Artigo Acesso aberto Revisado por pares

4-Hydroxy- l -Proline as a General Platform for Stereoregular Aliphatic Polyesters: Controlled Ring-Opening Polymerization, Facile Functionalization, and Site-Specific Bioconjugation

2020; Chinese Chemical Society; Volume: 2; Issue: 5 Linguagem: Inglês

10.31635/ccschem.020.201900119

ISSN

2096-5745

Autores

Jingsong Yuan, Dong Shi, Yi Zhang, Jianhua Lü, Letian Wang, Er‐Qiang Chen, Hua Lu,

Tópico(s)

Carbon dioxide utilization in catalysis

Resumo

Open AccessCCS ChemistryRESEARCH ARTICLE1 Oct 20204-Hydroxy-l-Proline as a General Platform for Stereoregular Aliphatic Polyesters: Controlled Ring-Opening Polymerization, Facile Functionalization, and Site-Specific Bioconjugation Jingsong Yuan, Dong Shi, Yi Zhang, Jianhua Lu, Letian Wang, Er-Qiang Chen and Hua Lu Jingsong Yuan Beijing National Laboratory for Molecular Sciences, Center for Soft Matter Science and Engineering, Key Laboratory of Polymer Chemistry and Physics of Ministry of Education, College of Chemistry and Molecular Engineering, Peking University, Beijing 100871 (China). , Dong Shi Beijing National Laboratory for Molecular Sciences, Center for Soft Matter Science and Engineering, Key Laboratory of Polymer Chemistry and Physics of Ministry of Education, College of Chemistry and Molecular Engineering, Peking University, Beijing 100871 (China). , Yi Zhang Beijing National Laboratory for Molecular Sciences, Center for Soft Matter Science and Engineering, Key Laboratory of Polymer Chemistry and Physics of Ministry of Education, College of Chemistry and Molecular Engineering, Peking University, Beijing 100871 (China). , Jianhua Lu Beijing National Laboratory for Molecular Sciences, Center for Soft Matter Science and Engineering, Key Laboratory of Polymer Chemistry and Physics of Ministry of Education, College of Chemistry and Molecular Engineering, Peking University, Beijing 100871 (China). , Letian Wang Beijing National Laboratory for Molecular Sciences, Center for Soft Matter Science and Engineering, Key Laboratory of Polymer Chemistry and Physics of Ministry of Education, College of Chemistry and Molecular Engineering, Peking University, Beijing 100871 (China). , Er-Qiang Chen *Corresponding authors: E-mail Address: [email protected], E-mail Address: [email protected] Beijing National Laboratory for Molecular Sciences, Center for Soft Matter Science and Engineering, Key Laboratory of Polymer Chemistry and Physics of Ministry of Education, College of Chemistry and Molecular Engineering, Peking University, Beijing 100871 (China). and Hua Lu *Corresponding authors: E-mail Address: [email protected], E-mail Address: [email protected] Beijing National Laboratory for Molecular Sciences, Center for Soft Matter Science and Engineering, Key Laboratory of Polymer Chemistry and Physics of Ministry of Education, College of Chemistry and Molecular Engineering, Peking University, Beijing 100871 (China). https://doi.org/10.31635/ccschem.020.201900119 SectionsSupplemental MaterialAboutAbstractPDF ToolsAdd to favoritesTrack Citations ShareFacebookTwitterLinked InEmail Degradable polyesters have long been regarded as eco-friendly materials, useful for various applications while meeting the growing needs of sustainability. However, it is still challenging to synthesize functional aliphatic polyesters from abundant and cheap renewable sources. Our present study reports a readily available and versatile platform for producing functional and stereoregular aliphatic polyesters from 4-hydroxy-l-proline(4-HYP). We synthesized a bicyclic bridged lactone monomer, namely, NR-PL, by a simple and scalable two-step process allowing facile side-chain functionalization and derivatization. The ring-opening homopolymerization and copolymerization for the generation of NR-PL were controlled fully by using organobases such as 1,8-diazabicyclo[5.4.0]-undec-7-ene (DBU) without any detectable epimerization. This process afforded stereoregular polyesters PNRPE with molar mass (Mn) up to 90 kg/mol and a narrow dispersity (Ð) generally below 1.10. The uniqueness of the backbone, which contains two chiral centers on a rigid propyl ring, together with the versatility of the side chain, offer tunable properties complementary to existing aliphatic polyesters. The utility of the polymers was showcased by the facile site-specific bioconjugation of PNEG3PE, a water-soluble polyester, to a protein. This work might open numerous opportunities in creating functional and sustainable polyesters for a wide range of applications, including degradable plastics, drug delivery, and