Toughening of amorphous poly(propylene carbonate) by rubbery CO2-based polyurethane: transition from brittle to ductile

Guanjie Renab, Yuyang Miaoa, Lijun Qiaoa, Yusheng Qin*a, Xianhong Wang*a and Fosong Wanga
aKey Laboratory of Polymer Ecomaterials, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun 130022, People's Republic of China. E-mail: ysqin@ciac.ac.cn; xhwang@ciac.ac.cn; Fax: +86 431 85262252, +86 431 85689095; Tel: +86 431 85262252, +86 431 85262250
bUniversity of Chinese Academy of Sciences, Beijing 100039, People's Republic of China

Received 20th April 2015 , Accepted 19th May 2015

First published on 21st May 2015


Abstract

Amorphous poly(propylene carbonate) (PPC) is brittle at room temperature, but the studies related to the toughening of PPC is rare. Herein, two types of polyurethane (PCO2PU) synthesized from a CO2-based diol and toluene diisocyanate were used as rubbery particles to toughen PPC. The notched impact strength of PPC increased from 20.8 J m−1 to 54.2 J m−1 at a PCO2PU loading of 20 wt%, comparable with that of neat nylon 6, and reached 228.3 J m−1 at a PCO2PU loading of 30 wt%, 10.9 fold that of neat PPC and even higher than bisphenol A polycarbonate. Matrix yielding as well as cavitation was observed during the impact process, which was responsible for the increase of impact strength. Moreover, the toughening efficiency was related with the carbonate content of PCO2PU, and the transition of fracture behavior from brittle to ductile occurred when the PCO2PU with a weight average diameter of 0.20 μm was uniformly dispersed in PPC substrate.


1. Introduction

Poly(propylene carbonate) (PPC) is an alternative copolymer of CO2 and propylene oxide that has received considerable attention in the past decades because it is biodegradable and has effective CO2 utilization.1 Now it has found applications not only in high value-added areas like tissue scaffolds2 or polymer electrolytes,3 but also as a low-cost biodegradable packaging material.4 However, PPC is amorphous and brittle with elongation at break below 10% and low notched impact strength of about 20 J m−1 at 20 °C. These values are comparable to polystyrene, which has a notched impact strength of 15.8 J m−1. Generally, the notched impact strengths of high density polyethylene (HDPE) and high impact strength polystyrene (HIPS) are 50–70 J m−1[thin space (1/6-em)]5 and 72–148 J m−1,6 respectively. Therefore, the notched impact strength of PPC is considerably low, which has severely limited its application.

Considerable effort has been made to overcome the brittleness of PPC. Our previous efforts indicate that PPC can be plasticized by diallyl phthalate7 or low molecular weight urethane compounds,8 or partly plasticized and enhanced via hydrogen bonding interaction by introducing a hydro-branched poly(ester amide).9 The elongation at break has been improved to over 700% while a completely miscible blend has been obtained.8 However, currently no report has been related to the notched impact test even though it represents the ability to absorb fracture energy under high loading in a notched state.

The notched impact strengths of general polymers like polyamide10 or polystyrene11 have been investigated for a long time. The introduction of elastomers (with appropriate domain size or interparticle distance) into the matrix can change the stress state around the particles and form microstructures that dissipate impact energy. These microstructures include intensified stable crazing in HIPS11b,11c and large area shear yielding10b,10c that accompany cavitation in polyamides.10a,10d,10e Both the notched impact strength and fracture morphology are significantly related to the rubber particle diameter and volume fraction.12 Increasing the content of small rubber particles favors local plane stress with the consequence of a transition from crazing to shear deformation. For a given polymer blend, maximum toughness can be obtained in a limited range of rubber particle sizes12b where microvoids are formed, thus avoiding premature fracture due to the extension of craze from the notch tip before significant energy has been dissipated, which facilitates the occurrence of shear yielding. For example, in a polyamide/polyolefin elastomer blend,12a energy dissipated by shear yielding was optimal at particle sizes between 0.1 μm and 0.3 μm.

