Zhanxin Jing*,
Jin Li,
Weiyu Xiao,
Hefeng Xu,
Pengzhi Hong and
Yong Li*
Department of Applied Chemistry, College of Chemistry and Environment, Guangdong Ocean University, Zhanjiang, Guangdong 524088, China. E-mail: jingzhan_xin@126.com; yongli6808@126.com
First published on 20th August 2019
In this study, we investigated the blending of poly(L-lactide) (PLLA) with supramolecular polymers based on poly(D-lactide)–poly(ε-caprolactone-co-δ-valerolactone)–poly(D-lactide) (PDLA–PCVL–PDLA) triblock copolymers as an efficient way to modify PLLA. The supramolecular polymers (SMP) were synthesized by the terminal functionalization of the PDLA–PCVL–PDLA copolymers with 2-ureido-4[1H]-pyrimidinone (UPy). The structure, thermal properties and rheological behavior of the synthesized supramolecular polymers were studied; we found that the formation of the UPy dimers expanded the molecular chain of the polymer and the incorporation of the UPy groups suppressed the crystallization of polymers. In addition, the synthesized supramolecular polymers had a low glass transition temperature of about −50 °C, showing the characteristics of elastomers. On this basis, superior properties such as a fast crystallization rate, high melt strength, and toughness of fully bio-based, i.e., PLA-based materials were achieved simultaneously by blending PLLA with the synthesized supramolecular polymers. In the PLLA/SMP blends, PLLA could form a stereocomplex with its enantiomeric PDLA blocks of supramolecular polymers, and the stereocomplex crystals with the cross-linking networks reinforced the melt strength of the PLLA/SMP blends. The influences of the SMP composition and the SMP content in the PLLA matrix on crystallization and mechanical properties were analyzed. The supramolecular polymers SMP0.49 and SMP1.04 showed a reverse effect on the crystallization of PLLA. Tensile tests revealed that the lower content of the synthesized supramolecular polymers could achieve toughening of the PLLA matrix. Therefore, the introduction of supramolecular polymers based on PDLA–PCVL–PDLA is an effective way to control the crystallization, rheology and mechanical properties of PLLA.
The stereocomplex technology is considered to be one of the most effective methods to improve the crystallization rate and heat resistance of PLA, which are important issues in expanding the applications of poly(lactide).20 Poly(L-lactide) (PLLA) and its enantiomer poly(D-lactide) (PDLA) have been reported to show a strong tendency to interact with each other to form stereocomplexes via the strong van der Waals forces of the CH3⋯CO interactions between enantiomeric molecular chains.21,22 The melting temperature of PLA-based stereocomplexes (sc-PLA) can reach 230 °C, which is higher than that of PLLA or PDLA (about 170 °C). In addition, sc-PLA also has a high crystallization rate and heat resistance compared to PLLA or PDLA. These are mainly attributed to the β-crystal of sc-PLA, which is different from the α-crystal of homo-crystallized PLA.23 To understand the mechanism of the stereocomplexation phenomenon, the effects of many parameters such as molecular weight, L/D ratios, optical purity, preparation method and crystallization conditions on the formation of stereocomplexes have been deeply studied.24–28 Recently, PLA stereocomplexes have been widely used to modify PLA. Our previous studies have reported29,30 that the stereocomplex interface formed between the PLLA matrix and nanoparticles not only promotes the dispersion of nanoparticles, but also increases the melt strength of the polymer matrix. This extensive research has demonstrated that a stereocomplex is an effective nucleating agent and a rheological modifier for PLLA, and it can also improve the thermodynamic properties of PLLA.31,32 However, the stereocomplex technology promotes the crystallization of PLA, which tends to make PLA more brittle. Therefore, this is also a problem that must be considered when the stereocomplex technology is applied to modify PLA.