protein therapeutics. Download figure Download PowerPoint Introduction The past half-century has witnessed the booming of petroleum-based polymeric materials. However, the everlasting life span and end-of-use issues of such plastics have caused vast environmental pollutions, making sustainable polymeric products derived from renewable feedstock increasingly desirable.1–10 Degradable aliphatic polyesters have long been regarded as eco-friendly materials useful for various applications, including food packing, containers, agricultural mulching film, and biomaterials.5,11 For instance, polylactide (PLA), produced by the ring-opening polymerization (ROP) of lactide, is an excellent material for surgical suturing and disposable packing owing to its biocompatibility, biodegradability, and mechanical properties.12–14 Additionally, polyhydroxyalkanoates (PHA) such as poly(3-hydroxybutyrate) (P3HB) either biosynthesized using fermentation or chemically synthesized via the ROP of lactones, have long been considered promising candidates for substituting petroleum-based plastics.15,16 Despite the immense promises, the current choices of bio-renewable aliphatic polyesters are still rare; therefore, there is a pressing need for an expanded repertoire of functionalizable polyesters.17–23 Accordingly, the recent development of the ROP of amino acid-derived O-carboxyanhydride (OCA) has gained considerable success in preparing functional poly(α-hydroxy acid)s.24–27 However, the handling of OCA requires sophisticated synthetic skills, in the sense that, α-amino acids need to be converted into their corresponding α-hydroxyl acids before the transformation into the unstable OCA monomers under anhydrous conditions. Moreover, the ROP of OCAs often leads to racemization due to the enhanced acidity of α-hydrogen, which was addressed only recently by using carefully designed organometallic or organic catalysts.26,28–30 As such, a novel monomer platform that could provide easier access to libraries of functional and stereoregular polyesters from bio-derived resources is still an urgent need. 4-Hydroxyl-l-proline (4-HYP) is abundant in collagen and many other proteins; it is a naturally occurring amino acid conveniently accessible at relatively low prices. Previously, 4-HYP-based polyesters with only low molar mass (Mn), typically < 12 kg·mol−1, and broad dispersity (Ð) > 1.5, were reported by Langer et al.31,32 and Park et al.33 by step-growth condensation polymerization, whereas a well-controlled ROP process has never been reported. We envisioned that 4-HYP is an ideal building block to construct functional polyesters via ROP approach for the following reasons: (1) 4-HYP could be convert readily to a bicyclic bridged lactone, namely NR-PL (Scheme 1), with the amine group as a convenient chemical handle for functionalization by using a well-established protocol.34 (2) In general, the α-hydrogen of α-amino acids is less acidic, compared with α-hydroxyl acids due to a weaker induction effect, thus making epimerization less likely. (3) Recently, we showed that 4-HYP-derived bridged bicyclic thiolactones (NR-PTL) could undergo well-controlled ROP and afford high Mn and narrow Ð polythioesters.35 The control over ROP was achieved through the judicious molecular design, which involved highly strained monomers for rapid chain propagation and inert proline–proline thioester junctions in the polymer backbone for minimized chain transfer.36 Owing to the structural similarity, we speculated that NR-PL might also generate an outstandingly controlled ROP, thereby, representing a modulable platform for fabricating numerous functional, sustainable polyesters with tunable properties. Scheme 1 | Synthesis and ring-opening polymerization of a bicyclic bridged lactone monomer, NR-PL, for the subsequent fabrication of functional, stereoregular aliphatic polyesters. Download figure Download PowerPoint Results and Discussion Synthesis and controlled ROP of NR-PL As depicted in Scheme 1, various alcohols were coupled to the amino group of 4-HYP through a urethane group. Next, the precursors NR-HYP were converted to corresponding monomers bearing different side chains via Mitsunobu reaction with typical yields ∼53–80%.34 NBoc-PL, NC6-PL, and NC12-PL were selected as monomers bearing alkyl side chains of varying lengths and branching degrees; NCbz-PL, NEG3-PL, and NEGene-PL were synthesized to represent monomers carrying aryl, hydrophilic, and modifiable moieties, respectively. The characterization data of the precursors and monomers were obtained by 1H, 13C nuclear magnetic