Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Functional biodegradable polymers via ring-opening polymerization of monomers without protective groups

Greta Becker ab and Frederik R. Wurm *a
aMax Planck Institute for Polymer Research, Ackermannweg 10, 55128 Mainz, Germany. E-mail: wurm@mpip-mainz.mpg.de
bGraduate School Materials Science in Mainz, Staudinger Weg 9, 55128 Mainz, Germany

Received 28th June 2018

First published on 17th September 2018


Abstract

Biodegradable polymers are of current interest and chemical functionality in such materials is often demanded in advanced biomedical applications. Functional groups often are not tolerated in the polymerization process of ring-opening polymerization (ROP) and therefore protective groups need to be applied. Advantageously, several orthogonally reactive functions are available, which do not demand protection during ROP. We give an insight into available, orthogonally reactive cyclic monomers and the corresponding functional synthetic and biodegradable polymers, obtained from ROP. Functionalities in the monomer are reviewed, which are tolerated by ROP without further protection and allow further post-modification of the corresponding chemically functional polymers after polymerization. Synthetic concepts to these monomers are summarized in detail, preferably using precursor molecules. Post-modification strategies for the reported functionalities are presented and selected applications highlighted.


image file: c8cs00531a-p1.tif

Greta Becker

Greta Becker studied biomedical chemistry at the University of Mainz, Germany, and at the Polymer Science and Engineering Department, the University of Massachusetts in Amherst, USA. She carried out her PhD research in the group of Dr Frederik R. Wurm at the Max Planck Institute for Polymer Research in Mainz, Germany and received her PhD degree in 2017. Great developed a series of degradable and polyfunctional polyphosphoesters for bioapplications. She was supported by a fellowship through funding of the Excellence Initiative (DFG/GSC 266) in the context of the graduate school of excellence “MAINZ” (Material Sciences in Mainz) and is currently working for Kuraray Europe GmbH.

image file: c8cs00531a-p2.tif

Frederik R. Wurm

Frederik R. Wurm (Priv.-Doz. Dr habil.) is currently heading the research group “Functional Polymers” at the Max Planck Institute for Polymer Research (MPIP), Mainz (D). In his interdisciplinary research, Frederik designs polymeric materials with molecular-defined functions. Controlling the monomer sequence and chemical functionality allowed designing materials for degradable polymers, nanocarriers with controlled blood interactions, and phosphorus flame-retardants. He has published more than 150 research articles to date. He was awarded the Reimund Stadler Award in 2016, the Lecturer Award of the German Chemical Industry Fund in 2017, the European Young Chemist Award and the Georg Manecke Prize of the German Chemical Society in 2014. Frederik received his PhD in 2009 (JGU Mainz, D). After a two-year stay at EPFL (CH) as a Humboldt fellow, he joined the department “Physical Chemistry of Polymers” at MPIP and finished his habilitation in Macromolecular Chemistry in 2016.


1. Introduction

Biodegradable polymers are of great current interest for biomedical applications, e.g. for drug and gene delivery systems, bioengineering scaffolds or as bioadhesives. They employ binding motifs within their backbone, inspired by natural biopolymers, e.g. polysaccharides, polyhydroxyalkanoates, polypeptides or (deoxy)ribonucleic acids (DNA and RNA). A broad range of synthetic biodegradable polymer classes has been developed so far, including polyesters, polyamides, polycarbonates, poly(phosphoester)s, polyphosphazenes, poly(ester amide)s, poly(ester-ether)s, poly(ester-anhydride)s, poly(ester urethane)s, poly(ester urea)s, polyacetals, polyorthoesters, polydioxanones and polyiminocarbonates, which can be obtained by step-growth polycondensation or -addition or chain-growth polymerization.1 Especially, when it comes to advanced applications, chemical functionality in such materials is demanded, e.g. to attach labels or other molecules along the backbone.

With a plethora of modern catalysts, the chain growth approach has a higher control over polymer molar masses and dispersities. Different mechanisms are available, including cationic, anionic, enzymatic, coordinative and radical ring-opening polymerization (ROP). Copolymerization of different cyclic monomers with pendant alkyl or aryl groups gives access to a variety of polymeric materials with a broad range of different physical properties, e.g. varying hydrophilicity/hydrophobicity, crystallinity, solubility, mechanical strength, degradation behavior or thermal stability. Such degradable polymers are also important for the future of sustainable polymers and plastics.2 Properties and features, as well as their advantages and drawbacks of the different classes of synthetic biodegradable polymers, are beyond our scope and are extensively discussed in several reviews.2–6

While copolymerization of alkylated and arylated monomers adjusts the materials properties, fine-tuning of the polymers is often demanded for specific applications: additional attachment of bioactive molecules, redox- or pH-sensitive functionalities or cross-linkable groups might be required for their applications. On the one hand, (especially) ionic ROP might be sensitive to impurities and tolerates only certain chemical functionalities. The sensitivity to moisture and thereby the exclusion of water as reaction solvent is a drawback. On the other hand, also bioactive molecules (e.g. carbohydrates, peptides or proteins) can be sensitive or undergo side-reactions that they do not tolerate the polymerization process or conditions, e.g. organic solvents, high temperatures or required catalysts. Great effort has been made in the last decades, developing cyclic monomers with orthogonal chemical functions, which do not interfere the polymerization process. These monomers can be divided into two groups: (I) orthogonally reactive groups that do not interfere with the polymerization but can be post-modified afterward; (II) active groups, e.g. photo- or redox-active.

In this review, we summarize synthetic strategies to orthogonally reactive cyclic monomers reported in the literature that allows subsequent post-polymerization modification. We highlight the general concepts, preferably using precursor molecules, which can be used to prepare these monomers and thereby chemically functional biodegradable polymers by ROP (Table 1). A comparison on the synthetic ease of the different monomer classes will be given, that helps to choose the polymer class of choice for the desired application. We further display post-modification strategies with selected applications.

Table 1 Overview of the monomers and polymer classes discussed in this review
Polymer class General structure Cyclic monomers
Polyesters image file: c8cs00531a-u1.tif Lactone
Macrolactone
Glycolide
Lactide
Hemilactide
O-Carboxyanhydride (OCA)
Polyamides image file: c8cs00531a-u2.tif Lactam
α-N-Carboxy anhydride (α-NCA)
γ-N-Carboxy anhydride (γ-NCA)
Poly(ester amide)s image file: c8cs00531a-u3.tif Esteramide
Polycarbonates image file: c8cs00531a-u4.tif Trimethylene carbonate (TMC)
Polyphosphoesters image file: c8cs00531a-u5.tif Phosphate
Phosphite
Phosphonate
Polyphosphazenes image file: c8cs00531a-u6.tif Hexachlorophosphazene


The scope of the review is to be a handbook on the preparation of orthogonally reactive cyclic monomers to deliver a “toolbox” on how functional synthetic biodegradable polymers are prepared and post-modified. Tables after each section summarize the monomers discussed in the text, together with literature references and some comments.

All the herein discussed polymer classes are potentially degradable or biodegradable, due to certain linkages in the backbone. The degradation profile is one of the most important features of these polymers, depending on their area of application. Several of the examples given in this review are claimed to be degradable, due to labile ester or amide linkages in the backbone, although degradation behavior was not studied in detail. Degradation is possible by acidic, alkaline, enzymatic, microbial or oxidative cleavage of ester/amide bonds. The comparison of degradation rates and conditions is difficult, as the degradation profiles depend on various factors: the hydrophilicity or hydrophobicity, water-solubility, crystallinity, glass transition, and/or glass transition temperature, processing, size, geometry (in bulk, as foams, thin fibers, nanoparticles, micelles, in solution, etc.), porosity and water diffusion (Table 2). In addition, the degree of polymerization, sterics of any substituents, polymer architecture, and solubility of degradation products have a strong impact on the degradation rates as well. Another factor that makes the comparison even in one polymer class difficult are additional post-modifications, e.g. with hydrophobic, polar or charged groups that further alter the degradation profiles.

Table 2 Overview of parameters influencing the degradability of polymers and polymeric materials
Degradation Influencing parameters of the…
Polymer Sample Procedure
– Hydrolytic: – Hydrophilicity/hydrophobicity – Processing – Choice of enzyme:
• Acidic – Water-solubility – Size/geometry: • Origin
• Basic – The degree of polymerization • Bulk • Activity
– Enzymatic – The glass transition temperature • Foam • Selectivity
– Microbial – Crystallinity • Fibers – Physiological/non-physiological conditions
– Oxidative – Sterics of substituents • Nanoparticles – pH:
– Architecture (linear/branched/cross-linked) • Micelles • Acidic
– The solubility of degradation products • In solution • Basic
– Post-modifications – Porosity • Molarity
– Water diffusion – Buffer:
• Buffer-system
• Capacity
– Concentration
– Temperature
In vitro/in vivo


The protocols for polymer degradation are diverse and lack standardized conditions, which makes most degradation studies non-comparable with each other (detailed information can be found in a recent review).7 Most common degradation mechanism for the polymers discussed herein is hydrolysis of the polymer backbone. In most cases either acidic or basic hydrolysis are conducted under non-physiological conditions, i.e. at very low or high pH values that do not occur in natural environments. Furthermore, the chemical nature of the buffer solution, buffer-capacity, temperature, and concentration of polymer (if water-soluble) or shape of the specimen is different for most studies. For the enzymatic degradation, different enzymes can be applied, which may stem from different organisms and vary in their activity. Even batch-to-batch variations of the very same enzyme makes standardization of in vitro degradations difficult (overview of parameters shown in Table 1).

Trying to summarize some general aspects of degradation profile, herein we give some examples of non-functionalized polymers. Hydrolysis or enzymatic degradation are the typical degradation mechanism for such materials, with kinetics being very dependent on the environment and the chemical structure and/or crystallinity of the polymers. While polycaprolactone shows rather a slow degradation rate (within 2–3 years), due to its crystallinity, polylactide (depending on the chirality and composition) undergoes loss of mass within 6–16 months; polyglycolide (45–55% crystallinity) is known to lose mass within 6–24 months. Copolymers of poly(D,L-lactide-co-glycolide) are reported to degrade faster, depending on the composition ratio, within 5–6 months. Polyesters hydrolyze under acidic and basic conditions;8 in contrast, some polyphosphoesters can be very stable under acidic conditions but degrade in the presence of a base. For polyphosphates, a typical water-soluble example is poly(methyl ethylene phosphate); while being stable at low pH, degradation of triester to diester bonds occurs under alkaline conditions within 5 h (at pH 12.3) to 21 months (pH 7.3). (Note: these are degradation times for 50% cleavage of the ester bonds in the main chain of the polymer.)9 Polyphosphonates with the P–C bond in the side chain show similar degradation profiles under neutral and basic conditions. Complete degradation was observed after 1 hour at pH 12.10 Contrary, polyphosphoramidates undergo hydrolysis in basic and acidic media.11–15 While hydrolysis almost exclusively proceeds at the P–N bond under acidic and nearly neutral conditions, P–O, as well as P–N bond cleavage, occurs under basic conditions, still with a higher probability for P–N cleavage.14 94% cleavage of main-chain polyphosphoramidates to diesters has been shown at pH 3.0 within 12.5 days.11 The degradation profile of polyphosphazenes strongly depends on the substituents and ranges from hydrolytically stable (with hydrophobic, bulky alkoxy side groups) to hydrolytically unstable (with hydrophilic amino substituents). Degradation of the P[double bond, length as m-dash]N-backbone is commonly accelerated in acidic media, but they are rather inert under basic conditions.16,17 The biodegradation of synthetic aliphatic polyamides is known to be low due to high crystallinity. Enzymatic or microbial degradation has been shown.18Fig. 1 gives a rough overview of the systematic order of degradability of the discussed polymer classes, however, the data has to be taken as an estimation only, as many factors as mentioned above may influence absolute values. A recent review summarizes the degradation of polylactides.19 We refer the interested reader to separate reviews concentrating on the degradation of synthetic polymers.3,4,16,17,20–23


image file: c8cs00531a-f1.tif
Fig. 1 Overview of the systematic order of degradability of the discussed polymer classes. *[thin space (1/6-em)]Note: degradation profiles of polyphosphazenes and polyphosphoesters depend strongly on the nature of the substituents.

2. Overview of orthogonally reactive groups

ROP only tolerates some additional chemical functionalities in the monomers. Since alcohols, thiols, amines or carboxylic acids interfere with the propagation and serve as initiators or terminating agents, they need to be protected before polymerization. Commonly used protecting groups, e.g. benzyl, benzoyl, ethers, silyl ethers, acetals, urethanes, sulfonamides or esters are applied in cyclic monomers for ROP. Removal of protection groups is conducted after the polymerization under alkaline or acidic conditions, or by hydrogenation and often demands harsh conditions, which might also degrade the polymer backbone. We do not further consider these protected monomers in the review and concentrate exclusively on orthogonally reactive functionalities (Fig. 2). For protected monomers, we refer to other reviews, specializing in the respective polymer class.24–28 The same functional groups can also be installed into the initiator structure, in order to prepare end-functionalized polymers (telechelics).29,30
image file: c8cs00531a-f2.tif
Fig. 2 Overview on cyclic monomers discussed in the review with orthogonally reactive groups.

For the orthogonal functions, alkynes and alkenes are by far most frequently reported in the literature and are used in the monomers and initiator structures. Alkynes undergo the copper(I)-catalyzed azide–alkyne cycloaddition (CuAAC, Huisgen 1,3-dipolar cycloaddition) with azides or can react in a thiol–yne reaction with thiols in often quantitative yield and mild conditions. Also, azides as the “counterpart” are reported as functionality in monomers. Besides CuAAC, they can be additionally modified with DBCO derivatives in a strain-promoted alkyne–azide cycloaddition (SPAAC),31 which turns the reaction into a copper-free functionalization and is especially interesting for biomedical applications. Alkenes are accessible for probably the most modification reactions: besides thiol–ene reaction and Michael addition, epoxidation (e.g. by mCPBA), dihydroxylation, hydroboration, ozonolysis, hydrazination, hydrogenation, bromination, hydrobromination, and others are applicable. Especially epoxidation opens a platform for diverse further reaction. Also, few monomers are reported that directly carry an epoxide, which under certain conditions does not interfere with ROP. Epoxides can cross-link the materials, react with thiols, be dihydroxylated or further polymerized. If alkene functions are vinylidenes, the cyclic monomers are bifunctional for radical polymerization or can serve as cross-linkers as well. They furthermore can be used for olefin cross-metathesis or Suzuki coupling. Acrylate, methacrylate and styrenic functions likewise can be radically polymerized, cross-link the materials, undergo thiol–ene reaction and Michael addition or are accessible for olefin cross-metathesis. Cinnamoyl groups serve as cross-linkers. Likewise, methylidene functions can be polymerized or cross-link materials and undergo thiol–ene reaction, which can be also achieved with norbornene groups, additionally suitable for 1,3-dipolar cycloadditions and ring-opening metathesis polymerization (ROMP). Completing the group of double bond-containing functionalities, internal double bonds are accessible for epoxidation, dihydroxylation, and cross-metathesis, while vinyl ethers are interesting reaction partners for thiol–ene reaction, acetal- and thioacetalisation.

Halogenated monomers are a second important category, especially with bromide or chloride substituents. Nucleophilic substitution e.g. with sodium azide and quaternization of tertiary amines or phosphines has been reported, as well as dehydrohalogenation or boration. Iodide substituted monomers play a minor part, but can also be used for nucleophilic substitution and quaternization of amines, or are used as a radio-opaque function, e.g. for contrast agents. Such halogenated polymers have also been used extensively as initiators for atom transfer radical polymerization (ATRP) to prepare graft or brush (co)polymers. Several bromo isobutyrate-containing monomers were developed for the same purpose, as well as trithiocarbonate monomers for reversible addition–fragmentation chain transfer polymerization (RAFT).

A third category includes more exotic, but at the same time very interesting and partly unexpected chemical functionality: besides the trithiocarbonate-containing monomers for RAFT polymerization, several further sulfur-containing monomers are introduced, bearing disulfide or S-sulfonyl groups. The functional groups do not interfere with ROP and can be considered as “protected thiols”. Functionalization is achieved with thiols by disulfide exchange reaction without any prior deprotection reaction: dynamic and redox-responsive cross-linking are accessible. Methyl-thioether functions can undergo reversible alkylation reaction, additionally implementing cationic charges. Vinyl sulfonyl moieties can react in Michael addition reactions. P–H bonds of cyclic H-phosphonate monomers are suitable for modification by esterification, amidation (after chlorination), hydrolysis and sulfurization. The P–H bond has not been reported yet in pendant chains. Ketones within the ring the cyclic monomer are accessible for reduction, hydrazination, and hydrazonation reactions. Benzophenone groups can be used as photo-cross-linkers by a C,H-insertion crosslinking reaction (CHic mechanism32,33) with CH groups. Cross-linking can also be achieved by catechol functions (1,2-dihydroxybenzene), either reversible by metal ion complexation or covalently by reaction with amine, thiols or other catechols after oxidation to quinone intermediates. In addition, active ester-containing monomers have been reported, such as trichloroethyl-, NHS- (N-hydroxysuccinimide) and pentafluorophenyl-ester groups, which undergo amidation and esterification reactions with alcohols or amines after polymerization.34,35 Finally, anthracene and furan derivatives are suitable for [4+2] cycloaddition Diels–Alder reactions. However, reports on this thermally reversible modification by additive/catalyst-free cycloaddition are rare, which might be a further potential for future applications.

In general, also orthogonal reactions have limitations. This starts with the degree of conversion, which is not always quantitative and leaves non-reacted groups behind. In addition, polymerization conditions of functional monomers need to be carefully selected. For instance, pentafluorophenyl esters can only be polymerized using acid catalysis, thiol derivatives in some cases are complicated to work with (for example due to formation of disulfides), catechol-functionalized compounds should be handled under oxygen-free atmosphere to avoid oxidation, and epoxides can also be ring-opened during the polymerization, if not carefully handled (e.g. at elevated temperatures). Moreover, the effectiveness of the post-functionalization methods can vary.36 Radical thiol–ene and thiol–yne reactions often need a large excess of the thiol to prevent unwanted crosslinking reactions. For click reactions, the removal of the copper catalyst needs to be taken into account, etc. Such issues need always to be considered and are sometimes not clearly mentioned in publications. We refer to some general reviews about post-polymerization modifications that can be considered as additional reading.25,36–38

3. Polyesters

Aliphatic polyesters from ROP probably display the broadest group of fully synthetic biodegradable polymers. A variety of lactones with different ring sizes are available resulting in poly(ε-caprolactone)s, poly(δ-valerolactone)s, poly(γ-butyrolactone)s, poly(β-butyrolactone)s, poly(β-propiolactone)s, and poly(α-propiolactone)s. Macrolactone monomers are likewise polymerizable. Furthermore, cyclic diesters (lactides, glycolides and O-carboxy anhydrides (OCAs)) produce the commercialized poly(α-hydroxy acid)s (PAHAs) polylactides (PLAs) and polyglycolides (PGAs) (Table 3). Several reviews about functional aliphatic poly(ester)s have been published and might be considered for further reading.26,39,40
Table 3 Orthogonally functional cyclic monomers for the synthesis of polyesters
Monomer R = No. Post-modification Ref.
ε-Caprolactones
α-Substituted
image file: c8cs00531a-u7.tif image file: c8cs00531a-u8.tif A1 – No modification 48
image file: c8cs00531a-u9.tif A2 – Nucleophilic substitution with sodium azide and click chemistry 46 and 50
– ATRP macroinitiator for “grafting from” of MMA
image file: c8cs00531a-u10.tif A3 – Nucleophilic substitution with sodium azide and click chemistry 49 and 51
– ATRP macroinitiator
image file: c8cs00531a-u11.tif A5 – Radio-opaque properties 53
image file: c8cs00531a-u12.tif A4 – Click chemistry 52
image file: c8cs00531a-u13.tif A7 – Click reaction with cyclodextrin or Gd3+-complexes 55, 58 and 59
image file: c8cs00531a-u14.tif A8 – Click reaction to core-cross-linked micelles 56
image file: c8cs00531a-u15.tif A6 – Thiol–ene reaction with amines, dyes, sugars or zwitterions 54 and 57
image file: c8cs00531a-u16.tif A9 – End-chain cross-linker for macrocyclic polyester 60
image file: c8cs00531a-u17.tif A10 – No modification 61
ε-Substituted
image file: c8cs00531a-u18.tif image file: c8cs00531a-u19.tif A7b – Click reaction to couple cyclodextrin 55 and 59
β-Substituted
image file: c8cs00531a-u20.tif image file: c8cs00531a-u21.tif A11 – No modification 63
image file: c8cs00531a-u22.tif A12 – Epoxidation and thiol–ene reaction to cross-link 64
γ-Substituted
image file: c8cs00531a-u23.tif image file: c8cs00531a-u24.tif A13 – Nucleophilic substitution with sodium azide and click chemistry with a cholesterol derivative for cell scaffolds and foams 65
image file: c8cs00531a-u25.tif A14 – Quaternization with pyridine 66–68
– Elimination, epoxidation, and ring-opening to diols
– Nucleophilic substitution with sodium azide and click chemistry
image file: c8cs00531a-u26.tif A15 – Hydrazination or hydrogenation 69–71, 151 and 152
– Reduction to alcohols and use as macroinitiator or coupling of maleic anhydride
image file: c8cs00531a-u27.tif A16 – Bifunctional polymerization 72–75
– Electrografting onto metal surfaces
– 2D- and 3D-microstructured resins
– Michael-addition of thiols
image file: c8cs00531a-u28.tif A17 – Photo-cross-linking 73 and 76
image file: c8cs00531a-u29.tif A18 – ATRP initiator or macroinitiator 77
image file: c8cs00531a-u30.tif A19 – Photo-cross-linking 78
δ-Valerolactones
image file: c8cs00531a-u31.tif image file: c8cs00531a-u32.tif A20 – Dihydroxylation with NMO/OsO4 80 and 85
image file: c8cs00531a-u33.tif A22 – Click chemistry with PEG-, GRGDS-, phosphorylcholine or benzophenone-azides 79, 82, 84 and 85
image file: c8cs00531a-u34.tif A21 – Dihydroxylation and PEG “grafting to” 81
image file: c8cs00531a-u35.tif A23 – Radical copolymerization with methacrylates to form networks 86
image file: c8cs00531a-u36.tif A24 – No polymerization 87
image file: c8cs00531a-u37.tif A25 – No polymerization 87
image file: c8cs00531a-u38.tif A26 – No polymerization 87
γ-Butyrolactones
image file: c8cs00531a-u39.tif image file: c8cs00531a-u40.tif A27 – Copolymerization with CL and cross-linking with methacrylate 92 and 93
image file: c8cs00531a-u41.tif A28 – Co- and terpolymerization with CL or TMCs and grafting from of methacrylates 95
β-Propiolactones
image file: c8cs00531a-u42.tif image file: c8cs00531a-u43.tif A29 – Tacticity studies 96
image file: c8cs00531a-u44.tif A30 – Tacticity studies 96
image file: c8cs00531a-u45.tif A31 – Tacticity studies 96 and 98
image file: c8cs00531a-u46.tif A32 – Tacticity and property studies 98
image file: c8cs00531a-u47.tif A33 – Tacticity and property studies 99
image file: c8cs00531a-u48.tif A34 – Tacticity and property studies 99
image file: c8cs00531a-u49.tif A39 – Selective polymerization studies 103
image file: c8cs00531a-u50.tif A40 – Selective polymerization studies 103
image file: c8cs00531a-u51.tif A41 – Selective polymerization studies 103
image file: c8cs00531a-u52.tif A42 – Hydroboration or olefin cross metathesis 104–107
image file: c8cs00531a-u53.tif A43 – Epoxidation and sulfonation to polymers inducing bone formation; thiol–ene reaction to macroinitiator for “grafting from” 108 and 110–112
image file: c8cs00531a-u54.tif A44 – Epoxidation 108
image file: c8cs00531a-u55.tif A45 – No modification 109
β-Butyrolactones
image file: c8cs00531a-u56.tif image file: c8cs00531a-u57.tif A35 – Tacticity and property studies 97
image file: c8cs00531a-u58.tif A36 – Tacticity and property studies 97
image file: c8cs00531a-u59.tif image file: c8cs00531a-u60.tif A37 – Quaternization with pyridine 100 and 101
image file: c8cs00531a-u61.tif A38 – Polymerization studies 102
image file: c8cs00531a-u62.tif A46 – Radical cross-linking or thiol–ene reaction 113 and 114
Macrolactones
image file: c8cs00531a-u63.tif A47 – Thiol–ene reaction for functionalization with pendant chains, cross-linking or functionalization with ATRP initiators for “grafting from” of tert-butyl acrylate 119 and 122
image file: c8cs00531a-u64.tif A48 – Epoxidation 125
image file: c8cs00531a-u65.tif A49 – No modification 125
image file: c8cs00531a-u66.tif A50 – Radical cross-linking 126
Mono-substituted glycolides
image file: c8cs00531a-u67.tif image file: c8cs00531a-u68.tif A51 – Epoxidation, dihydroxylation 127
image file: c8cs00531a-u69.tif A54 – Click chemistry with PEG–azide 128
Mono-substituted hemilactides
image file: c8cs00531a-u70.tif image file: c8cs00531a-u71.tif A52 – Thiol–ene reaction with amines for gene delivery 133 and 135
– Photo-cross-linking of nanoparticles/capsules
image file: c8cs00531a-u72.tif A55 – Click chemistry with PEG-/palitaxcel-azide 129
image file: c8cs00531a-u73.tif A58 – Click chemistry with dansyl-azide 138
image file: c8cs00531a-u74.tif A59 – Click chemistry with dansyl-alkyne 138
image file: c8cs00531a-u75.tif A60 – Click chemistry with dansyl-alkyne 138
image file: c8cs00531a-u76.tif A61 – Click chemistry with dye/cell internalizing peptide 139 and 140
– Staudinger condensation with Tap-GRGDS
image file: c8cs00531a-u77.tif A53 – Cross-linking 134
image file: c8cs00531a-u78.tif A62 – ROMP 130 and 131
– Click reaction with tetrazine derivatives
image file: c8cs00531a-u79.tif A64 – No modification 142
image file: c8cs00531a-u80.tif A65 – No modification 142
image file: c8cs00531a-u81.tif A63 – No modification 141
Di-substituted glycolides
image file: c8cs00531a-u82.tif image file: c8cs00531a-u83.tif A56 – Click chemistry with azide derivatives 136
image file: c8cs00531a-u84.tif A57 – Formation of double bonds 137
– Thiol–ene reaction
O-Carboxyanhydrides (OCAs)
image file: c8cs00531a-u85.tif image file: c8cs00531a-u86.tif A66 – Cross-linking with di-azides to redox- or light-responsive micelles 145 and 147–149
– Thiol–yne reaction to form polyelectrolytes for gene delivery
image file: c8cs00531a-u87.tif A67 – No modification 146
Morpholinones
image file: c8cs00531a-u88.tif – R = N-acyl morpholin-2-ones polymerize readily, but the N-aryl or N-alkyl substituted morpholin-2-ones do not polymerize. 144
image file: c8cs00531a-u89.tif A68 – After removal of BOC group water-soluble. 144


