate initiator.93 DA [4+2] cycloaddition between ω-maleimide-PMA and a ω-furan-poly(ethylene glycol) produced diblock copolymers with the furan-maleimide adduct at the center. The retro-Diels Alder reaction at 120 °C cleaved the adduct into the same ω-furan-poly(ethylene glycol) and ω-maleimide-PMA starting materials.93 Ultrasonication of PMA with the same DA adduct led to mechanochemical retro-DA of the polymer into two PMA chains with half the molecular weight of the original polymer.94
Multiblock Backbone Scission via RDRP then Step-Growth Polymerization
The chain-end functionality of polymers prepared by RDRP enables installation of various
functional groups by substitution chemistries. Multiblock copolymers can be prepared
by post-polymerization modification of bifunctional polymers into macromonomers, followed
by step-growth polymerization, to yield high molecular weight linear polymers (Scheme 7). The use of reversible, or degradable, bonds in the step-growth polymerization enables
degradation from high to lower molecular weight upon exposure to the appropriate chemical
or external stimuli.
Scheme 7 | RDRP with a bifunctional initiator or CTA installs end-group functionality (blue)
at both ends of the polymer. Modification of the end groups (red dots) can enable
step-growth polymerization to high molecular weight, with degradable bonds placed
between each macromer. Exposure to chemical or external stimuli can cleave the polymer
into linear polymers with comparable molecular weight to the precursor.
Multisegment degradable polymers were prepared by radical trap-assisted atom transfer radical coupling (RTA-ATRC) of halogenated styrenic and (meth)acrylic polymers with nitroxide coupling agents.95 ATRC is a method of coupling polymer chains by preference of crosstermination by combination.96,97 Step-growth ATRC of α,ω-dihalogenated PS was performed in the presence of a bifunctional (2,2,6,6-Tetramethylpiperidin-1-yl)oxyl (TEMPO) adduct with, and without, a central disulfide bond.95,98 PS prepared by ATRC was degraded thermally in the presence of excess TEMPO, due to exchange between the bifunctional crosslinker and the higher concentration of monofunctional TEMPO. High molecular weight polymers were also prepared by step-growth polymerization of a bifunctional 4-phenylene bis(2-bromoisobutyrate) ATRP initiator with a bifunctional TEMPO adduct by radical coupling between activated chain ends and TEMPO radical traps.99 Thermogravimetric analysis of the poly(alkoxyamines) showed considerable mass loss below 200 °C, suggesting poor thermal stability caused by scission of the initiator-alkoxyamine bonds.
Nitroso radical traps were also used to install thermally cleavable alkoxyamine functionalities
onto styrenic and acrylic polymers by ATRC.100 The coupling of polymer chains with nitroso compounds proceeds in two steps (Scheme 8). In the first step, the polymeric chain-end radical reacts with the nitroso functional
group to produce a polymer with nitroxide functionality. The chain-end nitroxide can
trap a second polymeric radical to produce alkoxyamine-linked polymer topologies.
Thermolysis of the alkoxyamine cleaves the bond and produces two polymer chains with
similar Mn and Ð as the starting material. This approach was first outlined in the RTA-ATRC
of a monobrominated PMMA-Br precursor with a nitrosobenzene radical trap.100 Brominated PS and PMA were also coupled by RTA-ATRC using nitrosobenzene radical
traps, and later work applied the concept to multiblock copolymers.101 Under dilute conditions, the coupling reaction produces cyclic polymers which can
be degraded back to linear polymers by thermolysis.102 Scheme 8 | ATRC of bifunctional PS with nitrosobenzene
Nucleophilic substitution of halogen chain-end functionality (CEF) can install thiol
linkages on telechelic α,ω-dibromo vinyl polymers.90,103 Telechelic polymers with thiol functionalities could also be prepared by aminolysis
of trithiocarbonate (TTC; Scheme 9)104 or xanthate RAFT CTAs.105 Oxidative coupling of linear α,ω-dithiol polymers produces high molecular weight
linear polymers connected by disulfide bonds, which can be reversibly degraded to
the low molecular weight telechelic α,ω-dithiol pre-polymer after reduction.
Scheme 9 | Aminolysis of RAFT chain ends enables the synthesis of bifunctional PS with thiol
chain ends. Two thiols can couple after oxidation to disulfide bonds to yield high
molecular weight PS. The polymer could be degraded back to low molecular weight after
reduction of the disulfide bonds.
Oxidative step-growth polymerization of α,ω-bis(thiol) polymers enabled the synthesis of degradable random multiblock copolymers. Multiblock telechelic copolymers were prepared by oxidative coupling of α,ω-bis(thiol)-poly(styrene)-b-poly(n-butyl acrylate)-b-poly(tert-butyl acrylate).105 Oxidative coupling of α,ω-bis(thiol) poly(N-isopropyl acrylamide) (PNIPAM) (Mn = 7860, Ð = 1.24) and α,ω-bis(thiol) poly(2-(dimethylamino)ethyl methacrylate) (PDMAEMA) (Mn = 7780, Ð = 1.18) at a 1:1 weight ratio produced a high molecular weight (Mn = 180,000 and Ð = 5.1) temperature and pH-responsive multiblock copolymer.106 The multiblock copolymers were degraded to lower molecular weight by reduction with DTT.
Block Copolymers
Degradable functionalities can also be installed using a degradable macroinitiator.
This has often involved the use of polymer macroinitiators prepared by ROP, such as
poly(lactic acid), polycaprolactone, polypeptides, or polyphosphates as the degradable
block (Scheme 10).107–109 Scheme 10 | Use of a degradable macroinitiator (red) enabled the synthesis of a block copolymer
with one degradable block (red) and one nondegradable vinyl polymer block (black).
The degradable block can be selectively etched to yield the nondegradable vinyl polymer.
Multifunctional initiators can be used to initiate ring-opening polymerization and controlled radical polymerization.110,111 Block copolymers of vinyl polymers and polyesters were prepared in one pot using a hydroxyl-functionalized TTC CTA as a dual RAFT and ROP initiator.112 Block copolymers of epoxides and acrylamides were prepared by a photoiniferter/organocatalyst dual catalysis approach using a hydroxyl-functionalized TTC as the ROP/RAFT initiator.113
Degradable Hyperbranched Polymers
Hyperbranched polymers are polymers with irregular branched topologies.114–117 Branching reduces the hydrodynamic size and entanglement between molecules in solution and the melt, resulting in reduced viscosity and a higher local concentration of functional groups relative to a linear polymer. These materials are commonly used in industrial coatings, resins, and adhesives as well as high performance applications like drug delivery and therapeutics.
Synthesis of hyperbranched polymers by self-condensing vinyl polymerization
Hyperbranched polymers can be prepared by polymerization, and copolymerization, of
vinyl “inimers” with initiator and monomer functional groups in a self-condensing
vinyl polymerization (SCVP). The degree of branching in a SCVP by ATRP is partially
determined by the difference in the activation-rate constant for the inimer (ka,s) and backbone (ka,b) dormant species (Figure 6).118 Predominantly linear polymers are produced if the activity of the dormant backbone
is larger than the activity of the inimer (ka,s << ka,b), or vice versa in the opposite case (ka,s >> ka,b). Comparable activity between inimer and backbone initiators can provide densely
branched copolymers.
Figure 6 | Possible structures of polymers derived from an inimer with a degradable functionality
and products of their complete degradation. Figure adapted with permission from ref 118. Copyright 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Thus, the size of a degraded hyperbranched polymer prepared by ATRP of a degradable inimer should be tunable by the reactivity of the inimer and backbone, respectively (Figure 6).118 Degradation of a hyperbranched polymer prepared by polymerization of a degradable inimer with a higher activity backbone and low activity initiator (ka,s << ka,b) yields a polymer with the majority of degradable bonds in the side chains and will not degrade to a significantly lower molecular weight. A degradable hyperbranched polymer prepared from inimers with comparable backbone and initiator reactivity will yield a highly branched polymer that can be degraded to lower molecular weight linear polymers (ka,s ∼ ka,b). Finally, polymerization of a degradable inimer with high initiator reactivity and low backbone reactivity produces mostly linear polymers with degradable bonds along the backbone. Degradation of this polymer produces low molecular weight fragments, analogous to the polymerization and degradation of vinyl chloropropionates by atom transfer radical addition.119
Highly branched degradable polymers were prepared by polymerization of inimers with redox-active disulfide and acid-sensitive acetal functionalities.118,120 Copolymerization of 2-(2-bromopropionyloxy)ethyl acrylate with styrene provided densely hyperbranched copolymers due to the large difference in activation between the initiator and backbone dormant species (i.e., ka,s >> ka,b). A disulfide-containing methacrylic inimer copolymerized with MMA provided well-defined branched copolymers with a lower degree of branching due to similar reactivities between the backbone and side chain (i.e., ka,s ∼ ka,b).118 The topology of the branches was confirmed by degradation of the inimer and analysis of the low molecular weight polymer by GPC.
A segmented hyperbranched poly(methyl methacrylate) (PMMA) copolymer with a furan-maleimide branch junctions was synthesized by SCVP by ATRP of MMA with a DA inimer.121 The retro-DA reaction at 120 °C cleaved polymer arms into linear PMMA segments with statistical distributions of furan in the side chains and maleimide chain ends. The PMMA segments could be grafted to single-chain nanoparticles by DA cycloaddition between the dissociated furan and maleimide macroinitiators.
Synthesis of hyperbranched polymers by (co)polymerization of a degradable crosslinker
Degradable branched polymers and single chain nanoparticles can also be prepared by
copolymerization of monofunctional vinyl monomers with degradable crosslinkers (Schemes 11a and 11b). The topology of the material and location of the degradable bonds is dictated by
the polymerization recipe. The copolymerization of MMA and HPMA with a disulfide-based
dimethacrylate (DSDMA) via RAFT-produced high molecular weight hyperbranched polymers
bridged by disulfide bonds between branch points.122–124 Reduction of the disulfide bonds degraded the high molecular weight branched polymer
to short oligomers.
Scheme 11 | Degradable hyperbranched polymers can be prepared by RDRP through (a) copolymerization
of a degradable crosslinker (red) with a spacer monomer (black), or (b) coupling of
multifunctional linear arms into randomly crosslinked/branched nanoparticles with
degradable bonds. The highly branched polymers have degradable bonds located at the
junctions of the branches. Thus, degradation of the branches provides linear polymers
of lower molecular weight. Branched polymers can be reformed in some cases.
Segregation of the polymerization into an immiscible phase, via RDRP in an emulsion or miniemulsion can yield densely branched nanoparticles. OEOMA was polymerized by activators regenerated by electron transfer ATRP in the presence of a disulfide crosslinker using the inverse miniemulsion technique, then chain-extended with styrene.125,126 The block copolymer nanoparticles were degraded to linear PS-b-P(OEOMA) block copolymers which self-assembled into micelles with a PS core and POEOMA corona in water. Degradable OEOMA nanogels had promising biomedical properties due to an affinity to biotin and a fast drug release profile in the presence of glutathione.127
The rate of branched polymer degradation varies by the crosslinker structure. Hyperbranched
polyacrylates with imine crosslinks hydrolyzed faster when less stable hydrazone functional
groups were installed in the branches, compared to more stable oxime functional groups.128 Mechanochemical scission of hyperbranched polymethacrylate arms with weaker disulfide
and sulfone functionalities at the junctions was ∼2 times faster than mechanochemical
scission of hyperbranched polymers with alkyl and ether crosslinks (Figure 7).129 Figure 7 | Mechanochemical degradation of single chain nanoparticles prepared by RAFT copolymerization
of MMA with 2-(methacryloyloxy)ethyl acetoacetate (Mn ∼ 100,000) followed by postpolymerization Michael addition of pendant 1,3-diketones
with divinyl crosslinkers. The fragmentation rate scaled with the calculated force
to break the crosslinker where weaker disulfide and sulfone bonds broke faster than
the stronger ether and carbon-carbon bonds. Reproduced with permission from ref 129. Copyright 2020 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Star Polymers
Polymer stars are branched polymers with a defined length and number of arms extending from a central core.130–132 Similar to other branched polymers, star polymers have a reduced molecular size which leads to less entanglement and lower viscosity in the melt and in solution.133 Star polymers are generally more well defined than hyperbranched polymers and could be used for similar applications in biomedicine, energy, and nanotechnology. Star polymers can be prepared by the “core-first” and “arm-first” approaches.130–132
Star polymers by the “core-first” approach
The core-first approach involves the synthesis of a multifunctional polymer core as
the macroinitiator, followed by polymerization of monomers from the core (Scheme 12). The core-first approach enables the synthesis of polymer stars with a precise number
and length of polymer arms. However, core-first synthesis by RDRP can lead to coupling
and eventual macroscopic gelation if a significant portion of stars terminate via
biradical combination. Polymer stars can be degraded to linear polymers if the multifunctional
core contains degradable functional groups.
Scheme 12 | Core-first synthesis of a star polymer. The use of a multifunctional degradable polymer
core enables cleavage of the polymer arms from the core. Some functional groups enable
association of degraded star arms to reform polymer star upon exposure to external
stimuli or via postdegradation modification.
Biobased α-d-glucose and β-cyclodextrin were used to prepare multifunctional star cores by base catalyzed esterification with 2-bromosiobutyryl bromide.134,135 A similar approach installed ATRP initiators onto PCL which was used as a star core in the synthesis of degradable core-shell polymers.136 Star polymer arms can be grafted from the cores via ATRP and cleaved from the core by transesterification of the ester with a strong acid catalyst. Star polymer arms grafted from tannic acid cores can be degraded in a mild methanol/bicarbonate solution due to the poorer stability of the phenyl ester bond tethering the arms to the core.137
Multifunctional CTAs with disulfide links were used to prepare redox degradable polymer
stars via the core-first RAFT approach (Figure 8).139,140 The cores were prepared such that the disulfide bonds and “R” groups of the CTA were
inside of the core, and the radical stabilizing Z-groups were on the outside. This
enabled polymerization from the core without dissociation of the arms. The linear
arms were cleaved from the star core after reduction of the disulfide bonds.
Figure 8 | A multifunctional RAFT agent was used to prepare polymer stars with degradable cores
by the transfer-from core-first approach. The RAFT CEF was reduced, then functionalized
with β-cyclodextrin at the arms. Reproduced with permission from ref 138. Copyright 2009 American Chemical Society.
Degradable star polymers by the “arm-first” approach
The arm-first approach involves the synthesis of a (block) copolymer arm (Scheme 13a), followed by chain extension with a divinyl crosslinker (Scheme 13b). This approach provides precise control over the length of polymer arms and is less
likely to reach macroscopic gelation than the core-fist approach. However, the method
provides less control over the number of arms in the final product. Miktoarm polymer
stars can be prepared by chain-extension from the polymer core via an “in-out” approach.
Scheme 13 | Degradable polymer stars can be synthesized via the arm-first approach by chain extension
with a degradable crosslinker. The degradable bonds at the core enable a change in
topology back to low molecular weight linear arms after scission. The cleaved arms
can be reconnected in some cases.
The use of a degradable crosslinker enables the synthesis of polymer stars with degradable
cores via the arm-first approach. Degradable divinyl crosslinkers with acetal, imine,
DA, and disulfide functional groups were reported in the literature (Figure 9). The arm-first method could also involve coupling of a functionalized-nonfunctionalized
diblock copolymer via other chemistry. Degradation of the polymer core liberates the
linear polymer arms (Scheme 13c). In some cases, the cleaved polymer arms could reassociate back to form the original
polymer star (Scheme 13d).