Recently, scientists have made many efforts to improve the toughness, crystallinity and melt strength of PLA at the same time. Liu et al.33 prepared PLLA/PDLA–PEG–PDLA blends; they found that the toughness and heat resistance of PLLA are improved due to the synergistic effects of stereocomplexation between enantiomeric PLA and plasticization of PEG blocks. Poly(D-lactide)–Pluronic F68–poly(D-lactide) multiblock copolymers were synthesized by Qi,34 who found that the synthesized multiblock copolymers can act as a tough agent of PLLA and PLLA/multiblock copolymer blends have a continuous amorphous phase with the crystalline regions being the discontinuous portion. In this study, we designed and synthesized supramolecular polymers containing PDLA hard segments and poly(ε-caprolactone-co-δ-valerolactone) (PCVL) soft segments via UPy-functionalized PDLA–PCVL–PDLA copolymers. On this basis, we prepared PLLA/supramolecular polymer blends and investigated their thermal properties, rheological behavior and mechanical properties. This work provides a new approach to prepare PLA-based materials with rapid crystallization capacity, high melt strength and excellent mechanical properties.
Then poly(D-lactide)–poly(ε-caprolactone-co-δ-valerolactone)–poly(D-lactide) triblock copolymers (PDLA–PCVL–PDLA) were synthesized. PCVL and D-lactide were added into a Schlenk flask, and the mixture was dried in a vacuum line at 60 °C for 30 min. After being purged with dry nitrogen, the mixture was heated to 130 °C and stirred. Then, Sn(Oct)2/toluene solution was added. The dosage of Sn(Oct)2 was 0.1 mol% with respect to the amount of D-lactide. The polymerization reaction was allowed to occur at 130 °C for 24 h. After the reaction, the obtained mixture was dissolved in dichloromethane and precipitated by adding excess of methanol. The obtained precipitate was dried in a vacuum oven at 30 °C for 48 h, and it was abbreviated as PDLA–PCVL–PDLAx, where x represents the molecular weight of PDLA block calculated using the 1H NMR spectrum.
Eventually, a supramolecular polymer based on PDLA–PCVL–PDLA was synthesized. A PDLA–PCVL–PDLA triblock copolymer and UPy-NCO were added into a dried Schlenk flask. Then, the mixture was further dried in a vacuum line at 60 °C for 30 min. After purging with nitrogen, toluene (50 mL) and Sn(Oct)2 (the dosage of Sn(Oct)2 was 3.0 wt% with respect to the amount of the PDLA–PCVL–PDLA triblock copolymer) were added, and the mixture solution was heated to 110 °C and kept at this temperature for 24 h. Then, toluene was removed by rotary evaporation under reduced pressure. The obtained crude product was dissolved in chloroform (100 mL), and silica and dibutyltin dilaurate were added to the above solution. The mixed solution was stirred at 60 °C for 2 h. The UPy-NCO-absorbed silica was removed through filtration. Chloroform in the obtained transparent solution was eliminated by rotary evaporation under reduced pressure, and the obtained supramolecular polymer was dried in a vacuum oven for 24 h. The synthesized supramolecular polymer was abbreviated as SMPx, where x represents the molecular weight of PDLA block.
Fig. 1 displays the 1H NMR spectra of the dihydroxyl-terminated PCVL macroinitiator, the PDLA–PCVL–PDLA triblock copolymer and the corresponding supramolecular polymer. In the 1H NMR spectrum of the PCVL macroinitiator, there are obvious proton peaks at 1.41 ppm, 1.59 ppm, 1.68 ppm, 2.33 ppm and 4.09 ppm, which are attributed to the methylene groups at different positions on the PCVL macromolecular chain. In addition, a weak proton peak is observed at 3.7 ppm, and it can be assigned to the methylene proton linked to the terminal hydroxyl group of PCVL.38 The number-average molecular weight of PCVL can be calculated by NMR, and the obtained result is 6.31 kg mol−1. The GPC curve of PCVL was also measured, as shown in Fig. 2, and it exhibited a single elution peak. The number-average molecular weight and polydispersity index were 7.48 kg mol−1 and 1.38, respectively. For the PDLA–PCVL–PDLA triblock copolymers, proton peaks appeared at 1.61 ppm, 5.18 ppm and 4.37 ppm with respect to PCVL, corresponding to the methyl, methine, and terminal methine protons of the PDLA blocks.36,38 The molecular weight of the PDLA–PCVL–PDLA triblock copolymers was measured by GPC and also calculated by NMR; the results are listed in Table 1. The synthesized PDLA–PCVL–PDLA triblock copolymers displayed molecular weight distribution with lower polydispersity indexes (<1.35), and the molecular weight increased as the D-lactide/PCVL feed ratio increased. These results indicated that the dihydroxyl-terminated PDLA–PCVL–PDLA triblock copolymers with a controlled molecular weight and composition were synthesized. The supramolecular polymer was synthesized by reacting the PDLA–PCVL–PDLA triblock copolymer and excess UPy-NCO. The synthesized supramolecular polymer shows weak peaks at 10.2, 11.9 and 13.1 ppm (Fig. 1(c)), which are the characteristic N–H proton peaks of the UPy motif.36,37,39 In addition, the weak peak (∼4.37 ppm) corresponding to the PDLA terminal methine protons disappeared completely. The GPC curve of the supramolecular polymer was also measured (Fig. 2), revealing that there appears a shoulder at the higher molecular weight side and larger PDI (about 2.0). These results demonstrated that the UPy groups were successfully introduced into the end of PDLA–PCVL–PDLA, and UPy dimerization could expand the polymer chain and increase the molecular weight. Therefore, we successfully synthesized the supramolecular polymer based on the PDLA–PCVL–PDLA triblock copolymer.