A less explored synthetic route to polyesters is the radical ring-opening polymerization (RROP) of cyclic ketene acetals (CKAs), which was summarized recently41 and is not further considered in this review. However, as radical polymerization is interesting also on the industrial scale, this strategy might be used also for the development of degradable functional polyesters. To the best of our knowledge, only methylidene-functionalized or chlorinated monomers are reported so far,41 resulting in halogenated polyesters or polymers with internal double bonds, but no postmodification was reported. In addition, less explored are poly(ester-ether)s and polythioesters. Five-, six- as well as seven-membered cyclic lactone ethers have been polymerized via ROP to poly(ester-ether)s. Reported substituents are mainly alkyl- or aryl chains or protected functions. To the best of our knowledge, orthogonally reactive monomers are not available so far but should be considered as a further development of lactone monomers. ε-Thiolactones and β-thiolactones can be polymerized by a base-catalyzed ring-opening polymerization to polythioesters. However, no functional reactive monomers have been reported to the best of our knowledge.42 The combination of polyaddition and ring-opening of different cyclic monomers offers a further strategy to novel functional materials with tunable properties: the reaction of lactone monomers with diamines has been reported.43 Polyesters can also be obtained by alternating ring-opening copolymerization (ROCOP) of epoxides and anhydrides. We exclude the technique of polyaddition of orthogonally reactive epoxides or anhydrides and point to several recent reviews.39,44,45

3.1 Lactones

3.1.1 ε-Caprolactones. Functional ε-caprolactones ε-(CL) can be divided into three subgroups, substituted in α-, β or γ-position, depending on the synthesis strategy for each monomer (overall yields: 19–70%). CLs are commonly polymerized with 2,2-dibutyl-2-stanna-1,3-dioxepane (DSDOP) as catalyst (in toluene at 20 °C for 24 h46 or at 60 °C for 2 h) or with tin(II) 2-ethyl hexanoate (SnOct2) in bulk or solution at 100–140 °C for 4–24 h (Scheme 1). For detailed polymerization conditions of lactones and lactides/glycolides and applied catalysts, we refer to separate literature.47
image file: c8cs00531a-s1.tif
Scheme 1 General polymerization scheme of caprolactones to polyesters (R1 and R2 represent non-specified alkyl or aryl substituents).

3.1.1.1 α-Substituted-ε-caprolactones. α-Halogenated caprolactones can be prepared by the Baeyer–Villiger oxidation of 2-halogenated-cyclohexanone with meta-chloroperoxybenzoic acid (mCPBA) (yield: 45–70%, Scheme 2A).46 Mohamod and coworkers48 polymerized α-fluoro caprolactone (A1) as homo- or copolymer with caprolactone. α-Chloro- (A2)46 and α-bromo-caprolactone (A3)49 was polymerized and substituted with azides, which allowed further postmodification with alkynes in a Huisgen 1,3-dipolar cycloaddition.49 These graft-polymers were used as macroinitiators for ATRP of methyl methacrylate (MMA)50 or hydroxyethyl methacrylate (HEMA)51 with the chlorides or bromides as initiating sites. Azide-functional poly(ε-CL) was also directly synthesized by an α-azido-ε-CL (A4),52 which was prepared by substitution of A2 or A3 with sodium azide.
image file: c8cs00531a-s2.tif
Scheme 2 Synthetic strategies for the synthesis of α-substituted ε-caprolactones: (A) by ring-expansion via the Baeyer–Villiger oxidation of cyclohexanones; (B) by substitution of ε-caprolactone.

A further general strategy is the functionalization of the α-position of ε-caprolactone by deprotonation with LDA (lithium diisopropylamide) and subsequent reaction with an electrophile (Scheme 2B). α-Iodo-caprolactone (A5) was obtained in this way by iodination with ICl (yield: 29%).53 The authors claimed the resulting copolymers to exhibit radio-opacity properties with potential application in temporary reconstructing material or drug delivery because of visualization via routine X-ray radioscopy.

α-Alkene and -alkyne functionalized caprolactones are used for the purpose of thiol–ene reaction and click reaction, introducing charged functionalities or bulky molecules, such as dyes or sugars. Following the described strategy, deprotonated ε-caprolactone reacts with allyl bromide, propargyl bromide or propargyl chloroformate to yield α-allyl-ε-caprolactone (A6, yield: 65%),54 α-propargyl-ε-caprolactone (A7)55 and α-propargyl carboxylate-ε-caprolactone (A8).56 After copolymerization of A6 with ε-CL, Coudane and coworkers54 attached Boc-protected-amines as the pendant chains by the radical thiol–ene reaction. They proved deprotection of the amine without degradation of the backbone and subsequent reaction with fluorescein isothiocyanate (FITC). They claimed the water-soluble cationic polyesters as interesting materials for gene delivery. Maynard and coworkers57 recently reported trehalose- and carboxybetaine-substituted poly(CL) and used it as a polymeric excipient for the stabilization of the therapeutic protein G-CSF for storage at 4 °C and at heat stressor temperatures of 60 °C. Copolymers of A7 were functionalized with the clinically used diethylenetriaminepentaacetic acid (DTPA)/Gd3+ complex, resulting in MRI-visible polymers as a hydrophobic contrast agent.58 PEG-block copolymers of α-propargyl carboxylate-ε-caprolactone formed micelles, which were core-cross-linked by a difunctional azide-cross-linker.56 An alternative strategy to produce α-propargyl-ε-caprolactone (A7, yield: 14%) starts with deprotonation of cyclohexanone and subsequent substitution with propargyl bromide, followed by the Baeyer–Villiger oxidation to expand the ring to ε-caprolactone (Scheme 2A).59 However, a mixture of α- and ε-substituted caprolactones (A7 and A7b, isomeric mixture yield: 30/70) were obtained. Ritter and coworkers prepared polymers and attached cyclodextrins via click reaction to form supramolecular organogels.55

Lecomte and coworkers60 reported an acrylate-substituted CL (A9) using it as end chain comonomer to form macrocyclic polyesters. The macrocycles were formed by UV-crosslinking of the acrylates. A9 was synthesized in three steps (Scheme 2B): deprotonation of caprolactone and addition of trimethylsilyl chloride formed a trimethylsilylketene acetal, which further reacted in a Mukaiyama aldol reaction with acetaldehyde. Esterification of the formed hydroxylactone with acryloyl chloride yielded the final monomer α-(1-acryloyloxyethyl)-ε-caprolactone (A9). Ritter and coworkers61 reported the ROP of α-methylidene-ε-caprolactone (A10), while the monomer has been polymerized at the vinyl functionality before. Homopolymerization yielded only low molar masses, copolymerization with caprolactone clearly higher molecular weight polymers. They claimed the polymers to be radically cross-linkable, however, did not report on further details. The monomer was synthesized by O-silylation, followed by thioalkylation with α-chloro thioanisole and completed by oxidative sulfur (Scheme 2B) removal (overall yield: 39%).62


3.1.1.2 β-Substituted-ε-caprolactones. Hillmyer and coworkers reported two β-substituted CLs derived from natural carvone: 7-methyl-4-(2-methyl oxirane-2-yl)oxepan-2-one (A11)63 and dihydrocarvide (A12)64 (Scheme 3). A11 was synthesized in two steps: hydrogenation of carvone resulted dihydrocarvone; epoxidation and ring-expansion yielded the final monomer (second step 25% yield). The monomer was homo- and copolymerized with CL, however, both rings reacted under the reported polymerization conditions (in bulk or solution, 20–120 °C, diethylzinc (ZnEt2) or tin(II) 2-ethyl hexanoate (SnOct2) catalyst). To obtain polymers with epoxides as pendant groups, monomer A12 was obtained by ring-expansion of dihydrocarvone using Oxone® (triple salt 2·KHSO5·KHSO4·K2SO4).64 Poly(dihydrocarvide-co-carvomenthide) was subsequently epoxidized and crosslinked with a dithiol to prove the possibility of post-modification.
image file: c8cs00531a-s3.tif
Scheme 3 Synthetic route to β-substituted CLs from carvone.

3.1.1.3 γ-Substituted-ε-caprolactones. γ-Halogenated caprolactones are synthesized starting from 4-halogenated-cyclohexanol (obtained by halogenation of 7-oxabicyclo[2.2.1]heptane); the alcohol was oxidized to a cyclic ketone to yield the monomer γ-chloro-ε-caprolactone (A13)65 or γ-bromo-ε-caprolactone (A14)66 after a Baeyer–Villiger oxidation (yield for the last step: 62%; Scheme 4A). Hegmann and coworkers65 used A13 in a copolymerization with caprolactone and lactide, substituted the chloride with an azide and attached cholesterol derivatives. The copolymers were used as cell scaffolds and foams. Jérôme and coworkers67 quaternized copolymers of caprolactone and A14 with pyridine, applied a debromination or epoxidation, and subsequent ring-opening to obtain hydrophilic poly(CL) substituted with diols in the backbone. P(A14) was substituted by an azide and amine-groups were then coupled by click chemistry to obtain pH-sensitive star-shaped polyesters.68 A γ-azide-ε-caprolactone has not been reported yet to the best of our knowledge. γ-Keto-ε-caprolactone (A15) was reported by the same group,69 obtained by ring extension of 1,4-cyclohexanedione (Scheme 4B). Qiao and coworkers70 functionalized the keto groups by hydrazine chemistry to introduce hydroxyl groups and coupled 4-nitrophenyl chloroformate. The activated ester was reacted with primary amines, e.g. of the cell adhesive peptide GRGDS. Lang and coworkers71 reduced the carbonyl group in copolymers to hydroxyl groups, using them as a macroinitiator for further grafting of lactide.
image file: c8cs00531a-s4.tif
Scheme 4 Synthetic strategies for the synthesis of γ-substituted ε-caprolactones: (A) for γ-halogenated CLs; (B) for γ-keto CL; (C) for γ-substituted CLs from cyclohexanols.

γ-Acryloyloxy ε-caprolactone (A16), a bifunctional monomer for ROP and radical polymerization,72 was prepared in two or three steps (Scheme 4C): 1,4-cyclohexanediol reacted with acryloyl chloride. The resulting monoalcohol was oxidized and the ring extended to yield the monomer (in an overall yield of 36%). A shorter alternative strategy started with the reaction of acryloyl chloride with 2-hydroxycyclohexan-1-one and subsequent ring extension (overall monomer yield 24%).73 Besides using the monomer for both ROP and ATRP, copolymers were grafted onto metal surfaces,74 used as 2D or 3D microstructured resins73 or were post-modified by Michael-addition of thiols.75 A similar monomer, γ-methacryloyloxy-ε-caprolactone (A17), was prepared by the same method, using methacryloyl chloride instead (overall monomer yield 29%),73 which were cross-linked upon UV-irradiation e.g. to form microparticles.76 γ-(2-Bromo-2-methyl propionate)-ε-caprolactone (A18) carries a classical ATRP initiating group and was presented by Hedrick and coworkers.77 A γ-cinnamate-modified caprolactone (A19) was recently reported by Budhlall and coworkers,78 which can undergo cis/trans isomerization and [2+2] cycloaddition upon UV-irradiation. Homo- or copolymers were used as thermoresponsive and semicrystalline networks after photochemical cross-linking.

3.1.2 δ-Valerolactones. Poly(δ-valerolactone)s have been much less studied compared to poly(ε-caprolactone)s and the number of orthogonally reactive δ-valerolactone monomers is very limited to a few examples of α-substituted δ-valerolactone. Examples for substitution in other positions are only available for alkylated substituents. δ-Valerolactones are polymerized under similar conditions as caprolactones, e.g. with and an alcohol as initiator and Sn(Oct)2 as a catalyst in bulk at 100 °C for 16 h,79 or with Sn(OTf)2 as a catalyst in bulk or THF at room temperature for 24 h.80

Emrick and coworkers80 reported the first functionally substituted monomer, α-allyl-δ-valerolactone (A20), synthesized by the same strategy as α-substituted ε-caprolactones: lithiation of δ-valerolactone in α-position with LDA and subsequent reaction with allyl bromide yielded A20 in one step (yield: 71%, Scheme 5). A20 was copolymerized with ε-caprolactone or δ-valerolactone, as well as homopolymerization obtained polymers in good conversion and narrow molecular weight distributions. The alkenes were quantitatively dihydroxylated with NMO/OsO4 to obtain more hydrophilic poly(ester)s. The group also introduced a α-cyclopentene-δ-valerolactone (A21):81A20 was lithiated and allylated to yield α,α-diallyl-δ-valerolactone. Ring-closing metathesis using a Grubbs catalyst gave A21. The cyclopentene substituted lactone was not able to homopolymerize; copolymerization with ε-caprolactone was realized with the incorporation of ca. 20% of A21. The pendant group was converted to cis-1,2-diols by dihydroxylation with OsO4 and showed longer bench-life stability compared to the diol-containing poly(ester)s from pendant allyl groups probably due to the higher rigidity of the monomer units. PEG was grafted onto the copolymers. Finally, Emrick and coworkers82 also used α-propargyl-δ-valerolactone (A22)83 (synthesized by the same strategy as A21) and obtained homo- as well as copolymers with ε-caprolactone. They functionalized the polymers by click chemistry with a PEG–azide, an oligopeptide-azide (GRGDS-N3), a phosphorylcholine derivative84 or a benzophenone group, to produce photopatternable aliphatic polyester.79 Harth and coworkers85 used A20 and A22 to form multifunctional polyester nanoparticles.


image file: c8cs00531a-s5.tif
Scheme 5 Synthetic strategy for the synthesis of substituted δ-valerolactones.

α-Methylidene-δ-valerolactone (A23) has usually been polymerized as “vinyl monomer”. Ritter and coworkers86 reported the first polymerization by ring-opening. Formylation of δ-valerolactone, subsequent formyl transfer and elimination of a carboxylate anion yielded A23 (yield: 57%, Scheme 5). The monomer was copolymerized with δ-valerolactone, and networks obtained by free radical polymerization of the methylidene functionality with different methacrylates.

Diaconescu and coworkers reported a series of three different α-ferrocenyl-δ-valerolactones (A24–A26) and six ferrocenyl-substituted trimethylene carbonate (TMC) monomers (C5–C10), all obtained by click chemistry of azide-functionalized ferrocene to A22, 5-(propynyl)-1,3-dioxane-2-one and propargyl 5-methyl-2-oxo-1,3-dioxane-5-carboxylate (C11) (see also below).87 While all TMC monomers were polymerizable with DBU/TU as the catalyst (1,8-diazabicycloundec-7-ene and 1-(3,5-bis(trifluoromethyl)phenyl)-3-cyclohexyl thiourea), A24–A26 were not able to be polymerized neither as homo- nor copolymers.

3.1.3 γ-Butyrolactones. γ-Butyrolactone is often considered to be the “non-polymerizable” lactone, due to its low ring strain.88 It can oligomerize using a lipase catalyst89 or under high pressure (20[thin space (1/6-em)]000 atm) and can be copolymerized with other lactone monomers.90 Chen and coworkers91 recently successfully obtained poly(γ-butyrolactone) via ROP with a La[N(SiMe3)2]3/R–OH catalyst system at −40 °C in THF with a number molecular weight of 30 kg mol−1, 90% monomer conversion and control over linear or cyclic topology.

Since polymerization of γ-butyrolactone remains difficult, only a few functional monomers have been reported so far. α-Methylidene-γ-butyrolactone (A27) is widely used as vinyl-comonomer; Ritter and coworkers92 reported copolymerization with caprolactone in a ROP for the first time, Chen and coworkers93 recently reported homopolymerization. They used the methylidene function for crosslinking of the polymers with a methacrylate to transparent polyester networks. The monomer was synthesized in two steps by the same strategy as A23 (Scheme 6): formylation of γ-butyrolactone, subsequent formyl transfer and elimination of a carboxylate anion yielded A27.94 Albertsson and coworkers95 recently reported the copolymerization of α-bromo-γ-butyrolactone (A28), which is commercially available at Sigma Aldrich. Due to the high selectivity and reactivity of modern organocatalysts at ambient reaction temperatures, the authors were able to polymerize co- and terpolymers, with trimethylene carbonate (TMC), C47 or ε-caprolactone with an alcohol as the initiator and diphenyl phosphate (DPP) as the catalyst at 30 °C for 48 h. Grafting of methyl acrylate via Cu(0)-mediated CRP (controlled radical polymerization) on the copolymers was demonstrated.


image file: c8cs00531a-s6.tif
Scheme 6 Synthetic strategy for the synthesis of α-methylidene-γ-butyrolactone.
3.1.4 β-Propiolactones and β-butyrolactones. Substituted β-propiolactones and β-butyrolactones are synthesized by three general synthetic strategies, following a “ketene”, “epoxide”, or “aspartic” route (Scheme 7).
image file: c8cs00531a-s7.tif
Scheme 7 Synthetic strategies to substituted of β-propio- and β-butyrolactones: (A) “ketene” route, e.g. for halogenated lactones; (B) “epoxide” route yielding β-heptenolactone; (C) “aspartic” route; (D) synthesis of α-methylene-β-butyrolactone; (E) synthesis of an α-disubstituted propiolactone.

Mono-, di-, and tri-halogenated propiolactones and their polymerization are reported extensively in the literature. Modification after polymerization has not been reported so far. Tani and coworkers96 synthesized β-chloromethyl- (A29), β-dichloromethyl- (A30) β-trichloromethyl-β-propiolactone (A31) by [2+2] cycloaddition from ketene and the corresponding mono-, di- or trichlorinated acetaldehyde and intensively investigated in their polymerization behavior and tacticity of obtained polymers (Schemes 7, 1A). Prud’Homme and coworkers further introduced several chlorinated and fluorinated propiolactones (A32–A36),97–99 partially being β-disubstituted at the lactone ring (Scheme 7, 1A–3A). For racemic mixtures of the monomers, they used the corresponding halogenated aldehyde (or halogenated acetone or butanone for β-di substituted lactones), acetyl chloride and triethylamine, for optically active monomers they used the synthetic route using ketene and the chiral catalyst quinidine. Li and coworkers100,101 copolymerized α-chloromethyl-α-methyl-propiolactone (A37) with caprolactone (Scheme 7, E). Quaternization with pyridine resulted in polymers with increased hydrophilicity. Chlorination of 2,2′-bis(hydroxymethyl)propionic acid with thionyl chloride, hydrolysis of the formed acyl chloride and cyclization under basic conditions yielded the monomer. A α,α-bis-chloromethyl-propiolactone monomer (A38) has been reported by Kuriyama and coworkers102 and copolymerized with β-propiolactone. Post-modification has not been shown.

Cherdron and coworkers103 presented several β-substituted lactones, carrying pendant groups suitable for other polymerization techniques (epoxide (A39), 3,4-dihydropyrane (A40), vinyl (A41)). They showed selective lactone polymerization, but also did not use the further functionality for post-modification reactions. They followed the general synthetic strategy using ketene and a corresponding aldehyde (Scheme 7, 1A).

The polymerization of β-heptenolactone (A42) (or also called allyl-β-butyrolactone) is only rarely reported in the literature, which is probably due to an inconvenient synthetic route of the monomer or use of special zinc or yttrium catalysts for polymerization. But-3-en-1-yl-epoxide reacted with carbon monoxide in the presence of a Co-based catalyst at 6.2 MPa/80 °C104 or an active Cr-catalyst at 1 atm/22 °C105 to the monomer (Scheme 7, B). While Guillaume and coworkers106 functionalized poly(β-heptenolactone) by hydroboration, Shaver and coworkers107 recently post-modified the polymer by olefin cross-metathesis with 15 different alkene cross-partners producing a whole library of poly(ester)s with different functionalities.

Guérin and coworkers108,109 developed three functional monomers for unsaturated poly(β-maleic acid) derivatives: allyl malolactonate (4-allyloxycarbonyl-2-oxetanone, A43), 3-methyl-3-butenyl malolactonate (4-[3-methyl-3-butenyloxycarbonyl]-2-oxetanone, A44) and 2-methylethenoyloxyethyl malolactonate (4-[2-methylethenoyloxyethyl-oxycarbonyl]-2-oxetanone, A45). While the ketene route gave only low yields, the “aspartic route” was applied (Scheme 7, C): aspartic acid was brominated and bromosuccinic acid anhydride formed. Esterification with an appropriate alcohol (allyl alcohol, 3-methyl-3-buten-1-ol or 2-hydroxyethyl methacrylate) opened the anhydride and formed a mixture of the corresponding mono-bromo succinic acid esters, and the major product was lactonizable. Epoxidation and subsequent sulfonation have been carried out. The copolymers were able to induce new bone formation and muscle regeneration in in vivo models.110,111 Thiol–ene reactions with mercaptoethanol converted the copolymers into macroinitiators to “graft from” polycaprolactone.112

Lu and coworkers113 recently reported a novel methylene functionalized monomer, α-methylidene-β-butyrolactone (A46), synthesized from carbon dioxide and 2-butyne in four steps (Scheme 7, D). After formation of tiglic acid, catalyzed by NiCl2*glyme and bathocuproine, an allylic peroxide was formed by photooxygenation. Dehydration formed a peroxylactone, which yielded A46 after deoxygenation. The vinylidene functional group was used for radical cross-linking or thiol–ene reaction.114

3.1.5 Macrolactones. Mainly two macrolactones, globalide (A47) and ambrettolide (Am, A48) are used to prepare long-chain aliphatic polyesters by ROP. A47 is a natural unsaturated 16-membered lactone, A48 a 17-membered lactone used in the fragrance industry. 14–19-membered lactones can be extracted from natural sources including angelica plant root. The ring-strain is the driving-force for ROP of smaller cycles and increases from 5- to 7-membered lactones, exhibiting the maximum for ε-caprolactone.115,116 Macrolactones have a low ring-strain and their ROP is entropy-driven instead of enthalpy-driven, as for the strained lactones.116 Macrolactones can be polymerized enzymatically by lipases, e.g. Novozyme 435 (Candida Antarctica lipase B (CALB) immobilized on acrylic resins).117 For further details we refer to excellent reviews of Kobayashi117,118 and the work of the Heise group.119–124

Heise and coworkers functionalized the olefins in polyglobalide via thiol–ene reaction with different thiols.120,122 In another study, dithiol-cross-linked polyglobalide films were further reacted with mercaptohexanol to attach ATRP initiators.119 Such films were further grafted with tert-butyl acrylate and proteins were conjugated to the deprotected grafts. Möller and coworkers125 polymerized A48 and oxidized the internal double bonds by Baeyer–Villiger oxidation using mCPBA to the epoxides. They showed, that a strategy vice versa is also possible: after epoxidation of A48, the monomer A49 (AmE) was polymerized with Novozyme 435, while the epoxides remained intact. Kobayashi and coworkers126 reported already in 2001 the enzymatic ROP of 2-methylene-4-oxa-12-dodecanolide (A50) by lipase and subsequent radical crosslinking of the polymers.

3.2 Cyclic diester monomers

Poly(α-hydroxy acid)s (PAHAs) are obtained from cyclic diester monomers (Scheme 8). Poly(lactic acid) (PLA), poly(glycolic acid) (PGA), and their copolymers (PLGA) are accessible from renewable resources. They are typically prepared by ROP of the cyclic dimers of lactic and glycolic acid (lactides and glycolides). However, the lack of structural diversity of lactides and glycolides limits the preparation of functional poly(α-hydroxy acid)s. Synthesis of substituted 1,4- dioxane-2,5-diones can be complicated and reactivity in ROP is often poor. They are usually polymerized with Sn(Oct)2 at 110–130 °C for 2–24 h in bulk127 or toluene,123 with 4-dimethylamino pyridine (DMAP) at 35 °C for 18–48 h in DCM,128,129 or with TBD or DBU at room temperature for 2 min to 24 h in DCM,122,130,131 using primary alcohols as initiator.
image file: c8cs00531a-s8.tif
Scheme 8 General scheme for the polymerization of lactides/glycolides and O-carboxy anhydrides (OCAs) to poly(α-hydroxy acid)s (PAHAs) (for definition of R-group, please see the main manuscript).