Figure 9 | Common degradable crosslinkers reported in the literature.
Redox-degradable star polymers were prepared by chain-extension of PMMA and PS-co-PMMA
block copolymers with a redox-degradable disulfide crosslinker via ATRP.141 Redox-degradable star polymers are promising for biomedical applications. Stars with
oligo(ethylene oxide) arms and cationic cores complexed with negatively charged siRNA,
which allowed for cellular internalization of siRNA using the polymer scaffold as
a biocompatible delivery vehicle (Figure 10).142 Cationic nanogels crosslinked with disulfide crosslinks were also effective at nucleic
acid delivery.143 Star polymers with POEOMA-b-PBMA arms self-assembled into nanocapsules with a diameter of 140∼195 nm after crosslinking
with a bis(2-methacryloyloxyethyl)disulfide (DSDMA) crosslinker.144 Figure 10 | Synthesis of star copolymers with PEG arms and a degradable cationic core by ATRP
through the arm-first approach. The star polymers complexed with siRNA. Reproduced
with permission from ref 142. Copyright 2011 American Chemical Society.
Copolymerization of two macroinitiators with a crosslinker yields mikto-arm polymer stars. ATRP and ROP were used to prepare a library of hydrolytically degradable star polymers by the arm-first approach.145 Copolymers were prepared by ROP of caprolactone from a 2-hydroxyethyl α-bromoisobutyrate initiator, followed by arm-first crosslinking by ATRP with divinyl benzene (DVB). Conversely, PMMA or PS arms could be prepared first by ATRP, then crosslinked by ROP with a bifunctional lactone, to yield star polymers with degradable cores. This approach was later expanded to the synthesis of core-shell block copolymers with PCL-b-PMMA arms.146 Block copolymer stars with PCL blocks on the outside were crosslinked by ATRP with ethylene glycol dimethacrylate (EDGMA) and could be degraded back to PMMA polymer stars via hydrolysis. Conversely, crosslinking the PCL blocks by ROP with a bifunctional lactone-produced star polymers with PCL cores.
Thermally degradable polymer stars were prepared by chain-extension of a polymer macroinitiator with a thermally unstable crosslinker. Polymer stars with 4,4Ȳ-azobis(4-cyanovaleric acid) crosslinks in the core irreversibly degraded upon exposure to high temperature and light.147 Temperature-responsive arms crosslinked by DA bonds reversibly aggregated into 18 nm aggregates and dissociated to 6 nm unimers by DA and retro-DA exchange.148
Reversible detachment of polymer arms from a star was also accomplished with alkoxyamine exchange chemistry. The arm-first synthesis of polymer stars using PMMA-TEMPO macromonomers via NMP with DVB was reported.149 The PMMA arms were designed such that the TEMPO adduct remained attached to PMMA after decomposition. Thus, heating the star polymer at 100 °C with a stoichiometric excess of alkoxyamines cleaved PMMA arms from the core by dynamic-covalent exchange. PS polymer stars were also crosslinked by dynamic-covalent exchange of alkoxyamine units in the side chain.150,151 Reflux of two complementary block copolymers liberated the nitroxide radical from one polymer backbone and produced nonfunctional PS radical in the other backbone. Trapping of the functional PS radicals in the side chain enabled reversible dynamic-covalent assembly of polymer stars at 100 °C. Mikto-arm star polymers were also prepared by heating the polymer stars in the presence of another alkoxyamine functionalized polymer.
Bottlebrushes
Molecular bottlebrushes are densely grafted molecules which consist of side chains attached to a polymer backbone. The steric repulsion between the side chains elongates the polymeric backbone and side chains, such that densely grafted molecular bottlebrushes with a high aspect ratio exist in rod-like conformations. Bottlebrushes have reduced viscosity and suppressed chain entanglement, which make them useful for potential applications in lubrication, soft materials, and biomedicine.21,152,153
The intrinsic tension of a densely grafted bottlebrush with side chains attached to every repeat unit in the backbone was on the order of f0Nα, where N is the side chain length and α = 3/8 in good solvents, 1/3 in theta solvents, and f0 in poor solvents and melts.154 The tension of a fully extended molecular bottlebrush in the melt is on the order of 1 pN. Tension increases when the polymer adsorbs to an attractive substrate, as the monomeric units try to maximize interaction with the surface despite the steric congestion and tension already present in the backbone and side chains.155,156 The attractive forces amplify the tension of the bottlebrushes up to the nN range.155
Scission of carbon–carbon bonds was observed when molecular bottlebrushes were adsorbed
onto attractive liquid and solid substrates (Figure 11).157 Atomic force microscopy (AFM) of poly(n-butyl acrylate) (PBA) bottlebrushes on a water/isopropanol surface showed a reduction
in contour length with time spent on the surface.157 The rate of backbone C–C bond scission was dependent on the substrate surface energy.
The rate of bond scission was six times faster when the surface energy of the substrate
was increased from 69.2 to 71.2 mN/m.158,159 Bond scission in PBA bottlebrushes had anti-Arrhenius behavior due to the decrease
in surface energy of the substrate upon heating, which resulted in an overall decrease
in tension of the adsorbed macromolecules.160,161 Figure 11 | (left) Height AFM micrographs of molecular bottlebrushes with long side chains measured
at different exposure times after adsorption on a water/propanol (99.8/0.2 w/w%) substrate.
(right) The number average contour length decreases with increasing exposure time
(white circles); the solid line is a fit to the experimental data assuming bond scission
as a first-order reaction. The experimentally determined dispersity (PDI) initially
increases and then decays and is also in agreement with the computer simulations (dashed
line). Reproduced with permission from ref 157. Copyright 2006 Springer Nature Limited.
The selectivity of molecular bottlebrush bond scission after adsorption on a surface was improved by installation of a weaker disulfide bond in the middle of the backbone.162 Similarly, spoked-wheel molecular bottlebrush star polymers showed preferential scission of the ester-linked arms before the scission of carbon–carbon covalent bonds in the bottlebrush backbone.163
Outside of the intrinsic tension in molecular bottlebrushes, degradable bonds can
be installed in the backbone and side chains, analogous to polymer stars. Molecular
bottlebrushes can be prepared by the grafting-through, grafting-from, and grafting-to
approaches (Scheme 14).
Scheme 14 | The three synthetic strategies used to prepare molecular bottlebrushes are the grafting-through,
grafting-to, and grafting-from approaches.
Degradable bottlebrushes by the “grafting-through” approach
The grafting-through approach involves the polymerization of “macromonomers,” which are oligomers with one polymerizable functional group.164–166 The grafting-through synthesis of bottlebrushes by RDRP is often limited by the dilute repeat unit concentration and high viscosity of the macromonomers. This lowers the bulk repeat unit concentration, leading to poorer kinetic control due to the low concentration of CTA and catalyst required to mediate polymerization and a lower ceiling temperature (Tc).25 The low Tc of methacrylic poly(macromonomer)s enables depolymerization of methacrylic molecular bottlebrushes back to macromonomers at elevated temperatures by RDRP mechanisms.167–169 The depolymerization of polymers by RDRP will be discussed in detail at the end of this review.
Degradable bottlebrushes by the “grafting-from” approach
The grafting-from approach involves the polymerization of small molecule monomers from a multifunctional backbone macroinitiator.170–172 The grafting-from approach enables the synthesis of molecular bottlebrushes with long backbones and high graft densities. However, grafting-from by RDRP may lead to coupling and macroscopic gelation if bimolecular combination is not adequately suppressed.171,172 Bottlebrush arms can be removed from the backbone by degradation of the backbone or by postpolymerization modification if they are attached by a degradable functional group. Degradable polymer backbones can enable a transition from bottlebrush to linear topology, with the molecular weight of the degraded product closely matching the separated side chains and backbone.
Side chains grafted from N-carboxy anhydride and cellulose backbones can be released
by hydrolysis along the polymer backbone.173,174 Poly(n-butyl acrylate) side chains grafted from poly(2-bromoisobutyryl ethyl methacrylate)
macroinitiator backbones are commonly etched from the backbone by transesterification
with butanol to calculate initiation efficiency (Scheme 15).172,174–176 Weaker silanol moieties installed along the backbone enabled hydrolysis of zwitterionic
poly(2-methacryloyloxyethyl phosphorylcholine) side chains from bottlebrush backbones
under milder conditions.177 Scheme 15 | Removal of side chains from a poly(2-bromoisobutyryl)oxyethyl methacrylate-graft-n-butyl
acrylate) bottlebrush by acid-catalyzed transesterification with butanol.
Degradable bottlebrushes by the “grafting-to” approach
The grafting-to approach involves a complementary reaction between a multifunctional backbone and polymeric side chains with one pendant functionality.178,179 An analogous approach to bottlebrush degradation can be accomplished if the grafting-to method relies on dynamic covalent, or degradable, functional groups to attach the arms to the backbone.
PS bottlebrushes with UV-cleavable side chains were prepared by grafting linear PS with terminal nitrobenzyloxycarbonyl and alkyne functionality onto a poly(3-azido-2-hydroxypropyl methacrylate) backbone by CuAAC.180 The side chains were cleaved from the polymer backbone after ten minutes of UV light exposure, which led to a decrease in the intrinsic viscosity ([η]), from a starting [η] = 12 mL/g to a final [η] = 7.39 mL/g.
ATRP of an alkoxyamine functional methacrylate and MMA at low temperature was used
to prepare a polymer backbone with dormant alkoxyamine functionalities.181 PS side chains were grafted onto the backbone via an alkoxyamine exchange reaction
between the alkoxyamines on the backbone and a stoichiometric excess of alkoxyamine-capped
PS (Figures 12a–12c). The side chains were removed from the backbone by heating them at 100 °C with an
excess of small molecule alkoxyamine.
Figure 12 | (a) Scheme of reversible synthesis of PMMA-g-PS bottlebrushes by the grafting-to alkoxyamine
exchange method, and side chain removal at 100 °C. (b) Dependence of Mn on reaction time for (circle) polymer reaction of the backbone (Mn = 11,800, Ð = 1.18, 92 mg) with the TEMPO-capped PS (Mn = 1700, Ð = 1.15, 918 mg, 5.0 equiv/alkoxyamine units) in anisole (1 wt % polymer
solution) at 100 °C. Triangles correspond to side chain removal of the fractionated
bottlebrush (Mn = 24,000, Ð = 1.16, 50.2 mg) with an excess if small molecule alkoxyamine (73 mg,
8.3 equiv/alkoxyamine units) in anisole (1 wt % polymer solution) at 100 °C. (c) GPC
traces for the removal of PS side chains by heating in the presence of a stoichiometric
excess of alkoxyamine. Reproduced with permission from ref 181. Copyright 2004 American Chemical Society.
Networks
Degradable polymer networks can be designed to contain degradable bonds placed at the junctions or in the middle of the mesh. The controlled radical copolymerization of vinyl monomers with a degradable crosslinker installs the degradable functional group only at the junctions of a network. Thus, degradation of the network yields mostly linear polymers because all branching points should contain the degradable bond. Multifunctional initiators containing degradable bonds could install degradable bonds at the center of a mesh if the initiator is bifunctional or at the junction if polymer stars are linked via irreversible coupling chemistry.
Redox degradable polymer networks were prepared by copolymerization of acrylic monomers
with a bis(2-methacryloyloxyethyl) disulfide (DSDMA) crosslinker.91 Similar to other redox degradable materials, reduction of the disulfide bonds led
to dissolution of polymer gels into soluble linear polymers which can be characterized
by gel permeation chromatography (Figures 13a–13c).182 The degraded thiol-functionalized polymers can be cured into polymer networks by
reformation of disulfide bonds.183 Segregation of the disulfide crosslinkers in an immiscible phase enabled the synthesis
of redox-degradable latex nanoparticles and high internal phase emulsion polymers.184–186 Figure 13 | Poly(hydroxyethyl acrylate) polymer networks were prepared by RAFT and ATRP copolymerization
with a bifunctional 2,2′-dithiodiethanol diacrylate (DSDA) crosslinker. The crosslinks
were reduced by DDT to liberate and dissolve linear chains within the mesh. The dissolved
polymers were comparable for RAFT and ATRP with (a) DP = 100 and (b) DP = 200 mesh,
but RAFT networks had a higher dispersity at a (c) DP = 500. GPC of the sol showed
comparable internal structure of polymer networks with primary chain lengths of DP = 100
and 200. However, the cleaved polymers from the ATRP network at a higher primary chain
length of 500 had lower Đ than the RAFT counterpart. The difference between the primary
chain lengths was observed as a higher swelling ratio and lower moduli in the DP = 500
mesh size ATRP networks. Reproduced with permission from ref 182. Copyright 2021 American Chemical Society.
Degradable model networks were synthesized by ATRP and click chemistry using two distinct
approaches. The first approach (Figure 14a) involved polymerization from a bifunctional ATRP initiator, followed by azide substitution
of the halogen chain ends, to yield a bifunctional azide functionalized macromonomer
with a photocleavable moiety at its center.188 Copper-catalyzed azide-alkyne click chemistry (CuAAC) with a tetrafunctional alkyne
provided model networks with pore sizes defined by the initial size of the macromonomer
and four-arm star polymer crosslinker. Degradation of the model network under UV irradiation
provided a mixture of star polymer and macromonomer. The second approach (Figure 14b) involves crosslinking of a tetrafunctional polymer star with photocleavable groups
within the core, followed by crosslinking via CuAAC.187,189 Degradation of this network produced linear polymers with half the molecular weight
of the star polymer (i.e., twice the size of the star arms).
Figure 14 | Design strategies used to prepare photodegradable model networks. Reproduced with
permission from ref 187. Copyright 2007 American Chemical Society.
Dynamic covalent exchange of functional groups enables self-healing under select conditions.
The degenerative transfer mechanism of the RAFT process was used to introduce self-healing
behavior in crosslinked polymer gels by reshuffling of TTCs at the junctions under
UV irradiation. Multiple rounds of self-healing and degradation of polymer networks
was accomplished by the photoiniferter exchange mechanism (Figures 15a–15c).190–193 This approach was recently applied to 3D printed networks. Application of a crosslinker
during the self-healing reaction led to an increase in tensile strength after self-healing
due to the increase in crosslinking density.194 Figure 15 | Synthetic scheme and photographs of PBA polymer gels crosslinked with bifunctional
TTCs at the junctions before and after self-healing reactions. The polymers were exposed
to repeated self-healing reactions under UV irradiation in acetonitrile (a) before
and after first reaction for 4 h, (b) before and after second reaction for 12 h, and
(c) after swelling test in anisole for 6 h. Reproduced with permission from ref 190. Copyright 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Copolymerization of a TEMPO-methacrylate, styrene, and a butadiene crosslinker with pendant vinyl functionalities provided polymer networks with dynamic-covalent character through alkoxyamine exchange.195 The high activation energy of the alkoxyamine degradation provides negligible creep at 80 °C but fast exchange at high temperature when dissociation of the alkoxyamine bond becomes favorable. Alkoxyamine exchange enabled reprocessing of materials.196 Model networks cured by RTA-ATRC of polymer stars with nitrosobenzene were degraded into soluble components resembling the original polymer star after decomposition of the alkoxyamines.197
Self-healing character was also introduced into methacrylic polymers via the Diels–Alder reaction. Furan functional groups in the backbone were used to crosslink networks by DA cycloaddition with a bis-maleimide crosslinker at ∼50 °C.93,198 The crosslinks could be reversibly cleaved via retro-DA of the DA adducts above 130 °C.