resonance (NMR), 1H–13C heteronuclear single quantum coherence (HSQC) NMR, and single-crystal X-ray diffraction (XRD), compiled in Supporting Information Figures S1–S21. We started our ROP investigation by screening proper organobases for the monomer NC12-PL using benzyl alcohol as a model initiator. We found that weak bases such as triethylamine (TEA), did not lead to any monomer conversion (entry 1, Supporting Information Table S1). Some phosphazene bases, such as 2-tert-butylimino-2-diethylamino-1,3-dimethylperhydro-1,3,2-diazaphosphorine (BEMP; BEMP−H+ pKaMeCN = 27.6) and the cyclic trimeric phosphazene base (CTPB; CTPB−H+, pKaMeCN = 33.3), gave fast ROP and over 95% monomer conversion in tetrahydrofuran (THF).37 However, both phosphazene superbases mediated ill-controlled ROP, as revealed by multimodal peaks, low Mn, and broad Ð in size-exclusion chromatography (SEC; entries 2 and 3, Supporting Information Table S1). In contrast, when the base 1,8-diazabicyclo[5.4.0]-undec-7-ene (DBU, DBU−H+ pKaMeCN = 24.3) was used for the ROP at an initial monomer/initiator/base ratio ([M]0/[I]0/[base]0) of 50/1/1 in deuterated chloroform (CDCl3), a > 95% monomer conversion was realized within 24 h, affording poly(N-carboxydodecane 4-hydroxyl-l-prolyl ester) (PNC12PE) with a sharp and unimodal peak in SEC. Also, the ROP was confirmed by 1H NMR spectroscopy ( Supporting Information Figure S22), and the shift of the ester carbonyl stretch peak from 1800 cm−1 of the lactone to the corresponding 1748 cm−1 of the open-chain ester was demonstrated by Fourier-transform infrared (FT-IR) spectroscopy ( Supporting Information Figure S23). The Mn and Ð measurements obtained were 17.9 kg·mol−1 (expected Mn = 16.3) and 1.09, respectively (entry 1, Table 1). Notably, prolonging the reaction time to 72 h did not cause broadening of the peak in the SEC, suggesting minimal transesterification-induced chain transfer, even after monomer consumption. Besides, increasing the equivalent of DBU to [M]0/[I]0/[base]0 = 50/1/5 led to a significantly enhanced ROP rate (> 95% monomer conversion within 5 h) without jeopardizing the degree of control of the reaction (entry 2, Table 1). Further, DBU-mediated ROP in other common organic solvents, such as dichloromethane (DCM) and THF, also gave similar well-controlled results (entries 3 and 4, Table 1; Supporting Information Figure S24). Kinetic studies of the DBU-catalyzed ROPs depicted first-order kinetics over monomer concentration in all three solvents, with those in DCM and CDCl3 being considerably faster than that in THF (Figure 1a). Moreover, the well-controlled chain-growth feature of DBU-mediated ROP of NC12-PL was confirmed by the linear relationship of Mn with the monomer conversion (Figure 1b) or the feeding [M]0/[I]0 ratio (Figure 1c). SEC analyses exhibited sharp and narrow unimodal peaks for all the polymers (inset of Figure 1b and c). The Mn of PNC12PE obtained were in < 10% deviation from expected values, and Ð were in the range of 1.06–1.18 (entries 5–10, Table 1). The highest Mn of PNC12PE obtained in this study was 89.6 kg·mol−1 at a feeding [M]0/[I]0 ratio of 300/1. Figure 1 | Controlled ROP of NR-PL. (a) ln([M]0/[M]) as a function of time for the NC12-PL ROP in different solvents at 30 °C, [M]0/[I]0/[base]0 = 50/1/5. (b) The plots of Mn and Ð as a function of monomer conversion for the NC12-PL ROP at 25 °C, [M]0/[I]0/[base]0 = 100/1/5. Inset: overlay of SEC traces at different monomer conversions. (c) The plots of Mn and Ð as a function of feeding [M]0/[I]0 ratio for NC12-PL ROP. Inset: overlay of SEC traces at different [M]0/[I]0 ratios. (d) MALDI-TOF mass spectrum of PNEG3PE initiated by benzyl alcohol, [M]0/[I]0/[base]0 = 10/1/5. ROP, ring-opening polymerization; NR-PL, a bicyclic bridged lactone monomer; SEC, size-exclusion chromatography; MALDI-TOF, matrix-assisted laser desorption/ionization-time of flight. Download figure Download PowerPoint Table 1 | Ring-Opening Polymerization Results of NR-PLa Formation Entry Monomer [M]0/[I]0/[base]0 Solvent Time (h) Mncal (kg·mol−1)b Mnobt (kg·mol−1)c Ðe 1 NC12-PL 50/1/1 CDCl3 24 16.3 17.9 1.09 2 NC12-PL 50/1/5 CDCl3 5 16.3 17.0 1.07 3 NC12-PL 50/1/5 DCM 4 16.3 14.0 1.10 4 NC12-PL 50/1/5 THF 24 16.3 17.9 1.11 5 NC12-PL 25/1/5 CDCl3 2 8.1 8.0 1.06 6 NC12-PL 75/1/5 CDCl3 8 24.4 21.8 1.08 7 NC12-PL 100/1/5 CDCl3 12 32.5 29.8 1.06 8 NC12-PL 150/1/5 CDCl3 18 48.8 48.4 1.09 9 NC12-PL 200/1/5 CDCl3 30 65.0 58.9 1.12 10 NC12-PL 300/1/5 CDCl3 48 97.5 89.6 1.18 11 NC6-PL 50/1/5 CDCl3 5 12.1 9.9 1.06 12 NC6-PL 100/1/5 CDCl3 12 24.1 21.2 1.08 13 NEG3-PL 