O-Carboxy anhydrides (OCAs) are suitable alternatives for the preparation of functionalized PAHAs under mild conditions and were recently summarized in an excellent article.132

3.2.1 Lactide and glycolide monomers. A general synthetic procedure to mono- or difunctional orthogonally reactive glycolide or lactide monomers was reported by Hennink and coworkers127 in three steps (Scheme 9A). Starting from an appropriate alkyl bromide (e.g. propargyl bromide), a Barbier-type addition to glyoxylic acid/ester (and cleavage of the ester, if used) resulted in a glycolic acid derivative which further reacted with 2-bromoacetyl bromide (or 2-bromopropanoyl bromide in case of a lactide monomer129,133–135). Intramolecular cyclization in diluted solution yielded the monomer (yields typically 15–45%). Difunctional monomers can be formed by dimerization and cyclization of the glycolic acid derivative.136
image file: c8cs00531a-s9.tif
Scheme 9 Synthetic strategies to functional lactides and glycolides: (A) by a Barbier-type addition; (B) by a Passerini-type reaction, (C) by functionalization of lactide.

Hennink and coworkers127 polymerized an allyl functional glycolide (A51) and showed epoxidation with NMO/OsO4 and subsequent hydrolysis to diols. The allyl lactide analog A52 has been reported by Cheng and coworkers.133 They photochemically crosslinked PEG–PLA block copolymers via thiol–ene reaction with a dithiol-crosslinker to obtain nanoparticles. Pfeifer and coworkers135 used cationic modified PEG–PLA-block copolymers for gene delivery. The monomer was also further functionalized by olefin cross-metathesis with an epoxy alkene and further hydrogenated to the saturated epoxy lactide (A53).134 An orthogonal reactive iron-based catalyst was applied for the polymerization of A53, which selectively polymerizes the diester cycle if the catalyst is in the iron(II) form. The oxidized catalyst (iron(III)-species) instead selectively polymerizes the epoxide. The bifunctional epoxy diester was selectively polymerized to an epoxy-functional polyester (Fig. 3). After oxidation of the catalyst and removal of solvent, the epoxy-functions were polymerized to cross-link the polymers.


image file: c8cs00531a-f3.tif
Fig. 3 Selective polymerization of a bifunctional monomer by the redox-controlled iron catalyst. Reproduced from ref. 134 with permission from The Royal Society of Chemistry, Copyright 2016.

Coudane and coworkers128 reported an alkyne functional glycolide (3-(2-propynyl)-1,4-dioxane-2,5-dione, A54), and modified PLGA-copolymers with PEG–azides. Cheng and coworkers129 reported the analogous alkyne lactide A55. They grafted PEG–paclitaxel–azide conjugates onto PLA-copolymers. A disubstituted alkynated glycolide (A56) has been used by Baker and coworkers136 for the polymerization of homopolymers and random or block copolymers, which were functionalized by click chemistry with PEG550–azide and azidododecane, to obtain thermoresponsive materials exhibiting lower critical solution temperatures (LCST) from room temperature to >60 °C. A facilitated synthesis of difunctional halogenide monomers was reported by Collard and coworkers to yielding 3,6-bis(chloromethyl)-1,4-dioxane-2,5-dione (A57):137 3-chloropropane-1,2-diol was oxidized to the glycolic acid derivative and subsequently dimerized and cyclized. Polymers were modified by dehydrochlorination to methylidene functions and further reacted with thiol derivatives by radical or nucleophilic thiol addition.

Yang and coworkers138 reported a further alkyne-functionalized lactide A58, synthesized by an alternative route: several commercially available aldehydes were reacted in a Passerini-type condensation to obtain the glycolic acid derivative (Scheme 9B). PLA-copolymers of A58 were modified with dansyl-azide as prove of concept. They additionally reported two azide-functionalized monomers (A59 and A60), synthesized by the same route. Copolymers were modified with dansyl alkyne. Overall yields for the synthetic strategy of A58–A60 were 6–16%. Weck and coworkers139 introduced an azido-tri(ethylene glycol) functional lactide (A61). Polymers were modified with a fluorescent dye (7-nitrobenzoxadiazole, NBD) and a cell internalization peptide gH625 by click chemistry, and proved cellular uptake. The group as well showed modification by Staudinger condensation with Tap-GRGDS.140

Modification of lactides without ring-opening is rare. Hillmyer and coworkers131 realized a bifunctional norbornene/lactide monomer (A62) suitable for ROP as well as ROMP by bromination and elimination of a lactide with an overall yield of 35% (Scheme 9C). The formed alkene reacted in a Diels–Alder reaction with cyclopentadiene and formed the bifunctional monomer. Dove and coworkers130 showed that the norbornenes in such copolymers were able to react with tetrazine derivatives. The norbornene-tetrazine reaction allowed post-modification under mild conditions at room temperature and without the addition of a catalyst or additives. The monomer was further functionalized with azide derivatives (A63), e.g. PEG-N3141 and polymerized. Two more lactides were realized by the same strategy using cyclohexa-1,3-diene (A64) and isoprene (A65) as diene for the Diels–Alder reaction, but the corresponding polymers were not further post-modified.142

3.2.2 O-Carboxyanhydrides (OCAs). In 1976, the first O-carboxy anhydride (OCA), 5-methyl-5-phenyl-1,3-dioxolan-2,4-dione, has been thermally polymerized.143 OCAs are readily available monomers from α-hydroxy acids (with yields up to 28–75%, depending on the number of synthesis steps) and are a suitable alternative to lactides and glycolides, yielding PAHAs. Preparation of OCA monomers follows a general synthetic strategy: α-hydroxy acids are carbonylated using phosgene, di- or triphosgene as a carbonylating agent. In case of the latter two agents, activated charcoal is often used to promote the decomposition to phosgene and a tertiary amine is added as an acid scavenger (Scheme 10). Thermodynamically the ROP of OCAs is favored compared to lactide, both enthalpically and entropically. During the ROP, CO2 is released by decarboxylation from every monomer unit and thus polymerization is more entropically driven than by release of ring strain.144 Bases and nucleophiles like pyridine, DMAP, NHCs or zinc complexes (with an external protic initiator) are able to promote the polymerization of OCAs in organic solvents as dichloromethane at room temperature in a few minutes to several hours, acid catalysts fail. Enzymatic polymerization showed higher polymerizability for OCAs than for lactides (polymer molar masses up to 104 g mol−1 within 24 h at 80 °C for OCAs, and 5–7 days at 80–130 °C for lactide).132
image file: c8cs00531a-s10.tif
Scheme 10 Synthetic route to O-carboxy anhydrides (OCAs) from amino acids.

The class of monomer has mainly been explored in the last decade and two orthogonally reactive OCAs have been reported so far: L-Tyr-alkynyl- (A66)145 and L-Tyr-allyl-OCA (A67).146 Cheng and coworkers used boc-protected L-Tyr-OH and reacted it with propargyl bromide to introduce the alkyne (or allyl bromide for the analogs alkene, Scheme 10).145 After the release of the amine group and formation of the α-hydroxy acid by diazotation with sodium nitrite, carbonylation, and cyclization yielded the final monomer. PEG-block copoly(ester) of A66 were core-crosslinked with a di-azide-cross-linker to redox-147 or light-responsive148 poly(ester) micelles; homopolymers were post-modified by thiol–yne reaction with cysteamine to polyelectrolytes for gene delivery and cell-penetration.149L-Tyr-allyl-OCA (A67) has not been used for post-modification so far.

3.2.3 Morpholinones. Waymouth and coworkers prepared N-substituted morpholin-2-ones by the oxidative lactonization of N-substituted diethanolamines with the Pd catalyst [LPd(OAc)]22+[OTf]2.150 The organocatalytic ring-opening polymerization of N-acyl morpholin-2-ones occurs readily to generate functionalized poly(amino esters) with N-acylated amines in the polyester backbone. The thermodynamics of the ring-opening polymerization depends sensitively on the hybridization of the nitrogen of the heterocyclic lactone. N-Acyl morpholin-2-ones polymerize readily to generate polymorpholinones, but the N-aryl or N-alkyl substituted morpholin-2-ones do not polymerize. Experimental and theoretical studies reveal that the thermodynamics of ring opening correlates to the degree of pyramidalization of the endocyclic N-atom. However, so far orthogonal monomers have not been prepared to the best of our knowledge. The deprotection of the poly(N-Boc-morpholin-2-one) (A68) produced a water-soluble, cationic polymorpholinone. We believe this area will continue to grow for future functional polymers.

4. Polyamides

Polyamides from cyclic monomers can be obtained from cyclic lactams (polyamide-3 to polyamide-6), from α-N-carboxy anhydrides (NCAs) (polypeptides or polyamide-2) and N-substituted glycine N-carboxy anhydrides (NNCAs) (polypeptoids), or from cyclic diamides and ester amides (Table 4). Besides synthetic polypeptides, most of them exhibit only poor biodegradability, but under harsh alkaline conditions or in the presence of special microorganisms can be degraded.
Table 4 Orthogonally functional cyclic monomers for the synthesis of polyamides and poly(ester amide)s
Monomer R = No. Post-modification Ref.
Lactams
image file: c8cs00531a-u90.tif image file: c8cs00531a-u91.tif B1 – No modification 155 and 156
image file: c8cs00531a-u92.tif image file: c8cs00531a-u93.tif B2 – Thermal or photochemical cross-linking 153
Glutamic acid-based NCAs
image file: c8cs00531a-u94.tif image file: c8cs00531a-u95.tif B3 – Click chemistry with PEG-, carbohydrate-, amine- or cyclodextrin-azides; photochemical thiol–yne reaction to introduce carboxy groups 165–169
image file: c8cs00531a-u96.tif B4 – Click chemistry to introduce alkyl chains of different lengths 170
image file: c8cs00531a-u97.tif B5 – Click reaction with amine/guanidines for gene delivery 171
image file: c8cs00531a-u98.tif B6 – Epoxidation and cross-linking; oxidation to carboxy functionalities; photochemical thiol–ene reaction to introduce carboxy groups 172 and 173
image file: c8cs00531a-u99.tif B7 – Epoxidation and cross-linking 172
– Oxidation to carboxy functionalities
image file: c8cs00531a-u100.tif B8 – Thiol–ene reaction with cysteamine to elongate the distance between charged groups and backbone 174
image file: c8cs00531a-u101.tif B9 – Radical cross-linking 175–178
– Ozonolysis to alcohols and aldehydes and hydroamination
– Oxidation to diols and carboxy groups
– Olefin metathesis
– Suzuki coupling
image file: c8cs00531a-u102.tif B10 – Formation of films by photo-cross-linking 179
image file: c8cs00531a-u103.tif B11 – Formation of films by photo-cross-linking 179
image file: c8cs00531a-u104.tif B12 – Formation of films by photo-cross-linking 179
image file: c8cs00531a-u105.tif B19 – Photo-cross-linking to stable micelles for drug delivery 180 and 181
image file: c8cs00531a-u106.tif B20 – No modification 182
image file: c8cs00531a-u107.tif B21 –ATRP macroinitiator 183 and 184
– Quaternization with diamines to form nanogels for drug delivery
image file: c8cs00531a-u108.tif B22 – Derivatization with NaN3 and click chemistry with carbohydrates, arginine or imidazolium 185–189
– Quaternization of phosphine and pyridinium salts
image file: c8cs00531a-u109.tif B23 – Derivatization with NaN3 and click chemistry with arginine 186, 188 and 190
– Quaternization of phosphine and pyridinium salts
image file: c8cs00531a-u110.tif B24 – Derivatization with NaN3 and click chemistry with arginine 186
image file: c8cs00531a-u111.tif B25 – Nucleophilic substitution with 1-alkylimidazolium salts to LCST- and UCST-type polypeptides 191
image file: c8cs00531a-u112.tif B26 – Amidation with amines for gene delivery 192 and 193
Tyrosine-based NCAs
image file: c8cs00531a-u113.tif image file: c8cs00531a-u114.tif B13 – Formation of films by photo-cross-linking 179
image file: c8cs00531a-u115.tif B14 – Formation of films by photo-cross-linking 179
image file: c8cs00531a-u116.tif B15 – Formation of films by photo-cross-linking 179
Lysine-based NCAs
image file: c8cs00531a-u117.tif image file: c8cs00531a-u118.tif B16 – Formation of films by photo-cross-linking 179 and 194
– Thiol–ene reaction for cross-linking
image file: c8cs00531a-u119.tif B17 – Formation of films by photo-cross-linking 179 and 195
– Thiol–ene reaction
image file: c8cs00531a-u120.tif B18 – Formation of films by photo-cross-linking 179
image file: c8cs00531a-u121.tif B28 –ATRP macroinitiator 196
image file: c8cs00531a-u122.tif B29 – Click chemistry 197
Ornithine-based NCAs
image file: c8cs00531a-u123.tif B30 – Click chemistry 197
Serine-based NCAs
image file: c8cs00531a-u124.tif image file: c8cs00531a-u125.tif B31 – Thiol–ene reaction with cysteamine to cell-penetrating peptides 198
image file: c8cs00531a-u126.tif B32 – Modification degrades the polymer 199
Homoserine-based NCAs
image file: c8cs00531a-u127.tif B33 – Amination to form poly(L-phosphorylchloline homoserine) 199
Cysteine-based NCAs
image file: c8cs00531a-u128.tif image file: c8cs00531a-u129.tif B34 – Michael-type addition of polar-, charged- or carbohydrate-thiols forming glycopeptides, coatings, and hydrogels 200
image file: c8cs00531a-u130.tif B35 – Nucleophilic substitution with imidazolium salts 201
image file: c8cs00531a-u131.tif B36 – Reaction with thiols to form asymmetric disulfides 202 and 203
image file: c8cs00531a-u132.tif B37 – Reaction with thiols to form asymmetric disulfides 202
Methionine-based NCAs
image file: c8cs00531a-u133.tif B38 – Alkylation with bromide, iodide and triflate derivatives and triggered dealkylation with sulfur nucleophiles 204–207
– Oxidation to sulfoxides causing a change of copolymer conformation
– Reaction with epoxides to β-alkyl-β-hydroxyethyl sulfonium products
DOPA-based NCAs
image file: c8cs00531a-u134.tif B39 – Oxidative cross-linking and tissue adhesion 194, 210 and 211
Unnatural amino acid-based NCAs
image file: c8cs00531a-u135.tif image file: c8cs00531a-u136.tif B40 – Reduction or bromination 212–214
– Glycosylation by thiol–ene reaction
image file: c8cs00531a-u137.tif B41 – No modification 215
image file: c8cs00531a-u138.tif B42 – Glycosylation by click chemistry 214 and 216–219
– Photochemical thiol–yne reaction
γ-NCAs
image file: c8cs00531a-u139.tif B27 – No modification 160
NNCAs
image file: c8cs00531a-u140.tif image file: c8cs00531a-u141.tif B43 – Thiol–ene reaction with thioglycerol and –glucose 224 and 225
image file: c8cs00531a-u142.tif B44 – Click chemistry with PEG–azide 226 and 227
– Thermal cross-linking
Cyclic esteramides
image file: c8cs00531a-u143.tif B45 – Thiol–ene reaction with charged or polar thiols 223


4.1 Lactams

β-Lactam (2-azetidinone), γ-butyrolactam (2-pyrrolidone), δ-valerolactam (2-piperidone) and ε-caprolactam (2-azepanone) are the main unsubstituted lactams used in the ROP. The number of orthogonally reactive lactam monomers is very limited so far. This might be attributed to the low solubility of the products in most common organic solvents, due to H-bonding or crystallization. ε-Caprolactam is typically polymerized in anionic ROP in bulk at temperatures of 140 °C above the melting point of the monomer (70 °C) within 15 min, with the polymer precipitating from the melt (Scheme 11).153N-Acyl or N-carbamoyl lactams as activators like hexamethylene-1,6-dicarbamoyl-caprolactam are commonly added to promote the polymerization. For more details, we refer to an excellent book chapter.154
image file: c8cs00531a-s11.tif
Scheme 11 General polymerization protocol for caprolactams to poly(amides)s with an N-acyl lactam as an activator (R1 and R2 represent non-specified alkyl substituents).

Vinyl lactam monomers are reported for all lactams, however, were only used for polymerization of the vinyl functionality. 4-Vinylazetidin-2-one (B1) is the only bifunctional monomer, whose anionic ROP was reported (in DMSO with potassium 2-pyrrolidone, at 25–30 °C, 2 h).155B1 was synthesized by reaction of 1,3-butadiene with chlorosulfonyl isocyanate and subsequent saponification (Scheme 12A).156 However, no post-modification has been reported so far.


image file: c8cs00531a-s12.tif
Scheme 12 Synthetic route to functional lactams: (A) 4-vinylazetidin-2-one B1 and (B) α-cinnamoylamido-ε-caprolactam B2.

While a few protected functional ε-caprolactam monomers (with an amine, carboxy, and carbonyl groups) are reported, Carlotti and coworkers153 synthesized a reactive monomer, bearing a cinnamoyl functionality. α-Cinnamoylamido-ε-caprolactam (B2) was synthesized in one step from cinnamoyl chloride and α-amino-ε-caprolactam (yield: 80%) (Scheme 12B). Copolymers with ε-caprolactam were cross-linked thermally (at 140 °C) or photochemically (at 364 nm) and again de-cross-linked photochemically (at 254 nm). α-Amino-ε-caprolactam might be an interesting precursor for future ε-caprolactams with other functional groups, e.g. to tailor the degradation rates. Polyamides can also be post-modified at the amide group by N-alkylation with formaldehyde,157 by epoxides or 2-bromoethylamine,158 isocyanates or acid chlorides.159

An interesting approach might be the synthesis of monomers from natural and renewable sources, e.g. macrolactones or carbohydrates. However, only a limited number of such examples is available up to date. Only a few functional OCA monomers for the synthesis of polyesters are reported, and expansion the monomer class is promising.

4.2 N-Carboxyanhydrides

Synthetic polypeptides are accessible by ROP of N-carboxy anhydrides (Leuchs’ anhydrides, NCAs), which are N-analogs to poly(α-hydroxy acid)s from O-carboxy anhydrides (OCAs). α-NCAs (5-membered rings) dominate the literature, while β-NCAs (6-membered rings) are only reported without additional functional groups to date. A single functional γ-NCA monomer (7-membered ring) has been reported in 1978,160 bearing a pendant carboxyl moiety (see below and Scheme 15).

A variety of protected, functionalized, and orthogonally reactive α-NCA monomers are reported. They are excellently summarized in two recent reviews.161,162 Primary amines are the common initiators163 for the ROP of NCAs, but also alcohols can be applied (as the initiation of the OH-group is slower than that of an amino-group, either the alcohol needs to be activated or broad molecular weight distributions will be obtained)164 (Schemes 13 and 14).


image file: c8cs00531a-s13.tif
Scheme 13 General protocol for the polymerization of NCAs to polypeptides with amines as the initiator (R1 and R2 represent non-specified substituents).

image file: c8cs00531a-s14.tif
Scheme 14 General synthetic routes to functional NCAs, depicted with glutamic acid.
4.2.1 Glutamic acid- and tyrosine-based NCAs. The majority of orthogonally functional NCA monomers are based on glutamic acid. Hammond and coworkers165 reported γ-propargyl-L-glutamate NCA (B3), with homopolymers being functionalized by click chemistry with PEG–azides of different chain lengths to form grafted brushes,165 different azide-functionalized carbohydrates,166 primary to quaternary amines to form antimicrobial polypeptides,167 or with cyclodextrin.168 Chen and coworkers169 introduced carboxy groups by the photochemical thiol–yne reaction. The polymers were used to control biomineralization of calcium carbonate.169 Tang and coworkers170 reported a second alkyne functionalized monomer, γ-(4-(propargoxycarbonyl)-benzyl)-L-glutamate NCA (B4), synthesized by the Cu(II)-complex method, using 4-(chloromethyl)benzoyl chloride (Scheme 14B). Obtained polypeptides were functionalized by click chemistry with alkyl chains of different lengths and the UCST-behaviour extensively analyzed in aqueous mixtures. Cheng and coworkers171 reported a similar monomer, γ-(4-propargyloxybenzyl)-L-glutamate NCA (B5), also obtained by the Cu(II)-complex method, using 4-propargyloxybenzyl chloride. The polypeptides were amine/guanidine functionalized by click chemistry and their transfection efficiency for gene delivery examined.

Several alkene-functionalized NCAs have been reported to date. Daly and coworkers172 prepared γ-allyl-L-glutamate (B6) and γ-(9-decenyl)-L-glutamate NCA (B7) and studied their homo- and copolymerization and epoxidation of the double bonds with mCPBA. Gelation was obtained by cross-linking of the epoxides with TFA, oxidation to carboxylic acids was shown with KMnO4/NaHCO3. Zhang and coworkers173 used B6 in copolymers for photochemical thiol–ene reaction and introduction of carboxy groups. γ-(4-Allyloxylbenzyl)-L-glutamate NCA (B8) has been reported by Cheng and coworkers,174 using the copper(II)-complex method and 4-allyloxylbenzyl chloride. Thiol–ene reaction in homopolymers of B8 with cysteamine to poly(γ-(4-aminoethylthiopropoxyl)benzyl-L-glutamate) exhibited a charge-backbone distance of 17 σ-bonds. Remarkably high helicity of 81% was observed for polypeptides with a DP of 10 at pH 2. Compared to the polymer with an elongated charge-backbone distance of 17 σ-bonds, poly(γ-(4-(1-hexanol-6-aminomethyl))benzyl-L-glutamate) with a DP of 10 and a charge-backbone distance of 11 σ-bonds showed mixed conformation of β-sheets and only 26% α-helices. γ-(4-Vinyl benzyl)-L-glutamate NCA was introduced by the group of Schouten.175 The olefins were used for radical cross-linking of terpolymer surface-grafted films,175 or transformation by ozonolysis into alcohols176 or aldehydes177 and subsequent hydroamination with primary amines, to yield polypeptides suitable for gene transfection.178 Oxidation to diols or carboxy functionalities by osmium tetroxide, olefin metathesis reaction with cis-1,4-dichlorobutene and Suzuki coupling were shown by Cheng and coworkers.176

Kamogawa and coworkers179 reported already in 1975 a whole library of photoreactive glutamic acid, tyrosine and lysine based NCAs with pendant acryloyl, methacryloyl and cinnamoyl groups (B10–B18). They showed polymerization of all NCAs and photochemical cross-linking of films, whereas the photosensitivity decreased for acryloyl ≈ methacryloyl > cinnamoyl, and polyglutamate > polylysine > polytyrosine. A further photo-cross-linkable monomer, also containing a cinnamoyl function, γ-cinnamoyl-L-glutamate NCA (B19), was already reported before by Iwakura and coworkers in 1974.180 Jing and coworkers181 photochemically crosslinked self-assembled polypeptide-block-PEG micelles of B19, loaded with paclitaxel as stable drug carriers. Iwakura and coworkers reported a further potentially photoactive benzophenone-containing NCA, γ-p-benzoyl benzyl-L-glutamate NCA and its homo- and copolymerization with γ-p-benzyl-L-glutamate NCA (B20).182 They studied the orientation of benzophenone groups in the side chains and conformation of polymers. The further photochemical reaction of the pendant groups has not been shown so far.

Several chlorinated monomers with varying length of alkyl spacers and suitable for nucleophilic substitution were reported: homo- and copolymers of a γ-(2-chloroethyl)-L-glutamate NCA (B21) were used as ATRP macroinitiator to graft oligoethylene methacrylate or to form nanogels for drug delivery.183 Cross-linking of the gels was achieved through quaternization reaction of 2,2′-dithiobis(N,N-dimethyl ethylamine) (dTbDEA) with the chloride functionalities.184 Polymers with the analogous γ-(3-chloropropyl)-L-glutamate (B22), γ-(6-chlorohexyl)-L-glutamate (B23), and γ-(8-chlorooctyl)-L-glutamate NCA (B24) were further derivatized to the respective azides and modified with carbohydrates,185 arginine,186 or imidazolium187 derivatives by click chemistry. These poly(arginine) mimics with hydrophobic side chains of different lengths exhibited helix-related cell-penetrating properties and high DNA and siRNA delivery efficiencies in various mammalian cells. Quaternization of the chloride substituted polypeptides with triethylphosphine yielded cell-penetrating peptides,188 quaternization with pyridinium salts UCST-type polypeptides,189 and with 1-methyl benzimidazole helical antimicrobial polypeptides.190 Tang and coworkers191 reported γ-(4-chloromethyl benzyl)-L-glutamate NCA (B25), synthesized by the copper(II) complex strategy. These chloride-substituted polypeptides were modified with 1-alkyl imidazolium (methyl or n-butyl) and various counter-anions (i.e. Cl, F, BF4). The polypeptides exhibited LCST- or UCST-type behavior in organic solvents or in water. Poly(peptide)s of an active ester functionalized monomer, γ-trichloroethyl-L-glutamate NCA (B26)192 were post-modified by amidation with different amine derivatives and their properties examined as gene delivery vectors.193 Higashi and coworkers reported already in 1978 a carboxylated γ-NCA (B27),160 after the reaction of the glutamic acid with diphosgene (Scheme 15). They proved successful polymerization by viscosity measurements and characteristic amide signals in IR spectra, which are also observed for Nylon-4. Modification of the carboxyl function has not been reported.


image file: c8cs00531a-s15.tif
Scheme 15 Synthetic strategy to a glutamic acid-based γ-NCA B27.
4.2.2 Lysine- and ornithine-based NCAs. Besides the photoreactive investigations of B16–B18 from Kamogawa179 (see above), Lei and coworkers194 recently used B16 for complex mussel-inspired thermoresponsive polypeptide-Pluronic copolymers as surgical adhesives and for hemostasis. Chen and coworkers195 prepared copolymers of B17 (Scheme 16A) and B3. Both groups applied thiol–ene reaction with high efficiency for post-modification.
image file: c8cs00531a-s16.tif
Scheme 16 Synthetic routes to lysine/ornithine-based NCAs.