PMMA vitrimers were prepared by curing of P(MMA-co-AEMA) copolymers with tris(2-aminoethyl)amine via condensation at room temperature.199 The vitrimers had a tensile moduli of 1800 MPa, and exhibited temperature-dependent stress relaxation at elevated temperature, with an activation energy (Ea) of crosslink exchange of 102 ± 8 kJ/mol. The vitrimers retained comparable mechanical properties to the virgin material after five cycles of breaking, followed by reprocessing via compression molding. Segregation of the dynamic-covalent adduct into a separate immiscible block led to reduced macroscopic flow and suppressed creep behavior at longer time scales and larger deformations.200
Biopolymer-Synthetic Polymer Hybrids
Polymers can be grafted from natural products, including cellulose derivatives and biological macromolecules. Wood-derived natural products are an important class of renewable materials which are not derived from fossil fuel feedstocks. These materials are commonly used as polymer backbones in the synthesis of composite materials by surface-initiated RAFT or ATRP, analogous to the grafting-from approach in the synthesis of molecular bottlebrushes or inorganic nanoparticles.
The synthesis of wood-polymer composites by RDRP was recently reviewed and will not be discussed in extensive detail in this article.201–204 Grafting nondegradable polymers from wood-derived materials endowed the biobased materials with properties unique to polymers from petroleum-based feedstocks but with additional structural integrity and degradability. Hybrid composites were prepared by SI-ATRP from diverse substrates such as filter paper,205 starch,206 chitin,207 and hyaluronic acid.208
Molecular poly(n-butyl acrylate) bottlebrushes grafted from a cellulose backbone had degradation rates
which depended on the graft density. Bottlebrushes with low grafting density degraded
to lower molecular weight by scission of the β-1,4-glycosidic bonds along the backbone
while bottlebrushes with high graft density primarily degraded by scission of the
side chains due to steric protection of the backbone by the side chains (Schemes 16a and 16b).174 Scheme 16 | Degradation of poly(n-butyl acrylate) bottlebrushes grafted from a cellulose backbone
were observed to degrade by hydrolysis along the backbone, or cleavage of the side
chains from the backbone. High graft density cellulose bottlebrushes preferentially
favored scission of the side chains from the backbone while low graft density bottlebrushes
degraded more readily along the backbone. Reproduced with permission from ref 174. Copyright 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Biomolecule-polymer conjugates are typically employed to improve the stability, solubility,
and half-lives of therapeutics for biomedical applications.209 Many conjugates are prepared by PEGylation with chain-end functionalized PEG in a
grafting-to approach. However, biomolecule-polymer conjugates prepared by RDRP provide
greater control over the topology, functionality, and degradability of the grafted
polymer chains.210–212 The grafting-from RDRP approach can be used to prepare protein-polymer conjugates
with the polymer side chains extending from a specific location on a protein backbone
(Figures 16a–16f). RDRP was used to graft from biomacromolecules such as proteins, DNA and RNA, and
exosomes.214–221 Biomolecules could be functionalized with pH-responsive, thermoresponsive, or even
multiblock or graft copolymer shells with degradable bonds installed at specific sites
along the polymeric ligands.218,222,223 The initiation efficiency of a grafting-from RDRP from biomacromolecule initiators
was calculated by selective etching of the degradable backbones in solutions of strong
acid or base (Figure 17).217,218,220,222,224–227 Figure 16 | Schematic representation of possible methods used to functionalize a streptavidin
protein with initiators or CTA for subsequent grafting-from synthesis of protein-polymer
hybrids. Initiators could be installed (a) at the N-terminus by transamination of
the primary amine at the N-terminus to an oxoamide, followed by oxime formation with
the initiator; (b) at any location along the backbone by genetically incorporating
a 4-(2′-bromoisobutyramido)phenylalanine ATRP initiator onto a peptide backbone; (c) at
sites along a Streptavidin backbone with strong binding affinity to biotin functional
groups; (d) intein-mediated protein ligation of the C-terminus into a thioester which
could be ligated with an amino-functionalized ATRP initiator; (e) at all amines along
a backbone by amination with an activated ester or acyl halide functionalized initiator;
and (f) at cysteine functional groups by thiolene click chemistry with maleimide functional
initiators,or oxidative coupling of deprotected thiol-functional initiators. Reproduced
with permission from ref 213. Copyright 2014 American Chemical Society.
The addition of a degradable bond between the polymeric ligand and the biomacromolecule
can provide targeted deshielding of the biomolecule upon cleavage of the ligands and
enhanced stability or stealth behavior when the ligand remains bound to the biopolymer.
Polymeric ligands bound to proteins and siRNA by disulfide bonds were released by
reduction to thiols, leaving both unattached linear polymer and intact biopolymers.216,219,228–231 Multiconjugate polymers containing siRNA and folate targeting moieties were prepared
for delivery of siRNA to cancer cells overexpressing folate receptors.232 A portion of the poly[N-(2-hydroxypropyl)methacrylamide-r-(N-(3-aminopropyl)methacrylamide)] copolymer side chains were functionalized with folate
reactive functional groups and thiol-functionalized si-RNA. The RNA side chains were
cleaved from the backbone after reduction with glutathione.
Figure 17 | Grafting-from ATRP of OEOMA from initiator-modified Bovine serum albumin protein (BSA)
and DNA macroinitiators by photo-ATRP under 450 nm irradiation. The biomacromolecule-polymer
conjugates were degraded in 5% NaOH. Reproduced with permission from ref 217. Copyright 2018 American Chemical Society.
Exosome polymer hybrids were prepared by grafting cholesterol functionalized DNA-polymer
hybrids, or 5′-α-bromoisobutyrate DNA macroinitiators, onto the surface of the membrane
(Figures 18a and 18b).221 The DNA-polymer conjugates significantly improved the stability of the exosome cargo
at ambient temperatures and improved the resilience of the particles against proteolytic
enzymes. Inclusion of a nitrophenyl group between the cholesterol and DNA segments
enabled selective cleavage of the DNA-polymer conjugates from the surface of the exosome
membrane without compromising the exosome or conjugate structure.
Figure 18 | (a) Exosome-polymer hybrids were prepared by the grafting-to and grafting-from approach.
The grafting-to approach involved grafting DNA-cholesterol hybrids onto the exosome,
followed by grafting DNA-P(OEOMA) hybrids onto the DNA segments on the surface of
the exosome. The grafting-from approach was achieved by grafting a cholesterol-DNA
macroinitiator with a 5′-α-bromoisobutyrate group onto the surface of the exosome,
then grafting polymers from the surface by photoinduced ATRP. (b) Incorporation of
a degradable nitrophenyl group between the cholesterol and DNA segments enable selective
release of DNA or DNA-polymer functionalities from the surface by cleavage under UV
light.221 Copyright 2021 National Academy of Sciences.
An arm-first approach was used to prepare OEOMA star polymers with terminal azido-functionalities. The star polymers were complexed with alkyne-functionalized DNA by CuAAC. The star polymers were annealed in the presence of complementary strands, which led to reversible complexation between complementary biopolymers-polymer conjugates into higher-order structures up to 75 nm in size.233
Inorganic Hybrid Materials
Inorganic nanoparticles are often prepared by the grafting-from or grafting-to approach
by RDRP (Scheme 17).8,19,22,24,26,27 In the grafting-from approach, initiators are functionalized on the surface of a
nanoparticle using complementary chemistry. The initiators are then used to grow small-molecule
monomer from the surface to yield a densely grafted nanoparticle or surface. The grafting-to
approach involves grafting premade polymer chains onto a substrate through complementary
chemistry. However, this approach typically yields sparsely grafted nanoparticles
due to blocking of neighboring functional groups. Cleavage of arms from the surface
can be accomplished by dissolution of the nanoparticle or by selective removal through
cleavage of a functional group. Indeed, the polymer fraction was selectively etched
from inorganic SiO2 particle brushes by dissolving the inorganic fraction in hydrofluoric acid to assess
the structure of polymers grafted from the particle.234–236 Scheme 17 | Synthesis of inorganic hybrid materials with cleavable arms via the grafting-from
or grafting-to approaches.
The grafting-to approach can install degradable functional groups anywhere along a polymeric ligand arm. Degradation of polymeric ligand arms altered the dispersibility of the nanoparticles within a matrix, resulting in a change in macroscopic properties. Cleavage of Diels–Alder adducts installed in the center of PS-b-PEG block copolymer ligands led to aggregation of gold nanoparticles within a PMMA-b-PS block polymer matrix.237 A combination of grafting-from ATRP and grafting-to alkoxyamine exchange was utilized to prepare dispersible comb-on-nanoparticle brushes with cleavable functional arms.238 A PMMA polymer brush with statistical incorporation of dormant alkoxyamine side chains was grown from a 100 nm SiO2 nanoparticle via a surface ATRP with an alkoxyamine inimer at low temperature. Alkoxyamine-terminated poly(4-vinyl pyridine) was grafted onto the particle and improved stability of the nanoparticles in water and methanol. The comb topology was reverted back to a linear brush topology by alkoxyamine exchange with other nitroxides. An analogous method was applied to reversibly tune the surface chemistry of a surface-grafted brush grown from a silicon wafer.239
Silicon oxide surfaces were modified with anthracene functional groups on the surface.240 Oligomeric PS or PBA were grafted onto the surface of the substrate via anthracene dimerization under visible light. The surface-grafted brushes could be selectively etched by exposure to low-intensity UV light in order to pattern the surface.
Depolymerization to Monomer
Recent work has mediated the depolymerization of vinyl polymers back to monomers via
a self-immolative approach.241,242 In this approach, polymer macroinitiators were prepared by RDRP with retained chain-end
functionality at low temperature. Activation of the polymer chain-end at elevated
temperature yields the chain-end radical which can either propagate, depropagate,
terminate, or be deactivated by the respective RDRP mechanism. The thermodynamics
of the reaction dictate the RDRP’s tendency towards propagation or depropagation.
Polymerizations of vinyl monomers are exothermic (ΔHp < 0), exoentropic (ΔSp < 0), and favorable at low temperatures typical for radical polymerization.243 The contribution of depropagation is generally negligible at lower temperature, unless
polymerizations are conducted under dilute conditions or at high temperature, when
the initial monomer concentration ([M]0) is close to the equilibrium monomer concentration ([M]eq) (Figure 19). Reported depolymerizations of vinyl polymers by an RDRP mechanism relied on elevated
temperature to increase the contribution of the entropic component (−TΔSp), leading to a higher free energy of polymerization (ΔGp) and tendency towards depropagation.
Figure 19 | Thermodynamic considerations for ideal living polymerizations with an initial monomer
concentration close to the equilibrium monomer concentration, and for depolymerization
reactions in absence of monomer in an ideal living process. Polymerizations proceed
at an apparent rate equal to the difference between the rate of propagation and rate
of depropagation until the rate of propagation equals the rate of depropagation at
the [M]eq. The depolymerization of polymers proceeds at an apparent rate equal to the difference
between the rate of depropagation and rate of propagation until equilibrium is established
at the same [M]eq. A polymerization and depolymerization by RDRP will be affected by loss of chain-end
functionalities, which may lead to a dead-end scenario where radical concentration
approaches zero and stops the reaction before it could reach equilibrium.
The polymerization of 1,1-disubstituted vinyl monomers is generally less exothermic
than mono-substituted acrylic and styrenic monomers. Methacrylates have an enthalpy
of polymerization low enough (ΔHp < 60 kJ/mol) for depropagation to become significant at temperatures above 120 °C.
Methacrylic monomers with bulkier side chains typically have less favorable thermodynamics
which lead to higher equilibrium monomer concentrations than smaller monomers (Figure 20). Additionally, the larger side chain of bulkier methacrylic monomers lowers the
repeat unit concentration such that the bulk monomer concentration is not significantly
higher than the [M]eq. These effects increase the impact of depropagation during both the polymerization
and depolymerization of bulky methacrylic macromonomers. In fact, the first reports
of polymethacrylate depolymerizations mediated by RDRP were part of kinetic studies
investigating the thermodynamic contribution of depropagation during the grafting-through
RDRP of macromonomers.
Figure 20 | The equilibrium monomer concentration of MMA, n-butyl methacrylate, n-dodecyl methacrylate,
POSSMA, PDMSMA, and OEOMA calculated between 25–300 °C using the scaling relationship
ln[M]eq = ΔH/RT − ΔS°/R. The thermodynamic parameters were gathered from references.167,168,244,245
Depolymerization of poly[polyhedral oligomeric silsesquioxane methacrylate] (P(POSSMA)) was observed after an ATRP reached thermodynamic equilibrium at ∼80% conversion at 60 °C was transferred to an oil bath at 90 °C.167 Depolymerization occurred until equilibrium was reestablished at ∼60% monomer conversion, in agreement with what would be observed in a depolymerization caused by an increase in [M]eq with temperature in a living polymerization.
Poly(dimethylsiloxane) methacrylate (P(PDMSMA)) and poly(oligoethylene oxide) methacrylate (P(OEOMA)) bottlebrushes were depolymerized from an initial repeat unit concentration ([P]0) of 100 mM to the [M]eq ∼ 30 mM by a RAFT-mediated process induced by decomposition of the CTA.168 The polymerizations conducted at a [M]0 = 100 mM stopped at a comparable [M]eq as the depolymerizations, which confirmed the reactions plateaued at the same equilibrium in both the forward and reverse directions.
A similar approach was optimized for the depolymerization of polymethacrylates by thermal decomposition of the terminal dithiobenzoate CTA.246 Depolymerizations of macroinitiators at a [P]0 = 5 mM (c.a. 0.05 wt % solid content) at 120 °C reached up to 92% conversion of polymer to monomer over 8 h. The reactions were conducted at a [P]0 far below the [M]eq of typical methacrylates at 120 °C, which highlights the use of dilution as a way to overcome thermodynamic constraints necessary to achieve high depolymerization yields at lower temperature. The high yield also suggests thermolysis of dithiobenzoates was not significant using this experimental set-up.247,248
Depolymerization of poly(methyl methacrylate) mediated by ATRP with a ruthenium (II) chloride catalyst reached a modest monomer recovery of 8% after 7 h of stirring at 120 °C, at a [P]0 ∼ 500 mM.249 The yield of this depolymerization agrees with the reaction stopping close to the [M]eq ∼ 50 mM for PMMA at this temperature. Up to 13.8% monomer recovery occurred after four rounds of reiterative depolymerizations at 100 °C for 10 h.
Degradation of halogen CEF was observed in depolymerizations of poly(n-butyl methacrylate) by ATRP with a copper(II) chloride/tris(2-pyridylmethyl) amine
catalyst at 170 °C.250 The depolymerizations reached up to 67% monomer recovery at a [P]0 = 750 mM (8 wt %) within 10 min. However, the reactions stopped below the theoretical
[M]eq. The plateau in conversion was attributed to loss of chlorine CEF by termination
and lactonization, which accumulated kinetically trapped chains which were unavailable
for depolymerization. Indeed, the importance of CEF preservation in depolymerizations
by ATRP was also emphasized in polymerization/depolymerizaton cycling experiments
with P(PDMS11MA) (Figures 21a and 21b).169 The first round of the cycle depolymerized 81% of the bottlebrush in 15 min. Subsequent
cycles of repolymerization and depolymerization isolated approximately half the amount
of monomer as the preceding cycle, leading to an accumulation of kinetically inactive
chains lacking halogen CEF. The chains lacking the relevant RDRP CEF cannot be activated
in a depolymerization or polymerization, emphasizing the importance of CEF retention
both during macroinitiator synthesis and in the depolymerization reaction.
