Functionalized Polyesters with Tunable Degradability Prepared by

Apr 11, 2017 - Enter the #ILoveBiochemistry Video Competition. Biochemistry is launching a video competition and YOU could win a $100 Amazon gift ... ...
0 downloads 19 Views 3MB Size
Article pubs.acs.org/Macromolecules

Functionalized Polyesters with Tunable Degradability Prepared by Controlled Ring-Opening (Co)polymerization of Lactones Yue-Chao Xu, Wei-Min Ren, Hui Zhou, Ge-Ge Gu, and Xiao-Bing Lu* State Key Laboratory of Fine Chemicals, Dalian University of Technology, 2 Linggong Road, Dalian 116024, China S Supporting Information *

ABSTRACT: Various well-defined diblock copolymers with poly(α-methylene−β-butyrolactone) (PMβBL) segment were prepared by salen−aluminum complexes or aminoalkoxybis(phenolate)yttrium amido complex mediated living ring-opening polymerization of racemic α-methylene-β-butyrolactone (rac-MβBL) and different lactones via one-pot, two-step method. These diblock copolymers all consist of semicrystalline and amorphous segments, possessing both a melting point and a glass transition temperature. With the use of simple salenAlMe with bulky (tert-butyldimethylsilyl) groups on the phenolate ortho position as catalyst in conjunction with equivalent benzyl alcohol, MβBL and β-butyrolactone (BL) have nearly identical reactivity in their copolymerization, affording the copolymers P(MβBL-ran-BL) with random distributions of the two monomers. Notably, these copolymers exhibited tunable degradability in the presence of Lewis base, dependent on the MβBL unit content in the random copolymers. Moreover, the vinylidene groups in P(MβBL-ran-BL) copolymers could be further functionalized through radicalinitiated cross-linking reactions and thiol−ene click chemistry, producing functionalized and cross-linking polymers with enhanced thermal property. These strategies will provide enormous possibilities to synthesize diverse well-defined block and random copolymers with designable segments or functional groups.



INTRODUCTION Aliphatic polyesters have received tremendous attention as environmentally friendly biomaterials in the past decades.1−4 P r o m i n e n t am o n g t h e s e p o l y es t e r s a r e p o l y ( β hydroxyalkanoate)s as suitable materials for medical applications due to their good biodegradability and biocompatibility5−8 as well as easily tunable substituent at β-position of the three-carbon backbone.9,10 Poly(β-hydroxybutyrate) (PHB), the simplest poly(βhydroxyalkanoate), can be produced by bacteria or ringopening polymerization (ROP) of racemic β-butyrolactone (rac-BL). PHB resulted from the fermentative process, possessing an isotactic structure with complete (R)-configuration, is a typical semicrystalline material with high melting temperature (Tm) and relatively low thermostability. As a consequence, the crystalline PHB is difficult to be processed, significantly limiting its applications.11−13 Contrarily, ROP of rac-BL usually affords atactic PHB with a low glass transition temperature (Tg) at 5 °C.12,14,15 Some well-defined catalyst systems have been developed to produce PHBs with various stereotacticity. Among them, yttrium complexes supported by aminoalkoxybis(phenolate) ligands,13,16−18 diamine bisphenolate Salan-type, and binaphthyl Salen-type ligands19−21 have shown great stereocontrol ability and reactivity to afford syndiotactic PHBs with different syndiotacticity (up to 0.94) © 2017 American Chemical Society

under mild reaction conditions. Tm of the resultant syndiotactic PHBs varied from 120 to 183 °C depending on their syndiotacticity. Besides, PHBs with high molecular weight and moderate syndiotacticity (0.55−0.75) could be obtained in the presence of racemic and enantiopure zinc catalysts supported by substituted diaminophenolate ancillary ligands.22 Isotactic-enriched high-molecular-weight PHBs can be obtained in the presence of chromium(III) salophen complexes.23,24 On the other hand, the copolymerization of rac-BL and other lactones by controlled ring-opening polymerization could afford PHB-based block and random copolymers with improved physical properties.25−31 For example, the well-defined PLLA− PHB−PLLA (PLLA, poly(L-lactic acid)) triblock polyester exhibited better mechanical properties than the corresponding pure PLLA and PHB polymers (Scheme 1, route A).32−34 In addition, the substituents at β-position of PHAs also influence their thermal properties and biocompatibilities dramatically. Benzyl β-malolactonate (MLABz) is a monomer for preparing water-insoluble poly(β-benzyl malolactonate) (PMLABz), which has a pendent benzyl functional group at β-position (Scheme 1, route B). Further postpolymerization through Received: February 3, 2017 Revised: April 4, 2017 Published: April 11, 2017 3131