Polymers of ε-N-bromoisobutyryl-L-lysine NCA (B28) were used as ATRP macroinitiators by Li and coworkers,196 and two polypeptide bottlebrushes with polystyrene or poly(oligoethylene glycol methacrylate) prepared, exhibiting an α-helical conformation in appropriate solvents.

Deming and coworkers197 reported azide-containing monomers: azido-norleucine (B29) and azido-norvaline NCAs (B30). α-N-Carboxybenzyl lysine or ornithine was reacted with imidazole-1-sulfonyl-azide·HCl, CuSO4, and K2CO3 to form the azide derivative (Scheme 16B). The derivatives were converted to NCAs, using the Ghosez's reagent (1-chloro-N,N′-2-trimethyl-1-propenylamine). Various alkyne derivatives were attached to azide-substituted pol(peptide)s through click chemistry with >95% conversion.

4.2.3 Serine- and homoserine-based NCAs. Cheng and coworkers198 synthesized O-pentenyl-L-serine NCA (B31) (Scheme 17A) and prepared block copolymers with PEG. Modification by thiol–ene reaction with cysteamine gave water-soluble copolymers with elongated and charged side chains exhibiting cell-penetrating properties. Deming and coworkers199 reported two phosphate containing monomers, O-2-bromo ethylbenzylphospho-L-serine NCA (B32) and O-2-bromo ethylbenzylphospho-L-homoserine NCA (B33) (Scheme 17B). While poly(L-phosphorylcholine serine) was not obtained after amination of the bromides in polymer P(B32) due to β-elimination and chain degradation, the use of poly(L-phosphorylcholine homoserine) overcame the side reaction. Both deprotection and amination were achieved in one step. B32 and B33 were synthesized from homoserine or serine and benzyl (2-bromo ethyl)diisopropylphosphoramidite. Cyclization was achieved by Ghosez's reagent.
image file: c8cs00531a-s17.tif
Scheme 17 Synthetic routes to serine/homoserine-based NCAs.
4.2.4 Cysteine-based NCAs. A vinyl sulfone-substituted L-cysteine NCA (B34) was reported by Zhong and coworkers,200 synthesized from divinyl sulfone and L-cysteine hydrochloride (Scheme 18A). Post-modification of vinyl sulfone-functionalized polypeptides with different thiols (polar, charged, carbohydrates) through nucleophilic addition produced glycopolypeptides, functional polypeptide coatings, or hydrogels. Tang and coworkers201 reported an S-(2-(3-chloropropoxycarbonyl)ethyl)-L-cysteine NCA (B35) from 3-chloropropyl acrylate. The polypeptides were substituted with imidazolium salts and their UCST-behavior was investigated.
image file: c8cs00531a-s18.tif
Scheme 18 Synthetic protocols for cysteine-based NCAs.

Barz and coworkers202 recently reported two interesting S-sulfonyl based NCA monomers: S-(ethylsulfonyl)-L-cysteine NCA (B36) and S-(isopropylsulfonyl)-L-cysteine NCA (B37) (Scheme 18B). Disulfides are reversible and therefore attractive for biomedical applications. They are stable under extracellular conditions, but cleavable inside of cells. Usually, disulfide formation has been achieved by oxidation of thiols (with long reaction times, often not 100% conversion and the lack of formation of asymmetric disulfides) or by the formation of reactive thiols (chlorinated or nitroso-thiols). A major drawback of activated thiols is the limited stability against aminolysis and hydrolysis. The group of Barz introduced two monomers with a protective and at the same time activating group, which was stable during ROP. Asymmetric disulfide formation was chemoselectively achieved by post-modification with appropriate thiols. An alkyl sulfonyl chloride was hydrolyzed to an alkyl sulfinic acid sodium salt and reacted with S-nitrosocysteine (generated in situ from L-cysteine) forming a thiosulfonate, which reacted with diphosgene to yield the final NCAs (B36 and B37) (Scheme 18B). Quantitative post-modification of polymers from B36 with benzylmercaptan was proven within 60 s, without degradation or side reactions. The approach opens a way to reversible conjugation of drugs as well as cross-linking to form nanostructures.203

4.2.5 Methionine-based NCAs. Deming and coworkers204 reported L-methionine NCA (B38). Beneficially, L-methionine NCA was synthesized in one step with high yield (91%)204 proceeding from L-methionine. L-Methionine is readily available and the amino acid with the highest production rate due to its use in animal feed. Methionine in homo- and polypeptides were chemoselectively and efficiently (>90%) alkylated with a broad range of bromide, iodide, triflate, and epoxide205 (pH < 3) derivatives yielding stable sulfonium or β-alkyl-β-hydroxyethyl sulfonium products. Polymethionine was compatible with the deprotection of other functional groups. Sulfur nucleophiles dealkylated the polypeptides again and triggered the release of therapeutics or tagged protein digests from affinity columns were achieved.206 Oxidation of sulfur in poly(L-methionine) to poly(L-methionine sulfoxide) or poly(L-methionine sulfone) influences the conformation of the polypeptides: poly(L-methionine) is hydrophobic forming an α-helix, poly(L-methionine sulfoxide) is water-soluble with a disordered conformation (Fig. 4); poly(L-methionine sulfone) exhibits a slightly water-soluble mainly α-helical structure. Deming and coworkers207 designed poly(L–methionine)-b-poly(L-leucine-stat-L-phenylalanine) copolymers and the oxidized sulfoxide copolymers self-assembled to stable vesicles. Upon reduction of the sulfoxides by methionine sulfoxide reductase A and B (MSR enzymes, found within human cells) and DTT as surrogate reductant, the hydrophilic disordered segments became hydrophobic and changed to rigid α-helical conformation. Change of conformation stiffened the vesicle membranes by forming a crumpled sheet-like morphology and eventually caused vesicle membrane rupture. Triggered release of an encapsulated model cargo (Texas Red labeled dextran) by enzymatic reduction of the sulfoxides segments has been demonstrated.
image file: c8cs00531a-f4.tif
Fig. 4 The enzyme-triggered release of cargos from methionine sulfoxide containing vesicles: (a) structure and redox properties of poly(L-methionine)-b-poly(L-leucine-stat-L-phenylalanine) peptides, (b) the possible effect of enzymatic reduction of vesicle surface of sulfoxide segments to methionine segments for a change of conformation and cause of vesicle ruptures. Adapted from ref. 207 with permission from the American Chemical Society, Copyright 2017.
4.2.6 DOPA-based NCAs. In recent years, catechol-containing polymers, inspired by mussel foot proteins (MFP) gained increasing interest for biomedical applications. Catechols complex metal-ions pH-dependently and reversible with high binding affinities, react with amines and thiols or covalently cross-link with each other under oxidizing conditions. O,O′-Dicarbobenzyloxy-208 and O,O′-acetyl-protected209L-DOPA NCAs were reported in the literature. The unprotected L-dihydroxyphenylalanine NCA (B39) was introduced from the group of Qiao210 by reaction of L-DOPA with triphosgene. They stated, that the phenolic hydroxyl groups did not initiate ROP of NCAs, however, used the NCA immediately after synthesis for polymerization, probably due to instability during storage. Block copolymers of L-glutamic acid and B39 showed reversible vesicle formation with ellipsoidal morphology at pH 3 and hollow vesicles at pH 12. Oxygen-mediated oxidation of DOPA groups at pH 12 stabilized the vesicles by cross-linking. Lei and coworkers194,211 recently applied the B39 in L-arginine- and L-DOPA-containing polypeptides as surgical adhesives.
4.2.7 Synthetic amino acid-based NCAs. In 1945, Schlögl et al.212 already introduced D,L-allylglycine NCA (B40) and its polymerization. Post-modification at that time was only conducted via hydrogenation, hydrobromination, and bromination. Schlaad and coworkers213,214 showed glycosylation by radical and photochemical thiol–ene reaction to enhance helical stability and solubility of poly(D,L-allylglycine) and poly(D,L-allylglycine-co-L-glutamate). A further unsaturated NCA, 5-pentenyl-D,L-glycine NCA (B41), was reported by Blanch and coworkers,215 postpolymerization modification has not been reported. In 1960, Schlögl et al.216 reported the synthesis of D,L-propylglycine NCA (B42) from D,L-propylglycine, and phosgene, and its polymerization. Post-modification was only reported decades later in 2010 by Heise and coworkers217 with glycosylation by click chemistry,218 and in 2012 by Schlaad and coworkers214 with glycosylation by the photochemical thiol–yne reaction. Bioactivity of the polymers, micelles or polymersomes was proven by selective lectin binding by the groups of Heise and Schlaad. Block copolypeptides with one glycosylated segment were found to be efficient stabilizers (completely based on renewable building blocks) in the emulsion polymerization of styrene.219

Alkene- or alkyne-substituted NCAs from unnatural amino acids compared to analog NCAs from natural amino acid display the distinct benefit, that ester or amide linkages are absent in the pendant chains. Pendant chains in polypeptides from naturally amino acid-based NCAs might be cleaved by hydrolysis or aminolysis at these linkages and thereby molecules introduced by post-modification detached. A challenging drawback of unnatural amino acid-based NCAs is the synthesis of the needed amino acids, which might be the reason for only a few reported monomers in literature.

4.3 Cyclic diamide monomers

To the best of our knowledge, no orthogonally functional cyclic diamides suitable for ROP have been reported so far. They are usually used for polycondensation reaction, where the ring remains in the backbone of the polymer. Elias and coworkers copolymerized the unfunctionalized monomer 2,5-dioxopiperazine (DOP) with caprolactam,220 Hildgen and coworkers with allyl glycidyl ether.221

4.4 Cyclic ester amide monomers

Morpholine-2,5-diones, resulting in polydepsipeptides, have been functionalized with different alkyl chains or protected groups. However, they are less reactive as lactides.222 To the best of our knowledge, Klok and coworkers223 reported the only orthogonally reactive ester amide monomer, L-allylglycine-morpholine-2,5-dione (B45). Further functional morpholine-2,5-diones employ protecting groups. A recent review summarizes the properties of poly(ester amide)s.43 The synthetic strategy is analog to the synthesis of functional glycolides and lactides: a corresponding α-amino acid and 2-bromo acetyl bromide reacted and the final monomer was realized by intramolecular cyclization of the intermediate in 13% yield (Scheme 19). Klok and coworkers showed modification by thiol–ene reaction with different fluorinated, carboxylated, aminated or dihydroxylated thiols with conversions between 15–100%.
image file: c8cs00531a-s19.tif
Scheme 19 Synthetic strategy to functional morpholino-2,5,-dione monomer (B45).

5. Polycarbonates

Aliphatic polycarbonates find broad application in biomedical devices and drug delivery. They are typically prepared by phosgene condensation or ester exchange, by addition polymerization of epoxides with carbon dioxide in ring-opening copolymerization (ROCOP)228 or by ROP of 6-membered cyclic carbonates, i.e. trimethylene carbonate (TMC). Poly(trimethylene carbonate) (PTMC) and poly(dimethyl trimethylene carbonate) (PDTC) display poor hydrophilicity and slow degradation rates. A broad range of functional TMCs, which we present here, has been explored to alter these properties and increase hydrophilicity and degradation. Polymerization of the five-membered cyclic carbonates is thermodynamically unfavorable.229 Polymerization above 150 °C results in poly(ether-carbonate)s and decarboxylation. Five-membered vinylene carbonate230 and vinylethylene carbonate231 are radically polymerized at the vinyl function, while the cyclic carbonate remains unaffected. Polymerization of seven-membered rings is possible but less relevant. Reported functional monomers are limited to alkyl- or aryl-pendant chains, or protected functionalities; reactive functionalities have not been reported up to date. For more details, we also refer to a review of Rokicki.229 The copolymerization of epoxides with anhydrides or carbon dioxide to polycarbonates is an attractive strategy. Functional epoxides and their synthetic protocols are well-established,232 and the potential of the combination with carbon dioxide228 or anhydrides45 are exploited (Table 5).
Table 5 Orthogonally functional cyclic TMCs for the synthesis of polycarbonates
Monomer R = No. Post-modification Ref.
5-Monosubstituted TMCs
image file: c8cs00531a-u144.tif image file: c8cs00531a-u145.tif C1 – Epoxidation with mCPBA, hydrazination and doxorubicin attachment for drug delivery 236 and 238–241
– Photochemical thiol–ene reaction with mercaptoethanol
image file: c8cs00531a-u146.tif C2 – No modification 237
image file: c8cs00531a-u147.tif C3 – No modification 237
image file: c8cs00531a-u148.tif C4 – Macroinitiator for ATRP polymerization of HEMA 242
image file: c8cs00531a-u149.tif C5 – Study of redox potential 87
image file: c8cs00531a-u150.tif C6 – Study of redox potential 87
image file: c8cs00531a-u151.tif C7 – Study of redox potential 87
Disubstituted TMCs
image file: c8cs00531a-u152.tif image file: c8cs00531a-u153.tif C11 – Click chemistry with carbohydrates, immobilization of TSP50 proteins or hemoglobin as oxygen carrier or reversible light-responsive micelles with spiropyran modification 243–248
– Thiol–yne reaction
– Functionalization with decaborane for BNCT
image file: c8cs00531a-u154.tif C13 – Epoxidation and hydrolyzation to diols or reaction with alcohols, amines or thiols 250–259
– Thiol–ene reaction to attach folic acid, nucleobases, for cross-linking and gelation
– Ozonolysis and reduction to aldehydes and aldehyde-aminooxy click reaction
image file: c8cs00531a-u155.tif C14 – Modification by Michael-addition with polar and charged groups 260–262
– Photo-cross-linking to obtain micelles and fibers
image file: c8cs00531a-u156.tif C15 – No modification 260
image file: c8cs00531a-u157.tif C16 – No modification 263
image file: c8cs00531a-u158.tif C17 – No modification 263
image file: c8cs00531a-u159.tif C18 – No modification 263
image file: c8cs00531a-u160.tif C19 – Photo-cross-linking 264
image file: c8cs00531a-u161.tif C20 – Michael-type reaction to attach polar or charged groups, GRGDC peptide or PEG-SH 265
image file: c8cs00531a-u162.tif C8 – Study of redox potential 87
image file: c8cs00531a-u163.tif C9 – Study of redox potential 87
image file: c8cs00531a-u164.tif C10 – Study of redox potential 87
image file: c8cs00531a-u165.tif C12 – PEG-copolymer nanoparticles as boron vector for BNCT 249
image file: c8cs00531a-u166.tif C21 – No modification 266
image file: c8cs00531a-u167.tif C28 – Nucleophilic substitution with bis-tertiary amines, pyridines or imidazoles as gene delivery vector and for antimicrobial applications 268, 269, 271, 272 and 274
image file: c8cs00531a-u168.tif C31 – Nucleophilic substitution with pyridines or imidazoles for antimicrobial applications 272
image file: c8cs00531a-u169.tif C22 – No modification 266
image file: c8cs00531a-u170.tif C27 – Nucleophilic substitution with bis-tertiary amines, pyridines or imidazoles as gene delivery vector and for antimicrobial applications 268, 270, 272 and 273
image file: c8cs00531a-u171.tif C29 – Nucleophilic substitution with pyridines or imidazoles for antimicrobial applications 272 and 273
image file: c8cs00531a-u172.tif C30 – Nucleophilic substitution with primary amines for antimicrobial applications 273
image file: c8cs00531a-u173.tif C32 – Nucleophilic substitution with primary amines for antimicrobial applications 274–277
image file: c8cs00531a-u174.tif C26 – Nucleophilic substitution with bis-tertiary amines as gene delivery vector 268
image file: c8cs00531a-u175.tif C25 – Click reaction with alkyne derivatives 267
– No polymerization is shown of azide monomer
image file: c8cs00531a-u176.tif C33 – Functionalization with different amines 277–281
image file: c8cs00531a-u177.tif C34 – Functionalization with amines to introduce charged groups or bioactive molecules 282
image file: c8cs00531a-u178.tif C35 – Functionalization with epoxide derivatives to introduce galactose, tocopherol or carbazols 283
image file: c8cs00531a-u179.tif C36 – Thiol–disulfide exchange reaction with thiols to reduction-sensitive self-assembled micelles or nanoparticles 284 and 285
image file: c8cs00531a-u180.tif C37 – Thiol–disulfide exchange reaction for formation of dynamic, stimuli-responsive and self-healing hydrogels 286 and 287
image file: c8cs00531a-u181.tif C38 – Thiol–disulfide exchange reaction for formation of core-cross-linked micelles or hydrogels 287
image file: c8cs00531a-u182.tif C39 – RAFT macroinitiator 288
image file: c8cs00531a-u183.tif C40 – RAFT macroinitiator 289
image file: c8cs00531a-u184.tif C41 – Diels–Alder reaction with maleimide-polymer derivatives to grafted copolymers 290
image file: c8cs00531a-u185.tif C42 – Cross-linking with maleimide-comonomers by Diels–Alder reaction to materials for nanoimprinting 291
image file: c8cs00531a-u186.tif C43 – Cross-linking with furfuryl-comonomers by Diels–Alder reaction to materials for nanoimprinting 291
image file: c8cs00531a-u187.tif C44 – Click reaction with azides 292
– Thiol–ene reaction
– Inverse electron demand Diels–Alder reaction with tetrazines
image file: c8cs00531a-u188.tif image file: c8cs00531a-u189.tif C45 – No modification 235
image file: c8cs00531a-u190.tif C46 – No modification 235
image file: c8cs00531a-u191.tif C47 – Thermal or photochemical cross-linking 293–295
– Thiol–ene reaction with PEG-SH for the formation of LCST-type polycarbonates
image file: c8cs00531a-u192.tif C49 – Cross-linking with styrene 297
image file: c8cs00531a-u193.tif C23 – No modification 266
image file: c8cs00531a-u194.tif C24 – No modification 266
image file: c8cs00531a-u195.tif C50 – Click reaction with alkynes 267
– No polymerization is shown of azide monomer
image file: c8cs00531a-u196.tif C51 – No modification 267
image file: c8cs00531a-u197.tif C48 – Functionalization with thiols or benzylamine 296
image file: c8cs00531a-u198.tif image file: c8cs00531a-u199.tif C52 – Quaternization with tertiary amines for antimicrobial applications 298–301
– Azidation and click reaction to form amphiphilic graft copolymers, micelles or nanoparticles for drug delivery
image file: c8cs00531a-u200.tif C53 – Click reaction with alkynes core-cross-linked micelles or hydrogels for drug delivery or cell tissue engineering 302–304
image file: c8cs00531a-u201.tif C54 – Self-cross-linking of micelles for drug delivery and intracellular de-cross-linking 305
image file: c8cs00531a-u202.tif C55 – Epoxidation with p-CPBA 306 and 307


5.1 Trimethylene carbonate

Various mono- and disubstituted trimethylene carbonate (TMC) monomers have been reported in the literature. Dove and coworkers published an excellent review about TMCs in 2013,25 including general synthetic strategies, functional monomers, and their polymerization, as well as post-modification reactions and applications. Here, we restrict to general synthesis strategies for orthogonally-reactive monomers and selected some representative examples for further applications.

While homopolymerization of the most functional monomers is possible, often copolymerization with alkyl-chain containing TMCs is conducted to adjust the functional group density. Furthermore, disubstituted monomers mostly contain a methyl or ethyl substituent, which has an impact on the thermal properties and solubility of the products. Alcohols are commonly used as initiators. A variety of synthetic protocols for the polymerization of TMCs are reported, including the polymerization with tin(II) octoate in bulk at ca. 110–120 °C for 4–48 h. Modern synthesis protocols rely on organocatalysis with 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) in bulk or solution at temperatures between 25 and 110 °C for 1–3 days, with a catalytic system of DBU or sparteine with a thiourea cocatalyst N-cyclohexyl-N′-(3,5-bis(trifluoromethyl)phenyl)thiourea (TU) at room temperature and in solution for 2–4 h or with the more reactive base 1,5,7-triazabicyclo[4.4.0]dec-5-ene (TBD) in bulk or solution at temperatures ranging from 0–25 °C over a period of 1 min to 2 h (Scheme 20). Commonly used solvents are dichloromethane, chloroform or toluene, depending on the solubility of TMCs and applied reaction temperatures.


image file: c8cs00531a-s20.tif
Scheme 20 General protocol for the organocatalytic polymerization of TMCs to poly(carbonates)s with alcohols as initiators.

A variety of synthetic strategies has been reported for the synthesis of substituted TMCs, including 2–5 steps with overall yields between 17–70%. A synthetic strategy relies on 2,2-bishydroxy(methyl)propionic acid (bis-MPA) as feedstock (Scheme 21), which was esterified to introduce the functional group, e.g. with alkylhalogenides (Scheme 21, route I). Following cyclization with ethyl chloroformate under basic conditions produced the TMCs. Hedrick and coworkers233 introduced an alternative route for a versatile precursor (5-methyl-5-carboxy-2-oxo-1,3-dioxane), yielding functional monomers in one further step (Scheme 21, route II). The carboxyl function of bis-MPA was protected with benzyl bromide, cyclization achieved, e.g. with triphosgene, and the carboxylic acid functionality deprotected again by hydrogenation (overall yield of precursor 59%). The precursor was either directly esterified again to introduce functionalities as pendant chains or transformed into a more reactive acid chloride and then esterified (overall yields: 15–41%). An alternative precursor, carrying a pentafluorophenyl ester protecting group,34,234 was synthesized on a gram to kilogram scale, easily handled and stored (yield: 75% Scheme 21, route III). It was reacted with suitable nucleophiles, as alcohols or amines in a transesterification reaction to functional TMCs. A further alternative is the use of imidazole intermediates using 1,1′-carbonyl diimidazole (CDI) as a key reagent (yield of intermediate: 67% Scheme 21, route IV). The intermediates were robust and bench stable.235 Alternatively, the hydroxyl functions of bis-MPA was first protected, e.g. with 2,2-dimethoxypropane (DMP) (Scheme 21, route V). Esterification of the carboxyl group introduced subsequently the pendant functionality. After deprotection of the hydroxyl groups, cyclization was achieved with ethyl chloroformate under basic conditions to yield the final monomers.


image file: c8cs00531a-s21.tif
Scheme 21 Synthetic strategies to trimethylene carbonates (TMCs) from 2,2-bishydroxy(methyl)propionic acid (bis-MPA).

Another general route uses glycerol or trimethylolalkanes (such as trimethylolpropane) as feedstock (Scheme 22), which are protected as acetals with benzaldehyde (Scheme 22, route VI), acetone, or 2,2-dimethoxypropane (DMP) (Scheme 22, route VII). The pendant hydroxyl group was substituted with alkylhalogenides or acid chlorides, the acetal hydrolyzed, and the final monomer formed by reaction with triphosgene or ethyl chloroformate under basic conditions.236 Alternatively, the starting compound was either directly transformed to the cyclic carbonate and afterward esterified to introduce the pendant functionality (Scheme 22, route VIII) or first, an oxetane-ring was generated (Scheme 22, routes IX–XI). After halogenation of the remaining hydroxyl group and ring-opening of the oxetane, cyclization was achieved, e.g. with ethyl chloroformate or 1,1′-carbonyldiimidazole (CDI) (overall yield: 31–47%, Scheme 22, route IX). Alternatively, the alkyl halogenide was reacted with sodium hydrogen sulfide and the resulting thiol modified either by thiol–ene reaction or disulfide formation (Scheme 22, routes X (yield: 16%) and XI (yield: 16%)). For more details, we also refer to the review of Dove.25


image file: c8cs00531a-s22.tif
Scheme 22 Synthetic strategies to trimethylene carbonates (TMCs) derived from trimethylolalkanes.

Yang and coworkers237 reported a strategy synthesizing monosubstituted TMCs from 2-aminopropane-1,3-diols (serinol) (Scheme 23, route XII; yields: 32–69%). Substitution of the amine with chloroformates, acyl halides or N-carbonyloxy succinimide derivatives introduced the pendant chain, followed by cyclization with ethyl chloroformate.


image file: c8cs00531a-s23.tif
Scheme 23 Synthetic strategy to trimethylene carbonates (TMCs) derived from serinol.
5.1.1 5-Monosubstituted trimethylene carbonates. Zhuo and coworkers236,238,239 reported an allyl-containing monomer, 5-(allyloxy)-1,3-dioxan-2-one (C1). Homo- or amphiphilic block copolymers with PEG were functionalized by photochemical thiol–ene reaction with cysteamine240 or by epoxidation, subsequent hydrazination and attachment of doxorubicin for drug delivery.241 Yang and coworkers237 reported allyl- (C2) and chloro- (C3) functional carbamate TMCs from serinol (Scheme 23, route XII), however, post-modification of polymers was not shown by the authors. Dai and coworkers242 reported recently a monomer with an ATRP initiating group, 2-(2-bromoisobutyrylamido) trimethylene carbonate (C4), using an analogous route with 2-bromoisobutyryl bromide. Hydroxyethyl methacrylate (HEMA) was grafted from copolymers as macroinitiator, and the amphiphilic polymers self-assembled in micelles.