Figure 21 | Depolymerization/repolymerization cycling of PDMS11MA by ATRP, starting from a depolymerization of P(PDMS11MA)-Cl macroinitiator. (a) Kinetic plot of the mol fraction of macromonomer measured
after each cycle. The red lines are markers for the beginning and end of a cycle.
The black arrows are guides for the eye. (b) Crude GPC traces of depolymerization/repolymerization
cycling experiments at each timepoint were taken relative to linear PMMA standards
in THF. The traces are normalized to the hexadecane peak and cut off at the initial
height of the macroinitiator. First depolymerization (D1) conditions: [P(PDMS11MA)38-Cl]0/[CuCl2]0/[BPMODA*]0 = 1/0.22/1.3 at a [P]0 = 275 mM, Solvent = 1,2,4-trichlorobenzene and 11.7 vol % hexadecane T = 170 °C.
Repolymerization reactions were conducted under UV light with 0.36 equiv EClPA initiator
(relative to the initial alkyl halide concentration) as an additive without prior
purification. All other rounds of depolymerization were accomplished by moving the
flask to the oil bath preheated to 170 °C without prior purification. Reproduced with
permission from ref 169. Copyright 2021 American Chemical Society.
Conclusion and Perspectives
This review summarized recent progress in synthetic strategies towards the topological transformation of vinyl polymers to polymers of lower molecular weight, or different topology, organized by topology of the polymer precursor and route of degradation. RDRP provides unparalleled control over polymer topology, composition, and functionalities. The ability to tether initiators to functionalized substrates, comprising biomolecules, inorganic surfaces, and to other polymers such as grafts and side chains, enabled the preparation of materials with complex topologies with embedded functional groups at the junctions. The use of (multi)functional macroinitiators with initiators tethered by dynamic covalent bonds enabled the synthesis of polymers with cleavable, and/or reversible, topologies. This included polymers which can shift from high to low molecular weight, branched polymers that could disconnect and reconnect side chains, and surfaces that can be reversibly etched or functionalized.
Future work should investigate the properties of polymeric materials before and after degradation to evaluate whether the technology could be feasibly adapted to a circular polymer economy for applications outside of academic laboratories. For example, the transformation of high molecular weight polymers to lower molecular weight oligomers, or the selective cleavage of polymer arms from a high molecular weight graft polymer to low molecular weight arms/branches, may reduce the viscosity of polymer and can have beneficial properties for processing (and reprocessing). Self-healing materials, and materials with dynamic-covalent bonds, may have applications as recyclable thermosets.
Polymer depolymerizations by RDRP highlight two general strategies to reach appreciable monomer recovery. The first method involves depolymerization at high dilution and low enough temperature to ensure that CEF remains intact. Dilution is required because the [M]eq of the reaction is low and the contribution of propagation needs to be mitigated for the reaction to reach completion. High monomer recovery can be achieved. However, separation of monomer from solvent and scalability may be challenging. The other strategy operates at higher concentration and higher temperature to raise [M]eq. The high temperature comes at a cost to livingness. CEF commonly used to regulate RDRP are prone to thermal degradation to inactive species which kinetically traps chain ends and diminishes depolymerization yields. Polymethacrylates with chlorine and bromine CEF were reported to degrade via lactonization above 150 °C.250,251 Dithiocarbonates and TTCs were also reported to decompose between 120–180 °C.247,248,252 NMPs of methacrylic polymers remain challenging to control. However, the rate of bond dissociation for an alkoxyamine increases with temperature.253 Thus, future work in this area should account for both the thermodynamic and kinetic favorability of RDRP reactions at elevated temperatures and high dilution. The use of midchain triggers, such as α-halo-acrylate or N-(acyloxy)phthalimide comonomers, may enable depolymerization through a combination of midchain scission and radical unzipping.88,254
The majority of plastics, including high-performance materials, are not recycled.3 This warrants a continued focus on incorporation of degradable bonds which are durable enough to last for the duration of a material’s usable lifetime but may be dissolved into low-toxicity oligomers if the materials are disposed of in a landfill or leech into the environment. These materials may be prepared by copolymerization of vinyl monomers and a cyclic monomer by RROP or through the use of degradable bonds installed at specific junctions on a polymeric backbone/side chain which can degrade under the appropriate conditions. Future work in RROP should aim to simplify synthetic preparation of (macro)cyclic monomers, and to improve reactivity ratios between vinyl monomers and cyclic monomers There should also be more emphasis on the characterization of degradation under ambient (i.e., nonaccelerated) conditions to evaluate whether the degradable bonds installed on polymeric materials are faster than state-of-the-art biopolymers and polyesters. The toxicity of degraded products should also be a priority in future work, particularly for materials with potential biomedical applications.
Conflict of Interest
There is no conflict of interest to report.
Acknowledgments
Financial support from NSF DMR 1921858 and NSF DMR 2202747 is acknowledged. MM acknowledges support from the Harrison Fellowship (CMU Department of Chemistry).
References
- 1. Plastics Europe. Plastics-the Facts 2019: An Analysis of European Plastics Production, Demand and Waste Data. https://plasticseurope.org/knowledge-hub/plastics-the-facts-2019/ (accessed March 2022). Google Scholar
- 2. Grand View Research. (2021). Market size of plastics worldwide from 2016 to 2028 (in billion U.S. dollars). Statista Inc. (accessed March 1, 2022). Google Scholar
- 3. Geyer R.; Jambeck J. R.; Law K. L.Production, Use, and Fate of All Plastics Ever Made.Sci. Adv.2017, 3, e1700782. Google Scholar
- 4. Dworakowska S.; Lorandi F.; Gorczyński A.; Matyjaszewski K.Toward Green Atom Transfer Radical Polymerization: Current Status and Future Challenges.Adv. Sci.2022, 2106076. Google Scholar
- 5. Fortman D. J.; Brutman J. P.; De Hoe G. X.; Snyder R. L.; Dichtel W. R.; Hillmyer M. A.Approaches to Sustainable and Continually Recyclable Cross-Linked Polymers.ACS Sustain. Chem. Eng.2018, 6, 11145–11159. Google Scholar
- 6. Schneiderman D. K.; Hillmyer M. A.50th Anniversary Perspective: There Is a Great Future in Sustainable Polymers.Macromolecules2017, 50, 3733–3749. Google Scholar
- 7. Shieh P.; Hill M. R.; Zhang W.; Kristufek S. L.; Johnson J. A.Clip Chemistry: Diverse (Bio)(macro)molecular and Material Function through Breaking Covalent Bonds.Chem. Rev.2021, 121, 7059–7121. Google Scholar
- 8. Braunecker W. A.; Matyjaszewski K.Controlled/Living Radical Polymerization: Features, Developments, and Perspectives.Prog. Polym. Sci.2007, 32, 93–146. Google Scholar
- 9. Corrigan N.; Jung K.; Moad G.; Hawker C. J.; Matyjaszewski K.; Boyer C.Reversible-Deactivation Radical Polymerization (Controlled/Living Radical Polymerization): From Discovery to Materials Design and Applications.Prog. Polym. Sci.2020, 111, 101311. Google Scholar
- 10. Matyjaszewski K.; Müller A. H.Controlled and Living Polymerizations: From Mechanisms to Applications; John Wiley & Sons: Hoboken, NJ, 2009. Google Scholar
- 11. Szwarc M.; Van Beylen M.Ionic Polymerization and Living Polymers; Springer: Dordrecht, The Netherlands, 1993. Google Scholar
- 12. Hadjichristidis N.; Pitsikalis M.; Pispas S.; Iatrou H.Polymers with Complex Architecture by Living Anionic Polymerization.Chem. Rev.2001, 101, 3747–3792. Google Scholar
- 13. Matyjaszewski K.; Gnanou Y.; Leibler L.Macromolecular Engineering: Precise Synthesis, Materials Properties, Applications; Wiley-VCH: Weinheim, Germany, 2011. Google Scholar
- 14. Matyjaszewski K.; Xia J.Atom Transfer Radical Polymerization.Chem. Rev.2001, 101, 2921–2990. Google Scholar
- 15. Wang J.-S.; Matyjaszewski K.Controlled/“Living” Radical Polymerization. Atom Transfer Radical Polymerization in the Presence of Transition-Metal Complexes.J. Am. Chem. Soc.1995, 117, 5614–5615. Google Scholar
- 16. Moad G.; Rizzardo E.; Thang S. H.Living Radical Polymerization by the RAFT Process.Aust. J. Chem.2005, 58, 379–410. Google Scholar
- 17. Nicolas J.; Guillaneuf Y.; Lefay C.; Bertin D.; Gigmes D.; Charleux B.Nitroxide-Mediated Polymerization.Prog. Polym. Sci.2013, 38, 63–235. Google Scholar
- 18. Hawker C. J.; Bosman A. W.; Harth E.New Polymer Synthesis by Nitroxide Mediated Living Radical Polymerizations.Chem. Rev.2001, 101, 3661–3688. Google Scholar
- 19. Matyjaszewski K.Atom Transfer Radical Polymerization (ATRP): Current Status and Future Perspectives.Macromolecules2012, 45, 4015–4039. Google Scholar
- 20. Oh J. K.; Drumright R.; Siegwart D. J.; Matyjaszewski K.The Development of Microgels/Nanogels for Drug Delivery Applications.Prog. Polym. Sci.2008, 33, 448–477. Google Scholar
- 21. Sheiko S. S.; Sumerlin B. S.; Matyjaszewski K.Cylindrical Molecular Brushes: Synthesis, Characterization, and Properties.Prog. Polym. Sci.2008, 33, 759–785. Google Scholar
- 22. Matyjaszewski K.; Tsarevsky N. V.Macromolecular Engineering by Atom Transfer Radical Polymerization.J. Am. Chem. Soc.2014, 136, 6513–6533. Google Scholar
- 23. Chong Y.; Le T. P.; Moad G.; Rizzardo E.; Thang S. H.A More Versatile Route to Block Copolymers and Other Polymers of Complex Architecture by Living Radical Polymerization: The RAFT Process.Macromolecules1999, 32, 2071–2074. Google Scholar
- 24. Brinks M. K.; Studer A.Polymer Brushes by Nitroxide-Mediated Polymerization.Macromol. Rapid Commun.2009, 30, 1043–1057. Google Scholar
- 25. Floyd T. G.; Häkkinen S.; Hartlieb M.; Kerr A.; Perrier S.
Complex Polymeric Architectures Synthesized through RAFT Polymerization . In RAFT Polymerization: Methods, Synthesis and Applications; Moad G., Rizzardo E., Eds.; Wiley-VCH: Weinheim, Germany, 2022; Vol. 2; pp 933–981. Google Scholar - 26. Pyun J.; Kowalewski T.; Matyjaszewski K.Synthesis of Polymer Brushes Using Atom Transfer Radical Polymerization.Macromol. Rapid Commun.2003, 24, 1043–1059. Google Scholar
- 27. Yan J.; Bockstaller M. R.; Matyjaszewski K.Brush-Modified Materials: Control of Molecular Architecture, Assembly Behavior, Properties and Applications.Prog. Polym. Sci.2020, 100, 101180. Google Scholar
- 28. Scholten P. B. V.; Moatsou D.; Detrembleur C.; Meier M. A. R.Progress toward Sustainable Reversible Deactivation Radical Polymerization.Macromol. Rapid Commun.2020, 41, 2000266. Google Scholar
- 29. Parkatzidis K.; Wang H. S.; Truong N. P.; Anastasaki A.Recent Developments and Future Challenges in Controlled Radical Polymerization: A 2020 Update.Chem2020, 6, 1575–1588. Google Scholar
- 30. Moad G.Reversible Addition–Fragmentation Chain Transfer (Co)polymerization of Conjugated Diene Monomers: Butadiene, Isoprene, and Chloroprene.Polym. Int.2017, 66, 26–41. Google Scholar
- 31. Lauterbach F.; Rubens M.; Abetz V.; Junkers T.Ultrafast PhotoRAFT Block Copolymerization of Isoprene and Styrene Facilitated through Continuous-Flow Operation.Angew. Chem. Int. Ed.2018, 57, 14260–14264. Google Scholar
- 32. Benoit D.; Harth E.; Fox P.; Waymouth R. M.; Hawker C. J.Accurate Structural Control and Block Formation in the Living Polymerization of 1,3-Dienes by Nitroxide-Mediated Procedures.Macromolecules2000, 33, 363–370. Google Scholar
- 33. De Bon F.; Ribeiro D. C.; Abreu C. M.; Rebelo R. A.; Isse A. A.; Serra A. C.; Gennaro A.; Matyjaszewski K.; Coelho J. F.Under Pressure: Electrochemically-Mediated Atom Transfer Radical Polymerization of Vinyl Chloride.Polym. Chem.2020, 11, 6745–6762. Google Scholar
- 34. Mendes J. P.; Branco F.; Abreu C. M.; Mendonca P. V.; Serra A. n. C.; Popov A. V.; Guliashvili T.; Coelho J. F.Sulfolane: An Efficient and Universal Solvent for Copper-Mediated Atom Transfer Radical (Co)polymerization of Acrylates, Methacrylates, Styrene, and Vinyl Chloride.ACS Macro Lett.2014, 3, 858–861. Google Scholar
- 35. Abreu C. M. R.; Mendonça P. V.; Serra A. C.; Coelho J. F. J.; Popov A. V.; Gryn’ova G.; Coote M. L.; Guliashvili T.Reversible Addition–Fragmentation Chain Transfer Polymerization of Vinyl Chloride.Macromolecules2012, 45, 2200–2208. Google Scholar
- 36. Sun Z.; Mi X.; Yu Y.; Shi W.; Feng A.; Moad G.; Thang S. H.“All-PVC” Flexible Poly(Vinyl Chloride): Nonmigratory Star-Poly(Vinyl Chloride) as Plasticizers for PVC by RAFT Polymerization.Macromolecules2021, 54, 5022–5032. Google Scholar
- 37. Szczepaniak G.; Fu L.; Jafari H.; Kapil K.; Matyjaszewski K.Making ATRP More Practical: Oxygen Tolerance.Acc. Chem. Res.2021, 54, 1779–1790. Google Scholar
- 38. Chen M.; Zhong M.; Johnson J. A.Light-Controlled Radical Polymerization: Mechanisms, Methods, and Applications.Chem. Rev.2016, 116, 10167–10211. Google Scholar
- 39. Lee Y.; Boyer C.; Kwon M. S.Visible-Light-Driven Polymerization towards the Green Synthesis of Plastics.Nat. Rev. Mater.2022, 7, 74–75. Google Scholar
- 40. Ribelli T. G.; Fantin M.; Daran J.-C.; Augustine K. F.; Poli R.; Matyjaszewski K.Synthesis and Characterization of the Most Active Copper ATRP Catalyst Based on Tris[(4-dimethylaminopyridyl)methyl]amine.J. Am. Chem. Soc.2018, 140, 1525–1534. Google Scholar
- 41. Moad G.