Fig. 1 1H NMR spectra of PCVL (a), PDLA–PCVL–PDLA triblock copolymer (b) and the corresponding supramolecular polymer (c). |
Fig. 2 GPC curves of PCVL, PDLA–PCVL–PDLA0.49 triblock copolymer and the corresponding supramolecular polymer (SMP0.49). |
Samples | [LA]/([CL] + [VL])a | Mn of each blockc (kg mol−1) | Mn,NMRb (kg mol−1) | Mn,GPCd (kg mol−1) | PDId | mPLAe (%) | Yield (%) |
---|---|---|---|---|---|---|---|
a [LA]/([CL] + [VL]) represents the molar ratio of lactide and the units of ε-caprolactone and δ-valerolactone of the PCVL macroinitiator during the synthesis of PLA–PCVL–PLA triblock copolymers.b Mn acquired from 1H NMR results.c The numerals denote Mn of the corresponding PLA and PCVL blocks, as derived from NMR data.d Mn and polydispersity index (PDI) acquired by GPC.e Mass fractions of PLA in the triblock copolymers calculated from 1H NMR spectroscopy. | |||||||
PCVL | — | — | 6.31 | 7.48 | 1.38 | — | 79.4 |
PDLA–PCVL–PDLA0.49 | 1:4 | 0.49–6.31–0.49 | 7.29 | 8.30 | 1.35 | 13.4 | 83.2 |
SMP0.49 | — | — | — | 23.0 | 2.01 | — | 81.4 |
PDLA–PCVL–PDLA1.04 | 2:4 | 1.04–6.31–1.04 | 8.37 | 8.95 | 1.28 | 32.6 | 66.3 |
SMP1.04 | — | — | — | 24.6 | 1.81 | — | 91.3 |
The effects of the UPy groups on the crystallization and thermal properties of the PDLA–PCVL–PDLA copolymers were studied by DSC (Fig. 3), and the obtained thermal parameters are listed in Table S1.† It is clear that the DSC cooling curve of the PDLA–PCVL–PDLA0.49 triblock copolymers shows an obvious exothermic peak at ∼5 °C, which is assigned to the crystallization of the PCVL segments. The corresponding reheating curve shows a melting peak of the PCVL segments at ∼30 °C. In the DSC cooling curve of the PDLA–PCVL–PDLA1.04 triblock copolymers, there appear two exothermic peaks at 7.9 and 72.4 °C, which are attributed to the crystallization of the PCVL and PDLA segments, respectively. It can also be observed from the corresponding reheating curve that there are two endothermic peaks at 15–40 and 100–140 °C, which correspond to the melting of the PCVL and PDLA segments. Compared with the result for the PDLA–PCVL–PDLA copolymer, the crystallization peak of the supramolecular polymer based on the PDLA–PCVL–PDLA triblock copolymer reduces or disappears and the corresponding crystallization temperature decreases, as shown in Fig. 3(b). It is clear that the synthesized supramolecular polymers have a low glass transition temperature (about −50 °C), indicating that the supramolecular polymers based on the PDLA–PCVL–PDLA triblock copolymers have good elasticity and can be used for toughening brittle materials. The glass transition temperature of the synthesized supramolecular polymer is related to the length of the PDLA block, and the increase in the length of the PDLA block leads to the increase in the glass transition temperature. In the reheating curve of the supramolecular polymer SMP0.49, there appears a significantly exothermic peak and an obviously endothermic peak at −12.9 and 24.6 °C, which correspond to the cold crystallization and melting of the PCVL segments. The reheating curve of the supramolecular polymer SMP1.04 shows a weak endothermic peak at 117.3 °C assigned to the melting of the PDLA segment. These phenomena revealed that the incorporation of the UPy groups suppressed the crystallization of the polymer. The possible reason is that the formation of self-complementing quadruple hydrogen bonds increases the molecular weight of the polymer. This can drastically restrain the freedom and mobility of the polymer chains, which would increase the difficulty of the regular alignment of polymer molecular chains during crystallization. This enabled the transition of the PDLA–PCVL–PDLA triblock copolymers from turbid, brittle solids to transparent, elastic solids.