Diaconescu and coworkers87 reported several ferrocene-containing TMCs (5-substituted (C5–C7) and 5-methyl-5′-substituted (C8–C10)), polymerized them and studied their redox potential for biological studies. Proceeding from 2-(prop-2-yn-1-yl)-propane-1,3-diol, cyclization was achieved with triphosgene. Ferrocene-azide derivatives were attached by click chemistry.

5.1.2 5-Methyl-5′-substituted trimethylene carbonates. A variety of alkyne-, alkene-, vinylidene-, and acrylate-functionalized 5-methyl-5′-substituted trimethylene carbonates (MTCs) has been reported with versatile applications so far. Jing and coworkers reported an alkyne monomer (C11), where lactide-copolymers were functionalized with sugars-azides243 and lectin-interactions has been proved, the protein TSP50 has been immobilized on polymer fibers,244 or amphiphilic triblock copolymers were self-assembled into core–shell micelles and conjugated with hemoglobin as artificial oxygen carriers.245 Reversibly light-responsive polycarbonate micelles functionalized with an azide-modified spiropyran were shown by Xu and coworkers.246 Dove and coworkers247 additionally showed quantitative radical thiol–yne functionalization. Carborane-conjugated amphiphilic PEG-block copolymers nanoparticles were used as drug delivery platform for doxorubicin and boron neutron capture therapy (BNCT) at the same time.248 The functionalization with decaborane was shown for the monomer (C12) and subsequent polymerization.249 An analogous allyl monomer (C13) was reported by Storey and coworkers250 and lactide-copolymers were epoxidized and hydrolyzed to diols, reacted with alcohols, polyethylenimine as gene delivery vector,251 or with diamines/dithiols to nanosponges for potentially controlled release.252 Thiol–ene reaction was applied to attach folic acid,253 or nucleobases254 (for core-crosslinking of micelles by base pairing)255 for drug delivery applications, to cross-link hydrogels,256 to microstereolithography resins as biocompatible 3D extracellular constructs,257 or to attach dopamine for self-healing gels via Fe3+-ion complexation.258 Wooley and coworkers polymerized 5-methyl-5-allyloxycarbonyl-1,3-dioxan-2-one (C13) to allyl-functionalized polycarbonates. The allyl bond was transformed into aldehyde-functional polycarbonates via ozonolysis and reductive work-up. Ozone efficiently cleaved the double bonds without damage to the polycarbonate backbone. The resulting aldehydes were functionalized with aminooxy compounds under mild conditions. Also, polycarbonates with both alkene and aldehyde groups were prepared by partial ozonolysis and functionalized by consecutive aldehyde–aminooxy and thiol–ene modifications in an orthogonal fashion.259

Zhong and coworkers reported (meth)acrylated MTCs C14 and C15. The acrylate was copolymerized with lactide or caprolactone and modified by Michael-addition with thiols of varying polar and charged groups,260 or folate-conjugated paclitaxel-loaded PEG–PC–PLA-triblock copolymer micelles photo-crosslinked for drug delivery.261 Amsden and coworkers262 electrospun lactide-copolymers of C14 and photo-crosslinked them yielding fibrous crimped scaffolds to culture cells. Hedrick and coworkers263 introduced additional (meth)acrylated and styrene functionalized monomers (C16–C18), synthesized from the pentafluorophenyl ester precursor (Scheme 22, route III). They showed polymerization of C16 and C18, but no further chemical modification. Jing and coworkers264 introduced a cinnamate-functionalized monomer (C19), which was photo-crosslinked after the ROP. A vinyl sulfone monomer (C20) was synthesized by Zhong and coworkers265 from 3-methyl-3-oxetanemethanol and divinyl sulfone (Scheme 22, route X). Copolymers and coatings were functionalized by selective Michael-type reaction with thiol-containing molecules to introduce polar and charged groups, GRGDC peptide, or thiolated poly(ethylene glycol) (PEG-SH).

Bowden and coworkers266 synthesized a series of 5-methyl or 5-ethyl-5′-halide-functional MTCs (halide = chloride or bromide, C21–C24) (Scheme 22, route IX), however, they did not perform post-functionalization. In another report, the bromide monomer C21 was transformed to an azide monomer (C25)267 and used as a precursor for modification by click reaction with alkyne derivatives and subsequent ROP. Hedrick and coworkers268 further reported 2-iodo-ethyl (C26), 3-chloro-propyl (C27) and 3-bromo-propyl (C28) MTC monomers, synthesized from the acid chloride MTC precursor (route II). Copolymers were functionalized with a bis-tertiary amine (TMEDA). While for the quaternization of the 3-chloropropyl substituted polymers from C27 90 °C was required, the 3-bromo-propyl (C28) and 2-iodo-ethyl (C26) substituted polymers were functionalized at room temperature in high conversion >90%. Chloride substituted polymers (C27) showed no gel formation because of the low reactivity of chlorine substituents. The iodide and bromide substituted polymers from C26 and C28 were more difficult to handle as undesired cross-linking occurred. Bromide-substituted copolymers from C28 with 50% 5-methyl-5-ethyloxycarboxyl-1,3-dioxan-2-one comonomer instead did not show cross-linking. The cationic polycarbonate was able to bind and complex DNA for generating nanoparticles and application as a gene delivery vector were studied.268 Coatings of surface grafted diblock copolymers of PEG and bromide substituted-polycarbonates from C28, quaternized with trimethylamine, exhibited antibacterial and antifouling properties and effectively killed Staphylococcus aureus (MSSA) and methicillin-resistant S. aureus (MRSA) on the coatings.269 The polymer coating prevented blood protein adhesion and no significant hemolysis was observed. Nanoparticles of cationic polycarbonates disrupted microbial walls/membranes selectively and efficiently and thus inhibited the growth of Gram-positive bacteria (MRSA and fungi).270 Galactose-functionalized cationic polycarbonates were applied for targeted gene delivery to hepatocytes.271 Yang and coworkers investigated in different pyridines and imidazoles as quaternizing agents272 for polycarbonates or the effect of hydrophobicity273 of the polymers from 3-chloropropyl (C27), -hexyl (C29) or -octyl (C30) and 3-bromo propyl (C28) or -hexyl (C31) MTCs, and their impact on antimicrobial properties.272 In other studies, cationic (co)polymers of a benzyl chloride-substituted MTC (5-methyl-5-(4-chloromethyl)benzyl carboxyl-1,3-dioxan-2-one, C32), again quaternized with different amines, were investigated with respect to their antimicrobial behavior,274 activity, and selectivity.275 Here, functionalization was faster compared to 3-bromo propyl substituted polymers from C28. Polymers were additionally functionalized by quaternization with phosphines, the chloride in polymers from C32 was substituted with NaN3 and subsequently clicked to various alkynes,276 or functionalized with boronic acid derivatives.277

Besides the use of a pentafluorophenyl ester MTC (C33) as a precursor for preparation of functionalized monomers (Scheme 22, route III),234C33 itself was polymerized to highly reactive PCs. Hedrick and coworkers proved post-modification with different amines278,279 to use them e.g. as stealth materials,280 with persistent radicals, or with boronic acid derivatives.277 They self-assembled them to nanoparticles or used them as MRI agents.281 Liu and coworkers282 reported a further active ester MTC monomer, functionalized with a NHS ester (5-methyl-5′-(succinimide-N-oxycarbonyl)-1,3-dioxan-2-one, C34), which was copolymerized with caprolactone and the copolymers modified by aminolysis with ethylene diamine to yield hydrophilic amido-amine pendant chains. They proposed facilitated attachment of bioactive molecules, targeting ligands, and covalent incorporation of prodrugs for the active ester copolymers.

A thioether-substituted monomer C35 was recently reported by Hedrick and coworkers,283 that was incorporated into homo-, di- and triblock copolymers, which were functionalized with epoxides, yielding sulfonium-functionalized PCs. Alkene functionalities, polar groups, galactose, tocopherol, and carbazols were introduced in the pendant chains. Zhong and coworkers reported a pyridyl disulfide-functionalized monomer (C36, Scheme 22, route XI), which was copolymerized with caprolactone. The polymers were functionalized in a thiol–disulfide exchange reaction with PEG-SH284 and self-assembled into reduction-sensitive micelles for active intracellular drug release or with thiolated lactobionic acid285 to reduction-sensitive shell-sheddable glyconanoparticles for efficient hepatoma-targeting delivery of doxorubicin. Two dithiolane-functionalized monomers C37 and C38 were recently synthesized by Waymouth and coworkers from 5-methyl-5′-carboxylic acid-trimethylene carbonate, oxalyl chloride, and 2-hydroxyethyl 4-carboxylate-4′-methyl-1,2-dithiolane286 (C37) or 2-hydroxyethyl-5-(1,2-dithiolan-3-yl)pentanoate (C38).287 Water-soluble triblock PEG–PC-copolymers were cross-linked with dithiols or with each other by reversible ring-opening of the pendant 1,2-dithiolanes and dynamic hydrogels obtained,286 or self-assembled into core-crosslinked (flower-bridged) micelles (Fig. 5).287 Depending on the pendant dithiolane chains, the hydrogels were dynamic, adaptable, and self-healing or rigid, resilient, and brittle. Cross-linked flower micelles dissociated upon the addition of acetone. Micelles cross-linked by a thiol and capped with maleimide persisted in acetone, and micelles cross-linked by a thiol, capped with maleimide, and then treated with dithiothreitol dissociated in acetone. Two trithiocarbonate functionalized monomers C39288 and C40289 were reported, serving as RAFT macroinitiator after ROP. N-Isopropylacrylamide (NiPAAm) (or methyl acrylate and tetrahydropyran acrylate)289 were grafted from the poly(carbonate)s.


image file: c8cs00531a-f5.tif
Fig. 5 Self-assembly of dithiolane-functionalized TMC–PEG–TMC block copolymers into flower micelles and thiol-initiated cross-linking of the micelles. Adapted from ref. 287 with permission from the American Chemical Society, Copyright 2017.

Tunca and coworkers290 reported an anthracene containing MTC monomer (C41), suitable for Diels–Alder reactions with dienophiles. Polymers were grafted with a furan-protected maleimide-terminated-poly(methyl methacrylate) or poly(ethylene glycol) or a mixture of both to yield well-defined polycarbonate graft or heterograft copolymers with an efficiency over 97%. Nelson and coworkers291 chose a similar approach, using a furfuryl-containing TMC monomer (C42). They provided copolymers with a second counterpart monomer: a Diels–Alder-protected maleimide-containing comonomer (C43). The maleimide was thermally deprotected and furan released at 130 °C. At 90 °C furfuryl functionalities in the copolymers reacted with the maleimide functions and covalently cross-linked the material. The polymer films were used for the thermally induced nanoimprinting process. Dove and coworkers292 introduced a monomer with a norbornene-functionality (C44). They presented different post-modification reactions on the homopolymers: (i) functionalization with azides via a 1,3-dipolar cycloaddition, (ii) inverse electron demand Diels–Alder reaction with tetrazines and (iii) radical thiol–ene coupling.

5.1.3 5-Ethyl-5′-substituted trimethylene carbonates. Malkoch and coworkers235 reported on alkyne- (C45) and alkene- (C46) functionalized 5-ethyl-5′-substituted TMCs, synthesized via imidazole intermediates (analog to Scheme 24, route IV with trimethylolpropane as starting material). They showed polymerization of the monomers, post-modification has not been reported so far. Höcker and coworkers293 studied the polymerization of 5-allyloxymethyl-5′-ethyl TMC (C47), cross-linked the polymers and investigated their thermal behavior. Brandell and coworkers294 used UV-cross-linked homopolymers as electrolytes in solid-state Li batteries. Dove and coworkers295 modified homopolymers of C47via photochemical thiol–ene reaction, e.g. with PEG-SH of different chain lengths, and showed a linear increase of LCST in correlation with increasing length of PEG chains. They further transformed the monomer to an oxirane ether carbonate (C48) by epoxidation with mCPBA.296 Homopolymers were functionalized in a thiol-epoxy reaction in presence of LiOH or DBU as a catalyst. While the degree of functionalization with aliphatic and benzylic thiols was rather low (<1–56%), thiophenol showed effective modification (50–94%) without polymer degradation in the presence of DBU as a catalyst. Functionalization with primary amines in the presence of Lewis acid catalysts was reported to be challenging because the Lewis acids promoted degradation of the polymers. Functionalization without a catalyst and increased temperature (50 °C) showed conversion >95% with benzylamine, but also the formation of some cross-linked material was observed. A styrene-functionalized TMC (C49) was introduced by Endo and coworkers,297 polymerized and radically cross-linked in the presence of styrene. The cross-linked polymer network was cleaved under basic conditions with potassium tert-butoxide and linear, soluble polystyrene with TMC functionalities was obtained.
image file: c8cs00531a-s24.tif
Scheme 24 The general protocol for the organocatalyzed ROP of cyclic phosphates to polyphosphoesters.

Bowden and coworkers266 reported two 5-ethyl-5′-halide-functional TMCs (halogenide = chloride (C23) or bromide (C24)). C24 was transformed to an azide functionalized monomer267 (C50) allowing further click chemistry. The authors showed selected monomers to be polymerizable, e.g. a α-methyl vinyl triazole monomer (C51). However, the post-polymerization modification has not been shown so far.

5.1.4 5,5′-Disubstituted trimethylene carbonates. Shen and coworkers298 reported a dibromo-substituted monomer (C52), obtained by transesterification between 2,2-bis(bromomethyl)propane-1,3-diol and ethyl chloroformate. Caprolactone–carbonate copolymers were functionalized by quaternization with tertiary amines and their antimicrobial properties were examined.299 Copolymers were quantitatively functionalized by azidation and click chemistry with PEG–alkyne of different chain length, to achieve amphiphilic graft copolymers with PEG as hydrophilic branches.298 From such materials, aggregation298,300 and paclitaxel-loaded redox-responsive core-crosslinked micelles for drug delivery were investigated.301 In contrast to Shen and coworkers, Zhuo and coworkers302 reported only poor conversions for the azidation of bromide copolymers. They introduced a diazido-substituted TMC (C53, 2,2-bis(azidomethyl)trimethylene carbonate) by transesterification of 2,2-bis(azidomethyl)propane-1,3-diol with ethyl chloroformate. Functionalization of polymers with different alkynes was achieved in quantitative yields. Amphiphilic PEG-b-P(C53-r-DTC) copolymers were cross-linked with disulfide-containing dialkyne-crosslinkers. The core-crosslinked redox-responsive micelles were investigated with respect to their drug release of methotrexate (MTX).303 Song and coworkers304 prepared hydrogels from PEG–PC–PEG triblock copolymer, containing azide substituents, and crosslinked them via copper-free ring-strain driven click reaction (SPAAC). Encapsulated bone marrow stromal cells exhibited high cell viability in the hydrogels and they exhibited higher cytocompatibility than hydrogels cross-linked by methacrylate functionalities.

Meng and coworkers305 introduced a dithiolane TMC (C54), synthesized from 2,2-bis(bromomethyl)-1,3-propanediol and NaSH and subsequent cyclization with ethyl chloroformate. PEG–PC copolymers self-assembled to micelles and crosslinked by disulfide exchange reaction, encapsulating doxorubicin (DOX). In vivo studies of the micelles in malignant B16 melanoma-bearing C57BL/6 mice with a dosage of 30 mg DOX eq. per kg effectively suppressed tumor growth, prolonged mice survival time and did not cause systemic toxicity. The drug was released within the tumor cells after reductive cleavage by glutathione.

Gross and coworkers306 reported 2,2-(2-pentene-1,5-diyl)trimethylene carbonate (C55), synthesized in one step from cyclohexene-4,4-dimethanol and ethyl chloroformate. They extensively studied the polymerization behavior and showed epoxidation of the pendant vinyl group with pCPBA (22–95% conversion, <2% hydrolysis of epoxides to diols). They suggested further hydrolysis to diols, reaction with alcohols or amines or initiation of ring-opening polymerization.307

5.1.5 Other polycarbonates. Venkataraman et al. prepared polycarbonates from N-substituted functional eight-membered cyclic carbonates.308 Organocatalytic ring-opening polymerization (ROP) produced well-defined polymers carrying different pendant groups at the in-chain tertiary amine. The copolymerization of the eight-membered monomers with 6-membered cyclic comonomers including commercially available L-lactide and trimethylene carbonate produced novel copolymers. Difunctional analogs produced antimicrobial hydrogels after quaternization of the amine groups using methyl iodide to confer antimicrobial properties. Stable hydrogels were obtained only when the bifunctional monomer concentration was equal to or higher than 12 mol%. In vitro and in vivo antimicrobial studies revealed that all quaternized hydrogels exhibited broad-spectrum antimicrobial activity against Gram-positive and Gram-negative bacteria and fungi, while the antimicrobial activity of the non-quaternized hydrogels was negligible. Moreover, the gels showed fast degradation at room temperature (4–6 days), which makes them ideal candidates for wound healing and implantable biomaterials.309,310

6. Polyphosphoesters

Polyphosphoesters (PPEs) have gained interest in the last two decades for biomedical applications due to their biocompatibility and -degradability. Depending on the number of methylene spacers in the backbone, PPEs with a broad range of hydrophobic to hydrophilic properties have been reported with various degradation behavior. The binding motif around the phosphorus atom (P–O, P–H, P–C or P–N) divides PPEs into further subclasses of polyphosphates, -phosphites, -phosphonates, or -phosphoramidates. The pentavalency of the phosphorus atom allows the facile introduction of alkyl-, aryl- or functional side chains and thus further tuning of the polymer properties. PPEs can be obtained by ROP from 5- and 6-membered cyclic phosphates, H-phosphonates, phosphonates or phosphoramidates (Table 6). Orthogonally functional cyclic phosphoramidates has not been reported so far. We refer to two recent reviews for detailed information on PPEs.311,312
Table 6 Orthogonally functional cyclic phosphorus-containing monomers for the synthesis of polyphosphoesters or polyphosphazenes
Monomer R = No. Post-modification Ref.
Phosphates
image file: c8cs00531a-u203.tif image file: c8cs00531a-u204.tif D1 – Click chemistry with PEG–azide to self-assembled block copolymers 315
image file: c8cs00531a-u205.tif D2 – Click chemistry or thiol–yne reaction to form charged micelles, silver- or paclitaxel-loaded nanoparticles, or block copolymers as a templating agent for calcium carbonate particles 316–325
image file: c8cs00531a-u206.tif D3 – Thiol–ene reaction to attach cysteamine for siRNA delivery or subsequent attachment of doxorubicin by hydrazine chemistry and dimethyl maleic anhydride for pH-responsivity 326–328
image file: c8cs00531a-u207.tif D4 – Thiol–ene reaction to conjugate paclitaxel with an acid-labile linker for drug delivery 314, 329 and 330
image file: c8cs00531a-u208.tif D5 – Thiol–ene reaction, acetalization or thio-acetalization 331
image file: c8cs00531a-u209.tif D6 – Michael-type addition of charged thiols 332
image file: c8cs00531a-u210.tif D7 – Copolymers as cross-linker for copolymerization with 2-methacryloyloxyethyl phosphorylcholine (MPC) to hydrogels 333–336
– Photo-cross-linking to hydrogels
image file: c8cs00531a-u211.tif D8 –ATRP-macroinitiator to graft poly(2-methacryloyloxyethyl phosphorylcholine) 337 and 338
image file: c8cs00531a-u212.tif image file: c8cs00531a-u213.tif D9 – No modification 339
image file: c8cs00531a-u214.tif D10 – No modification 339
image file: c8cs00531a-u215.tif D11 – No modification 339
H-Phosphonates
image file: c8cs00531a-u216.tif D12 – Oxidation with N2O4 to poly(phosphoric acid)s 340–347
– Reaction with amines to poly(phosphoramidate)s as a gene carrier
– Chlorination and reaction with alcohols to polyphosphates as a gene carrier
image file: c8cs00531a-u217.tif D13 – Oxidation with N2O4 to poly(phosphoric acid)s 348–353
– Chlorination and reaction with amines to polyphosphoramidates
– Sulfurization to polyphosphorothioates
Phosphonates
image file: c8cs00531a-u218.tif image file: c8cs00531a-u219.tif D14 – Thiol–ene reaction with cysteine or 3-mercaptopropionic acid to form thermoresponsive terpolymers 10 and 354
Phostones D15 – Potential for functional R groups (to date only ethyl and butyl have been reported) 357
image file: c8cs00531a-u220.tif
Phosphazenes
image file: c8cs00531a-u221.tif image file: c8cs00531a-u222.tif D16 – Nucleophilic substitution with alcohols and amines 16 and 358
image file: c8cs00531a-u223.tif D17 – Nucleophilic substitution with alcohols and amines 361 and 363
image file: c8cs00531a-u224.tif D18 – Nucleophilic substitution with alcohols and amines 323 and 362
image file: c8cs00531a-u225.tif D19 – Nucleophilic substitution with alcohols and amines 365 and 366
image file: c8cs00531a-u226.tif D20 – Nucleophilic substitution with alcohols and amines 365 and 366


6.1 Cyclic phosphates

PPEs can be prepared from 5-membered cyclic phosphates. Most monomers are prepared via 2-chloro-2-oxo-1,3,2-dioxaphospholane (COP) as the precursor. COP is either commercially available (but relatively expensive) or can be synthesized in two steps by the reaction of phosphorus trichloride with ethylene glycol, followed by oxidation (Scheme 26A). The oxidation procedure, however, was proven to be time-consuming and care has to be taken that the P–Cl-intermediate does not hydrolyze, e.g. if the oxygen gas is not dry. We recently optimized the reaction, using CoCl2 as catalyst and oxygen from the dried air instead of elemental oxygen, making the synthesis process more convenient and shorten the reaction time from days to hours.313 The final monomers are obtained by esterification of COP with an appropriate alcohol (yield for esterification step 30–70%). Their purification is sometimes challenging due to the high ring-strain: column chromatography over deactivated silica or distillation at reduced pressure are typical purification methods, but sometimes reduce the yields. If the starting COP is highly pure and the alcohol used in the final step can be removed in vacuo, the monomers are obtained pure enough without additional purification steps. Some orthogonally reactive monomers have been reported in literature up to date. Because of a growing interest in PPEs for biomedical applications, we expect the report of several further monomers in the common years.

Modern polymerization techniques for PPEs rely on the use of organocatalysts such as TBD, DBU, or an additional use of a thiourea cocatalyst (TU), but also classical basic, acidic or metal-based catalysts (Sn(Oct)2) may be applied.314 Primary alcohols are typically used as the initiator. Depending on the reactivity of catalyst, polymerizations can be conducted in dichloromethane at 0 °C for a period of 1 min to several hours. The reaction is terminated after ca. 80% conversion of monomers to avoid transesterification side reactions during a later stage of polymerization (Scheme 24).

A propargyl-functionalized phosphate monomer (2-(prop-2-yn-1-yloxy)-2-oxo-1,3,2-dioxaphospholane, D1) was reported by Wang and coworkers,315 block copolymers with caprolactone were synthesized and PEG-N3 clicked on the alkyne. The polymers self-assembled to nanoparticles with high biocompatibility. Contrary, Wooley and coworkers316 reported the unsuccessful synthesis of D1 with isolated yield <20% and suspected the reaction via a SN2′ mechanism to be responsible for decomposition of the monomer (nucleophilic attack at the acetylene proton and subsequent loss of the pendant chain at the phosphate group). They pointed out the absence or loss of the terminal acetylene proton in the 1H-NMR spectra provided by Wang and coworkers and suggested loss or partial loss of the alkynyl functionality. Instead, Wooley et al. installed an additional methylene spacer and reported the butynyl monomer 2-(but-3-yn-1-yloxy)-2-oxo-1,3,2-dioxaphospholane (D2) with high yield and purity.316 The monomer was polymerized (homopolymers or copolymers with caprolactone or cyclic phosphates) and post-modification by click chemistry or thiol–yne reaction was applied to prepare non-ionic, cationic, anionic, and zwitterionic micelles,317,318 covalently labeled core–shell polymeric nanoparticles with fluorescent contrast agents for theranostic applications,319 PEG-b-PPE-based paclitaxel conjugates for ultra-high paclitaxel-loaded multifunctional nanoparticles,320,321 or silver-bearing degradable polymeric nanoparticles showing in vitro antimicrobial activity (Fig. 6).322,323 For this purpose, silver cations were chelated into the corona or incorporated into the core using AgOAc, or SCC22 or SCC10. The placements of the silver species within the nanoparticle frameworks are proposed locations that have not been confirmed experimentally. Frank and coworkers used D2 to form anionic hydrophilic PPE-copolymers as efficient templating agent for calcium carbonate particles.324,325


image file: c8cs00531a-f6.tif
Fig. 6 Synthetic route to silver-bearing degradable polymeric nanoparticles: post-polymerization modification of alkyne-functionalized PPE-b-PLLA via thiol–yne “click” reaction to prepare anionic amphiphilic diblock copolymer, reproduced from ref. 322 with permission from the American Chemical Society, Copyright 2017.