25/1/5 CDCl3 2 7.6 7.1d 1.09 14 NEG3-PL 50/1/5 CDCl3 5 15.2 14.1d 1.06 15 NEG3-PL 75/1/5 CDCl3 8 22.7 19.3d 1.09 16 NEGene-PL 25/1/1 CDCl3 10 6.0 6.1d 1.08 17 NBoc-PL 50/1/1 CDCl3 24 10.6 13.4d 1.05 18 NEG3-PL-r-NCbz-PL 25/25/1/5 CDCl3 5 13.8 13.0d 1.05 19 NEG3-PL-b-NC12-PL 25/25/1/5 CDCl3 2/2 15.7 14.6 1.09 aAll the polymerizations were conducted at 25 °C in a glove box with benzyl alcohol as the initiator, DBU as the base, [NR-PL]0 = 1.4 M, quenched upon >95% monomer conversion. bCalculated molar mass based on feeding [M]0/[I]0 ratio. cObtained relative molar mass determined by SEC in THF using polystyrene as the standard; PNC6PE and PNC12PE were insoluble in DMF. dObtained absolute molar mass determined by SEC in DMF containing 0.1 M LiBr equipped with a multiangle laser light scattering (MALLS) detector. eDispersity, determined by SEC. To expand the monomer scope, we explored the ROP of NBoc-PL, NCbz-PL, NC6-PL, NEG3-PL, and NEGene-PL, respectively, under the reaction conditions mentioned above. Four out of the five monomers, namely NBoc-PL, NC6-PL, NEG3-PL, and NEGene-PL gave over 95% conversion and controlled the homopolymerization process satisfactorily (entries 11–17, Table 1; Supporting Information Figures S25–S29). For example, NEG3-PL was polymerized in a fully controlled manner and gave water-soluble PNEG3PE with linear growth of Mn as a function of feeding [M]0/[I]0 ratio up to 75/1 as unimodal SEC peaks ( Supporting Information Figure S30). Matrix-assisted laser desorption ionization-time of flight mass spectrometry (MALDI-TOF MS) of benzyl alcohol-initiated ROP of NEG3-PL at a feeding [M]0/[I]0 ratio of 10/1 displayed a group of Poisson-distributed peaks, which were assigned to oligomeric PNEG3PE bearing the C6H5CH2–/–OH end groups (Figure 1d). Homopolymerization of NCbz-PL in most common organic solvents, including DCM, CDCl3, THF, and DMF, led to evident precipitation due to the poor solubility of resultant PNCbzPE. Nevertheless, a random copolymerization of NEG3-PL and NCbz-PL gave copolymers with good solubility and excellent control, shown by both Mn and Ð (entry 18, Table 1). Sequential ROP of NEG3-PL and NC12-PL in one-pot synthesis gave a well-defined block copolymer P(NEG3PE-b-NC12PE), once again highlighting the controllability of the ROP (entry 19, Table 1; Supporting Information Figure S31). Moreover, the structural diversity of the polymer was tailorable by postpolymerization modification of PNEGenePE via highly efficient UV-triggered thiol–ene reaction, which enabled the facile introduction of cationic, anionic, as well as zwitterionic moieties in an almost quantitative fashion ( Supporting Information Figures S32–S35). Overall, the ROP of NR-PL offered outstanding control and tunability over Mn, Ð, end groups, and side-chain functionalities. Stereoregularity of PNRPE The ROP of OCA could undergo epimerization easily due to the deprotonation of the acidic α-hydrogen. Thus, we conducted comprehensive NMR and hydrolysis studies to investigate the stereoregularity of NR-PL ROP (Figure 2). Unfavorably, 13C NMR spectrum of PNBocPE in CDCl3 at room temperature (Figure 2a) exhibited complicated splitting and multimodal peaks. Switching the solvent to DMF-d7, the same polymer gave a considerably cleaner 13C NMR spectrum pattern than that in CDCl3 (Figure 2a), but the peaks for carbon (c, d, and b) were still broad and difficult to assign. We reasoned that the broadening and splitting could be a combined effect of the trans–cis isomerization of the urethane carbonyl and the endo–exo conformation isomerization of the prolyl ring.38–40 This notion was later confirmed by the 13C NMR spectroscopy measurements of PNBocPE in DMF-d7 at 90 °C, which showed sharp and single peaks for major prolyl carbons (Figure 2a). We next hydrolyzed the monomer NEG3-PL and corresponding polymer PNEG3PE in alkaline D2O solution at 4 °C for 2 h to obtain more conclusive evidence regarding the tacticity of the polymers. Interestingly, completion of both hydrolysis reactions were attained, giving almost identical 1H NMR spectra, assignable to the pure enantiomer (2S,4S)-N-(2,5,8,11-tetraoxadodecanoyl)-4-hydroxypyrrolidine-2-carboxylic acid (top and middle spectra in Figure 2b), which was distinct from the 1H NMR of (2S,4R)-N-(2,5,8,11-tetraoxadodecanoyl)-4-hydroxypyrrolidine-2-carboxylic acid (trans-NEG3-HYP; bottom spectrum in Figure 2b). Together, we inferred from the solvent and temperature-dependent 13C NMR spectroscopy and the 1H NMR analysis of the hydrolysis product that there was indeed no epimerization during the polymerization