DOI: 10.1021/acs.macromol.7b00239 Macromolecules 2017, 50, 3131−3142

Article

Macromolecules

Scheme 1. Polyesters Based on PHB or β-Functionalized PHB (Routes A−C) and α-Functionalized Polyesters from rac-MβBL (Route D)

due to the absence of functionality at the α-position which allows for further postpolymerization. In this contribution, we introduce racemic α-methylene-βbutyrolactone (rac-MβBL)66 with a vinylidene functional group at the α-position and copolymerized with other lactones to produce various well-defined block copolymers or random copolymers through sequential or simultaneous copolymerization (Scheme 1, route D). Previously, Salen aluminum complexes and aminoalkoxybis(phenolate) yttrium amido complexes were proved to be efficient complexes in living ROP of rac-MβBL and other lactones.17,66,67 The living polymerization character offers advantages to make various well-defined block copolymers with different lactones. In this work, rac-butyrolactone (rac-BL), rac-lactide (rac-LA), εcaprolactone (ε-CL), and macrolactone ω-pentadecalactone (ω-PDL) were employed as comonomers to build P(MβBL)based diblock copolymers. Because of the distinct stereoselectivity of two monomers, the tacticity of each block unit is varied depending on the catalytic system. Herein, several

hydrogenolysis of benzyl groups in PMLABz leads to an hydrophilic poly(malicacid) (PMLA).9,10,35−40 Copolymerization of rac-BL and racemic allyl-β-butyrolactone (rac-BLallyl) can afford PHB-based copolyesters with pendant allyl groups at βposition which have great potential to functionalization (Scheme 1, route C).41−43 Besides the aforementioned works, many papers have illustrated synthesis of various functional aliphatic polyesters through ROP of α-functionalized lactides44−52 or lactones.53−59 However, for poly(β-hydroxyalkanoate)s, most works focus on the functionalization at the β-position. It is highly desired to develop new analogue β-lactones and further copolymerize with other lactones to provide novel polyesters with various properties. Accordingly, modifications at the α-position of PHB should extend the scope of functional polyesters. Although some α,α′,β-trisubstituted β-lactones60−62 and Ntritylated-serine-β-propiolactone63−65 have been employed to synthesize novel polyesters, research in this field is still limited 3132

DOI: 10.1021/acs.macromol.7b00239 Macromolecules 2017, 50, 3131−3142

Article

Macromolecules

Scheme 2. Synthesis of Diblock and Random Copolymers by Metal Complexes Mediated ROP of rac-MβBL and Various Lactones

prepared by the copolymerization of rac-MβBL and rac-BL. The degradability of the resultant copolymers improved dramatically with the increasing MβBL unit content in the random copolymers. Moreover, the vinylidene groups in the backbones could be functionalized by thiol−ene click chemistry as well as cross-linking74−80 through the inter- or intramolecular process.

complexes are employed to obtain diblock copolymers and random copolymers (Scheme 2). Degradability is one of the most important properties for polyesters. For PHB, this process always proceed slowly in microorganisms or neutral pH environment.68 Therefore, some methods for chemical degradation of PHB have been reported, such as hydrolysis69,70 and alcoholysis.71−73 However, certain amounts of strong acid were required as a protic reagent in the reaction. On the other hand, the electricity of carbonyl group in the PHB backbone is not positive enough to be attacked by a nucleophilic reagent. Thus, the degradation of PHB promoted by Lewis base, especially tertiary amine, still remains unsuccessful. Herein, by the introduction of vinylidene group at the α-position of PHB, the positive electricity of carbonyl group in the PMβBL backbone is improved, and the carbonyl group can be attacked by nucleophilic reagents and thus lead to the degradation of polymer through random chain scission. Subsequently, copolymers with tunable degradability can be