Wang and coworkers326 reported a cyclic phosphate carrying an allyl pendant group (2-(prop-2-en-1-yloxy)-2-oxo-1,3,2-dioxaphospholane, D3), which was polymerized with a PEG-macroinitiator. The block copolymers self-assembled into nanoparticles and the allyl groups were post-modified by thiol–ene reaction with cysteamine and partially further reacted with dimethyl maleic anhydride to introduce pH-responsivity. Additionally, doxorubicine was conjugated to the polymers by hydrazine chemistry or the polymers were used for siRNA delivery.327 In contrast, Lecomte and coworkers328 used the pendant allyl chain of D3 as a protective group to produce polyphosphodiesters after selective deprotection with 1.5 eq. sodium benzene-thiolate. Quantitative deprotection without degradation was reported within 2 h in DMF/H2O (50[thin space (1/6-em)]:[thin space (1/6-em)]50, v/v) at room temperature. A butenyl monomer (2-butenoxy-2-oxo-1,3,2-dioxaphospholane, D4) was reported by Lecomte and coworkers314 and block copolymers with 2-isobutoxy-2-oxo-1,3,2-dioxaphospholane were prepared to demonstrate the control over copolymerizations of cyclic phosphates. Further post-modification has not been shown. Wooley and coworkers329 synthesized PEG-b-PPE of D4 and conjugated Paclitaxel with acid-labile linkages by thiol–ene reaction for drug delivery. Junkers and coworkers330 developed a flow microreactor to attach thiols upon UV-irradiation to butenyl-containing PPEs of D4. Wooley and coworkers331 also introduced a monomer with a reactive vinyl ether moiety (D5), which was either reacted in a thiol–ene reaction, thio-acetalization with thiols or acetalization with alcohols. While they reported quantitative conversion for thiol–ene reaction, acetalization, and thio- acetalization resulted in rather low conversions of 18% and 8%, respectively. Ni and coworkers332 reported a functional phosphate monomer carrying an acrylate group (D6, care has to be taken as the monomer undergoes quick radical polymerization during workup) and prepared block copolymers with caprolactone. The pendant acrylates were quantitatively functionalized by nucleophilic addition chemistry with thiols to introduce hydrophilic chains with hydroxyl, carboxyl, amine, and amino acid functionalities with low cytotoxicity. Stable micelles in aqueous solution, loaded with doxorubicin, opened the way to drug delivery carriers. While after 2 days at 37 °C, 30% release of cargo was observed in PBS-buffer (pH 7.4, 0.01 M), over 70% was reported in the presence of 0.2 mg mL−1 phosphodiesterase I. Iwasaki et al.333 used a methacryloyl phosphate monomer (D7) to obtain copolymers with 2-isopropyl-2-oxo-1,3,2-dioxaphospholane. They were used as cross-linker and copolymerized with 2-methacryloyloxyethyl phosphorylcholine (MPC) to prepare hydrogels. 100% degradation of the PPE-crosslinker polymer was observed within 6 days at pH 11, and 50% within 15 days at pH 7.4. 50% of the hydrogels degraded after 44 days at pH 7.4 (PBS-buffer).334 As highly porous hydrogels, they are suitable for 3D cell cultivation and increased proliferation of mouse osteoblastic cell (MC3T3-E1) was reported with increasing amount of PPEs in the hydrogels.335 The methacrylates as well were photo-cross-linked to obtain hydrogels.336

Iwasaki et al.337,338 copolymerized a cyclic bromoisobutylate phosphate (2-(2-oxo-1,3,2-dioxaphosphoroyloxyethyl-2-bromoisobutylate), D8) with 2-isopropyl-2-oxo-1,3,2-dioxaphospholane. The polymer was used as a macroinitiator for ATRP polymerization, to form amphiphilic PPEs with poly(2-methacryloyloxyethyl phosphorylcholine)-grafted chains.

6-Membered phosphate monomers are synthesized by a similar strategy, using 2-chloro-2-oxo-1,3,2-dioxaphosphorinane as the precursor, which is further substituted with an appropriate alcohol (Scheme 25B). Penczek and coworkers339 reported in 1977 three potentially orthogonally reactive monomers with 2,2,2-trichloroethyl (D9), 2,2,2-trifluoroethyl (D10) or 2-cyanoethyl (D11) functionalities as pendant chains (yields were not reported), but did not further modified the corresponding polymers, which were obtained by cationic polymerization in bulk. 6-Membered cyclic phosphates as monomers are less considered in the literature compared to 5-membered phosphates, due to a lower ring strain.


image file: c8cs00531a-s25.tif
Scheme 25 Synthetic strategies for cyclic phosphate, phosphonate, and H-phosphonate monomers.

6.2 Cyclic H-phosphonates

A cyclic 5-membered H-phosphonate monomer, 4-methyl-2-oxo-2-hydro-1,3,2-dioxaphospholane (D12), was polymerized by Vogl and coworkers,340 starting from 1,2-propandiol and PCl3. 4-Methyl-2-chloro-1,3,2-dioxaphospholane was oxidized by a mixture of water/dioxane to obtain the phosphite (yield has not been reported) (Scheme 25C). The monomer was also synthesized from optically active propandiol. The polymer p(D12) was converted by various strategies to (i) poly(phosphoric acid)s by oxidation with dinitrogen tetroxide, (ii) poly(phosphoramidate)s341,342 by an Atherton–Todd reaction with primary or secondary amine (30–60% conversion), and (iii) polyphosphates343 by chlorination and subsequent substitution with an appropriate alcohol (quantitative conversion), e.g. to poly(2-aminoethyl propylene phosphate) (PPE-EA). Leong and Mao used the polyphosphates343,344 and polyphosphoarmidates342,345 (with charged groups346) in several reports as gene carrier and transfection agent, or to form plasmid-templated DNA-block copolymer nanoparticles.347

A 6-membered H-phosphonate monomer, 2-hydro-2-oxo-1,3,2-dioxaphosphorinane (D13),348 was polymerized by Penczek and coworkers349 (Scheme 25D). It was prepared by transesterification from dimethyl phosphite and 1,3-propylene glycol (yield: 79%348). Penczek showed conversion to poly(phosphoric acid)s by oxidation with dinitrogen tetroxide, to polyphosphoramidates by chlorination with gaseous chlorine and subsequent reaction with primary amines (C- or N-substituted imidazoles, adenine or uracil),350–352 or to polyphosphorothioates353 by sufurization with sulfur in the presence of lutidine.

6.3 Cyclic phosphonates

A 5-membered cyclic allyl phosphonate monomer (2-allyl-2-oxo-1,3,2-dioxaphospholane, D14) was recently reported by Wurm and coworkers.10,354 For cyclic phosphonates the pendant group was introduced in the first step of a three-step reaction (Scheme 25E): after reaction of allyl bromide with triethyl phosphite, allyl phosphonic acid dichloride was formed by reaction with chlorotrimethylsilane and oxalyl chloride. Ring-closing reaction with ethylene glycol formed the final phosphonate monomer with an overall yield of 28% after distillation (as for cyclic phosphates, purification of the highly strained monomers can reduce the yield). Terpolymers were quantitatively modified with cysteine by a thiol–ene reaction and their thermoresponsive behavior analyzed.354 PEG-b-PPEs were modified with 3-mercaptopropionic acid. The block copolymers exhibited UCST behavior and were self-assembled into temperature-dependent dynamic aggregates.10 Polyphosphonates are prepared by using the organocatalyst DBU and primary alcohols as an initiator in dichloromethane at 0 °C for 16 hours with monomer conversions of 90–96%.354–356 Very recently also phostones have been polymerized to polyphosphonates with the P–C bond in the main-chain, which might be extended to functional derivatives.357 Their great benefit compared to the other cyclic phosphates or phosphonates is the relatively easy one-step synthesis, which might be expanded to allyl-functionalized phostones, for example (D15). It was proven that their degradation proceeds much slower than the corresponding side-chain polyphosphonates.

7. Polyorganophosphazenes

Polyorganophosphazenes are a polymer class of inorganic–organic hybrid polymers and exhibit a high diversity of properties due to the high versatility of possible organic substituents (Table 6). The phosphorus-nitrogen backbone is substituted by two organic side groups on the phosphorus atoms. Polymerization techniques and application of polyorganophosphazenes were recently summarized in excellent reviews.16,358,359

The most widely used method to prepare polyorganophosphazenes is the thermal ROP of hexachlorophosphazene (D16) with no control over the polymer molar mass and thus generally high molecular weights (>106 Da) and broad molecular weight distributions to the precursor polydichloro-phosphazene [NPCl2]. Living cationic polymerization of trichlorophosphoranimine (Cl3PNSiMe3) with two equivalents of phosphorus pentachloride by a chain growth mechanism in solution at room temperature results in higher control of polymer molar mass via the feed monomer to initiator ratio with lower dispersities. We refer the reader to a separate review.16

D15 is commercially available (at Sigma Aldrich) or can be synthesized from phosphorous pentachloride and ammonium chloride and subsequent sublimation from a mixture of oligomers (Scheme 26). The precursor rapidly hydrolyzes in the presence of water but can be completely post-modified under dry conditions with amines, alcohols, thiols or alkylation agents such as RMgX or RLi. Mixed, stepwise substitution with two or more organic substituents on the same macromolecule was possible and gives access to a library of polymers with varied properties.16 The complete modification is important to prevent uncontrolled cross-linking or degradation due to unreacted P–Cl groups. ROP was achieved with partially or completely substituted cyclophosphazene, but side reactions as ring expansion, decomposition or no reaction at all were observed. Generally, the probability for polymerization decreased with increasing number of organic substituents.360


image file: c8cs00531a-s26.tif
Scheme 26 Synthetic route to hexachlorophosphazene and polyorganophosphazenes.

Several further cyclic phosphazene monomers with a substitution of a phosphorus atom by carbon or sulfur were reported yielding cyclocarbophosphazene N3P2CCl5 (D17) from phosphorous pentachloride, ammonium chloride and cyanamide (10% yield)361 or cyclothiophosphazene N3P2SCl5 (D18) (yield 40–50%).362 Polychlorocarbophosphazenes363 and polychlorothiophosphazenes323 were obtained by thermal polymerization (120 °C, 4–6 h and 90 °C, 4 h, respectively), and were more reactive to nucleophilic substitution by aryloxides than classical polyphosphazene.361,362,364 Polycarbophosphazenes exhibited higher glass transition temperatures than polyphosphazene analog due to less torsional mobility in the backbone,361,364 polythiophosphazenes362 less hydrolytic stability. Polymerization (165–180 °C, 4 h) of cyclothionylphosphazenes N3P2SOCl5 (D19) and N3P2SOFCl4 (D20) yielded hydrolytically stable polythionylphosphazenes.365,366

8. Summary & outlook

The ring-opening polymerization gives access to well-defined polymers with narrow molecular weight distributions. However, as the reaction conditions for ROP typically do not tolerate nucleophiles and need inert conditions (in most cases), many functional groups in the monomers need to be protected. The deprotection step after the polymerization can be challenging, when also the polymer backbone can be degraded. In contrast, orthogonally reactive groups that do not interfere with the polymerization conditions are efficient alternatives to protective group chemistry. They can be post-modified by efficient reactions, which do not degrade the polymer backbone. Typical functional groups are alkenes, alkynes, and halogens, but several further functions have been summarized in this review.

Most functional cyclic monomers are synthesized in several reaction steps: the orthogonal functional groups are mostly introduced in the first step, making the synthesis of most functional monomers lengthy with variable yields. The use of “precursor monomers” would be beneficial, introducing the desired functionality in the last reaction step, but such strategies are missing. It is however possible for the preparation of functional phosphate monomers, that rely on the commercially available precursor 2-chloro-2-oxo-1,3,2-dioxaphospholane. The design of similar precursor strategies for other material classes would give faster access to multifunctional polymers with tailored properties (such as solubility or degradation profile).

The presented general concepts of post-modifications can be used for all summarized polymer classes, which are:

– nucleophilic substitution of halides,

– alkyne–azide click chemistry,

– thiol–ene and –yne reactions

– Diels–Alder reactions

They often yield quantitative conversion under mild conditions (e.g. excluding acidic or alkaline conditions or heavy-metal catalysts).

A broad variety of orthogonally reactive functionalities for cyclic monomers for the ROP and post-modification opportunities has been reported so far, which give access to diverse chemically functional biodegradable polymers and promising applications. There is still plenty of scope for further developments.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

G. B. is recipient of a fellowship through funding of the Excellence Initiative (DFG/GSC 266) in the context of the graduate school of excellence “MAINZ” (Material Sciences in Mainz). Funding of the Deutsche Forschungsgemeinschaft (DFG) to F. R. W. (WU750/6-1) is gratefully acknowledged. Open Access funding provided by the Max Planck Society.