An Industrial History of RAFT Polymerization . In RAFT Polymerization: Methods, Synthesis and Applications; Moad G., Rizzardo E., Eds.; Wiley-VCH: Weinheim, Germany, 2022; Vol. 2; pp 1077–1169. Google Scholar - 42. Destarac M.Controlled Radical Polymerization: Industrial Stakes, Obstacles and Achievements.Macromol. React. Eng.2010, 4, 165–179. Google Scholar
- 43. Destarac M.Industrial Development of Reversible-Deactivation Radical Polymerization: Is the Induction Period Over?Polym. Chem.2018, 9, 4947–4967. Google Scholar
- 44. Serrano D. P.; Aguado J.; Escola J. M.Developing Advanced Catalysts for the Conversion of Polyolefinic Waste Plastics into Fuels and Chemicals.ACS Catal.2012, 2, 1924–1941. Google Scholar
- 45. Chu M.; Liu Y.; Lou X.; Zhang Q.; Chen J.Rational Design of Chemical Catalysis for Plastic Recycling.ACS Catal.2022, 12, 4659–4679. Google Scholar
- 46. Castro-Aguirre E.; Iñiguez-Franco F.; Samsudin H.; Fang X.; Auras R.Poly(lactic acid)—Mass Production, Processing, Industrial Applications, and End of Life.Adv. Drug Deliv. Rev.2016, 107, 333–366. Google Scholar
- 47. Jem K. J.; Tan B.The Development and Challenges of Poly(Lactic Acid) and Poly (Glycolic Acid).Adv. Ind. Eng. Polym. Res.2020, 3, 60–70. Google Scholar
- 48. Häußler M.; Eck M.; Rothauer D.; Mecking S.Closed-Loop Recycling of Polyethylene-Like Materials.Nature2021, 590, 423–427. Google Scholar
- 49. De Hoe G. X.; Zumstein M. T.; Tiegs B. J.; Brutman J. P.; McNeill K.; Sander M.; Coates G. W.; Hillmyer M. A.Sustainable Polyester Elastomers from Lactones: Synthesis, Properties, and Enzymatic Hydrolyzability.J. Am. Chem. Soc.2018, 140, 963–973. Google Scholar
- 50. Cywar R. M.; Rorrer N. A.; Hoyt C. B.; Beckham G. T.; Chen E. Y. X.Bio-Based Polymers with Performance-Advantaged Properties.Nat. Rev. Mater.2022, 7, 83–103. Google Scholar
- 51. Jehanno C.; Alty J. W.; Roosen M.; De Meester S.; Dove A. P.; Chen E. Y. X.; Leibfarth F. A.; Sardon H.Critical Advances and Future Opportunities in Upcycling Commodity Polymers.Nature2022, 603, 803–814. Google Scholar
- 52. Cywar R. M.; Rorrer N. A.; Mayes H. B.; Maurya A. K.; Tassone C. J.; Beckham G. T.; Chen E. Y. X.Redesigned Hybrid Nylons with Optical Clarity and Chemical Recyclability.J. Am. Chem. Soc.2022, 144, 5366–5376. Google Scholar
- 53. Abel B. A.; Snyder R. L.; Coates G. W.Chemically Recyclable Thermoplastics from Reversible-Deactivation Polymerization of Cyclic Acetals.Science2021, 373, 783–789. Google Scholar
- 54. Shi C.; Reilly L. T.; Kumar V. S. P.; Coile M. W.; Nicholson S. R.; Broadbelt L. J.; Beckham G. T.; Chen E. Y.-X.Design Principles for Intrinsically Circular Polymers with Tunable Properties.Chem2021, 7, 2896–2912. Google Scholar
- 55. Vogt B. D.; Stokes K. K.; Kumar S. K.Why Is Recycling of Postconsumer Plastics so Challenging?ACS Appl. Polym. Mater.2021, 3, 4325–4346. Google Scholar
- 56. Coates G. W.; Getzler Y. D.Chemical Recycling to Monomer for an Ideal, Circular Polymer Economy.Nat. Rev. Mater.2020, 5, 501–516. Google Scholar
- 57. Rahimi A.; García J. M.Chemical Recycling of Waste Plastics for New Materials Production.Nat. Rev. Chem.2017, 1, 1–11. Google Scholar
- 58. Hong M.; Chen E. Y. X.Future Directions for Sustainable Polymers.Trends Chem.2019, 1, 148–151. Google Scholar
- 59. Tardy A.; Nicolas J.; Gigmes D.; Lefay C.; Guillaneuf Y.Radical Ring-Opening Polymerization: Scope, Limitations, and Application to (Bio) Degradable Materials.Chem. Rev.2017, 117, 1319–1406. Google Scholar
- 60. Gigmes D.; Van Steenberge P. H.; Siri D.; D’hooge D. R.; Guillaneuf Y.; Lefay C.Simulation of the Degradation of Cyclic Ketene Acetal and Vinyl-Based Copolymers Synthesized via a Radical Process: Influence of the Reactivity Ratios on the Degradability Properties.Macromol. Rapid Commun.2018, 39, 1800193. Google Scholar
- 61. De Smit K.; Marien Y. W.; Van Geem K. M.; Van Steenberge P. H.; D’hooge D. R.Connecting Polymer Synthesis and Chemical Recycling on a Chain-by-Chain Basis: A Unified Matrix-Based Kinetic Monte Carlo Strategy.React. Chem. Eng.2020, 5, 1909–1928. Google Scholar
- 62. Barbon S. M.; Carter M. C. D.; Yin L.; Whaley C. M.; Albright III V. C.; Tecklenburg R. E.Synthesis and Biodegradation Studies of Low-Dispersity Poly(acrylic Acid).Macromol. Rapid Commun.2022, 2100773. Google Scholar
- 63. Hedir G. G.; Bell C. A.; Ieong N. S.; Chapman E.; Collins I. R.; O’Reilly R. K.; Dove A. P.Functional Degradable Polymers by Xanthate-Mediated Polymerization.Macromolecules2014, 47, 2847–2852. Google Scholar
- 64. Hedir G. G.; Bell C. A.; O’Reilly R. K.; Dove A. P.Functional Degradable Polymers by Radical Ring-Opening Copolymerization of MDO and Vinyl Bromobutanoate: Synthesis, Degradability and Post-Polymerization Modification.Biomacromolecules2015, 16, 2049–2058. Google Scholar
- 65. Hedir G.; Stubbs C.; Aston P.; Dove A. P.; Gibson M. I.Synthesis of Degradable Poly(Vinyl Alcohol) by Radical Ring-Opening Copolymerization and Ice Recrystallization Inhibition Activity.ACS Macro Lett.2017, 6, 1404–1408. Google Scholar
- 66. He T.; Zou Y.-F.; Pan C.-Y.Controlled/“Living” Radical Ring-Opening Polymerization of 5,6-Benzo-2-Methylene-1,3-Dioxepane Based on Reversible Addition-Fragmentation Chain Transfer Mechanism.Polym. J.2002, 34, 138–143. Google Scholar
- 67. Yuan J.-Y.; Pan C.-Y.; Tang B. Z.“Living” Free Radical Ring-Opening Polymerization of 5,6-Benzo-2-methylene-1,3-dioxepane Using the Atom Transfer Radical Polymerization Method.Macromolecules2001, 34, 211–214. Google Scholar
- 68. Tardy A.; Delplace V.; Siri D.; Lefay C.; Harrisson S.; de Fatima Albergaria Pereira B.; Charles L.; Gigmes D.; Nicolas J.; Guillaneuf Y.Scope and Limitations of the Nitroxide-Mediated Radical Ring-Opening Polymerization of Cyclic Ketene Acetals.Polym. Chem.2013, 4, 4776. Google Scholar
- 69. Decker C. G.; Maynard H. D.Degradable PEGylated Protein Conjugates Utilizing RAFT Polymerization.Eur. Polym. J.2015, 65, 305–312. Google Scholar
- 70. Lai H.; Ouchi M.Backbone-Degradable Polymers via Radical Copolymerizations of Pentafluorophenyl Methacrylate with Cyclic Ketene Acetal: Pendant Modification and Efficient Degradation by Alternating-Rich Sequence.ACS Macro Lett.2021, 10, 1223–1228. Google Scholar
- 71. Delplace V.; Tardy A.; Harrisson S.; Mura S.; Gigmes D.; Guillaneuf Y.; Nicolas J.Degradable and Comb-Like PEG-Based Copolymers by Nitroxide-Mediated Radical Ring-Opening Polymerization.Biomacromolecules2013, 14, 3769–3779. Google Scholar
- 72. Delplace V.; Guegain E.; Harrisson S.; Gigmes D.; Guillaneuf Y.; Nicolas J.A Ring to Rule Them All: A Cyclic Ketene Acetal Comonomer Controls the Nitroxide-Mediated Polymerization of Methacrylates and Confers Tunable Degradability.Chem. Commun. (Camb.)2015, 51, 12847–12850. Google Scholar
- 73. Tran J.; Guégain E.; Ibrahim N.; Harrisson S.; Nicolas J.Efficient Synthesis of 2-Methylene-4-phenyl-1,3-dioxolane, a Cyclic Ketene Acetal for Controlling the NMP of Methyl Methacrylate and Conferring Tunable Degradability.Polym. Chem.2016, 7, 4427–4435. Google Scholar
- 74. Guégain E.; Michel J.-P.; Boissenot T.; Nicolas J.Tunable Degradation of Copolymers Prepared by Nitroxide-Mediated Radical Ring-Opening Polymerization and Point-by-Point Comparison with Traditional Polyesters.Macromolecules2018, 51, 724–736. Google Scholar
- 75. Hill M. R.; Guégain E.; Tran J.; Figg C. A.; Turner A. C.; Nicolas J.; Sumerlin B. S.Radical Ring-Opening Copolymerization of Cyclic Ketene Acetals and Maleimides Affords Homogeneous Incorporation of Degradable Units.ACS Macro Lett.2017, 6, 1071–1077. Google Scholar
- 76. Smith Q.; Huang J.; Matyjaszewski K.; Loo Y.-L.Controlled Radical Polymerization and Copolymerization of 5-Methylene-2-phenyl-1,3-dioxolan-4-one by ATRP.Macromolecules2005, 38, 5581–5586. Google Scholar
- 77. Smith R. A.; Fu G.; McAteer O.; Xu M.; Gutekunst W. R.Radical Approach to Thioester-Containing Polymers.J. Am. Chem. Soc.2019, 141, 1446–1451. Google Scholar
- 78. Spick M. P.; Bingham N. M.; Li Y.; de Jesus J.; Costa C.; Bailey M. J.; Roth P. J.Fully Degradable Thioester-Functional Homo- and Alternating Copolymers Prepared through Thiocarbonyl Addition–Ring-Opening RAFT Radical Polymerization.Macromolecules2020, 53, 539–547. Google Scholar
- 79. Paulusse J. M. J.; Amir R. J.; Evans R. A.; Hawker C. J.Free Radical Polymers with Tunable and Selective Bio- and Chemical Degradability.J. Am. Chem. Soc.2009, 131, 9805–9812. Google Scholar
- 80. Evans R. A.; Moad G.; Rizzardo E.; Thang S. H.New Free-Radical Ring-Opening Acrylate Monomers.Macromolecules1994, 27, 7935–7937. Google Scholar
- 81. Ratcliffe L. P. D.; Couchon C.; Armes S. P.; Paulusse J. M. J.Inducing an Order–Order Morphological Transition via Chemical Degradation of Amphiphilic Diblock Copolymer Nano-Objects.Biomacromolecules2016, 17, 2277–2283. Google Scholar
- 82. Huang H.; Sun B.; Huang Y.; Niu J.Radical Cascade-Triggered Controlled Ring-Opening Polymerization of Macrocyclic Monomers.J. Am. Chem. Soc.2018, 140, 10402–10406. Google Scholar
- 83. Wang W.; Zhou Z.; Sathe D.; Tang X.; Moran S.; Jin J.; Haeffner F.; Wang J.; Niu J.Degradable Vinyl Random Copolymers via Photocontrolled Radical Ring-Opening Cascade Copolymerization.Angew. Chem. Int. Ed.2022, 61, e202113302. Google Scholar
- 84. Reeves J. A.; De Alwis Watuthanthrige N.; Boyer C.; Konkolewicz D.Intrinsic and Catalyzed Photochemistry of Phenylvinylketone for Wavelength-Sensitive Controlled Polymerization.ChemPhotoChem2019, 3, 1171–1179. Google Scholar
- 85. Nwoko T.; Watuthanthrige N. D. A.; Parnitzke B.; Yehl K.; Konkolewicz D.Tuning Molecular Weight Distributions of Vinylketone-based Polymers using RAFT Photopolymerization and UV Photodegradation.Polym. Chem.2021, 12, 6761–6770. Google Scholar
- 86. Cheng C.; Sun G.; Khoshdel E.; Wooley K. L.Well-Defined Vinyl Ketone-Based Polymers by Reversible Addition−Fragmentation Chain Transfer Polymerization.J. Am. Chem. Soc.2007, 129, 10086–10087. Google Scholar
- 87. De Alwis Watuthanthrige N.; Reeves J. A.; Dolan M. T.; Valloppilly S.; Zanjani M. B.; Ye Z.; Konkolewicz D.Wavelength-Controlled Synthesis and Degradation of Thermoplastic Elastomers Based on Intrinsically Photoresponsive Phenyl Vinyl Ketone.Macromolecules2020, 53, 5199–5207. Google Scholar
- 88. Kimura T.; Kuroda K.; Kubota H.; Ouchi M.Metal-Catalyzed Switching Degradation of Vinyl Polymers via Introduction of an “In-Chain” Carbon–Halogen Bond as the Trigger.ACS Macro Lett.2021, 10, 1535–1539. Google Scholar
- 89. Rikkou M. D.; Patrickios C. S.Polymers Prepared Using Cleavable Initiators: Synthesis, Characterization and Degradation.Prog. Polym. Sci.2011, 36, 1079–1097. Google Scholar
- 90. Tsarevsky N. V.; Matyjaszewski K.Reversible Redox Cleavage/Coupling of Polystyrene with Disulfide or Thiol Groups Prepared by Atom Transfer Radical Polymerization.Macromolecules2002, 35, 9009–9014. Google Scholar
- 91. Tsarevsky N. V.; Matyjaszewski K.Combining Atom Transfer Radical Polymerization and Disulfide/Thiol Redox Chemistry: A Route to Well-Defined (Bio)degradable Polymeric Materials.Macromolecules2005, 38, 3087–3092. Google Scholar
- 92. Fritze U. F.; Craig S. L.; von Delius M.Disulfide-Centered Poly(Methyl Acrylates): Four Different Stimuli to Cleave a Polymer.J. Polym. Sci. A: Polym. Chem.2018, 56, 1404–1411. Google Scholar
- 93. Sun H.; Kabb C. P.; Dai Y.; Hill M. R.; Ghiviriga I.; Bapat A. P.; Sumerlin B. S.Macromolecular Metamorphosis via Stimulus-Induced Transformations of Polymer Architecture.Nat. Chem.2017, 9, 817–823. Google Scholar
- 94. Duan H.-Y.; Wang Y.-X.; Wang L.-J.; Min Y.-Q.; Zhang X.-H.; Du B.-Y.An Investigation of the Selective Chain Scission at Centered Diels–Alder Mechanophore under Ultrasonication.Macromolecules2017, 50, 1353–1361. Google Scholar