RESULTS AND DISCUSSION Synthesis of Diblock Copolymers. As shown in Table 1, various diblock copolymers containing PMβBL segment were synthesized by the one-pot, two-step method, using different catalytic systems. In a previous contribution, we have demonstrated that the ROP of rac-MβBL catalyzed by complex 1/benzyl alcohol exhibited a living polymerization character, affording syndiotactic-enriched semicrystalline PMβBL with Tm around 110 °C.66 Although complex 1 also showed great catalytic activity for the ROP of rac-BL, the resultant PHB is a typical amorphous and atactic material (Supporting Informa3133

DOI: 10.1021/acs.macromol.7b00239 Macromolecules 2017, 50, 3131−3142

Article

Macromolecules Table 1. Sequential Copolymerization of Different Monomers Initiated by Different Metal Complex/BnOH Systems entrya

complex

monomer

temp (°C)

convb (%)

Mn,theoryc (kg mol−1)

Mn,expd (kg mol−1)

PDId

Tge (°C)

Tme (°C)

Hme (J/g)

1

1

2

2

3

3

4

3

5

3

MβBL BL BL MβBL MβBL rac-LA MβBL ε-CL MβBL ω-PDL

100 100 25 25 75 75 100 100 100 100

>99 93 >99 43 >99 >99 >99 >99 >99 77

5.0 9.0 4.4 8.6 5.0 12.2 5.0 10.8 5.0 14.3

6.6 12.7 4.9 7.2 5.9 16.6 5.6 13.9 5.5 14.8

1.05 1.06 1.15 1.29 1.16 1.20 1.20 1.34 1.17 1.35

16.6 11.0 1.9 9.6 17.4 43.1 19.2 n.d. 18.7 n.d.

115.5 101.1 135.1 128.2 n.d. 185.3 n.d. 53.4 n.d. 90.5

25.6 9.1 19.2 15.3 n.d. 34.8 n.d. 46.1 n.d. 86.4

a

Sequential copolymerizations were performed through one-pot, two-step processes with feed ratio of M1:M2:initiator = 50:50:1, [M]0 = 2.5 M in toluene (except for entry 3 in THF). For example, the first monomer (50 equiv) was allowed to react completely to prepare first segment, and then the second monomer (50 equiv) was added to without any treatment of the first reaction to obtain the block copolymer. The reaction time of every step was 24 h which was not optimized. bDetermined by 1H NMR spectroscopy. cCalculated from the equation Mn,theory = (M1/I) × conversion of M1 × 98.04 + (M2/I) × conversion of M2 × 86.04 + 108. dDetermined by GPC in THF, polystyrene as standards. eInvestigated by differential scanning calorimetry (DSC).

Figure 1. Representative GPC profiles of (A) first PMβBL segments (black) and PMβBL-b-PHB (red), Table 1, entry 1. (B) first PMβBL segments (black) and PMβBL-b-PLA (red), Table 1, entry 3.

heat fusion decreased slightly due to the introduction of amorphous PMβBL segment. Complex 3 is also an atactic catalyst for ROP of MβBL, affording amorphous PMβBL segment in the PMβBL-b-PLA, PMβBL-b-PCL, and PMβBL-b-PPDL diblock copolymers. The representative GPC profiles of first PMβBL segment and PMβBL-b-PLA diblock copolymer are shown in Figure 1B. Molecular weight of the resultant PMβBL-b-PLA increased without increasing in polydispersity. 1H NMR spectra in Figure 3 revealed the composition of the resultant diblock copolymers. The ratio fits well with the conversion of monomers in Table 1. Thermal analysis indicated amorphous PMβBL segment influenced the crystallization of PMβBL-b-PLA dramatically (Figure 2C). In the second heating process, PLA homopolymer has shown a melting peak without cold crystallization, indicating PLA has crystallized completely during the cooling process. PMβBL-b-PLA exhibited an obvious cold crystallization peak in the second heating run, suggesting its relative lower crystallization capacity compared to PLA homopolymer. However, the crystallization behavior of PMβBL-b-PPDL has little changes compared to PPDL (Figure 2D) because PPDL are polyethylene-like polyesters with extremely high crystallinity. In addition, polarizing microscope (POM) tests revealed these different crystallization behaviors visually (Supporting Information, Figure S1). Photographs recorded at different temperatures have proved PMβBL-b-PLA barely crystallize in the cooling process at 10 °C/min, while PMβBL-b-PPDL crystallized rapidly at the same cooling rate.