Fig. 3 DSC cooling and heating curves of PDLA–PCVL–PDLA triblock copolymers (a) and the corresponding supramolecular polymers (b). |
Fig. 4 Variation in viscoelastic moduli of supramolecular polymers as a function of temperature: (a) SMP0.49; (b) SMP1.04. |
Fig. 5 shows the variation in the storage modulus (G′) and loss modulus (G′′) of the supramolecular polymers melted at 100 °C as a function of frequency. It is clear that the G′ and G′′ values of the supramolecular polymers increase as the frequency increases, and the G′ values are larger than G′′. For an ideal polymer melt, the variation in the moduli as a function of frequency follows the following equations: logG′ ∝ logω and logG′′ ∝ logω. However, it can be observed from Fig. 5 that the moduli of the synthesized supramolecular polymers deviate significantly from the above equation, and the melt exhibits a solid-like viscoelastic behavior. For neat PCVL, its melting temperature is about 30–40 °C. However, the DSC curves of the supramolecular polymers based on the PDLA–PCVL–PDLA triblock copolymers show the melting peak of PCVL, and the melting peak corresponding to the PDLA block is observed in a high-temperature region (Fig. 3). This reveals that the supramolecular polymer still exists in the PDLA crystalline region after melting at 100 °C, which can act as the crosslinking point to restrain the movement of polymer chains. Therefore, the synthesized supramolecular polymers melted at 100 °C showed a solid-like viscoelastic behavior; the higher the PDLA content, the more pronounced the solid-like viscoelastic behavior.
Fig. 5 Variation in storage modulus (G′) and loss modulus (G′′) of supramolecular polymers as a function of frequency (100 °C, strain = 1%). |
The DSC curves of the PLLA/supramolecular polymer blends are displayed in Fig. 6, and the obtained thermal parameters are listed in Table S2.† In Fig. 6(a), neat PLLA shows a weak crystallization peak at 99.4 °C in the DSC cooling curve. The corresponding heating curve presents an exothermic peak at 112.2 °C and an endothermic peak at 155–175 °C (Fig. 6(b)), corresponding to cold crystallization and melting of PLLA. A weak peak is also detected at 60.0 °C, and it is assigned to the glass transition of PLLA. For the blend PLLA/SMP0.49-10%, the cooling curves show weak crystallization of PLLA at 99.3 °C. With the increase in the SMP0.49 content, the crystallization peak of PLLA disappears. As shown in Fig. 6(b), with the increase in the SMP0.49 content, an endothermic peak is observed in a lower temperature region (about 30 °C), which can be attributed to the melting of the PCVL segments of supramolecular polymers. In addition, the glass transition temperature of the PLLA/SMP0.49 blends decreased as the content of SMP0.49 increased. The synthesized supramolecular polymer SMP0.49 contained a higher PCVL content with a low glass transition temperature. When the supramolecular polymer SMP0.49 was introduced into the PLLA matrix, the PLLA/SMP0.49 blends could present a lower glass transition temperature with respect to that of PLLA. It can be observed in Fig. 6(b) that the PLLA/SMP0.49 blends show a significantly exothermic peak at about 122 °C due to the cold crystallization of PLLA. As listed in Table S2,† the cold crystallization enthalpy of the blend PLLA/SMP0.49-10% SMPs is larger than that of neat PLLA, while this area decreases as the SMP content further increases. With the increase in the SMP0.49 content, the melting enthalpy of PLLA decreased, but a weak melting peak assigned to the PLA stereocomplex was observed at about 182 °C.