- 95. Nicolay R.; Marx L.; Hemery P.; Matyjaszewski K.Synthesis of Multisegmented Degradable Polymers by Atom Transfer Radical Cross-Coupling.Macromolecules2007, 40, 9217–9223. Google Scholar
- 96. Wang G.; Huang J.Versatility of Radical Coupling in Construction of Topological Polymers.Polym. Chem.2014, 5, 277–308. Google Scholar
- 97. Sarbu T.; Lin K.-Y.; Ell J.; Siegwart D. J.; Spanswick J.; Matyjaszewski K.Polystyrene with Designed Molecular Weight Distribution by Atom Transfer Radical Coupling.Macromolecules2004, 37, 3120–3127. Google Scholar
- 98. Nicolaÿ R.; Matyjaszewski K.Synthesis of Cyclic (Co)polymers by Atom Transfer Radical Cross-Coupling and Ring Expansion by Nitroxide-Mediated Polymerization.Macromolecules2011, 44, 240–247. Google Scholar
- 99. Wang X.; Huang J.; Chen L.; Liu Y.; Wang G.Synthesis of Thermal Degradable Poly(alkoxyamine) through a Novel Nitroxide Radical Coupling Step Growth Polymerization Mechanism.Macromolecules2014, 47, 7812–7822. Google Scholar
- 100. Valente C. J.; Schellenberger A. M.; Tillman E. S.Dimerization of Poly(methyl methacrylate) Chains Using Radical Trap-Assisted Atom Transfer Radical Coupling.Macromolecules2014, 47, 2226–2232. Google Scholar
- 101. Butcher W. E.; Radzinski S. C.; Tillman E. S.Selective Formation of Diblock Copolymers Using Radical Trap-Assisted Atom Transfer Radical Coupling.J. Polym. Sci. A: Polym. Chem.2013, 51, 3619–3626. Google Scholar
- 102. Voter A. F.; Tillman E. S.; Findeis P. M.; Radzinski S. C.Synthesis of Macrocyclic Polymers Formed via Intramolecular Radical Trap-Assisted Atom Transfer Radical Coupling.ACS Macro Lett.2012, 1, 1066–1070. Google Scholar
- 103. Whittaker M. R.; Goh Y.-K.; Gemici H.; Legge T. M.; Perrier S.; Monteiro M. J.Synthesis of Monocyclic and Linear Polystyrene Using the Reversible Coupling/Cleavage of Thiol/Disulfide Groups.Macromolecules2006, 39, 9028–9034. Google Scholar
- 104. You Y.-Z.; Manickam D. S.; Zhou Q.; Oupicky D.A Versatile Approach to Reducible Vinyl Polymers via Oxidation of Telechelic Polymers Prepared by Reversible Addition Fragmentation Chain Transfer Polymerization.Biomacromolecules2007, 8, 2038–2044. Google Scholar
- 105. Hakobyan K.; McErlean C. S. P.; Müllner M.RAFT Without an “R-Group”: From Asymmetric Homo-Telechelics to Multiblock Step-Growth and Cyclic Block Copolymers.Macromolecules2021, 54, 7732–7742. Google Scholar
- 106. You Y.-Z.; Zhou Q.-H.; Manickam D. S.; Wan L.; Mao G.; Oupicky D.Dually Responsive Multiblock Copolymers via Reversible Addition-Fragmentation Chain Transfer Polymerization: Synthesis of Temperature- and Redox-Responsive Copolymers of Poly(N-isopropylacrylamide) and Poly(2-(dimethylamino)ethyl methacrylate).Biomacromolecules2007, 40, 8617–8624. Google Scholar
- 107. Ko N. R.; Yao K.; Tang C.; Oh J. K.Synthesis and Thiol-Responsive Degradation of Polylactide-Based Block Copolymers Having Disulfide Junctions Using ATRP and ROP.J. Polym. Sci. A: Polym. Chem.2013, 51, 3071–3080. Google Scholar
- 108. Cunningham A.; Oh J. K.New Design of Thiol-Responsive Degradable Polylactide-Based Block Copolymer Micelles.Macromol. Rapid Commun.2013, 34, 163–168. Google Scholar
- 109. Clement B.; Trimaille T.; Alluin O.; Gigmes D.; Mabrouk K.; Feron D.; Decherchi P.; Marqueste T.; Bertin D.Convenient Access to Biocompatible Block Copolymers from SG1-Based Aliphatic Polyester Macro-Alkoxyamines.Biomacromolecules2009, 10, 1436–1445. Google Scholar
- 110. Resendiz-Lara D. A.; Wurm F. R.Polyphosphonate-Based Macromolecular RAFT-CTA Enables the Synthesis of Well-Defined Block Copolymers Using Vinyl Monomers.ACS Macro Lett.2021, 10, 1273–1279. Google Scholar
- 111. Liu X.; Ni P.; He J.; Zhang M.Synthesis and Micellization of pH/Temperature-Responsive Double-Hydrophilic Diblock Copolymers Polyphosphoester-block-poly[2-(dimethylamino)ethyl methacrylate] Prepared via ROP and ATRP.Macromolecules2010, 43, 4771–4781. Google Scholar
- 112. Kang H. U.; Yu Y. C.; Shin S. J.; Youk J. H.One-Step Synthesis of Block Copolymers Using a Hydroxyl-Functionalized Trithiocarbonate RAFT Agent as a Dual Initiator for RAFT Polymerization and ROP.J. Polym. Sci. A: Polym. Chem.2013, 51, 774–779. Google Scholar
- 113. Xia Y.; Scheutz G.; Easterling C.; Zhao J.; Sumerlin B. S.Hybrid Block Copolymer Synthesis by Merging Photoiniferter and Organocatalytic Ring-Opening Polymerizations.Angew. Chem. Int. Ed.2021, 60, 18537–18541. Google Scholar
- 114. Voit B. I.; Lederer A.Hyperbranched and Highly Branched Polymer Architectures—Synthetic Strategies and Major Characterization Aspects.Chem. Rev.2009, 109, 5924–5973. Google Scholar
- 115. Zheng Y.; Li S.; Weng Z.; Gao C.Hyperbranched Polymers: Advances from Synthesis to Applications.Chem. Soc. Rev.2015, 44, 4091–4130. Google Scholar
- 116. Jiang W.; Zhou Y.; Yan D.Hyperbranched Polymer Vesicles: From Self-Assembly, Characterization, Mechanisms, and Properties to Applications.Chem. Soc. Rev.2015, 44, 3874–3889. Google Scholar
- 117. Caminade A.-M.; Yan D.; Smith D. K.Dendrimers and Hyperbranched Polymers.Chem. Soc. Rev.2015, 44, 3870–3873. Google Scholar
- 118. Tsarevsky N. V.; Huang J.; Matyjaszewski K.Synthesis of Hyperbranched Degradable Polymers by Atom Transfer Radical (Co)polymerization of Inimers with Ester or Disulfide Groups.J. Polym. Sci. A: Polym. Chem.2009, 47, 6839–6851. Google Scholar
- 119. Mizutani M.; Satoh K.; Kamigaito M.Metal-Catalyzed Radical Polyaddition for Aliphatic Polyesters via Evolution of Atom Transfer Radical Addition into Step-Growth Polymerization.Macromolecules2009, 42, 472–480. Google Scholar
- 120. Graff R. W.; Wang X.; Gao H.Exploring Self-Condensing Vinyl Polymerization of Inimers in Microemulsion To Regulate the Structures of Hyperbranched Polymers.Macromolecules2015, 48, 2118–2126. Google Scholar
- 121. Sun H.; Kabb C. P.; Sumerlin B. S.Thermally-Labile Segmented Hyperbranched Copolymers: Using Reversible-Covalent Chemistry to Investigate the Mechanism of Self-Condensing Vinyl Copolymerization.Chem. Sci.2014, 5, 4646–4655. Google Scholar
- 122. Rosselgong J.; Armes S. P.; Barton W.; Price D.Synthesis of Highly Branched Methacrylic Copolymers: Observation of Near-Ideal Behavior Using RAFT Polymerization.Macromolecules2009, 42, 5919–5924. Google Scholar
- 123. Vo C. J.; Rosselgong J.; Armes S. P.; Billingham N. C.RAFT Synthesis of Branched Acrylic Copolymers.Macromolecules2007, 40, 7119–7125. Google Scholar
- 124. Wang L.; Li C.; Ryan A. J.; Armes S. P.Synthesis and Peptide-Induced Degradation of Biocompatible Fibers Based on Highly Branched Poly(2-hydroxyethyl methacrylate).Adv. Mater.2006, 18, 1566–1570. Google Scholar
- 125. Oh J. K.; Tang C.; Gao H.; Tsarevsky N.; Matyjaszewski K.Inverse Miniemulsion ATRP: A New Method for Synthesis and Functionalization of Well-Defined Water-Soluble/Cross-Linked Polymeric Particles.J. Am. Chem. Soc.2006, 128, 5578–5584. Google Scholar
- 126. Oh J. K.; Siegwart D. J.; Matyjaszewski K.Synthesis and Biodegradation of Nanogels as Delivery Carriers for Carbohydrate Drugs.Biomacromolecules2007, 8, 3326–3331. Google Scholar
- 127. Oh J. K.; Siegwart D. J.; Lee H.; Sherwood G.; Peteanu L.; Hollinger J. O.; Kataoka K.; Matyjaszewski K.Biodegradable Nanogels Prepared by Atom Transfer Radical Polymerization as Potential Drug Delivery Carriers: Synthesis, Biodegradation, in Vitro Release, and Bioconjugation.J. Am. Chem. Soc.2007, 129, 5939–5945. Google Scholar
- 128. Sims M. B.; Patel K. Y.; Bhatta M.; Mukherjee S.; Sumerlin B. S.Harnessing Imine Diversity to Tune Hyperbranched Polymer Degradation.Macromolecules2018, 51, 356–363. Google Scholar
- 129. Levy A.; Goldstein H.; Brenman D.; Diesendruck C. E.Effect of Intramolecular Crosslinker Properties on the Mechanochemical Fragmentation of Covalently Folded Polymers.J. Polym. Sci.2020, 58, 692–703. Google Scholar
- 130. Blencowe A.; Tan J. F.; Goh T. K.; Qiao G. G.Core Cross-Linked Star Polymers via Controlled Radical Polymerisation.Polymer2009, 50, 5–32. Google Scholar
- 131. Ren J. M.; McKenzie T. G.; Fu Q.; Wong E. H. H.; Xu J.; An Z.; Shanmugam S.; Davis T. P.; Boyer C.; Qiao G. G.Star Polymers.Chem. Rev.2016, 116, 6743–6836. Google Scholar
- 132. Gao H.; Matyjaszewski K.Synthesis of Functional Polymers with Controlled Architecture by CRP of Monomers in the Presence of Cross-Linkers: From Stars to Gels.Prog. Polym. Sci.2009, 34, 317–350. Google Scholar
- 133. Goh T. K.; Coventry K. D.; Blencowe A.; Qiao G. G.Rheology of Core Cross-Linked Star Polymers.Polymer2008, 49, 5095–5104. Google Scholar
- 134. Chmielarz P.Synthesis of α-d-Glucose-Based Star Polymers through Simplified Electrochemically Mediated ATRP.Polymer2016, 102, 192–198. Google Scholar
- 135. Chmielarz P.; Park S.; Sobkowiak A.; Matyjaszewski K.Synthesis of β-Cyclodextrin-Based Star Polymers via a Simplified Electrochemically Mediated ATRP.Polymer2016, 88, 36–42. Google Scholar
- 136. Zhu W.; Nese A.; Matyjaszewski K.Thermoresponsive Star Triblock Copolymers by Combination of ROP and ATRP: From Micelles to Hydrogels.J. Polym. Sci. A: Polym. Chem.2011, 49, 1942–1952. Google Scholar
- 137. Cuthbert J.; Yerneni S. S.; Sun M.; Fu T.; Matyjaszewski K.Degradable Polymer Stars Based on Tannic Acid Cores by ATRP.Polymers (Basel)2019, 11, 752. Google Scholar
- 138. Setijadi E.; Tao L.; Liu J.; Jia Z.; Boyer C.; Davis T. P.Biodegradable Star Polymers Functionalized with β-Cyclodextrin Inclusion Complexes.Biomacromolecules2009, 10, 2699–2707. Google Scholar
- 139. Liu J.; Liu H.; Jia Z.; Bulmus V.; Davis T. P.An Approach to Biodegradable Star Polymeric Architectures Using Disulfide Coupling.Chem. Commun. (Camb.)2008, 48, 6582–6584. Google Scholar
- 140. Rosselgong J.; Williams E. G. L.; Le T. P.; Grusche F.; Hinton T. M.; Tizard M.; Gunatillake P.; Thang S. H.Core Degradable Star RAFT Polymers: Synthesis, Polymerization, and Degradation Studies.Macromolecules2013, 46, 9181–9188. Google Scholar
- 141. Gao H.; Tsarevsky N. V.; Matyjaszewski K.Synthesis of Degradable Miktoarm Star Copolymers via Atom Transfer Radical Polymerization.Macromolecules2005, 38, 5995–6004. Google Scholar
- 142. Cho H. Y.; Srinivasan A.; Hong J.; Hsu E.; Liu S.; Shrivats A.; Kwak D.; Bohaty A. K.; Paik H. J.; Hollinger J. O.; Matyjaszewski K.Synthesis of Biocompatible PEG-Based Star Polymers with Cationic and Degradable Core for siRNA Delivery.Biomacromolecules2011, 12, 3478–3486. Google Scholar
- 143. Averick S. E.; Paredes E.; Irastorza A.; Shrivats A. R.; Srinivasan A.; Siegwart D. J.; Magenau A. J.; Cho H. Y.; Hsu E.; Averick A. A.; Kim J.; Liu S.; Hollinger J. O.; Das S. R.; Matyjaszewski K.Preparation of Cationic Nanogels for Nucleic Acid Delivery.Biomacromolecules2012, 13, 3445–3449. Google Scholar
- 144. Li W.; Yoon J. A.; Martjaszewski K.Dual-Reactive Surfactant Used for Synthesis of Functional Nanocapsules in Miniemulsion.J. Am. Chem. Soc.2010, 132, 7823–7825. Google Scholar
- 145. Wiltshire J. T.; Qiao G. G.Selectively Degradable Core Cross-Linked Star Polymers.Macromolecules2006, 39, 9018–9027. Google Scholar
- 146. Wiltshire J. T.; Qiao G. G.Synthesis of Core Cross-Linked Star Polymers with Adjustable Coronal Properties.Macromolecules2008, 41, 623–631. Google Scholar
- 147. Dai Y.; Sun H.; Pal S.; Zhang Y.; Park S.; Kabb C. P.; Wei W. D.; Sumerlin B. S.Near-IR-Induced Dissociation of Thermally-Sensitive Star Polymers.Chem. Sci.2017, 8, 1815–1821. Google Scholar
- 148. Bapat A. P.; Ray J. G.; Savin D. A.; Hoff E. A.; Patton D. L.; Sumerlin B. S.Dynamic-Covalent Nanostructures Prepared by Diels–Alder Reactions of Styrene-Maleic Anhydride-Derived Copolymers Obtained by One-Step Cascade Block Copolymerization.Polym. Chem.2012, 3, 3112. Google Scholar