process. Figure 2 | (a) Overlay of 13C NMR spectra of PNBocPE in CDCl3 (25 °C) and DMF-d7 (25 and 90°C). (b) Overlay of 1H NMR spectra of trans-NEG3-HYP and the hydrolyzed product of NEG3-PL and PNEG3PE in alkaline D2O (pD = 10 for NEG3-PL pD = 13 for PNEG3PE, respectively, 4 °C, 2 h). Download figure Download PowerPoint Thermal and mechanical properties of PNRPE We conducted thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC) experiments to evaluate the thermal property and phase transition of PNRPE. The decomposition temperature (Td) to a 5% weight loss of PNEG3PE, PNC6PE, and PNC12PE were ∼300 °C, close to the Td of most aliphatic polyesters, while the Td of PNBocPE and PNEGenePE were 210 °C and 252 °C, respectively, likely due to the thermal instability of the Boc and allyl carbamate groups ( Supporting Information Figure S36 and Supporting Information Table S2). The DSC traces acquired during the second heating at a rate of 10 °C/min of PNRPE are depicted in Figure 3a. The glass transition temperatures (Tg) of PNEG3PE and PNEGenePE were detected at 40 °C and 64 °C, respectively. PNC12PE and PNC6PE showed complex transition behaviors; in the low-temperature range, both samples likely underwent the glass transition immediately, followed by a cold-crystallization process, resulting in an overall apparent endothermic process. Moreover, upon heating PNC12PE to high temperature (190 °C), sequential small endotherm, an exotherm, and a larger endotherm were observed, suggesting a minor melting/recrystallization at 196 °C and 198 °C, respectively occurred, proceeded by a major melting process with a melting temperature (Tm) at 202 °C. For PNC6PE, the larger endothermic peak was apparent at 221 °C, right after the small one at 206 °C. The DSC results implied that PNC12PE and PNC6PE might exhibit crystalline phase at low temperatures, and can be fabricated into a transparent film by hot compression at 180 °C ( Supporting Information Figure S37). Besides, the strain–stress curve of PNC12PE revealed a 6.1% elongation at break and Young's modulus of 143 MPa at 20°C ( Supporting Information Figure S38). Figure 3 | (a) DSC traces of PNRPE obtained during the second heating at a rate of 10 °C/min; Inset: expanded DSC traces of PNEG3PE and PNEGenePE demonstrating the glass transformation. (b) One-dimensional-XRD profiles of PNRPE recorded at room temperature. DSC, differential scanning calorimetry; PNRPE, bicyclic bridged lactone monomer. Download figure Download PowerPoint Figure 3b depicted the one-dimensional (1D) XRD profiles of PNRPE in which the samples were cooled at 1.0 °C/min from the molten state to room temperature. The XRD results indicated that both PNEG3PE and PNEGenePE were amorphous, in agreement with the DSC results. Two scattering halos were observed in the low- and high-angle regions, of which the former was believed to associate with the lateral dimension of the chain bearing rigid backbone and the relatively large pendants. In contrast, both PNC12PE and PNC6PE showed several sharp diffraction peaks. For PNC12PE, three low-angle peaks following a ratio of scattering vector (q-ratio) of 1∶2∶3 were detected, with the first peak at 0.260 Å−1. This observation was an indication of a layer structure with a layer period of 24.2 Å. Similar low-angle diffraction was observed with PNC6PE, having a shorter side chain of hexyl, and presenting a smaller layer period of 15.8 Å. In the high-angle region, PNC12PE and PNC6PE showed similar diffraction patterns, confirming the existence of crystalline structures. The crystallization behaviors were different between PNEG3PE, PNEGenePE and PNC12PE, PNC6PE, which might result from the differences in both their main-chain and side-chain compatibility. Collectively, the above results suggested that the thermal and mechanical properties of PNRPE could be tailored readily in a broad range by simply adjusting the side chains, thereby, offering tremendous potential for materials optimization in the future. Site-specific protein conjugation Previously, we employed trimethylsilyl phenylsulfide as an initiator for the ROP of amino acid N-carboxyanhydride (NCA) to install in situ a reactive phenyl thioester group at the end of synthetic polypeptides. Then the polymers were attached to enhanced green fluorescent protein bearing an N-terminal Cysteine (Cys-eGFP) successfully via a chemoselective native chemical ligation (NCL).41,42 Here, we envisaged that a similar strategy could be applicable for protein–PNEG3PE conjugation if a phenyl thioester could be installed at the α-chain end