References

  1. M. Okada, Prog. Polym. Sci., 2002, 27, 87–133 CrossRef.
  2. Y. Zhu, C. Romain and C. K. Williams, Nature, 2016, 540, 354 CrossRef PubMed.
  3. B. D. Ulery, L. S. Nair and C. T. Laurencin, J. Polym. Sci., Part B: Polym. Phys., 2011, 49, 832–864 CrossRef PubMed.
  4. L. S. Nair and C. T. Laurencin, Prog. Polym. Sci., 2007, 32, 762–798 CrossRef.
  5. B. Guo and P. X. Ma, Sci. China: Chem., 2014, 57, 490–500 CrossRef.
  6. A. Tardy, J. Nicolas, D. Gigmes, C. Lefay and Y. Guillaneuf, Chem. Rev., 2017, 117, 1319–1406 CrossRef PubMed.
  7. T. Haider, C. Völker, J. Kramm, K. Landfester and F. R. Wurm, Angew. Chem., Int. Ed., 2018 DOI:10.1002/anie.201805766.
  8. M. Rahman, in Polyester, ed. H. E.-D. M. Saleh, InTech, Rijeka, 2012, ch. 05,  DOI:10.5772/47765.
  9. J. Baran and S. Penczek, Macromolecules, 1995, 28, 5167–5176 CrossRef.
  10. T. Wolf, T. Rheinberger, J. Simon and F. R. Wurm, J. Am. Chem. Soc., 2017, 139, 11064–11072 CrossRef PubMed.
  11. H. Wang, L. Su, R. Li, S. Zhang, J. Fan, F. Zhang, T. P. Nguyen and K. L. Wooley, ACS Macro Lett., 2017, 6, 219–223 CrossRef.
  12. S. Zhang, H. Wang, Y. Shen, F. Zhang, K. Seetho, J. Zou, J.-S. A. Taylor, A. P. Dove and K. L. Wooley, Macromolecules, 2013, 46, 5141–5149 CrossRef PubMed.
  13. A. Cankaya, M. Steinmann, Y. Bulbul, I. Lieberwirth and F. R. Wurm, Polym. Chem., 2016, 7, 5004–5010 RSC.
  14. J. Baran, K. Kaluzynski, R. Szymanski and S. Penczek, Biomacromolecules, 2004, 5, 1841–1848 CrossRef PubMed.
  15. M. Steinmann, M. Wagner and F. R. Wurm, Chem. – Eur. J., 2016, 22, 17329–17338 CrossRef PubMed.
  16. S. Rothemund and I. Teasdale, Chem. Soc. Rev., 2016, 45, 5200–5215 RSC.
  17. I. Teasdale and O. Brüggemann, Polymers, 2013, 5, 161 CrossRef PubMed.
  18. J. Rydz, W. Sikorska, M. Kyulavska and D. Christova, Int. J. Mol. Sci., 2015, 16, 564 CrossRef PubMed.
  19. M. A. Elsawy, K.-H. Kim, J.-W. Park and A. Deep, Renewable Sustainable Energy Rev., 2017, 79, 1346–1352 CrossRef.
  20. K. Leja and G. Lewandowicz, Pol. J. Environ. Stud., 2010, 19, 255–266 Search PubMed.
  21. M. A. Woodruff and D. W. Hutmacher, Prog. Polym. Sci., 2010, 35, 1217–1256 CrossRef.
  22. A. Banerjee, K. Chatterjee and G. Madras, Mater. Sci. Technol., 2014, 30, 567–573 CrossRef.
  23. V. M. Pathak and Navneet, Bioresour. Bioprocess., 2017, 4, 15 CrossRef.
  24. H. Tian, Z. Tang, X. Zhuang, X. Chen and X. Jing, Prog. Polym. Sci., 2012, 37, 237–280 CrossRef.
  25. S. Tempelaar, L. Mespouille, O. Coulembier, P. Dubois and A. P. Dove, Chem. Soc. Rev., 2013, 42, 1312–1336 RSC.
  26. C. Jérôme and P. Lecomte, Adv. Drug Delivery Rev., 2008, 60, 1056–1076 CrossRef PubMed.
  27. X. Lou, C. Detrembleur and R. Jérôme, Macromol. Rapid Commun., 2003, 24, 161–172 CrossRef.
  28. C. K. Williams, Chem. Soc. Rev., 2007, 36, 1573–1580 RSC.
  29. S. M. Guillaume, Eur. Polym. J., 2013, 49, 768–779 CrossRef.
  30. S. M. Guillaume, Handbook of Telechelic Polyesters, Polyethers and Polycarbonates, Pan Stanford Publishing, 2017 Search PubMed.
  31. J. Dommerholt, F. P. J. T. Rutjes and F. L. van Delft, Top. Curr. Chem., 2016, 374, 16 CrossRef PubMed.
  32. C. K. Pandiyarajan, O. Prucker and J. Rühe, Macromolecules, 2016, 49, 8254–8264 CrossRef.
  33. O. Prucker, C. A. Naumann, J. Rühe, W. Knoll and C. W. Frank, J. Am. Chem. Soc., 1999, 121, 8766–8770 CrossRef.
  34. J. M. W. Chan, H. Sardon, A. C. Engler, J. M. García and J. L. Hedrick, ACS Macro Lett., 2013, 2, 860–864 CrossRef.
  35. P. Theato, J. Polym. Sci., Part A: Polym. Chem., 2008, 46, 6677–6687 CrossRef.
  36. M. A. Gauthier, M. I. Gibson and H.-A. Klok, Angew. Chem., Int. Ed., 2009, 48, 48–58 CrossRef PubMed.
  37. A. Das and P. Theato, Chem. Rev., 2016, 116, 1434–1495 CrossRef PubMed.
  38. F. Moldenhauer and P. Theato, in Multi-Component and Sequential Reactions in Polymer Synthesis, ed. P. Theato, 2015, vol. 269, pp. 133–162 Search PubMed.
  39. R. Tong, Ind. Eng. Chem. Res., 2017, 56, 4207–4219 CrossRef.
  40. H. Seyednejad, A. H. Ghassemi, C. F. van Nostrum, T. Vermonden and W. E. Hennink, J. Controlled Release, 2011, 152, 168–176 CrossRef PubMed.
  41. S. Agarwal, Polym. Chem., 2010, 1, 953–964 RSC.
  42. P. Espeel and F. E. Du Prez, Eur. Polym. J., 2015, 62, 247–272 CrossRef.
  43. M. Winnacker and B. Rieger, Polym. Chem., 2016, 7, 7039–7046 RSC.
  44. S. Paul, Y. Zhu, C. Romain, R. Brooks, P. K. Saini and C. K. Williams, Chem. Commun., 2015, 51, 6459–6479 RSC.
  45. J. M. Longo, M. J. Sanford and G. W. Coates, Chem. Rev., 2016, 116, 15167–15197 CrossRef PubMed.
  46. R. Riva, S. Schmeits, C. Jérôme, R. Jérôme and P. Lecomte, Macromolecules, 2007, 40, 796–803 CrossRef.
  47. A. Duda, in Polymer Science: A Comprehensive Reference, ed. M. Möller, Elsevier, Amsterdam, 2012, pp. 213–246,  DOI:10.1016/B978-0-444-53349-4.00104-7.
  48. T. F. Al-Azemi and A. A. Mohamod, Polymer, 2011, 52, 5431–5438 CrossRef.
  49. N. Xu, R. Wang, F.-S. Du and Z.-C. Li, J. Polym. Sci., Part A: Polym. Chem., 2009, 47, 3583–3594 CrossRef.
  50. S. Lenoir, R. Riva, X. Lou, C. Detrembleur, R. Jérôme and P. Lecomte, Macromolecules, 2004, 37, 4055–4061 CrossRef.
  51. B. Massoumi, Polym. Sci., Ser. B, 2015, 57, 85–92 CrossRef.
  52. R. Riva, S. Schmeits, F. Stoffelbach, C. Jerome, R. Jerome and P. Lecomte, Chem. Commun., 2005, 5334–5336,  10.1039/B510282K.
  53. S. E. Habnouni, V. Darcos and J. Coudane, Macromol. Rapid Commun., 2009, 30, 165–169 CrossRef PubMed.
  54. V. Darcos, S. Antoniacomi, C. Paniagua and J. Coudane, Polym. Chem., 2012, 3, 362–368 RSC.
  55. O. Jazkewitsch, A. Mondrzyk, R. Staffel and H. Ritter, Macromolecules, 2011, 44, 1365–1371 CrossRef.
  56. S. M. Garg, X.-B. Xiong, C. Lu and A. Lavasanifar, Macromolecules, 2011, 44, 2058–2066 CrossRef.
  57. E. M. Pelegri-O’Day, S. J. Paluck and H. D. Maynard, J. Am. Chem. Soc., 2017, 139, 1145–1154 CrossRef PubMed.
  58. S. El Habnouni, B. Nottelet, V. Darcos, B. Porsio, L. Lemaire, F. Franconi, X. Garric and J. Coudane, Biomacromolecules, 2013, 14, 3626–3634 CrossRef PubMed.
  59. O. Jazkewitsch and H. Ritter, Macromolecules, 2011, 44, 375–382 CrossRef.
  60. H. Li, R. Jéröme and P. Lecomte, Polymer, 2006, 47, 8406–8413 CrossRef.
  61. S. Sinnwell and H. Ritter, Macromolecules, 2006, 39, 2804–2807 CrossRef.
  62. K. Takeda, I. Imaoka and E. Yoshii, Tetrahedron, 1994, 50, 10839–10848 CrossRef.
  63. J. R. Lowe, W. B. Tolman and M. A. Hillmyer, Biomacromolecules, 2009, 10, 2003–2008 CrossRef PubMed.
  64. J. R. Lowe, M. T. Martello, W. B. Tolman and M. A. Hillmyer, Polym. Chem., 2011, 2, 702–708 RSC.
  65. Y. Gao, T. Mori, S. Manning, Y. Zhao, A. d. Nielsen, A. Neshat, A. Sharma, C. J. Mahnen, H. R. Everson, S. Crotty, R. J. Clements, C. Malcuit and E. Hegmann, ACS Macro Lett., 2016, 5, 4–9 CrossRef.
  66. C. Detrembleur, M. Mazza, O. Halleux, P. Lecomte, D. Mecerreyes, J. L. Hedrick and R. Jérôme, Macromolecules, 2000, 33, 14–18 CrossRef.
  67. C. Detrembleur, M. Mazza, X. Lou, O. Halleux, P. Lecomte, D. Mecerreyes, J. L. Hedrick and R. Jérôme, Macromolecules, 2000, 33, 7751–7760 CrossRef.
  68. R. Riva, W. Lazzari, L. Billiet, F. Du Prez, C. Jérôme and P. Lecomte, J. Polym. Sci., Part A: Polym. Chem., 2011, 49, 1552–1563 CrossRef.
  69. J.-P. Latere, P. Lecomte, P. Dubois and R. Jérôme, Macromolecules, 2002, 35, 7857–7859 CrossRef.
  70. E. L. Prime, Z. A. A. Hamid, J. J. Cooper-White and G. G. Qiao, Biomacromolecules, 2007, 8, 2416–2421 CrossRef PubMed.
  71. W. Dai, J. Zhu, A. Shangguan and M. Lang, Eur. Polym. J., 2009, 45, 1659–1667 CrossRef.
  72. D. Mecerreyes, J. Humes, R. D. Miller, J. L. Hedrick, C. Detrembleur, P. Lecomte, R. Jérôme and J. San Roman, Macromol. Rapid Commun., 2000, 21, 779–784 CrossRef.
  73. C. Vaida, P. Mela, H. Keul and M. Möller, J. Polym. Sci., Part A: Polym. Chem., 2008, 46, 6789–6800 CrossRef.
  74. X. Lou, C. Jérôme, C. Detrembleur and R. Jérôme, Langmuir, 2002, 18, 2785–2788 CrossRef.
  75. J. Rieger, K. Van Butsele, P. Lecomte, C. Detrembleur, R. Jerome and C. Jerome, Chem. Commun., 2005, 274–276,  10.1039/B411565A.
  76. C. Vaida, P. Mela, K. Kunna, K. Sternberg, H. Keul and M. Möller, Macromol. Biosci., 2010, 10, 925–933 CrossRef PubMed.
  77. D. Mecerreyes, B. Atthoff, K. A. Boduch, M. Trollsås and J. L. Hedrick, Macromolecules, 1999, 32, 5175–5182 CrossRef.
  78. A. Garle, S. Kong, U. Ojha and B. M. Budhlall, ACS Appl. Mater. Interfaces, 2012, 4, 645–657 CrossRef PubMed.
  79. D. Chen, C.-C. Chang, B. Cooper, A. Silvers, T. Emrick and R. C. Hayward, Biomacromolecules, 2015, 16, 3329–3335 CrossRef PubMed.
  80. B. Parrish, J. K. Quansah and T. Emrick, J. Polym. Sci., Part A: Polym. Chem., 2002, 40, 1983–1990 CrossRef.
  81. B. Parrish and T. Emrick, Macromolecules, 2004, 37, 5863–5865 CrossRef.
  82. B. Parrish, R. B. Breitenkamp and T. Emrick, J. Am. Chem. Soc., 2005, 127, 7404–7410 CrossRef PubMed.
  83. J. L. Herrmann and R. H. Schlessinger, J. Chem. Soc., Chem. Commun., 1973, 711–712,  10.1039/C39730000711.
  84. B. M. Cooper, D. Chan-Seng, D. Samanta, X. Zhang, S. Parelkar and T. Emrick, Chem. Commun., 2009, 815–817,  10.1039/B817600K.
  85. A. E. v. d. Ende, E. J. Kravitz and E. Harth, J. Am. Chem. Soc., 2008, 130, 8706–8713 CrossRef PubMed.
  86. O. Stöhr and H. Ritter, Macromol. Chem. Phys., 2014, 215, 426–430 CrossRef.
  87. B. M. Upton, R. M. Gipson, S. Duhovic, B. R. Lydon, N. M. Matsumoto, H. D. Maynard and P. L. Diaconescu, Inorg. Chem. Front., 2014, 1, 271–277 RSC.
  88. K. N. Houk, A. Jabbari, H. K. Hall and C. Alemán, J. Org. Chem., 2008, 73, 2674–2678 CrossRef PubMed.
  89. G. A. R. Nobes, R. J. Kazlauskas and R. H. Marchessault, Macromolecules, 1996, 29, 4829–4833 CrossRef.
  90. K. Tada, Y. Numata, T. Saegusa and J. Furukawa, Makromol. Chem., 1964, 77, 220–228 CrossRef.
  91. M. Hong and E. Y. X. Chen, Nat. Chem., 2016, 8, 42–49 CrossRef PubMed.
  92. J. Zhou, A. M. Schmidt and H. Ritter, Macromolecules, 2010, 43, 939–942 CrossRef.
  93. X. Tang, M. Hong, L. Falivene, L. Caporaso, L. Cavallo and E. Y. X. Chen, J. Am. Chem. Soc., 2016, 138, 14326–14337 CrossRef PubMed.
  94. A. W. Murray and R. G. Reid, Synthesis, 1985, 35–38 CrossRef.
  95. J. Undin, P. Olsén, J. Godfrey, K. Odelius and A.-C. Albertsson, Polymer, 2016, 87, 17–25 CrossRef.
  96. M. Iida, T. Araki, K. Teranishi and H. Tani, Macromolecules, 1977, 10, 275–284 CrossRef.
  97. C. Lavallée, A. Leborgne, N. Spassky and R. E. Prud'homme, J. Polym. Sci., Part A: Polym. Chem., 1987, 25, 1315–1328 CrossRef.
  98. C. Lavallee, G. Lemay, A. Leborgne, N. Spassky and R. E. Prud'homme, Macromolecules, 1984, 17, 2457–2462 CrossRef.
  99. R. Voyer and R. E. Prud'Homme, J. Polym. Sci., Part A: Polym. Chem., 1986, 24, 2773–2787 CrossRef.
  100. X.-Q. Liu, Z.-C. Li, F.-S. Du and F.-M. Li, Macromol. Rapid Commun., 1999, 20, 470–474 CrossRef.
  101. X.-Q. Liu, M.-X. Wang, Z.-C. Li and F.-M. Li, Macromol. Chem. Phys., 1999, 200, 468–473 CrossRef.
  102. Y. Yamashita, T. Tsuda, H. Ishida, A. Uchikawa and Y. Kuriyama, Makromol. Chem., 1968, 113, 139–146 CrossRef.
  103. V. H. Ohse and H. Cherdron, Makromol. Chem., 1966, 95, 283–295 CrossRef.
  104. J. T. Lee, P. J. Thomas and H. Alper, J. Org. Chem., 2001, 66, 5424–5426 CrossRef PubMed.
  105. J. W. Kramer, E. B. Lobkovsky and G. W. Coates, Org. Lett., 2006, 8, 3709–3712 CrossRef PubMed.
  106. C. Guillaume, N. Ajellal, J.-F. Carpentier and S. M. Guillaume, J. Polym. Sci., Part A: Polym. Chem., 2011, 49, 907–917 CrossRef.
  107. F. Sinclair, L. Chen, B. W. Greenland and M. P. Shaver, Macromolecules, 2016, 49, 6826–6834 CrossRef.
  108. M.-A. Leboucher-Durand, V. Langlois and P. Guérin, React. Funct. Polym., 1996, 31, 57–65 CrossRef.
  109. R. Bizzarri, F. Chiellini, R. Solaro, E. Chiellini, S. Cammas-Marion and P. Guérin, Macromolecules, 2002, 35, 1215–1223 CrossRef.
  110. V. Jeanbat-Mimaud, C. Barbaud, J.-P. Caruelle, D. Barritault, S. Cammas-Marion, V. Langlois and P. Guérin, C. R. Acad. Sci., Ser. IIc: Chim., 1999, 2, 393–401 CrossRef.
  111. J.-P. Caruelle, D. Barritault, V. Jeanbat-Mimaud, S. Cammas-Marion, V. Langlois, P. Guerin and C. Barbaud, J. Biomater. Sci., Polym. Ed., 2000, 11, 979–991 CrossRef.
  112. O. Coulembier, P. Degée, P. Gerbaux, P. Wantier, C. Barbaud, R. Flammang, P. Guérin and P. Dubois, Macromolecules, 2005, 38, 3141–3150 CrossRef.
  113. Y.-C. Xu, H. Zhou, X.-Y. Sun, W.-M. Ren and X.-B. Lu, Macromolecules, 2016, 49, 5782–5787 CrossRef.
  114. Y.-C. Xu, W.-M. Ren, H. Zhou, G.-G. Gu and X.-B. Lu, Macromolecules, 2017, 50, 3131–3142 CrossRef.
  115. W. Saiyasombat, R. Molloy, T. M. Nicholson, A. F. Johnson, I. M. Ward and S. Poshyachinda, Polymer, 1998, 39, 5581–5585 CrossRef.
  116. A. Duda and A. Kowalski, in Handbook of Ring-Opening Polymerization, Wiley-VCH Verlag GmbH & Co. KGaA, 2009, ch. 1, pp. 1–51,  DOI:10.1002/9783527628407.
  117. S. Kobayashi, Proc. Jpn. Acad., Ser. B, 2010, 86, 338–365 CrossRef PubMed.
  118. S. Kobayashi, Polym. Adv. Technol., 2015, 26, 677–686 CrossRef.
  119. Z. Ates, F. Audouin, A. Harrington, B. O'Connor and A. Heise, Macromol. Biosci., 2014, 14, 1600–1608 CrossRef PubMed.
  120. Z. Ates and A. Heise, Polym. Chem., 2014, 5, 2936–2941 RSC.
  121. M. Claudino, I. van der Meulen, S. Trey, M. Jonsson, A. Heise and M. Johansson, J. Polym. Sci., Part A: Polym. Chem., 2012, 50, 16–24 CrossRef.
  122. Z. Ates, P. D. Thornton and A. Heise, Polym. Chem., 2011, 2, 309–312 RSC.
  123. I. van der Meulen, Y. Li, R. Deumens, E. A. J. Joosten, C. E. Koning and A. Heise, Biomacromolecules, 2011, 12, 837–843 CrossRef PubMed.
  124. I. van der Meulen, M. de Geus, H. Antheunis, R. Deumens, E. A. J. Joosten, C. E. Koning and A. Heise, Biomacromolecules, 2008, 9, 3404–3410 CrossRef PubMed.
  125. C. Vaida, H. Keul and M. Möller, Green Chem., 2011, 13, 889–899 RSC.
  126. H. Uyama, S. Kobayashi, M. Morita, S. Habaue and Y. Okamoto, Macromolecules, 2001, 34, 6554–6556 CrossRef.
  127. M. Leemhuis, N. Akeroyd, J. A. W. Kruijtzer, C. F. van Nostrum and W. E. Hennink, Eur. Polym. J., 2008, 44, 308–317 CrossRef.
  128. F. Coumes, V. Darcos, D. Domurado, S. Li and J. Coudane, Polym. Chem., 2013, 4, 3705–3713 RSC.
  129. Y. Yu, J. Zou, L. Yu, W. Ji, Y. Li, W.-C. Law and C. Cheng, Macromolecules, 2011, 44, 4793–4800 CrossRef.
  130. I. A. Barker, D. J. Hall, C. F. Hansell, F. E. Du Prez, R. K. O'Reilly and A. P. Dove, Macromol. Rapid Commun., 2011, 32, 1362–1366 CrossRef PubMed.
  131. F. Jing and M. A. Hillmyer, J. Am. Chem. Soc., 2008, 130, 13826–13827 CrossRef PubMed.
  132. B. Martin Vaca and D. Bourissou, ACS Macro Lett., 2015, 4, 792–798 CrossRef.
  133. J. Zou, C. C. Hew, E. Themistou, Y. Li, C.-K. Chen, P. Alexandridis and C. Cheng, Adv. Mater., 2011, 23, 4274–4277 CrossRef PubMed.
  134. K. R. Delle Chiaie, L. M. Yablon, A. B. Biernesser, G. R. Michalowski, A. W. Sudyn and J. A. Byers, Polym. Chem., 2016, 7, 4675–4681 RSC.
  135. C. H. Jones, C.-K. Chen, M. Chen, A. Ravikrishnan, H. Zhang, A. Gollakota, T. Chung, C. Cheng and B. A. Pfeifer, Mol. Pharmaceutics, 2015, 12, 846–856 CrossRef PubMed.
  136. X. Jiang, E. B. Vogel, M. R. Smith and G. L. Baker, Macromolecules, 2008, 41, 1937–1944 CrossRef.
  137. P. P. Kalelkar, G. R. Alas and D. M. Collard, Macromolecules, 2016, 49, 2609–2617 CrossRef.
  138. M. Rubinshtein, C. R. James, J. L. Young, Y. J. Ma, Y. Kobayashi, N. C. Gianneschi and J. Yang, Org. Lett., 2010, 12, 3560–3563 CrossRef PubMed.
  139. D. E. Borchmann, R. Tarallo, S. Avendano, A. Falanga, T. P. Carberry, S. Galdiero and M. Weck, Macromolecules, 2015, 48, 942–949 CrossRef.
  140. D. E. Borchmann, N. ten Brummelhuis and M. Weck, Macromolecules, 2013, 46, 4426–4431 CrossRef PubMed.
  141. J. A. Castillo, D. E. Borchmann, A. Y. Cheng, Y. Wang, C. Hu, A. J. García and M. Weck, Macromolecules, 2012, 45, 62–69 CrossRef PubMed.
  142. G. L. Fiore, F. Jing, J. V. G. Young, C. J. Cramer and M. A. Hillmyer, Polym. Chem., 2010, 1, 870–877 RSC.
  143. I. J. Smith and B. J. Tighe, J. Polym. Sci., Polym. Chem. Ed., 1976, 14, 949–960 CrossRef.
  144. R. J. Pounder and A. P. Dove, Polym. Chem., 2010, 1, 260–271 RSC.
  145. Z. Zhang, L. Yin, C. Tu, Z. Song, Y. Zhang, Y. Xu, R. Tong, Q. Zhou, J. Ren and J. Cheng, ACS Macro Lett., 2013, 2, 40–44 CrossRef PubMed.
  146. R. Wang, J. Zhang, Q. Yin, Y. Xu, J. Cheng and R. Tong, Angew. Chem., Int. Ed., 2016, 55, 13010–13014 CrossRef PubMed.
  147. H. Wang, L. Tang, C. Tu, Z. Song, Q. Yin, L. Yin, Z. Zhang and J. Cheng, Biomacromolecules, 2013, 14, 3706–3712 CrossRef PubMed.
  148. Y. Li, Y. Niu, D. Hu, Y. Song, J. He, X. Liu, X. Xia, Y. Lu and W. Xu, Macromol. Chem. Phys., 2015, 216, 77–84 CrossRef.
  149. Z. Zhang, L. Yin, Y. Xu, R. Tong, Y. Lu, J. Ren and J. Cheng, Biomacromolecules, 2012, 13, 3456–3462 CrossRef PubMed.
  150. T. R. Blake and R. M. Waymouth, J. Am. Chem. Soc., 2014, 136, 9252–9255 CrossRef PubMed.
  151. J. Li, Y. Zhang, J. Chen, C. Wang and M. Lang, J. Macromol. Sci., Part A: Pure Appl. Chem., 2009, 46, 1103–1113 CrossRef.
  152. J. B. Gilroy, T. Gädt, G. R. Whittell, L. Chabanne, J. M. Mitchels, R. M. Richardson, M. A. Winnik and I. Manners, Nat. Chem., 2010, 2, 566 CrossRef PubMed.
  153. D. Tunc, C. Le Coz, M. Alexandre, P. Desbois, P. Lecomte and S. Carlotti, Macromolecules, 2014, 47, 8247–8254 CrossRef.
  154. S. Russo and E. Casazza, in Polymer Science: A Comprehensive Reference, ed. M. Möller, Elsevier, Amsterdam, 2012, pp. 331–396,  DOI:10.1016/B978-0-444-53349-4.00109-6.
  155. V. E. Schmidt, Angew. Makromol. Chem., 1970, 14, 185–202 CrossRef.
  156. R. Graf, Angew. Chem., 1968, 80, 179–189 CrossRef.
  157. J. Lin, C. Winkelman, S. D. Worley, R. M. Broughton and J. F. Williams, J. Appl. Polym. Sci., 2001, 81, 943–947 CrossRef.
  158. X. Jia, M. Herrera-Alonso and T. J. McCarthy, Polymer, 2006, 47, 4916–4924 CrossRef.
  159. E. Perry and J. Savory, J. Appl. Polym. Sci., 1967, 11, 2473–2483 CrossRef.
  160. N. Honda, T. Kawai and F. Higashi, Makromol. Chem., 1978, 179, 1643–1646 CrossRef.
  161. D. Huesmann, K. Klinker and M. Barz, Polym. Chem., 2017, 8, 957–971 RSC.
  162. T. J. Deming, Chem. Rev., 2016, 116, 786–808 CrossRef PubMed.
  163. N. Hadjichristidis, H. Iatrou, M. Pitsikalis and G. Sakellariou, Chem. Rev., 2009, 109, 5528–5578 CrossRef PubMed.
  164. Š. Gradišar, E. Žagar and D. Pahovnik, ACS Macro Lett., 2017, 6, 637–640 CrossRef.
  165. A. C. Engler, H.-i. Lee and P. T. Hammond, Angew. Chem., Int. Ed., 2009, 48, 9334–9338 CrossRef PubMed.
  166. C. Xiao, C. Zhao, P. He, Z. Tang, X. Chen and X. Jing, Macromol. Rapid Commun., 2010, 31, 991–997 CrossRef PubMed.
  167. A. C. Engler, A. Shukla, S. Puranam, H. G. Buss, N. Jreige and P. T. Hammond, Biomacromolecules, 2011, 12, 1666–1674 CrossRef PubMed.
  168. Y.-C. Lin, P.-I. Wang and S.-W. Kuo, Soft Matter, 2012, 8, 9676–9684 RSC.
  169. Y. Huang, Y. Zeng, J. Yang, Z. Zeng, F. Zhu and X. Chen, Chem. Commun., 2011, 47, 7509–7511 RSC.
  170. W. Liu, M. Zhu, J. Xiao, Y. Ling and H. Tang, J. Polym. Sci., Part A: Polym. Chem., 2016, 54, 3425–3435 CrossRef.
  171. R. Zhang, N. Zheng, Z. Song, L. Yin and J. Cheng, Biomaterials, 2014, 35, 3443–3454 CrossRef PubMed.
  172. D. S. Poché, S. J. Thibodeaux, V. C. Rucker, I. M. Warner and W. H. Daly, Macromolecules, 1997, 30, 8081–8084 CrossRef.
  173. H. Tang and D. Zhang, Polym. Chem., 2011, 2, 1542–1551 RSC.
  174. Y. Zhang, H. Lu, Y. Lin and J. Cheng, Macromolecules, 2011, 44, 6641–6644 CrossRef PubMed.
  175. J. Luijten, D. Y. Groeneveld, G. W. Nijboer, E. J. Vorenkamp and A. J. Schouten, Langmuir, 2007, 23, 8163–8169 CrossRef PubMed.
  176. H. Lu, Y. Bai, J. Wang, N. P. Gabrielson, F. Wang, Y. Lin and J. Cheng, Macromolecules, 2011, 44, 6237–6240 CrossRef PubMed.
  177. H. Lu, J. Wang, Y. Bai, J. W. Lang, S. Liu, Y. Lin and J. Cheng, Nat. Commun., 2011, 2, 206 CrossRef PubMed.
  178. N. P. Gabrielson, H. Lu, L. Yin, D. Li, F. Wang and J. Cheng, Angew. Chem., Int. Ed., 2012, 51, 1143–1147 CrossRef PubMed.
  179. N. Masato and K. Hiroyoshi, Bull. Chem. Soc. Jpn., 1975, 48, 2588–2591 CrossRef.
  180. Y. Kadoma, A. Ueno, K. Takeda, K. Uno and Y. Iwakura, J. Polym. Sci., Part B: Polym. Lett., 1974, 12, 715–718 Search PubMed.
  181. L. Yan, L. Yang, H. He, X. Hu, Z. Xie, Y. Huang and X. Jing, Polym. Chem., 2012, 3, 1300–1307 RSC.
  182. A. Ueno, F. Toda and Y. Iwakura, Chem. Lett., 1974, 97–100 CrossRef.
  183. J. Ding, C. Xiao, Z. Tang, X. Zhuang and X. Chen, Macromol. Biosci., 2011, 11, 192–198 CrossRef PubMed.
  184. J. Ding, W. Xu, Y. Zhang, D. Sun, C. Xiao, D. Liu, X. Zhu and X. Chen, J. Controlled Release, 2013, 172, 444–455 CrossRef PubMed.
  185. H. Tang and D. Zhang, Biomacromolecules, 2010, 11, 1585–1592 CrossRef PubMed.
  186. H. Tang, L. Yin, K. H. Kim and J. Cheng, Chem. Sci., 2013, 4, 3839–3844 RSC.
  187. Q. Hu, Y. Deng, Q. Yuan, Y. Ling and H. Tang, J. Polym. Sci., Part A: Polym. Chem., 2014, 52, 149–153 CrossRef.
  188. Z. Song, N. Zheng, X. Ba, L. Yin, R. Zhang, L. Ma and J. Cheng, Biomacromolecules, 2014, 15, 1491–1497 CrossRef PubMed.
  189. Y. Wu, X. Wang, Y. Ling and H. Tang, RSC Adv., 2015, 5, 40772–40778 RSC.