Fig. 6 DSC cooling (a and c) and reheating curves (b and d) of poly(L-lactide)/supramolecular polymer blends: (a and b) PLLA/SMP0.49; (c and d) PLLA/SMP1.04. |
In Fig. 6(c), the cooling curves of the PLLA/SMP1.04 blends show two weak endothermic peaks: the peak in the higher temperature region is assigned to the stereocomplex crystallization between PLLA and PDLA blocks of supramolecular polymers; the peak corresponding to the lower temperature region is attributed to the homo-crystallization of PLLA. It can also be found from Fig. 6(c) that the area and temperature of the peak corresponding to stereocomplex crystallization increase as the content of SMP1.04 increases. The supramolecular polymer SMP1.04 exhibited a higher PDLA content; thus, the increase in the SMP content led to an increase in the PDLA content in the PLLA/SMP1.04 blends, which was beneficial for stereocomplex crystallization. The peak corresponding to PLLA homo-crystallization also shifted to the higher temperature region as the content of SMP1.04 increased, and the peak area was enhanced. This is because the formed stereocomplex could act as a heterogeneous nucleating agent for PLLA homo-crystallization.46 In Fig. 6(d), the PLLA/SMP1.04 blends show similar reheating curves to that of the PLLA/SMP1.04 blends. The melting of the PCVL block and cold crystallization of PLLA were observed at about 30 °C and 110 °C, respectively. However, with the increase in the SMP1.04 content, the melting peak of the PCVL block and the cold crystallization peak weakened and eventually disappeared. The PLLA/SMP1.04 blends showed multi-step melting at 150–170 °C, which was assigned to the melting–recrystallization–remelting process of PLLA. It is also clear that a weak melting peak of the stereocomplex is observed at about 190 °C, and this melting area increased as the SMP1.04 content increased. Similar results have been reported for the asymmetric PLLA/PDLA blends and the PCL/PLA alternating multiblock supramolecular polymers.37,48,50 It can be found from Table S2† that the crystallinity of PLLA/SMP0.49 decreased with the content of the supramolecular polymers, while the crystallinity of PLLA/SMP1.04 increased. This may be attributed to the different compositions of supramolecular polymers. For the PCVL block, a low glass transition temperature (about −50 °C) was observed and thus, it could accelerate the mobility of polymer chains. The PDLA block could form a PLA stereocomplex, which could act as the heterogeneous nucleating agent for PLLA crystallization. When a supramolecular polymer with a shorter PDLA block is blended with PLLA, it is difficult to achieve stereocomplex formation. Because the supramolecular polymer shows good movement capacity, the short PDLA block is not beneficial to the alternating arrangement of PLLA and PDLA due to the weak hydrogen bond. For the supramolecular polymer with a longer PDLA block, it is easy to form a stereocomplex with PLLA. This is because the longer PDLA block could form strong hydrogen bonds, accelerating the alternating arrangement of the PLLA and PDLA chains. The formed stereocomplex could act as heterogeneous nucleation sites, and the PCVL block would facilitate the favorable arrangement of PLLA around the nucleation sites. Therefore, the synthesized supramolecular polymers SMP0.49 and SMP1.04 showed the reverse effect on the crystallization of PLLA.