tion, Table S1, entry 1). These characters provide a route to synthesize diblock copolymer with crystalline PMβBL segment and amorphous PHB segment. Syndiotactic-enriched semicrystalline PMβBL was first prepared by complex 1/BnOH mediated ROP of rac-MβBL. After complete conversion of racMβBL, equivalent rac-BL was added to the reaction system, in which PMβBL with a hydroxyl end group plays as a macromolecular initiator for further polymer-chain growth to afford PMβBL-b-PHB diblock copolymer with narrow molecular weight distribution (PDI) of 1.06 (Figure 1A), which is close to that of PMβBL segment (PDI = 1.05). (Table 1, entry 1) The PMβBL-b-PHB copolymer has Tm at 101.1 °C and Tg at 11.0 °C, which is between that of PMβBL and PHB homopolymers. Similarly, the use of complex 2 or 3, various diblock copolymers, PHB-b-PMβBL, PMβBL-b-PLA, PMβBL-b-PCL, and PMβBL-b-PPDL, were prepared as the same procedure. Yttrium complexes are efficient catalysts for controlled-living ROP of rac-BL and highly syndiotactic-enriched PHB could be made under mild reaction conditions.3,16,17 However, yttrium complex 2 exhibited no stereoselectivity in the ROP of racMβBL to produce atactic PMβBL (Table S1, entry 4), though it is a syndiotactic-specific catalyst for the ROP of rac-BL. Therefore, the resulting PHB-b-PMβBL diblock copolymer consists of semicrystalline PHB and amorphous PMβBL segments, possessing a Tg of 9.6 °C and a Tm of 128.2 °C (Table 1, entry 2). Figure 2B shows the comparative thermal analysis of PHB-b-PMβBL, PHB, and PMβBL homopolymers produced by complex 2 (Table S1, entries 3 and 4). Tm and 3134

DOI: 10.1021/acs.macromol.7b00239 Macromolecules 2017, 50, 3131−3142

Article

Macromolecules

Figure 2. Comparative DSC profiles (second heating runs) of homopolymers and corresponding diblock copolymers: (A) 1. first PMβBL segment (Table 1, entry 1); 2. PHB homopolymer (Table S1, entry 1); 3. PMβBL-b-PHB (Table 1, entry 1); 4. PMβBL-b-PHB (Table 1, entry 1, crystallized at 90 °C). (B) 1. 1st PHB segment (Table 1, entry 2); 2. Atactic PMβBL (Table S1, entry 4); 3. PHB-b-PMβBL (Table 1, entry 2). (C) 1. PLA homopolymer (Table S1, entry 5); 2. Atactic PMβBL (Table 1, entry 3); 3. PMβBL-b-PLA (Table 1, entry 3). (D) 1. PPDL homopolymer (Table S1, entry 6); 2. Atactic PMβBL (Table 1, entry 5); 3. PMβBL-b-PPDL (Table 1, entry 5).

Figure 3. 1H NMR spectra of the well-defined diblock copolymers initiated by Salen−Al complexes.

3135

DOI: 10.1021/acs.macromol.7b00239 Macromolecules 2017, 50, 3131−3142

Article

Macromolecules

Figure 4. Fineman−Ross plot and copolymerization kinetic study initiated by the Salen−Al complex 4/BnOH system.

Table 2. Homopolymerization and Simultaneous Copolymerization of rac-MβBL (M1) and rac-BL (M2) Initiated by the Salen− Al Complex 4/BnOH convb (%) entrya

M1:M2:I

time (h)

1 2 3 4 5

100:0:1 0:100:1 10:90:1 30:70:1 50:50:1

16 16 60 60 60

M1

M2

vinyli dene groupc (%)

Mn,theoryd (kg mol−1)

Mn,expe (kg mol−1)

PDIe

Tgf (°C)

Tmf (°C)

hetero-sequencesg (% in mol)

100 0 9 27 46

4.0 3.1 8.6 8.7 8.8

4.3 4.0 8.9 8.2 7.9

1.12 1.19 1.25 1.21 1.22

16.5 2.1 2.7 5.7 5.2

93.2 n.d. n.d. n.d. n.d.