- 149. Amamoto Y.; Kikuchi M.; Otsuka H.; Takahara A.Arm-Replaceable Star-Like Nanogels: Arm Detachment and Arm Exchange Reactions by Dynamic Covalent Exchanges of Alkoxyamine Units.Polym. J.2010, 42, 860–867. Google Scholar
- 150. Amamoto Y.; Kikuchi M.; Otsuka H.; Takahara A.Solvent-Controlled Formation of Star-like Nanogels via Dynamic Covalent Exchange of PSt-b-PMMA Diblock Copolymers with Alkoxyamine Units in the Side Chain.Macromolecules2010, 43, 5470–5473. Google Scholar
- 151. Amamoto Y.; Higaki Y.; Matsuda Y.; Otsuka H.; Takahara A.Programmed Thermodynamic Formation and Structure Analysis of Star-Like Nanogels with Core Cross-Linked by Thermally Exchangeable Dynamic Covalent Bonds.J. Am. Chem. Soc.2007, 129, 13298–13304. Google Scholar
- 152. Xie G.; Martinez M. R.; Olszewski M.; Sheiko S. S.; Matyjaszewski K.Molecular Bottlebrushes as Novel Materials.Biomacromolecules2019, 20, 27–54. Google Scholar
- 153. Lee H.-i.; Pietrasik J.; Sheiko S. S.; Matyjaszewski K.Stimuli-Responsive Molecular Brushes.Prog. Polym. Sci.2010, 35, 24–44. Google Scholar
- 154. Panyukov S.; Zhulina E. B.; Sheiko S. S.; Randall G. C.; Brock J.; Rubinstein M.Tension Amplification in Molecular Brushes in Solutions and on Substrates.J. Phys. Chem. B2009, 113, 3750–3768. Google Scholar
- 155. Panyukov S. V.; Sheiko S. S.; Rubinstein M.Amplification of Tension in Branched Macromolecules.Phys. Rev. Lett.2009, 102, 148301. Google Scholar
- 156. Zheng Z.; Müllner M.; Ling J.; Müller A. H. E.Surface Interactions Surpass Carbon–Carbon Bond: Understanding and Control of the Scission Behavior of Core–Shell Polymer Brushes on Surfaces.ACS Nano2013, 7, 2284–2291. Google Scholar
- 157. Sheiko S. S.; Sun F. C.; Randall A.; Shirvanyants D.; Rubinstein M.; Lee H. I.; Matyjaszewski K.Adsorption-Induced Scission of Carbon-Carbon Bonds.Nature2006, 440, 191–194. Google Scholar
- 158. Lebedeva N. V.; Sun F. C.; Lee H.-i.; Matyjaszewski K.; Sheiko S. S.“Fatal Adsorption” of Brushlike Macromolecules: High Sensitivity of C−C Bond Cleavage Rates to Substrate Surface Energy.J. Am. Chem. Soc.2008, 130, 4228–4229. Google Scholar
- 159. Park I.; Shirvanyants D.; Nese A.; Matyjaszewski K.; Rubinstein M.; Sheiko S. S.Spontaneous and Specific Activation of Chemical Bonds in Macromolecular Fluids.J. Am. Chem. Soc.2010, 132, 12487–12491. Google Scholar
- 160. Lebedeva N. V.; Nese A.; Sun F. C.; Matyjaszewski K.; Sheiko S. S.Anti-Arrhenius Cleavage of Covalent Bonds in Bottlebrush Macromolecules on Substrate.Proc. Natl. Acad. Sci. U. S. A.2012, 109, 9276–9280. Google Scholar
- 161. Li Y.; Nese A.; Matyjaszewski K.; Sheiko S. S.Molecular Tensile Machines: Anti-Arrhenius Cleavage of Disulfide Bonds.Macromolecules2013, 46, 7196–7201. Google Scholar
- 162. Park I. S.; Nese A.; Matyjaszewski K.Molecular Tensile Testing Machines: Breaking a Specific Covalent Bond by Adsorption-Induced Tension in Brushlike Macromolecules.Macromolecules2009, 42, 1805–1807. Google Scholar
- 163. Burdyńska J.; Li Y.; Aggarwal A. V.; Höger S.; Sheiko S. S.; Matyjaszewski K.Synthesis and Arm Dissociation in Molecular Stars with a Spoked Wheel Core and Bottlebrush Arms.J. Am. Chem. Soc.2014, 136, 12762–12770. Google Scholar
- 164. Martinez M. R.; Cong Y.; Sheiko S. S.; Matyjaszewski K.A Thermodynamic Roadmap for the Grafting-through Polymerization of PDMS11MA.ACS Macro Lett.2020, 9, 1303–1309. Google Scholar
- 165. Cho H. Y.; Krys P.; Szcześniak K.; Schroeder H.; Park S.; Jurga S.; Buback M.; Matyjaszewski K.Synthesis of Poly (OEOMA) Using Macromonomers via “Grafting-Through” ATRP.Macromolecules2015, 48, 6385–6395. Google Scholar
- 166. Shinoda H.; Miller P. J.; Matyjaszewski K.Improving the Structural Control of Graft Copolymers by Combining ATRP with the Macromonomer Method.Macromolecules2001, 34, 3186–3194. Google Scholar
- 167. Raus V.; Čadová E.; Starovoytova L.; Janata M.ATRP of POSS Monomers Revisited: Toward High-Molecular Weight Methacrylate–POSS (Co)Polymers.Macromolecules2014, 47, 7311–7320. Google Scholar
- 168. Flanders M. J.; Gramlich W. M.Reversible-Addition Fragmentation Chain Transfer (RAFT) Mediated Depolymerization of Brush Polymers.Polym. Chem.2018, 9, 2328–2335. Google Scholar
- 169. Martinez M. R.; Dadashi-Silab S.; Lorandi F.; Zhao Y.; Matyjaszewski K.Depolymerization of P(PDMS11MA) Bottlebrushes via Atom Transfer Radical Polymerization with Activator Regeneration.Macromolecules2021, 54, 5526–5538. Google Scholar
- 170. Beers K. L.; Gaynor S. G.; Matyjaszewski K.; Sheiko S. S.; Möller M.The Synthesis of Densely Grafted Copolymers by Atom Transfer Radical Polymerization.Macromolecules1998, 31, 9413–9415. Google Scholar
- 171. Xie G.; Martinez M. R.; Daniel W. F. M.; Keith A. N.; Ribelli T. G.; Fantin M.; Sheiko S. S.; Matyjaszewski K.Benefits of Catalyzed Radical Termination: High-Yield Synthesis of Polyacrylate Molecular Bottlebrushes without Gelation.Macromolecules2018, 51, 6218–6225. Google Scholar
- 172. Martinez M. R.; Sobieski J.; Lorandi F.; Fantin M.; Dadashi-Silab S.; Xie G.; Olszewski M.; Pan X.; Ribelli T. G.; Matyjaszewski K.Understanding the Relationship between Catalytic Activity and Termination in photoATRP: Synthesis of Linear and Bottlebrush Polyacrylates.Macromolecules2020, 53, 59–67. Google Scholar
- 173. Liu Y.; Chen P.; Li Z.Molecular Bottlebrushes with Polypeptide Backbone Prepared via Ring-Opening Polymerization of NCA and ATRP.Macromol. Rapid Commun.2012, 33, 287–295. Google Scholar
- 174. Olszewski M.; Li L.; Xie G.; Keith A.; Sheiko S. S.; Matyjaszewski K.Degradable Cellulose-Based Polymer Brushes with Controlled Grafting Densities.J. Polym. Sci. A: Polym. Chem.2019, 57, 2426–2435. Google Scholar
- 175. Neugebauer D.; Sumerlin B. S.; Matyjaszewski K.; Goodhart B.; Sheiko S. S.How Dense Are Cylindrical Brushes Grafted from a Multifunctional Macroinitiator?Polymer2004, 45, 8173–8179. Google Scholar
- 176. Sumerlin B. S.; Neugebauer D.; Matyjaszewski K.Initiation Efficiency in the Synthesis of Molecular Brushes by Grafting from via Atom Transfer Radical Polymerization.Macromolecules2005, 38, 702–708. Google Scholar
- 177. Raj W.; Jerczynski K.; Rahimi M.; Przekora A.; Matyjaszewski K.; Pietrasik J.Molecular Bottlebrush with pH-Responsive Cleavable Bonds as a Unimolecular Vehicle for Anticancer Drug Delivery.Mater. Sci. Eng. C2021, 130, 112439. Google Scholar
- 178. Gao H.; Matyjaszewski K.Synthesis of Molecular Brushes by “Grafting Onto” Method: Combination of ATRP and Click Reactions.J. Am. Chem. Soc.2007, 129, 6633–6639. Google Scholar
- 179. Navarro L. A.; Shah T. P.; Zauscher S.Grafting to of Bottlebrush Polymers: Conformation and Kinetics.Langmuir2020, 36, 4745–4756. Google Scholar
- 180. Zhu W.; Zhang L.; Chen Y.; Zhang K.A UV-Cleavable Bottlebrush Polymer with o-Nitrobenzyl-Linked Side Chains.Macromol. Rapid Commun.2017, 38, 1700007. Google Scholar
- 181. Higaki Y.; Otsuka H.; Takahara A.Dynamic Formation of Graft Polymers via Radical Crossover Reaction of Alkoxyamines.Macromolecules2004, 37, 1696–1701. Google Scholar
- 182. Cuthbert J.; Wanasinghe S. V.; Matyjaszewski K.; Konkolewicz D.Are RAFT and ATRP Universally Interchangeable Polymerization Methods in Network Formation?Macromolecules2021, 54, 8331–8340. Google Scholar
- 183. Kamada J.; Koynov K.; Corten C.; Juhari A.; Yoon J. A.; Urban M. W.; Balazs A. C.; Matyjaszewski K.Redox Responsive Behavior of Thiol/Disulfide-Functionalized Star Polymers Synthesized via Atom Transfer Radical Polymerization.Macromolecules2010, 43, 4133–4139. Google Scholar
- 184. Tsarevsky N. V.; Min K.; Jahed N. M.; Gao H.; Matyjaszewski K.Functional Degradable Polymeric Materials Prepared by Atom Transfer Radical Polymerization (ATRP). ACS Symp. Ser.2006, 939, 184–200. Google Scholar
- 185. Lamson M.; Epshtein-Assor Y.; Silverstein M. S.; Matyjaszewski K.Synthesis of Degradable polyHIPEs by AGET ATRP.Polymer2013, 54, 4480–4485. Google Scholar
- 186. Dong H.; Mantha V.; Matyjaszewski K.Thermally Responsive PM(EO)2MA Magnetic Microgels via Activators Generated by Electron Transfer Atom Transfer Radical Polymerization in Miniemulsion.Chem. Mater.2009, 21, 3965–3972. Google Scholar
- 187. Johnson J. A.; Finn M. G.; Koberstein J. T.; Turro N. J.Synthesis of Photocleavable Linear Macromonomers by ATRP and Star Macromonomers by a Tandem ATRP-Click Reaction: Precursors to Photodegradable Model Networks.Macromolecules2007, 40, 3589–3598. Google Scholar
- 188. Johnson J. A.; Lewis D. R.; Diaz D. D.; Finn M. G.; Koberstein J. T.; Turro N. J.Synthesis of Degradable Model Networks via ATRP and Click Chemistry.J. Am. Chem. Soc.2006, 128, 6564–6565. Google Scholar
- 189. Johnson J. A.; Baskin J. M.; Bertozzi C. R.; Koberstein J. T.; Turro N. J.Copper-Free Click Chemistry for the In Situ Crosslinking of Photodegradable Star Polymers.Chem. Commun. (Camb.)2008, 26, 3064–3066. Google Scholar
- 190. Amamoto Y.; Kamada J.; Otsuka H.; Takahara A.; Matyjaszewski K.Repeatable Photoinduced Self-Healing of Covalently Cross-Linked Polymers through Reshuffling of Trithiocarbonate Units.Angew. Chem. Int. Ed. Engl.2011, 50, 1660–1663. Google Scholar
- 191. Scheutz G. M.; Lessard J. J.; Sims M. B.; Sumerlin B. S.Adaptable Crosslinks in Polymeric Materials: Resolving the Intersection of Thermoplastics and Thermosets.J. Am. Chem. Soc.2019, 141, 16181–16196. Google Scholar
- 192. Chen M.; Gu Y.; Singh A.; Zhong M.; Jordan A. M.; Biswas S.; Korley L. T. J.; Balazs A. C.; Johnson J. A.Living Additive Manufacturing: Transformation of Parent Gels into Diversely Functionalized Daughter Gels Made Possible by Visible Light Photoredox Catalysis.ACS Cent. Sci.2017, 3, 124–134. Google Scholar
- 193. Ida S.; Kimura R.; Tanimoto S.; Hirokawa Y.End-Crosslinking of Controlled Telechelic Poly(N-Isopropylacrylamide) toward a Homogeneous Gel Network with Photo-Induced Self-Healing.Polym. J.2017, 49, 237–243. Google Scholar
- 194. Zhang Z.; Corrigan N.; Boyer C.A Photoinduced Dual-Wavelength Approach for 3D Printing and Self-Healing of Thermosetting Materials.Angew. Chem. Int. Ed.2022, 61, e202114111. Google Scholar
- 195. Li L.; Chen X.; Jin K.; Rusayyis M. B.; Torkelson J. M.Arresting Elevated-Temperature Creep and Achieving Full Cross-Link Density Recovery in Reprocessable Polymer Networks and Network Composites via Nitroxide-Mediated Dynamic Chemistry.Macromolecules2021, 54, 1452–1464. Google Scholar
- 196. Su J.; Amamoto Y.; Nishihara M.; Takahara A.; Otsuka H.Reversible Cross-Linking of Hydrophilic Dynamic Covalent Polymers with Radically Exchangeable Alkoxyamines in Aqueous Media.Polym. Chem.2011, 2, 2021–2026. Google Scholar
- 197. Martinez M. R.; Zhuang Z.; Treichel M.; Cuthbert J.; Sun M.; Pietrasik J.; Matyjaszewski K.Thermally Degradable Poly(n-butyl acrylate) Model Networks Prepared by PhotoATRP and Radical Trap-Assisted Atom Transfer Radical Coupling.Polymers2022, 14, 713. Google Scholar
- 198. Pramanik N. B.; Nando G. B.; Singha N. K.Self-Healing Polymeric Gel via RAFT Polymerization and Diels–Alder Click Chemistry.Polymer2015, 69, 349–356. Google Scholar
- 199. Lessard J. J.; Garcia L. F.; Easterling C. P.; Sims M. B.; Bentz K. C.; Arencibia S.; Savin D. A.; Sumerlin B. S.Catalyst-Free Vitrimers from Vinyl Polymers.Macromolecules2019, 52, 2105–2111. Google Scholar
- 200. Lessard J. J.; Scheutz G. M.; Sung S. H.; Lantz K. A.; Epps T. H.; Sumerlin B. S.Block Copolymer Vitrimers.J. Am. Chem. Soc.2020, 142, 283–289. Google Scholar
- 201. Zaborniak I.; Chmielarz P.; Matyjaszewski K.Modification of Wood-Based Materials by Atom Transfer Radical Polymerization Methods.Eur. Polym. J.2019, 120, 109253. Google Scholar
- 202. Wang J.; Zhang D.; Chu F.Wood-Derived Functional Polymeric Materials.Adv. Mater.2021, 33, 2001135. Google Scholar
- 203. Kumar M.; Gehlot P. S.; Parihar D.; Surolia P. K.; Prasad G.Promising Grafting Strategies on Cellulosic Backbone through Radical Polymerization Processes – A Review.Eur. Polym. J.2021, 152, 110448. Google Scholar