  190. M. Xiong, M. W. Lee, R. A. Mansbach, Z. Song, Y. Bao, R. M. Peek, C. Yao, L.-F. Chen, A. L. Ferguson, G. C. L. Wong and J. Cheng, Proc. Natl. Acad. Sci. U. S. A., 2015, 112, 13155–13160 CrossRef PubMed.
  191. Y. Deng, Y. Xu, X. Wang, Q. Yuan, Y. Ling and H. Tang, Macromol. Rapid Commun., 2015, 36, 453–458 CrossRef PubMed.
  192. K. Ishikawa and T. Endo, J. Am. Chem. Soc., 1988, 110, 2016–2017 CrossRef.
  193. P. Dubruel, L. Dekie and E. Schacht, Biomacromolecules, 2003, 4, 1168–1176 CrossRef PubMed.
  194. D. Lu, H. Wang, T. E. Li, Y. Li, F. Dou, S. Sun, H. Guo, S. Liao, Z. Yang, Q. Wei and Z. Lei, ACS Appl. Mater. Interfaces, 2017, 9, 16756–16766 CrossRef PubMed.
  195. Q. Xu, C. He, C. Xiao, S. Yu and X. Chen, Polym. Chem., 2015, 6, 1758–1767 RSC.
  196. Y. Liu, P. Chen and Z. Li, Macromol. Rapid Commun., 2012, 33, 287–295 CrossRef PubMed.
  197. A. J. Rhodes and T. J. Deming, ACS Macro Lett., 2013, 2, 351–354 CrossRef.
  198. H. Tang, L. Yin, H. Lu and J. Cheng, Biomacromolecules, 2012, 13, 2609–2615 CrossRef PubMed.
  199. I. Yakovlev and T. J. Deming, J. Am. Chem. Soc., 2015, 137, 4078–4081 CrossRef PubMed.
  200. J. Zhou, P. Chen, C. Deng, F. Meng, R. Cheng and Z. Zhong, Macromolecules, 2013, 46, 6723–6730 CrossRef.
  201. J. Xiao, M. Li, W. Liu, Y. Li, Y. Ling and H. Tang, Eur. Polym. J., 2017, 88, 340–348 CrossRef.
  202. O. Schäfer, D. Huesmann, C. Muhl and M. Barz, Chem. – Eur. J., 2016, 22, 18085–18091 CrossRef PubMed.
  203. O. Schäfer, D. Huesmann and M. Barz, Macromolecules, 2016, 49, 8146–8153 CrossRef.
  204. J. R. Kramer and T. J. Deming, Biomacromolecules, 2012, 13, 1719–1723 CrossRef PubMed.
  205. E. G. Gharakhanian and T. J. Deming, Biomacromolecules, 2015, 16, 1802–1806 CrossRef PubMed.
  206. J. R. Kramer and T. J. Deming, Chem. Commun., 2013, 49, 5144–5146 RSC.
  207. A. R. Rodriguez, J. R. Kramer and T. J. Deming, Biomacromolecules, 2013, 14, 3610–3614 CrossRef PubMed.
  208. M. Yu and T. J. Deming, Macromolecules, 1998, 31, 4739–4745 CrossRef PubMed.
  209. W. D. Fuller, M. S. Verlander and M. Goodman, Biopolymers, 1978, 17, 2939–2943 CrossRef.
  210. A. Sulistio, A. Blencowe, J. Wang, G. Bryant, X. Zhang and G. G. Qiao, Macromol. Biosci., 2012, 12, 1220–1231 CrossRef PubMed.
  211. D. Lu, H. Wang, T. E. Li, Y. Li, X. Wang, P. Niu, H. Guo, S. Sun, X. Wang, X. Guan, H. Ma and Z. Lei, Chem. Mater., 2017, 29, 5493–5503 CrossRef.
  212. K. Schlögl and H. Fabitschowitz, Monatsh. Chem. Verw. Teile Anderer Wiss., 1954, 85, 1060–1076 CrossRef.
  213. J. Sun and H. Schlaad, Macromolecules, 2010, 43, 4445–4448 CrossRef.
  214. K.-S. Krannig and H. Schlaad, J. Am. Chem. Soc., 2012, 134, 18542–18545 CrossRef PubMed.
  215. R. M. Guinn, A. O. Margot, J. R. Taylor, M. Schumacher, D. S. Clark and H. W. Blanch, Biopolymers, 1995, 35, 503–512 CrossRef PubMed.
  216. K. Schlögl and H. Pelousek, Monatsh. Chem. Verw. Teile Anderer Wiss., 1960, 91, 227–237 CrossRef.
  217. J. Huang, G. Habraken, F. Audouin and A. Heise, Macromolecules, 2010, 43, 6050–6057 CrossRef.
  218. J. Huang, C. Bonduelle, J. Thévenot, S. Lecommandoux and A. Heise, J. Am. Chem. Soc., 2012, 134, 119–122 CrossRef PubMed.
  219. J. Jacobs, N. Gathergood, J. P. A. Heuts and A. Heise, Polym. Chem., 2015, 6, 4634–4640 RSC.
  220. U. Haberthür and H.-G. Elias, Makromol. Chem., 1971, 144, 193–212 CrossRef.
  221. V. Nadeau, G. Leclair, S. Sant, J.-M. Rabanel, R. Quesnel and P. Hildgen, Polymer, 2005, 46, 11263–11272 CrossRef.
  222. A. C. Fonseca, M. H. Gil and P. N. Simões, Prog. Polym. Sci., 2014, 39, 1291–1311 CrossRef.
  223. N. Franz and H.-A. Klok, Macromol. Chem. Phys., 2010, 211, 809–820 CrossRef.
  224. J. W. Robinson and H. Schlaad, Chem. Commun., 2012, 48, 7835–7837 RSC.
  225. L. Guo, S. H. Lahasky, K. Ghale and D. Zhang, J. Am. Chem. Soc., 2012, 134, 9163–9171 CrossRef PubMed.
  226. J. W. Robinson, C. Secker, S. Weidner and H. Schlaad, Macromolecules, 2013, 46, 580–587 CrossRef.
  227. S. H. Lahasky, W. K. Serem, L. Guo, J. C. Garno and D. Zhang, Macromolecules, 2011, 44, 9063–9074 CrossRef.
  228. M. Taherimehr and P. P. Pescarmona, J. Appl. Polym. Sci., 2014, 131, 41141 CrossRef.
  229. G. Rokicki, Prog. Polym. Sci., 2000, 25, 259–342 CrossRef.
  230. L. Ding, Y. Li, Y. Li, Y. Liang and J. Huang, Eur. Polym. J., 2001, 37, 2453–2459 CrossRef.
  231. D. C. Webster and A. L. Crain, Functional Polymers, 1998, ch. 21, vol. 704, pp. 303–320 Search PubMed.
  232. J. Herzberger, K. Niederer, H. Pohlit, J. Seiwert, M. Worm, F. R. Wurm and H. Frey, Chem. Rev., 2016, 116, 2170–2243 CrossRef PubMed.
  233. R. C. Pratt, F. Nederberg, R. M. Waymouth and J. L. Hedrick, Chem. Commun., 2008, 114–116,  10.1039/B713925J.
  234. D. P. Sanders, K. Fukushima, D. J. Coady, A. Nelson, M. Fujiwara, M. Yasumoto and J. L. Hedrick, J. Am. Chem. Soc., 2010, 132, 14724–14726 CrossRef PubMed.
  235. J. V. Olsson, D. Hult, Y. Cai, S. Garcia-Gallego and M. Malkoch, Polym. Chem., 2014, 5, 6651–6655 RSC.
  236. G. Liu, F. He, Y. P. Wang, J. Feng and R. X. Zhuo, Chin. Chem. Lett., 2006, 17, 137–139 Search PubMed.
  237. S. Venkataraman, N. Veronica, Z. X. Voo, J. L. Hedrick and Y. Y. Yang, Polym. Chem., 2013, 4, 2945–2948 RSC.
  238. F. He, Y.-P. Wang, G. Liu, H.-L. Jia, J. Feng and R.-X. Zhuo, Polymer, 2008, 49, 1185–1190 CrossRef.
  239. T. Jiang, Y.-M. Li, Y. Lv, Y.-J. Cheng, F. He and R.-X. Zhuo, Colloids Surf., B, 2013, 111, 542–548 CrossRef PubMed.
  240. P. G. Parzuchowski, M. Jaroch, M. Tryznowski and G. Rokicki, Macromolecules, 2008, 41, 3859–3865 CrossRef.
  241. T. Jiang, Y. Li, Y. Lv, Y. Cheng, F. He and R. Zhuo, J. Mater. Sci.: Mater. Med., 2014, 25, 131–139 CrossRef PubMed.
  242. X. Zhang, H. Wang and Y. Dai, Mater. Lett., 2017, 198, 144–147 CrossRef.
  243. C. Lu, Q. Shi, X. Chen, T. Lu, Z. Xie, X. Hu, J. Ma and X. Jing, J. Polym. Sci., Part A: Polym. Chem., 2007, 45, 3204–3217 CrossRef.
  244. Q. Shi, X. Chen, T. Lu and X. Jing, Biomaterials, 2008, 29, 1118–1126 CrossRef PubMed.
  245. Q. Shi, Y. Huang, X. Chen, M. Wu, J. Sun and X. Jing, Biomaterials, 2009, 30, 5077–5085 CrossRef PubMed.
  246. D. Hu, H. Peng, Y. Niu, Y. Li, Y. Xia, L. Li, J. He, X. Liu, X. Xia, Y. Lu and W. Xu, J. Polym. Sci., Part A: Polym. Chem., 2015, 53, 750–760 CrossRef.
  247. S. Tempelaar, I. A. Barker, V. X. Truong, D. J. Hall, L. Mespouille, P. Dubois and A. P. Dove, Polym. Chem., 2013, 4, 174–183 RSC.
  248. H. Xiong, D. Zhou, Y. Qi, Z. Zhang, Z. Xie, X. Chen, X. Jing, F. Meng and Y. Huang, Biomacromolecules, 2015, 16, 3980–3988 CrossRef PubMed.
  249. H. Xiong, X. Wei, D. Zhou, Y. Qi, Z. Xie, X. Chen, X. Jing and Y. Huang, Bioconjugate Chem., 2016, 27, 2214–2223 CrossRef PubMed.
  250. B. D. Mullen, C. N. Tang and R. F. Storey, J. Polym. Sci., Part A: Polym. Chem., 2003, 41, 1978–1991 CrossRef.
  251. C.-F. Wang, Y.-X. Lin, T. Jiang, F. He and R.-X. Zhuo, Biomaterials, 2009, 30, 4824–4832 CrossRef PubMed.
  252. D. M. Stevens, S. Tempelaar, A. P. Dove and E. Harth, ACS Macro Lett., 2012, 1, 915–918 CrossRef PubMed.
  253. X. Hu, X. Chen, S. Liu, Q. Shi and X. Jing, J. Polym. Sci., Part A: Polym. Chem., 2008, 46, 1852–1861 CrossRef.
  254. H. Kuang, S. Wu, Z. Xie, F. Meng, X. Jing and Y. Huang, Biomacromolecules, 2012, 13, 3004–3012 CrossRef PubMed.
  255. H. Kuang, S. Wu, F. Meng, Z. Xie, X. Jing and Y. Huang, J. Mater. Chem., 2012, 22, 24832–24840 RSC.
  256. V. X. Truong, I. A. Barker, M. Tan, L. Mespouille, P. Dubois and A. P. Dove, J. Mater. Chem. B, 2013, 1, 221–229 RSC.
  257. I. A. Barker, M. P. Ablett, H. T. J. Gilbert, S. J. Leigh, J. A. Covington, J. A. Hoyland, S. M. Richardson and A. P. Dove, Biomater. Sci., 2014, 2, 472–475 RSC.
  258. K. Olofsson, M. Malkoch and A. Hult, J. Polym. Sci., Part A: Polym. Chem., 2016, 54, 2370–2378 CrossRef.
  259. G. S. Heo, S. Cho and K. L. Wooley, Polym. Chem., 2014, 5, 3555–3558 RSC.
  260. W. Chen, H. Yang, R. Wang, R. Cheng, F. Meng, W. Wei and Z. Zhong, Macromolecules, 2010, 43, 201–207 CrossRef.
  261. J. Xiong, F. Meng, C. Wang, R. Cheng, Z. Liu and Z. Zhong, J. Mater. Chem., 2011, 21, 5786–5794 RSC.
  262. F. Chen, J. W. S. Hayami and B. G. Amsden, Biomacromolecules, 2014, 15, 1593–1601 CrossRef PubMed.
  263. D. P. Sanders, D. J. Coady, M. Yasumoto, M. Fujiwara, H. Sardon and J. L. Hedrick, Polym. Chem., 2014, 5, 327–329 RSC.
  264. X. Hu, X. Chen, H. Cheng and X. Jing, J. Polym. Sci., Part A: Polym. Chem., 2009, 47, 161–169 CrossRef.
  265. R. Wang, W. Chen, F. Meng, R. Cheng, C. Deng, J. Feijen and Z. Zhong, Macromolecules, 2011, 44, 6009–6016 CrossRef.
  266. J. Mindemark and T. Bowden, Polymer, 2011, 52, 5716–5722 CrossRef.
  267. J. Mindemark and T. Bowden, Polym. Chem., 2012, 3, 1399–1401 RSC.
  268. Z. Y. Ong, K. Fukushima, D. J. Coady, Y.-Y. Yang, P. L. R. Ee and J. L. Hedrick, J. Controlled Release, 2011, 152, 120–126 CrossRef PubMed.
  269. X. Ding, C. Yang, T. P. Lim, L. Y. Hsu, A. C. Engler, J. L. Hedrick and Y.-Y. Yang, Biomaterials, 2012, 33, 6593–6603 CrossRef PubMed.
  270. F. Nederberg, Y. Zhang, J. P. K. Tan, K. Xu, H. Wang, C. Yang, S. Gao, X. D. Guo, K. Fukushima, L. Li, J. L. Hedrick and Y.-Y. Yang, Nat. Chem., 2011, 3, 409–414 CrossRef PubMed.
  271. Z. Y. Ong, C. Yang, S. J. Gao, X.-Y. Ke, J. L. Hedrick and Y. Y. Yang, Macromol. Rapid Commun., 2013, 34, 1714–1720 CrossRef PubMed.
  272. V. W. L. Ng, J. P. K. Tan, J. Leong, Z. X. Voo, J. L. Hedrick and Y. Y. Yang, Macromolecules, 2014, 47, 1285–1291 CrossRef.
  273. A. C. Engler, J. P. K. Tan, Z. Y. Ong, D. J. Coady, V. W. L. Ng, Y. Y. Yang and J. L. Hedrick, Biomacromolecules, 2013, 14, 4331–4339 CrossRef PubMed.
  274. V. W. L. Ng, X. Ke, A. L. Z. Lee, J. L. Hedrick and Y. Y. Yang, Adv. Mater., 2013, 25, 6730–6736 CrossRef PubMed.
  275. W. Chin, C. Yang, V. W. L. Ng, Y. Huang, J. Cheng, Y. W. Tong, D. J. Coady, W. Fan, J. L. Hedrick and Y. Y. Yang, Macromolecules, 2013, 46, 8797–8807 CrossRef.
  276. R. J. Ono, S. Q. Liu, S. Venkataraman, W. Chin, Y. Y. Yang and J. L. Hedrick, Macromolecules, 2014, 47, 7725–7731 CrossRef.
  277. N. H. Park, Z. X. Voo, Y. Y. Yang and J. L. Hedrick, ACS Macro Lett., 2017, 6, 252–256 CrossRef.
  278. A. C. Engler, J. M. W. Chan, D. J. Coady, J. M. O’Brien, H. Sardon, A. Nelson, D. P. Sanders, Y. Y. Yang and J. L. Hedrick, Macromolecules, 2013, 46, 1283–1290 CrossRef.
  279. J. M. W. Chan, X. Ke, H. Sardon, A. C. Engler, Y. Y. Yang and J. L. Hedrick, Chem. Sci., 2014, 3294–3300,  10.1039/C4SC00789A.
  280. A. C. Engler, X. Ke, S. Gao, J. M. W. Chan, D. J. Coady, R. J. Ono, R. Lubbers, A. Nelson, Y. Y. Yang and J. L. Hedrick, Macromolecules, 2015, 48, 1673–1678 CrossRef.
  281. J. M. W. Chan, R. J. Wojtecki, H. Sardon, A. L. Z. Lee, C. E. Smith, A. Shkumatov, S. Gao, H. Kong, Y. Y. Yang and J. L. Hedrick, ACS Macro Lett., 2017, 6, 176–180 CrossRef.
  282. Y. Zhou, R.-X. Zhuo and Z.-L. Liu, Macromol. Rapid Commun., 2005, 26, 1309–1314 CrossRef.
  283. N. H. Park, M. Fevre, Z. X. Voo, R. J. Ono, Y. Y. Yang and J. L. Hedrick, ACS Macro Lett., 2016, 5, 1247–1252 CrossRef.
  284. W. Chen, Y. Zou, J. Jia, F. Meng, R. Cheng, C. Deng, J. Feijen and Z. Zhong, Macromolecules, 2013, 46, 699–707 CrossRef.
  285. W. Chen, Y. Zou, F. Meng, R. Cheng, C. Deng, J. Feijen and Z. Zhong, Biomacromolecules, 2014, 15, 900–907 CrossRef PubMed.
  286. G. A. Barcan, X. Zhang and R. M. Waymouth, J. Am. Chem. Soc., 2015, 137, 5650–5653 CrossRef PubMed.
  287. X. Zhang and R. M. Waymouth, J. Am. Chem. Soc., 2017, 139, 3822–3833 CrossRef PubMed.
  288. L. Mespouille, F. Nederberg, J. L. Hedrick and P. Dubois, Macromolecules, 2009, 42, 6319–6321 CrossRef.
  289. R. J. Williams, R. K. O'Reilly and A. P. Dove, Polym. Chem., 2012, 3, 2156–2164 RSC.
  290. A. Dag, M. Aydin, H. Durmaz, G. Hizal and U. Tunca, J. Polym. Sci., Part A: Polym. Chem., 2012, 50, 4476–4483 CrossRef.
  291. W. Thongsomboon, M. Sherwood, N. Arellano and A. Nelson, ACS Macro Lett., 2013, 2, 19–22 CrossRef.
  292. R. J. Williams, I. A. Barker, R. K. O’Reilly and A. P. Dove, ACS Macro Lett., 2012, 1, 1285–1290 CrossRef.
  293. S. Kühling, H. Keul and H. Höcker, Makromol. Chem., 1990, 191, 1611–1622 CrossRef.
  294. J. Mindemark, L. Imholt, J. Montero and D. Brandell, J. Polym. Sci., Part A: Polym. Chem., 2016, 54, 2128–2135 CrossRef.
  295. A. W. Thomas, P. K. Kuroishi, M. M. Perez-Madrigal, A. K. Whittaker and A. P. Dove, Polym. Chem., 2017, 8, 5082–5090 RSC.
  296. P. K. Kuroishi, M. J. Bennison and A. P. Dove, Polym. Chem., 2016, 7, 7108–7115 RSC.
  297. T. Miyagawa, M. Shimizu, F. Sanda and T. Endo, Macromolecules, 2005, 38, 7944–7949 CrossRef.
  298. Q. J. Zhang, W. P. Zhu and Z. Q. Shen, Chin. Chem. Lett., 2010, 21, 1255–1258 CrossRef.
  299. W. Zhu, Y. Wang, S. Sun, Q. Zhang, X. Li and Z. Shen, J. Mater. Chem., 2012, 22, 11785–11791 RSC.
  300. W. Zhu, Y. Wang, Q. Zhang and Z. Shen, J. Polym. Sci., Part A: Polym. Chem., 2011, 49, 4886–4893 CrossRef.
  301. W. Zhu, Y. Wang, X. Cai, G. Zha, Q. Luo, R. Sun, X. Li and Z. Shen, J. Mater. Chem. B, 2015, 3, 3024–3031 RSC.
  302. X. Zhang, Z. Zhong and R. Zhuo, Macromolecules, 2011, 44, 1755–1759 CrossRef.
  303. X. Zhang, H. Dong, S. Fu, Z. Zhong and R. Zhuo, Macromol. Rapid Commun., 2016, 37, 993–997 CrossRef PubMed.
  304. J. Xu, T. M. Filion, F. Prifti and J. Song, Chem. – Asian J., 2011, 6, 2730–2737 CrossRef PubMed.
  305. Y. Zou, Y. Fang, H. Meng, F. Meng, C. Deng, J. Zhang and Z. Zhong, J. Controlled Release, 2016, 244, 326–335 CrossRef PubMed.
  306. X. Chen, S. P. McCarthy and R. A. Gross, Macromolecules, 1997, 30, 3470–3476 CrossRef.
  307. X. Chen, S. P. McCarthy and R. A. Gross, Macromolecules, 1998, 31, 662–668 CrossRef.
  308. S. Venkataraman, V. W. L. Ng, D. J. Coady, H. W. Horn, G. O. Jones, T. S. Fung, H. Sardon, R. M. Waymouth, J. L. Hedrick and Y. Y. Yang, J. Am. Chem. Soc., 2015, 137, 13851–13860 CrossRef PubMed.
  309. A. Pascual, J. P. K. Tan, A. Yuen, J. M. W. Chan, D. J. Coady, D. Mecerreyes, J. L. Hedrick, Y. Y. Yang and H. Sardon, Biomacromolecules, 2015, 16, 1169–1178 CrossRef PubMed.
  310. J. P. K. Tan, D. J. Coady, H. Sardon, A. Yuen, S. Gao, S. W. Lim, Z. C. Liang, E. W. Tan, S. Venkataraman, A. C. Engler, M. Fevre, R. Ono, Y. Y. Yang and J. L. Hedrick, Adv. Healthcare Mater., 2017, 6, 1601420 CrossRef PubMed.
  311. T. Steinbach and F. R. Wurm, Angew. Chem., Int. Ed., 2015, 54, 6098–6108 CrossRef PubMed.
  312. K. N. Bauer, H. T. Tee, M. M. Velencoso and F. R. Wurm, Prog. Polym. Sci., 2017, 73, 61–122 CrossRef.
  313. G. Becker and F. R. Wurm, Tetrahedron, 2017, 73, 3536–3540 CrossRef.
  314. B. Clément, B. Grignard, L. Koole, C. Jérôme and P. Lecomte, Macromolecules, 2012, 45, 4476–4486 CrossRef.
  315. Y.-C. Wang, Y.-Y. Yuan, F. Wang and J. Wang, J. Polym. Sci., Part A: Polym. Chem., 2011, 49, 487–494 CrossRef.
  316. S. Zhang, A. Li, J. Zou, L. Y. Lin and K. L. Wooley, ACS Macro Lett., 2012, 1, 328–333 CrossRef PubMed.
  317. S. Zhang, J. Zou, F. Zhang, M. Elsabahy, S. E. Felder, J. Zhu, D. J. Pochan and K. L. Wooley, J. Am. Chem. Soc., 2012, 134, 18467–18474 CrossRef PubMed.
  318. Y. H. Lim, G. S. Heo, S. Cho and K. L. Wooley, ACS Macro Lett., 2013, 2, 785–789 CrossRef PubMed.
  319. T. P. Gustafson, Y. H. Lim, J. A. Flores, G. S. Heo, F. Zhang, S. Zhang, S. Samarajeewa, J. E. Raymond and K. L. Wooley, Langmuir, 2014, 30, 631–641 CrossRef PubMed.
  320. F. Zhang, S. Zhang, S. F. Pollack, R. Li, A. M. Gonzalez, J. Fan, J. Zou, S. E. Leininger, A. Pavía-Sanders, R. Johnson, L. D. Nelson, J. E. Raymond, M. Elsabahy, D. M. P. Hughes, M. W. Lenox, T. P. Gustafson and K. L. Wooley, J. Am. Chem. Soc., 2015, 137, 2056–2066 CrossRef PubMed.
  321. S. Zhang, J. Zou, M. Elsabahy, A. Karwa, A. Li, D. A. Moore, R. B. Dorshow and K. L. Wooley, Chem. Sci., 2013, 4, 2122–2126 RSC.
  322. Y. H. Lim, K. M. Tiemann, G. S. Heo, P. O. Wagers, Y. H. Rezenom, S. Zhang, F. Zhang, W. J. Youngs, D. A. Hunstad and K. L. Wooley, ACS Nano, 2015, 9, 1995–2008 CrossRef PubMed.
  323. T. A. Aweda, S. Zhang, C. Mupanomunda, J. Burkemper, G. S. Heo, N. Bandara, M. Lin, C. S. Cutler, C. L. Cannon, W. J. Youngs, K. L. Wooley and S. E. Lapi, J. Labelled Compd. Radiopharm., 2015, 58, 234–241 CrossRef PubMed.
  324. E. Y. Zeynep, D. Antoine, C. Brice, B. Frank and J. Christine, J. Mater. Chem. B, 2015, 3, 7227–7236 RSC.
  325. Z. Ergul Yilmaz, T. Cordonnier, A. Debuigne, B. Calvignac, C. Jerome and F. Boury, Int. J. Pharm., 2016, 513, 130–137 CrossRef PubMed.
  326. J.-Z. Du, X.-J. Du, C.-Q. Mao and J. Wang, J. Am. Chem. Soc., 2011, 133, 17560–17563 CrossRef PubMed.
  327. X.-Z. Yang, J.-Z. Du, S. Dou, C.-Q. Mao, H.-Y. Long and J. Wang, ACS Nano, 2012, 6, 771–781 CrossRef PubMed.
  328. B. Clément, D. G. Molin, C. Jérôme and P. Lecomte, J. Polym. Sci., Part A: Polym. Chem., 2015, 53, 2642–2648 CrossRef.
  329. J. Zou, F. Zhang, S. Zhang, S. F. Pollack, M. Elsabahy, J. Fan and K. L. Wooley, Adv. Healthcare Mater., 2014, 3, 441–448 CrossRef PubMed.
  330. E. Baeten, S. Vanslambrouck, C. Jérôme, P. Lecomte and T. Junkers, Eur. Polym. J., 2016, 80, 208–218 CrossRef.
  331. Y. H. Lim, G. S. Heo, Y. H. Rezenom, S. Pollack, J. E. Raymond, M. Elsabahy and K. L. Wooley, Macromolecules, 2014, 47, 4634–4644 CrossRef PubMed.
  332. H. Shao, M. Zhang, J. He and P. Ni, Polymer, 2012, 53, 2854–2863 CrossRef.
  333. Y. Iwasaki, S. Komatsu, T. Narita, K. Akiyoshi and K. Ishihara, Macromol. Biosci., 2003, 3, 238–242 CrossRef.
  334. Y. Iwasaki, C. Nakagawa, M. Ohtomi, K. Ishihara and K. Akiyoshi, Biomacromolecules, 2004, 5, 1110–1115 CrossRef PubMed.
  335. C. Wachiralarpphaithoon, Y. Iwasaki and K. Akiyoshi, Biomaterials, 2007, 28, 984–993 CrossRef PubMed.
  336. J.-Z. Du, T.-M. Sun, S.-Q. Weng, X.-S. Chen and J. Wang, Biomacromolecules, 2007, 8, 3375–3381 CrossRef PubMed.
  337. Y. Iwasaki and K. Akiyoshi, Macromolecules, 2004, 37, 7637–7642 CrossRef.
  338. Y. Iwasaki and K. Akiyoshi, Biomacromolecules, 2006, 7, 1433–1438 CrossRef PubMed.
  339. G. Lapienis and S. Penczek, Macromolecules, 1977, 10, 1301–1306 CrossRef.
  340. T. Biela, S. Penczek, S. Slomkowski and O. Vogl, Makromol. Chem., Rapid Commun., 1982, 3, 667–671 CrossRef.
  341. J.-C. Brosse, D. Derouet, L. Fontaine and S. Chairatanathavorn, Makromol. Chem., 1989, 190, 2339–2345 CrossRef.
  342. J. Wang, P.-C. Zhang, H.-F. Lu, N. Ma, S. Wang, H.-Q. Mao and K. W. Leong, J. Controlled Release, 2002, 83, 157–168 CrossRef PubMed.
  343. J. Wang, H.-Q. Mao and K. W. Leong, J. Am. Chem. Soc., 2001, 123, 9480–9481 CrossRef PubMed.
  344. S.-W. Huang, J. Wang, P.-C. Zhang, H.-Q. Mao, R.-X. Zhuo and K. W. Leong, Biomacromolecules, 2004, 5, 306–311 CrossRef PubMed.
  345. J. Wang, P.-C. Zhang, H.-Q. Mao and K. W. Leong, Gene Ther., 2002, 9, 1254–1261 CrossRef PubMed.
  346. P.-C. Zhang, J. Wang, K. W. Leong and H.-Q. Mao, Biomacromolecules, 2005, 6, 54–60 CrossRef PubMed.
  347. X. Jiang, W. Qu, D. Pan, Y. Ren, J.-M. Williford, H. Cui, E. Luijten and H.-Q. Mao, Adv. Mater., 2013, 25, 227–232 CrossRef PubMed.
  348. A. A. Oswald, Can. J. Chem., 1959, 37, 1498–1504 CrossRef.
  349. K. Kaluzynski, J. Libisowski and S. Penczek, Macromolecules, 1976, 9, 365–367 CrossRef.
  350. J. Pretula, K. Kaluzynski and S. Penczek, Macromolecules, 1986, 19, 1797–1799 CrossRef.
  351. G. Łapienis, S. Penczek, G. P. Aleksiuk and V. A. Kropachev, J. Polym. Sci., Part A: Polym. Chem., 1987, 25, 1729–1736 CrossRef.
  352. G. Łapienis and S. Penczek, J. Polym. Sci., Part A: Polym. Chem., 1990, 28, 1519–1526 CrossRef.
  353. O. Kuznetsova, K. Kaluzynski, G. Lapienis, J. Pretula and S. Penczek, J. Bioact. Compat. Polym., 1999, 232–242 CrossRef.
  354. T. Wolf, T. Rheinberger and F. R. Wurm, Eur. Polym. J., 2017, 95, 756–765,  DOI:10.1016/j.eurpolymj.2017.05.048.
  355. T. Wolf, T. Steinbach and F. R. Wurm, Macromolecules, 2015, 48, 3853–3863 CrossRef.
  356. T. Steinbach, S. Ritz and F. R. Wurm, ACS Macro Lett., 2014, 3, 244–248 CrossRef.
  357. K. N. Bauer, L. Liu, D. Andrienko, M. Wagner, E. K. Macdonald, M. P. Shaver and F. R. Wurm, Macromolecules, 2018, 51, 1272–1279 CrossRef.
  358. I. Teasdale and O. Brüggemann, Polymers, 2013, 5, 161–187 CrossRef PubMed.
  359. R. S. Ullah, L. Wang, H. Yu, N. M. Abbasi, M. Akram, Z. ul-Abdin, M. Saleem, M. Haroon and R. U. Khan, RSC Adv., 2017, 7, 23363–23391 RSC.
  360. M. Gleria and R. De Jaeger, in New Aspects in Phosphorus Chemistry, ed. J.-P. Majoral, Springer Berlin Heidelberg, Berlin, Heidelberg, 2005, pp. 165–251,  DOI:10.1007/b100985.
  361. H. R. Allcock, S. M. Coley, I. Manners, O. Nuyken and G. Renner, Macromolecules, 1991, 24, 2024–2028 CrossRef.
  362. J. A. Dodge, I. Manners, H. R. Allcock, G. Renner and O. Nuyken, J. Am. Chem. Soc., 1990, 112, 1268–1269 CrossRef.
  363. I. Manners, H. R. Allcock, G. Renner and O. Nuyken, J. Am. Chem. Soc., 1989, 111, 5478–5480 CrossRef.
  364. H. R. Allcock, S. M. Coley and C. T. Morrissey, Macromolecules, 1994, 27, 2904–2911 CrossRef.
  365. M. Liang and I. Manners, J. Am. Chem. Soc., 1991, 113, 4044–4045 CrossRef.
  366. Y. Ni, A. Stammer, M. Liang, J. Massey, G. J. Vancso and I. Manners, Macromolecules, 1992, 25, 7119–7125 CrossRef.

This journal is © The Royal Society of Chemistry 2018