of PNEG3PE (Figure 4a). For this reason, thiophenol and DBU were selected to mediate the ROP of NEG3-PL simultaneously, which afforded PNEG3PE bearing terminal phenyl thioester (PNEG3PE-SPh) with a Mn of 6.0 kg·mol−1 and a Ð of 1.07 in SEC ( Supporting Information Figure S39). Remarkably, MALDI-TOF MS analysis of PNEG3PE-SPh revealed exclusively, the corresponding oligomer bearing the desirable phenyl thioester group on the chain end (Figure 4b). Next, we pursued the conjugation of PNEG3PE-SPh to a Cys-eGFP. Room temperature incubation of the two substrates at a molar ratio of 3/1 in 50 mM Tris·HCl buffer (pH = 7.0) for 12 h, afforded a 90% Cys-eGFP conversion ( Supporting Information Figure S40) and a purification yield of 47% of the corresponding conjugate, PNEG3PE-eGFP (Figure 4c). It is noteworthy that alkyl prolyl thioester was notoriously sluggish in NCL. Nonetheless, by using an in situ installed phenyl prolyl thioester, we demonstrated that effective NCL could take place at a reasonably good rate and yield without additives or catalysts promotion. Figure 4 | (a) Synthesis of PNEG3PE-SPh and subsequent NCL reaction with Cys-eGFP. (b) MALDI-TOF mass spectrum of PNEG3PE initiated by thiophenol, [M]0/[I]0/[base]0 = 10/1/1. (c) SDS-PAGE of PNEG3PE-eGFP conjugate. NCL, native chemical ligation; Cys-eGFP, cysteine-rich tagged green fluorescent protein MALDI-TOF; matrix-assisted laser desorption/ionization-time of flight. Download figure Download PowerPoint Experimental Methods General procedure for the synthesis of NR-PL We present NC12-PL as a model for the synthesis reaction, as follows: A solution of diisopropyl azodicarboxylate (DIAD; 4.1 g, 20.4 mmol, 1.1 equiv.) dissolved in THF (40 mL) was added dropwise to a solution mixture of NC12-HYP (6.4 g, 18.5 mmol, 1.0 equiv.) and triphenylphosphine (5.3 g, 20.4 mmol, 1.1 equiv.) dissolved in THF (250 mL) over 20 min period at 0°C. Subsequently, the reaction was stirred at room temperature for 4 h, following solvent removal by rotary evaporation. We purified the crude product by column chromatography (DCM:EA = 40∶1) to afford a white solid (4.5 g, yield 74%). The conditions for the crystal characterization and the results are indicated, as follows: 1H NMR (400 MHz, CDCl3): δ 5.11 (s, 1H), 4.61 (br, 1H), 4.12 (t, J = 6.7 Hz, 2H), 3.59 (d, J = 10.8 Hz, 1H), 3.50 (d, J = 10.6 Hz, 1H), 2.24 (d, J = 10.8 Hz, 1H), 2.03 (d, J = 10.6 Hz, 1H), 1.64 (m, 2H), 1.26 (m, 18H), 0.88 (t, J = 6.6 Hz, 3H). 13 C NMR (126 MHz, CDCl3): δ 170.7, 154.7, 78.3, 66.3, 57.4, 50.0, 39.1, 31.9, 29.6, 29.5, 29.3, 29.2, 28.9, 25.8, 22.7, 14.1. Electrospray ionization mass spectrometry (ESI-MS): Calculated m/z = 325.2; found m/z = 348.1 [M + Na]+. General procedure for the ROP of NR-PL We use NC12-PL as an example for the ROP, as follows: In a glovebox, the solution of NC12-PL (32.5 mg, 0.1 mmol, 50 equiv.) in CDCl3 (22 μL) was mixed with benzyl alcohol (2.0 μL × 1.0 M, 1.0 equiv.) in CDCl3, followed by the addition of 1,8-Diazabicyclo[5.4.0]undec-7-ene (DBU; 10.0 μL × 1.0 M, 5.0 equiv.) in CDCl3, and stirring the reaction at room temperature for 5 h. Then an aliquot of the solution was quenched by acetic acid (10.0 μ × 1.0 M) in CDCl3 before analysis by 1H NMR for the calculations of the monomer conversion. The polymer was purified initially by precipitation in hexane (10 mL) and centrifuged (4000 g for 5 min). After drying, the average yield obtained was ∼ 80–90%. The polymer was characterized as follows: Mn and Ð were measured by size-exclusion chromatography (SEC). 1H NMR (400 MHz, CDCl3): δ 7.38 (br, 0.11H), 5.24 (br, 1H), 4.86–4.35 (m, 1H), 4.25–3.90 (m, 2H), 3.85–3.47 (m, 2H), 2.80–2.22 (m, 2H), 1.61 (br, 2H), 1.26 (br, 18H), 0.88 (m, 3H). Conclusions We have demonstrated here the facile synthesis of a series of 4-HYP-derived functional aliphatic polyesters, (NR-PL), bearing tunable side chains. The polymerization process of the NR-PL was carefully controlled by using DBU as the catalyst and alcohol/thiol as initiators, affording stereoregular PNRPE polyesters with controlled Mn up to ∼90 kg·mol−1 and well-defined end groups. The solubility and thermomechanical properties of the resulting PNRPE were tailored easily by the introduction of side-chain functionalization either directly from the monomer or indirectly via postpolymerization modification. One polymer, PNEG3PE, demonstrated excellent water solubility and was amenable to site-specific protein conjugation via highly efficient NCL. With the controlled ROP and tunability of NR-PL, this approach has the potential to open opportunities in creating functional polyesters from renewable sources for a wide range of applications, including special plastics, drug delivery, and protein therapeutics. Supporting Information Supporting Information is available. Conflicts of Interest There is no conflict of interest to report. Acknowledgments We thank Prof. Eugene Y.