- 204. Zhang Z.; Sèbe G.; Hou Y.; Wang J.; Huang J.; Zhou G.Grafting Polymers from Cellulose Nanocrystals via Surface-Initiated Atom Transfer Radical Polymerization.J. Appl. Polym. Sci.2021, 138, 51458. Google Scholar
- 205. Wu H.; Wu L.; Lu S.; Lin X.; Xiao H.; Ouyang X.; Cao S.; Chen L.; Huang L.Robust Superhydrophobic and Superoleophilic Filter Paper via Atom Transfer Radical Polymerization for Oil/Water Separation.Carbohydr. Polym.2018, 181, 419–425. Google Scholar
- 206. Wang L.; Wu Y.; Men Y.; Shen J.; Liu Z.Thermal-Sensitive Starch-g-PNIPAM Prepared by Cu (0) Catalyzed SET-LRP at Molecular Level.RSC Adv.2015, 5, 70758–70765. Google Scholar
- 207. Yamamoto K.; Yoshida S.; Mine S.; Kadokawa J.-i.Synthesis of Chitin-Graft-Polystyrene via Atom Transfer Radical Polymerization Initiated from a Chitin Macroinitiator.Polym. Chem.2013, 4, 3384–3389. Google Scholar
- 208. Pitarresi G.; Fiorica C.; Licciardi M.; Palumbo F. S.; Giammona G.New Hyaluronic Acid Based Brush Copolymers Synthesized by Atom Transfer Radical Polymerization.Carbohydr. Polym.2013, 92, 1054–1063. Google Scholar
- 209. Baker S. L.; Kaupbayeva B.; Lathwal S.; Das S. R.; Russell A. J.; Matyjaszewski K.Atom Transfer Radical Polymerization for Biorelated Hybrid Materials.Biomacromolecules2019, 20, 4272–4298. Google Scholar
- 210. Gong Y.; Leroux J.-C.; Gauthier M. A.Releasable Conjugation of Polymers to Proteins.Bioconjugate Chem.2015, 26, 1172–1181. Google Scholar
- 211. Averick S.; Mehl R. A.; Das S. R.; Matyjaszewski K.Well-Defined Biohybrids Using Reversible-Deactivation Radical Polymerization Procedures.J. Control. Release2015, 205, 45–57. Google Scholar
- 212. Messina M. S.; Messina K. M. M.; Bhattacharya A.; Montgomery H. R.; Maynard H. D.Preparation of Biomolecule-Polymer Conjugates by Grafting-from Using ATRP, RAFT, or ROMP.Prog. Polym. Sci.2020, 100, 101186. Google Scholar
- 213. Pelegri-O’Day E. M.; Lin E.-W.; Maynard H. D.Therapeutic Protein–Polymer Conjugates: Advancing Beyond PEGylation.J. Am. Chem. Soc.2014, 136, 14323–14332. Google Scholar
- 214. Carmali S.; Murata H.; Cummings C.; Matyjaszewski K.; Russell A. J.Chapter Fifteen - Polymer-Based Protein Engineering: Synthesis and Characterization of Armored, High Graft Density Polymer–Protein Conjugates. Methods Enzymol.2017, 590, 347–380. Google Scholar
- 215. De P.; Li M.; Gondi S. R.; Sumerlin B. S.Temperature-Regulated Activity of Responsive Polymer–Protein Conjugates Prepared by Grafting-from via RAFT Polymerization.J. Am. Chem. Soc.2008, 130, 11288–11289. Google Scholar
- 216. Heredia K. L.; Bontempo D.; Ly T.; Byers J. T.; Halstenberg S.; Maynard H. D.In Situ Preparation of Protein–“Smart” Polymer Conjugates with Retention of Bioactivity.J. Am. Chem. Soc.2005, 127, 16955–16960. Google Scholar
- 217. Fu L.; Wang Z.; Lathwal S.; Enciso A. E.; Simakova A.; Das S. R.; Russell A. J.; Matyjaszewski K.Synthesis of Polymer Bioconjugates via Photoinduced Atom Transfer Radical Polymerization under Blue Light Irradiation.ACS Macro Lett.2018, 7, 1248–1253. Google Scholar
- 218. Tucker B. S.; Coughlin M. L.; Figg C. A.; Sumerlin B. S.Grafting-From Proteins Using Metal-Free PET–RAFT Polymerizations under Mild Visible-Light Irradiation.ACS Macro Lett.2017, 6, 452–457. Google Scholar
- 219. Liu J.; Bulmus V.; Herlambang D. L.; Barner-Kowollik C.; Stenzel M. H.; Davis T. P.In Situ Formation of Protein–Polymer Conjugates through Reversible Addition Fragmentation Chain Transfer Polymerization.Angew. Chem. Int. Ed.2007, 46, 3099–3103. Google Scholar
- 220. Averick S.; Simakova A.; Park S.; Konkolewicz D.; Magenau A. J. D.; Mehl R. A.; Matyjaszewski K.ATRP under Biologically Relevant Conditions: Grafting from a Protein.ACS Macro Lett.2012, 1, 6–10. Google Scholar
- 221. Lathwal S.; Yerneni S. S.; Boye S.; Muza U. L.; Takahashi S.; Sugimoto N.; Lederer A.; Das S. R.; Campbell P. G.; Matyjaszewski K.Engineering Exosome Polymer Hybrids by Atom Transfer Radical Polymerization.Proc. Natl. Acad. Sci.2021, 118, e2020241118. Google Scholar
- 222. Murata H.; Cummings C. S.; Koepsel R. R.; Russell A. J.Polymer-Based Protein Engineering Can Rationally Tune Enzyme Activity, pH-Dependence, and Stability.Biomacromolecules2013, 14, 1919–1926. Google Scholar
- 223. Kaupbayeva B.; Murata H.; Matyjaszewski K.; Russell A. J.; Boye S.; Lederer A.A Comprehensive Analysis in One Run–in-Depth Conformation Studies of Protein–Polymer Chimeras by Asymmetrical Flow Field-Flow Fractionation.Chem. Sci.2021, 12, 13848–13856. Google Scholar
- 224. Baker S. L.; Munasinghe A.; Kaupbayeva B.; Rebecca Kang N.; Certiat M.; Murata H.; Matyjaszewski K.; Lin P.; Colina C. M.; Russell A. J.Transforming Protein-Polymer Conjugate Purification by Tuning Protein Solubility.Nat. Commun.2019, 10, 4718. Google Scholar
- 225. Baker S. L.; Murata H.; Kaupbayeva B.; Tasbolat A.; Matyjaszewski K.; Russell A. J.Charge-Preserving Atom Transfer Radical Polymerization Initiator Rescues the Lost Function of Negatively Charged Protein–Polymer Conjugates.Biomacromolecules2019, 20, 2392–2405. Google Scholar
- 226. Kaupbayeva B.; Murata H.; Lucas A.; Matyjaszewski K.; Minden J. S.; Russell A. J.Molecular Sieving on the Surface of a Nano-Armored Protein.Biomacromolecules2019, 20, 1235–1245. Google Scholar
- 227. Murata H.; Carmali S.; Baker S. L.; Matyjaszewski K.; Russell A. J.Solid-Phase Synthesis of Protein-Polymers on Reversible Immobilization Supports.Nat. Commun.2018, 9, 845. Google Scholar
- 228. Heredia K. L.; Nguyen T. H.; Chang C.-W.; Bulmus V.; Davis T. P.; Maynard H. D.Reversible siRNA–Polymer Conjugates by RAFT Polymerization.Chem. Commun.2008, 28, 3245–3247. Google Scholar
- 229. Bontempo D.; Heredia K. L.; Fish B. A.; Maynard H. D.Cysteine-Reactive Polymers Synthesized by Atom Transfer Radical Polymerization for Conjugation to Proteins.J. Am. Chem. Soc.2004, 126, 15372–15373. Google Scholar
- 230. Vázquez-Dorbatt V.; Tolstyka Z. P.; Chang C.-W.; Maynard H. D.Synthesis of a Pyridyl Disulfide End-Functionalized Glycopolymer for Conjugation to Biomolecules and Patterning on Gold Surfaces.Biomacromolecules2009, 10, 2207–2212. Google Scholar
- 231. Gunasekaran K.; Nguyen T. H.; Maynard H. D.; Davis T. P.; Bulmus V.Conjugation of siRNA with Comb-Type PEG Enhances Serum Stability and Gene Silencing Efficiency.Macromol. Rapid Commun.2011, 32, 654–659. Google Scholar
- 232. York A. W.; Huang F.; McCormick C. L.Rational Design of Targeted Cancer Therapeutics through the Multiconjugation of Folate and Cleavable siRNA to RAFT-Synthesized (HPMA-s-APMA) Copolymers.Biomacromolecules2010, 11, 505–514. Google Scholar
- 233. Averick S.; Paredes E.; Li W.; Matyjaszewski K.; Das S. R.Direct DNA Conjugation to Star Polymers for Controlled Reversible Assemblies.Bioconjugate Chem.2011, 22, 2030–2037. Google Scholar
- 234. Pyun J.; Jia S.; Kowalewski T.; Patterson G. D.; Matyjaszewski K.Synthesis and Characterization of Organic/Inorganic Hybrid Nanoparticles: Kinetics of Surface-Initiated Atom Transfer Radical Polymerization and Morphology of Hybrid Nanoparticle Ultrathin Films.Macromolecules2003, 36, 5094–5104. Google Scholar
- 235. von Werne T.; Patten T. E.Preparation of Structurally Well-Defined Polymer−Nanoparticle Hybrids with Controlled/Living Radical Polymerizations.J. Am. Chem. Soc.1999, 121, 7409–7410. Google Scholar
- 236. Wang Z.; Fantin M.; Sobieski J.; Wang Z.; Yan J.; Lee J.; Liu T.; Li S.; Olszewski M.; Bockstaller M. R.; Matyjaszewski K.Pushing the Limit: Synthesis of SiO2-g-PMMA/PS Particle Brushes via ATRP with Very Low Concentration of Functionalized SiO2–Br Nanoparticles.Macromolecules2019, 52, 8713–8723. Google Scholar
- 237. Costanzo P. J.; Beyer F. L.Thermally Driven Assembly of Nanoparticles in Polymer Matrices.Macromolecules2007, 40, 3996–4001. Google Scholar
- 238. Sato T.; Ohishi T.; Higaki Y.; Takahara A.; Otsuka H.Radical Crossover Reactions of Alkoxyamine-Based Dynamic Covalent Polymer Brushes on Nanoparticles and the Effect on Their Dispersibility.Polym. J.2016, 48, 147–155. Google Scholar
- 239. Sato T.; Amamoto Y.; Yamaguchi H.; Ohishi T.; Takahara A.; Otsuka H.Dynamic Covalent Polymer Brushes: Reversible Surface Modification of Reactive Polymer Brushes with Alkoxyamine-Based Dynamic Covalent Bonds.Polym. Chem.2012, 3, 3077–3083. Google Scholar
- 240. Claus T. K.; Telitel S.; Welle A.; Bastmeyer M.; Vogt A. P.; Delaittre G.; Barner-Kowollik C.Light-Driven Reversible Surface Functionalization with Anthracenes: Visible Light Writing and Mild UV Erasing.Chem. Commun.2017, 53, 1599–1602. Google Scholar
- 241. Yardley R. E.; Kenaree A. R.; Gillies E. R.Triggering Depolymerization: Progress and Opportunities for Self-Immolative Polymers.Macromolecules2019, 52, 6342–6360. Google Scholar
- 242. Sagi A.; Weinstain R.; Karton N.; Shabat D.Self-Immolative Polymers.J. Am. Chem. Soc.2008, 130, 5434–5435. Google Scholar
- 243. Sawada H.Chapter 3. Thermodynamics of Radical Polymerization.J. Macromol. Sci. C: Polym. Rev.1969, 3, 357–386. Google Scholar
- 244. Odian G.Principles of Polymerization, 4th ed.; John Wiley & Sons, Inc.: Hoboken, NJ, 2004. Google Scholar
- 245. Hutchinson R. A.; Paquet D. A.; Beuermann S.; McMinn J. H.Investigation of Methacrylate Free-Radical Depropagation Kinetics by Pulsed-Laser Polymerization.Ind. Eng. Chem. Res.1998, 37, 3567–3574. Google Scholar
- 246. Wang H. S.; Truong N. P.; Pei Z.; Coote M. L.; Anastasaki A.Reversing RAFT Polymerization: Near-Quantitative Monomer Generation Via a Catalyst-Free Depolymerization Approach.J. Am. Chem. Soc.2022, 144, 4678–4684. Google Scholar
- 247. Xu J.; He J.; Fan D.; Tang W.; Yang Y.Thermal Decomposition of Dithioesters and Its Effect on RAFT Polymerization.Macromolecules2006, 39, 3753–3759. Google Scholar
- 248. Chong B.; Moad G.; Rizzardo E.; Skidmore M.; Thang S. H.Thermolysis of RAFT-Synthesized Poly(methyl Methacrylate).Aust. J. Chem.2006, 59, 755–762. Google Scholar
- 249. Sano Y.; Konishi T.; Sawamoto M.; Ouchi M.Controlled Radical Depolymerization of Chlorine-Capped PMMA via Reversible Activation of the Terminal Group by Ruthenium Catalyst.Eur. Polym. J.2019, 120, 109181. Google Scholar
- 250. Martinez M. R.; De Luca Bossa F.; Olszewski M.; Matyjaszewski K.Copper(II) Chloride/Tris(2-pyridylmethyl)amine-Catalyzed Depolymerization of Poly(n-butyl methacrylate).Macromolecules2022, 55, 78–87. Google Scholar
- 251. Jackson A. T.; Bunn A.; Priestnall I. M.; Borman C. D.; Irvine D. J.Molecular Spectroscopic Characterisation of Poly(methyl Methacrylate) Generated by Means of Atom Transfer Radical Polymerisation (ATRP).Polymer2006, 47, 1044–1054. Google Scholar
- 252. Bekanova M. Z.; Neumolotov N. K.; Jablanović A. D.; Plutalova A. V.; Chernikova E. V.; Kudryavtsev Y. V.Thermal Stability of RAFT-Based Poly(methyl Methacrylate): A Kinetic Study of the Dithiobenzoate and Trithiocarbonate End-Group Effect.Polym. Degrad. Stab.2019, 164, 18–27. Google Scholar
- 253. Li I.; Howell B. A.; Matyjaszewski K.; Shigemoto T.; Smith P. B.; Priddy D. B.Kinetics of Decomposition of 2,2,6,6-Tetramethyl-1-(1-phenylethoxy)piperidine and Its Implications on Nitroxyl-Mediated Styrene Polymerization.Macromolecules1995, 28, 6692–6693. Google Scholar
- 254. Garrison J. B.; Hughes R. W.; Sumerlin B. S.Backbone Degradation of Polymethacrylates via Metal-Free Ambient-Temperature Photoinduced Single-Electron Transfer.ACS Macro Lett.2022, 11, 441–446. Google Scholar
- 255. Pesenti T.; Nicolas J.100th Anniversary of Macromolecular Science Viewpoint: Degradable Polymers from Radical Ring-Opening Polymerization: Latest Advances, New Directions, and Ongoing Challenges.ACS Macro Lett.2020, 9, 1812–1835. Google Scholar