Fig. 7 Variation in storage modulus (G′), loss modulus (G′′), loss tangent (tanδ) and complex viscosity (η*) as a function of frequency for PLLA/SMP0.49 blends at 175 °C. |
The loss tangent (tanδ) is more sensitive to the relaxation behavior of a polymer melt with respect to G′ and G′′. To further analyze the rheological behavior of the PLLA/SMP blends, the variation in the loss tangent as a function of frequency was also measured, as shown in Fig. 7(c). The curves of neat PLLA show an obvious peak (Fig. S3†), exhibiting a typical behavior of viscoelastic liquids. With the increase in the content of SMP, the value corresponding to this peak obviously decreased, revealing a more significant elastic response. Especially for the blends with more than 30% SMP, this peak completely disappeared. This indicated that the melt of the PLLA/SMP blends achieved a transition from liquid-like to solid-like viscoelastic behaviors when the SMP0.49 content exceeded 30%, which was because the long-range polymer chain movement was restrained by the stereocomplex crystals. The complex viscosity (η*) was also used to investigate the rheological behavior of the PLLA/SMP melt, as shown in Fig. 7(d). For neat PLLA, the complex viscosity (η*) exhibited no tolerance to frequency at a lower frequency (<10 rad s−1) (Fig. S4†), showing a typical Newtonian fluid behavior. At a high frequency (>10 rad s−1), the complex viscosity (η*) was reduced due to shear thinning. However, the complex viscosity curves of the PLLA/SMP melt exhibited a significant slope, and the slope increased as the SMP content increased. This indicates that the PLLA/SMP melt exhibits a non-Newtonian fluid behavior. Therefore, there may be a structure similar to the cross-linked network in the melt of the PLLA/SMP blends.
Our previous studies48,50,51 found that the stereocomplex crystal is different from an inorganic particle, and it has a certain tolerance to temperature, causing the effect of the stereocomplex crystal as a rheological modifier to be affected by the thermal treatment temperature. Therefore, the rheological behavior of the PLLA/SMP blends at different thermal treatment temperatures was measured, and the obtained results are displayed in Fig. 8. For the PLLA/SMP0.49-50% blends, the G′ and G′′ values obviously decreased as the thermal treatment temperature increased. This was mainly because the stereocomplex crystals began to melt as the thermal treatment temperature increased, resulting in the stereocomplex crystals becoming smaller. This not only directly affected the stereocomplex crystals existed in the polymer melt, but also destroyed the network structure based on the stereocomplex crystals in the melt. Therefore, the increased thermal treatment temperature resulted in the weakening of the strength of the PLLA/SMP melt. It could be observed that the G′ and G′′ values of the PLLA/SMP1.04-50% blends melted at 185 °C were larger than those of the PLLA/SMP0.49-50% blends. This was attributed to the high stereocomplex content. The supramolecular polymer SMP1.04 contained more D-lactide repeating units with respect to SMP0.49. This not only led to an increase in the content of the stereocomplex crystals, but also made the formed stereocomplex crystals more perfect. The rheological results demonstrated that a similar cross-linked network was formed in the melt of the PLLA/SMP blends, resulting in a transition from liquid-like to solid-like viscoelastic behaviors, and the network existing in the PLLA matrix was closely related to the content of the PDLA blocks and the thermal treatment temperature.
Fig. 8 Variation in storage modulus (G′) and loss modulus (G′′) as a function of frequency for PLLA/SMP blends at different temperatures. |
Samples | Strength at yielding (MPa) | Strength at break (MPa) | Elongation at break (%) |
---|---|---|---|
PLLA | 51.33 ± 4.4 | 32.82 ± 3.1 | 24.1 ± 6.1 |
PLLA/SMP0.49-10% | 22.00 ± 1.02 | 24.28 ± 2.83 | 330.4 ± 35.3 |
PLLA/SMP0.49-30% | 15.49 ± 0.97 | 19.90 ± 1.56 | 372.4 ± 76.5 |
PLLA/SMP0.49-50% | 10.33 ± 2.47 | 9.90 ± 1.45 | 91.4 ± 15.4 |
PLLA/SMP1.04-10% | 33.62 ± 6.23 | 30.02 ± 5.29 | 143.1 ± 17.3 |
PLLA/SMP1.04-30% | 16.70 ± 2.45 | 14.13 ± 1.64 | 23.08 ± 2.54 |
PLLA/SMP1.04-50% | 9.71 ± 0.48 | 9.71 ± 0.48 | 5.38 ± 0.24 |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c9ra04283k |
This journal is © The Royal Society of Chemistry 2019 |