-X. Chen, Prof. Zi-Chen Li, Prof. Youhua Tao, and Prof. Rong Tong for inspiring discussions. This research was made possible as a result of a generous grant from the National Natural Science Foundation of China (21722401 for H. Lu and 21634001 for E.Q. Chen). References 1. Schneiderman D. K.; Hillmyer M. A.50th Anniversary Perspective: There Is a Great Future in Sustainable Polymers.Macromolecules2017, 50, 3733–3749. Google Scholar 2. Hong M.; Chen E. Y. X.Chemically Recyclable Polymers: A Circular Economy Approach to Sustainability.Green Chem.2017, 19, 3692–3706. Google Scholar 3. Lu X. B.; Liu Y.; Zhou H.Learning Nature: Recyclable Monomers and Polymers.Chemistry2018, 24, 11255–11266. Google Scholar 4. Zhu Y.; Romain C.; Williams C. K.Sustainable Polymers from Renewable Resources.Nature2016, 540, 354–362. Google Scholar 5. Tang X. Y.; Chen E. Y. X.Toward Infinitely Recyclable Plastics Derived from Renewable Cyclic Esters.Chem2019, 5, 284–312. Google Scholar 6. Trotta J. T.; Watts A.; Wong A. R.; LaPointe A. M.; Hillmyer M. A.; Fors B. P.Renewable Thermosets and Thermoplastics from Itaconic Acid.ACS Sustain. Chem. Eng.2018, 7, 2691–2701. Google Scholar 7. Christensen P. R.; Scheuermann A. M.; Loeffler K. E.; Helms B. A.Closed-Loop Recycling of Plastics Enabled by Dynamic Covalent Diketoenamine Bonds.Nat. Chem.2019, 11, 442–448. Google Scholar 8. Chen X. S.; Chen G. Q.; Tao Y. H.; Wang Y. Z.; Lv X. B.; Zhang L. Q.; Zhu J.; Zhang J.; Wang X. H.Research Progress in Eco-Polymers.Acta Polym. Sin.2019, 50, 1068–1082. Google Scholar 9. Chen Y.-J.; Huang X.; Chen Y.; Wang Y.-R.; Zhang H.; Li C.-X.; Zhang L.; Zhu H.; Yang R.; Kan Y.-H.; Li S. L.Polyoxometalate-Induced Efficient Recycling of Waste Polyester Plastics into Metal–Organic Frameworks.CCS Chem.2019, 1, 561–570. Abstract, Google Scholar 10. Zhang X.; Fevre M.; Jones G. O.; Waymouth R. M.Catalysis as an Enabling Science for Sustainable Polymers.Chem. Rev.2018, 118, 839–885. Google Scholar 11. Hillmyer M. A.; Tolman W. B.Aliphatic Polyester Block Polymers: Renewable, Degradable, and Sustainable.Acc. Chem. Res.2014, 47, 2390–2396. Google Scholar 12. Nagarajan V.; Mohanty A. K.; Misra M.Perspective on Polylactic Acid (PLA) based Sustainable Materials for Durable Applications: Focus on Toughness and Heat Resistance.ACS Sustain. Chem. Eng.2016, 4, 2899–2916. Google Scholar 13. Auras R.; Harte B.; Selke S.An Overview of Polylactides as Packaging Materials.Macromol. Biosci.2004, 4, 835–864. Google Scholar 14. Zhang L.; Pratt R. C.; Nederberg F.; Horn H. W.; Rice J. E.; Waymouth R. M.; Wade C. G.; Hedrick J. L.Acyclic Guanidines as Organic Catalysts for Living Polymerization of Lactide.Macromolecules2010, 43, 1660–1664. Google Scholar 15. Tang X.; Chen E. Y. X.Chemical Synthesis of Perfectly Isotactic and High Melting Bacterial Poly(3-hydroxybutyrate) from Bio-Sourced Racemic Cyclic Diolide.Nat. Commun.2018, 9, 2345. Google Scholar 16. Muhammadi ; Shabina ; Afzal M.; Hameed S.Bacterial Polyhydroxyalkanoates-Eco-Friendly Next Generation Plastic: Production, Biocompatibility, Biodegradation, Physical Properties and Applications.Green Chem. Lett. Rev.2015, 8, 56–77. Google Scholar 17. Shi C.-X.; Guo Y.-T.; Wu Y.-H.; Li Z.-Y.; Wang Y.-Z.; Du F.-S.; Li Z.-C.Synthesis and Controlled Organobase-Catalyzed Ring-Opening Polymerization of Morpholine-2,5-Dione Derivatives and Monomer Recovery by Acid-Catalyzed Degradation of the Polymers.Macromolecules2019, 52, 4260–4269. Google Scholar 18. Yu L.; Zhang M.; Du F.-S.; Li Z.-C.ROS-Responsive Poly(ɛ-caprolactone) with Pendent Thioether and Selenide Motifs.Polym. Chem.2018, 9, 3762–3773. Google Scholar 19. Sanford M. J.; Van Zee N. J.; Coates G. W.Reversible-Deactivation Anionic Alternating Ring-Opening Copolymerization of Epoxides and Cyclic Anhydrides: Access to Orthogonally Functionalizable Multiblock Aliphatic Polyesters.Chem. Sci.2018, 9, 134–142. Google Scholar 20. Hu S.; Zhao J.; Zhang G.Noncopolymerization Approach to Copolymers via Concurrent Transesterification and Ring-Opening Reactions.ACS Macro Lett.2015, 5, 40–44. Google Scholar 21. Ji H.-Y.; Wang B.; Pan L.; Li Y.