0

35 >99 >99 >99

40 91 92 90

91 69 48

a

All polymerization were carried out in toluene solution using benzyl alcohol as an initiator to generate aluminum benzyloxides, [lactone]0 = 2.5 M. Determined by 1H NMR from each monomer. cDetermined by 1H NMR from vinylidene groups of the resultant copolymers. dCalculated from the equation Mn, theory = (M1/I) × conversion of M1 × 98.04 + (M2/I) × conversion of M2 × 86.04 + 108. eDetermined by GPC in THF, polystyrene as standards. fDetermined by DSC. gThe percent mole ratio of heterosequences in the resultant copolymers, determined by 1H NMR from vinylidene groups, calculated from [MβBL-BL]/[MβBL-MβBL + MβBL-BL]. b

Figure 5. Comparative 1H NMR spectra of the resultant random copolymers in Table 2: (A) PMβBL, entry 1; (B) PHB, entry 2; (C) P9, entry 3; (D) P27, entry 4; (E) P46, entry 5.

Synthesis of Random Copolymers. The homopolymerizations of MβBL and BL were performed in the presence of various complexes 1−4 in conjunction with equivalent BnOH. Complex 4 showed nearly identical reactivity for both

monomers, while other catalytic systems exhibited great disparities in the ROP polymerization of rac-MβBL and racBL (Table S1). It means the copolymerization of rac-MβBL and rac-BL might proceed by random insertions of two monomers, 3136

DOI: 10.1021/acs.macromol.7b00239 Macromolecules 2017, 50, 3131−3142

Article

Macromolecules Table 3. Degradation of PMβBL and P(MβBL-ran-BL) in the Presence of Amine entry

parent polymer

Lewis basea (equiv)

T (°C)

reaction time (h)

Mn,parentb (kg mol−1)

PDIb

Mn,degradation.c (kg mol−1)

PDIc

1 2 3 4 5 6 7 8 9 10 11 12

PHB PMβBL PMβBL PMβBL P46 P27 P9 P27 P9 PMβBL PMβBL PMβBL

TEA, 0.1 TEA, 0.1 TEA, 0.1 TEA, 0.01 TEA, 0.1 TEA, 2.0 TEA, 2.0 TEA, 2.0 TEA, 2.0 DIPEA, 0.1 DIPEA, 0.01 TEA, 0.1

80 80 25 80 80 80 80 25 25 80 80 80

12 12 12 144 12 12 12 12 12 144 144 144

11.2 8.1 5.7 8.1 7.9 8.2 8.9 8.2 8.9 8.1 8.1 8.1

1.25 1.15 1.08 1.15 1.22 1.21 1.25 1.21 1.25 1.15 1.15 1.15

10.1 1.0 2.2 0.9 0.9 1.6 5.0 5.6 6.2 1.0 1.1 0.4

1.32 2.20 1.72 2.10 2.79 2.15 1.77 1.83 2.27 2.24 2.03 1.5

a

Amine was added according to the MβBL unit content. bMolecular weight and distribution of parent polymers before degradation. Determined by GPC in THF, with polystyrene as standard. cMolecular weight and distribution of degradation products. Determined by GPC in THF, with polystyrene as standard.

Figure 6. ESI-MS spectra of the degradation products mediated by amine in Table 3: (A) entry 4; (B) entry 11.

amorphous without Tm. The copolymers contain various contents of vinylidene groups are marked as P9, P27, P46, and P100. The composition of copolymers can be determined by NMR investigations according the previous work of Pappalardo.28 Comparative 1H NMR spectra of the resultant random copolymers P(M β BL-ran-BL) are shown in Figure 5. Fortunately, the vinylidene hydrogen of MβBL unit is very sensitive to chemical condition. Compared with the 1H NMR spectra of PMβBL and PHB homopolymer (Figure 5A,B), MβBL−MβBL homosequences and MβBL-BL heterosequences can be assigned to the peaks at chemical shifts of 6.31 and 6.23 ppm, respectively (Figure 5C−E). The percent mole ratio of heterosequences in different copolymers can be calculated clearly through 1H NMR investigation. As expected, the content of heterosequences decreased with the increasing racMβBL feed ratio (Table 2), which matches well with the kinetic study results. Degradation Studies of PMβBL and P(MβBL-ran-BL)S. Surprisingly, we found PMβBL homopolymer degraded rapidly in the presence of Lewis base. As shown in Table 3, trimethylamine (TEA) was employed to perform the degradation reactions of PMβBL and P(MβBL-ran-BL). The molecular weight of PHB homopolymer decreased slightly after 12 h in the presence of 0.1 equiv of TEA at 80 °C, while PMβBL degraded dramatically under the same condition. This implies that PMβBL has better degradability than PHB (Table 3, entries 1 and 2). The molecular weight distribution

using complex 4/BnOH catalytic system. In order to understand the copolymerization process, kinetic study was carried out by sampling at various time points. As shown in Table S2, five groups of feed ratios were investigated. In all reactions, the conversion of monomers was below 20% to optimize the determination of copolymerization parameters. The conversion and composition in the resultant copolyesters were determined by 1H NMR spectroscopy (using 1,2,4,5tetramethylbenzene as an internal standard; see Figure S2). Figure 4 shows the copolymerization kinetic study initiated by the Salen−Al complex 4/BnOH system; the copolymerization reactivity ratio of two monomers could be fitted from the slope and intercept of the straight line (rMβBL = 0.992 and rBL = 0.923). These results suggest rac-MβBL and rac-BL have equal probability to insert into the polymer main chain. The resultant copolymer exhibits practically random distributions of the two monomers. Moreover, by changing the reaction time, the consumption of each monomer and formation of copolymers were monitored. When the feed ratio was fixed (rac-MβBL:racBL:complex 4 = 50:50:1), the consumption rates of both monomers remained nearly equal with respect to the reaction time (Table S3 and Figure S3). A series of random copolymers with different rac-MβBL contents have been produced by changing the feed ratio of two monomers (Table 2). Molecular weight of the random copolymers is close to the theoretical value and the molecular weight distribution remains narrow, suggesting a controllable polymerization process. All the random copolymers are 3137

DOI: 10.1021/acs.macromol.7b00239 Macromolecules 2017, 50, 3131−3142

Article

Macromolecules Scheme 3. Proposed Mechanism of Amine-Mediated Degradation

Scheme 4. Postpolymerization of P(MβBL-ran-BL) via Thiol−Ene Click Chemistry and Radical-Initiated Cross-Linking Reaction

broadened significantly indicates degradation occurred through random chain scission. Representative GPC profiles are shown in Figure S4. Degradation rate decreased with the reaction temperature (Table 3, entry 3). A catalytic amount of TEA can also trigger the degradation by prolonging the reaction time (Table 3, entry 4). Taking the advantage of excellent degradability of MβBL unit, random copolymers with different MβBL unit content have tunable degradability. Degradability increased with the increase of MβBL unit content (Table 3, entries 5−7). Degradation of copolymers occurred even at room temperature, which proved their good degradability (Table 3, entries 8 and 9). In order to get more information on the degradation products, diisopropylethylamine (DIPEA) was used as a nucleophilic reagent instead of TEA. Molecular weight of the resultant degradation products was higher than that of the use of TEA. This might ascribe to the steric effect of the bulky isopropyl substituents of DIPEA (Table 3, entries 10−12). The degradation products of PMβBLs were investigated by electrospray ionization mass spectrometry (ESI-MS). As shown in Figure 6, ESI-MS results indicated the degradation products were mainly inner salts with several repeating MβBL units. The chain ends on both sides are quaternary ammonium and negative oxygen, respectively. Besides the major product quaternary ammonium salts, another minor products were confirmed as cyclic oligomers. Thus, a possible degradation mechanism is proposed in Scheme 3. The quaternary ammonium salts were formed by random nucleophilic attack of organic amine at the carbonyl group. Cyclic oligomers generated through the backbiting reaction of inner salts with negative oxygen. Functionalization of Random Copolymers. The vinylidene functional groups at the α-position provide enormous possibilities of functionalization by radical-initiated click chemistry and cross-linking reactions (Scheme 4). As shown in Table 4, the first route for postpolymerization was intra- or inter-cross-linking reactions initiated by azodiisobutyronitrile (AIBN) (entries 1−3). Under the same reaction condition, functionalization of P9 gave a complete conversion of vinylidene groups. The resulting cross-linking copolymer was designated as CP9, having a higher Tg than that of the parent polymer (Figure 7). The cross-linking points should play an

Table 4. Summary of Postpolymerization of P(MβBL-ranBL)S via Radical-Initiated Cross-Linking Reaction (Entries 1−3) and Thiol−Ene Click Chemistry (Entry 4) entrya

parent polymer

Tgb (°C)

resultant polymer

vinylidene remainingc (%)

Tgd (°C)

ΔTge (°C)

1 2 3f 4g

P9 P27 P100 P27

1.9 5.2 15.6 5.2

CP9 CP27 CP100 TP27