-S.Lewis Pairs for Ring-Opening Alternating Copolymerization of Cyclic Anhydrides and Epoxides.Green Chem.2018, 20, 641–648. Google Scholar 22. Hu S.; Dai G.; Zhao J.; Zhang G.Ring-Opening Alternating Copolymerization of Epoxides and Dihydrocoumarin Catalyzed by a Phosphazene Superbase.Macromolecules2016, 49, 4462–4472. Google Scholar 23. Hu S.; Zhao J.; Zhang G.; Schlaad H.Macromolecular Architectures Through Organocatalysis.Prog. Polym. Sci.2017, 74, 34–77. Google Scholar 24. du Boullay O. T.; Marchal E.; Martin-Vaca B.; Cossio F. P.; Bourissou D.An Activated Equivalent of Lactide Toward Organocatalytic Ring-Opening Polymerization.J. Am. Chem. Soc.2006, 128, 16442–16443. Google Scholar 25. Martin Vaca B.; Bourissou D.O-Carboxyanhydrides: Useful Tools for the Preparation of Well-Defined Functionalized Polyesters.ACS Macro Lett.2015, 4, 792–798. Google Scholar 26. Wang R.; Zhang J.; Yin Q.; Xu Y.; Cheng J.; Tong R.Controlled Ring-Opening Polymerization of O-Carboxyanhydrides Using a β-Diiminate Zinc Catalyst.Angew. Chem. Int. Ed.2016, 55, 13010–13014. Google Scholar 27. Yin Q.; Yin L.; Wang H.; Cheng J.Synthesis and Biomedical Applications of Functional Poly(α-hydroxy acids) via Ring-Opening Polymerization of O-Carboxyanhydrides.Acc. Chem. Res.2015, 48, 1777–1787. Google Scholar 28. Feng Q.; Tong R.Controlled Photoredox Ring-Opening Polymerization of O-Carboxyanhydrides.J. Am. Chem. Soc.2017, 139, 6177–6182. Google Scholar 29. Feng Q. Y.; Yang L.; Zhong Y. L.; Guo D.; Liu G. L.; Xie L. H.; Huang W.; Tong R.Stereoselective Photoredox Ring-Opening Polymerization of O-Carboxyanhydrides.Nat. Commun.2018, 9. Google Scholar 30. Li M.; Tao Y.; Tang J.; Wang Y.; Zhang X.; Tao Y.; Wang X.Synergetic Organocatalysis for Eliminating Epimerization in Ring-Opening Polymerizations Enables Synthesis of Stereoregular Isotactic Polyester.J. Am. Chem. Soc.2019, 141, 281–289. Google Scholar 31. Kohn J.; Langer R.Polymerization Reactions Involving the Side Chains of α-l-Amino Acids.J. Am. Chem. Soc.1987, 109, 817–820. Google Scholar 32. Kwon H. Y.; Langer R.Pseudopoly(amino acids): A Study of the Synthesis and Characterization of Poly(trans-4-hydroxy-N-acyl-l-proline esters).Macromolecules1989, 22, 3250–3255. Google Scholar 33. Lim Y. B.; Choi Y. H.; Park J. S.A Self-Destroying Polycationic Polymer: Biodegradable Poly(4-hydroxy-l-proline ester).J. Am. Chem. Soc.1999, 121, 5633–5639. Google Scholar 34. Bowers-Nemia M. M.; Joullié M. M.A Short Improved Synthesis of N-Substituted 5-aza-2-oxa-3-oxo-bicyclo[2.2.1]heptanes.Heterocycles1983, 20, 817. Google Scholar 35. Yuan J.; Xiong W.; Zhou X.; Zhang Y.; Shi D.; Li Z.; Lu H.4-Hydroxyproline-Derived Sustainable Polythioesters: Controlled Ring-Opening Polymerization, Complete Recyclability, and Facile Functionalization.J. Am. Chem. Soc.2019, 141, 4928–4935. Google Scholar 36. Gui Y.; Qiu L. Q.; Li Y. H.; Li H. X.; Dong S. W.Internal Activation of Peptidyl Prolyl Thioesters in Native Chemical Ligation.J. Am. Chem. Soc.2016, 138, 4890–4899. Google Scholar 37. Zhao N.; Ren C. L.; Li H. K.; Li Y. X.; Liu S. F.; Li Z. B.Selective Ring-Opening Polymerization of Non-Strained γ-Butyrolactone Catalyzed by A Cyclic Trimeric Phosphazene Base.Angew. Chem. Int. Ed.2017, 56, 12987–12990. Google Scholar 38. Bartuschat A. L.; Wicht K.; Heinrich M. R.Switching and Conformational Fixation of Amides Through Proximate Positive Charges.Angew. Chem. Int. Ed.2015, 54, 10294–10298. Google Scholar 39. Hinderaker M. P.; Raines R. T.An Electronic Effect on Protein Structure.Protein Sci.2003, 12, 1188–1194. Google Scholar 40. Hodges J. A.; Raines R. T.Energetics of an n→π* Interaction That Impacts Protein Structure.Org. Lett.2006, 8, 4695–4697. Google Scholar 41. Yuan J.; Sun Y.; Wang J.; Lu H.Phenyl Trimethylsilyl Sulfide-Mediated Controlled Ring-Opening Polymerization of α-Amino Acid N-Carboxyanhydrides.Biomacromolecules2016, 17, 891–896. Google Scholar 42. Hou Y.; Yuan J.; Zhou Y.; Yu J.; Lu H.A Concise Approach to Site-Specific Topological Protein-Poly(amino acid) Conjugates Enabled by In Situ-Generated Functionalities.J. Am. Chem. Soc.2016, 138, 10995–11000. Google Scholar Previous articleNext article FiguresReferencesRelatedDetails Issue AssignmentVolume 2Issue 5Page: 236-244Supporting Information Copyright & Permissions© 2020 Chinese Chemical SocietyKeywordsROPbioconjugationpolyesterstereoregular4-hydroxyprolinesustainableAcknowledgmentsWe thank Prof. Eugene Y.-X. Chen, Prof. Zi-Chen Li, Prof. Youhua Tao, and Prof. Rong Tong for inspiring discussions. This research was made possible as a result of a generous grant from the National Natural Science Foundation of China (21722401 for H. Lu and 21634001 for E.Q. Chen). Downloaded 1,420 times Loading ...

Referência(s)
Altmetric
PlumX