The rubbery CO2-based polyurethane (PCO2PU) (Scheme 1) was synthesized from the chain extending reaction of CO2-based diols with toluene diisocyanate (TDI). CO2-based diols with different molecular weights and carbonate content were prepared in our previous study by the copolymerization of propylene oxide and CO2[thin space (1/6-em)]13 to obtain low molecular weight (number average molecular weight of ca. 1000–1500 g mol−1) polymers with perfect OH functionalities. It is noteworthy that such a CO2-based diol has a 20–30% lower cost than conventional polypropylene glycol (PPG). It may be a competitive raw material for polyurethane formation. Most importantly, the presence of a carbonate unit in both PCO2PU and PPC may enhance the miscibility of the two components. Moreover, the formation of intermolecular hydrogen bonding between N–H in the main chain of PCO2PU and carbonyl group in PPC may prove to be beneficial. Therefore, two types of rubbery PCO2PU with different carbonate content were used to toughen PPC. A sudden increase in notched impact strength compared to the value of neat polyamide (PA6) was observed accompanying the fracture behaviour transition from brittle to ductile, and matrix shear yielding appeared for the domain size of PCO2PU at 0.20 μm.


image file: c5ra07142a-s1.tif
Scheme 1 Synthesis of PCO2PU with CO2-based diol.

2. Experiment

2.1 Materials

Toluene diisocyanate (TDI) and dibutyltin dilaurate were purchased from Tianjin Guangfu Chemical Research Institute (China) and used as received. CO2-based diols with carbonate unit content of 50.4% and 62.3% were prepared in our laboratory according to the literature.13 The number-average molecular weight (Mn) of the CO2-based diol with 50.4% carbonate unit content was 1377 g mol−1. The number-average molecular weight (Mn) of the CO2-based diol with 62.3% carbonate unit content was 1245 g mol−1.

PPC was supplied by Zhejiang Bangfeng Plastic Co. (China), whose technique was licensed under our laboratory. To remove residue rare earth metal ternary catalyst, the copolymer was purified twice by a repeated dissolution/precipitation procedure with dimethyl carbonate as solvent and ethanol as a precipitant. The number-average molecular weight (Mn) and the polydispersity index (PDI) of the purified PPC were determined by gel permeation chromatography (GPC) to be 17.3 × 104 g mol−1 and 3.65, respectively. The carbonate unit content of the purified PPC was 92%, as estimated from its 1H NMR spectrum according to the literature.14

2.2 Synthesis of PCO2PUs

PCO2PU was synthesized by the polyaddition of CO2-based diols and TDI under N2 protection. Briefly, CO2-based diols (0.035 mol) were dried at 80 °C under vacuum for 40 min to complete dehydration. TDI (0.035 mol) and dibutyltin dilaurate (3.0 μL) were injected into the reaction vessel and stirred at 110 °C for 30 min. The resultant PCO2PUs from CO2-based diol with 50.3% and 62.4% carbonate unit were denoted as PCO2PU1 and PCO2PU2, respectively. The 1H NMR (d6-CHCl3, TMS, 300 MHz) data were listed as follows.
PCO2PU1. δ (ppm) = 7.61, 7.21, 7.07 (–ArH), 5.17–4.76 (–CH2CH(CH3)OCOO–), 3.97–4.34 (–CH2CH(CH3)OCOO–), 3.25–3.81 (–CH2CH(CH3)O–), 2.30 (–CH2COO–), 2.20 (–ArCH3), 1.29 (–CH2CH(CH3)OCOO–, –CH2CH2CH2CH2COO–), 1.14 (–CH2CH(CH3)O–).
PCO2PU2. δ (ppm) = 7.67, 7.20, 7.06 (–ArH), 5.20–4.78 (–CH2CH(CH3)OCOO–), 3.93–4.36 (–CH2CH(CH3)OCOO–), 3.30–3.88 (–CH2CH(CH3)O–), 2.30 (–CH2COO–), 2.20 (–ArCH3), 1.28 (–CH2CH(CH3)OCOO–, –CH2CH2CH2CH2COO–), 1.16 (–CH2CH(CH3)O–).

2.3 Melt-blending procedure

Prior to the blending, PPC and PCO2PUs were dried at 45 °C in vacuum for 24 h. Then, PPC and PCO2PU were mixed in calculated weight ratio (PCO2PU/PPC = 2.5/97.5, 5/95, 10/90, 15/85, 20/80, 30/70). The mixing was operated on a Haake batch-intensive mixer (Haake Rheomix 600) at a speed of 60 rpm for 5 min at 140 °C. All the blends were placed in a desiccator before use.

2.4 Characterization

Tensile performance was evaluated using a dumbbell-shaped sample punched out from the molded sheet in a screw-driven universal testing machine (Z010, Zwick Co., Germany) equipped with a 10 kN electronic load cell and mechanical grips. The test was conducted at 20 °C using a cross-head rate of 20 mm min−1 according to the ASTM standard, and the data reported were the mean of the parallel values in five determinations.

The Izod notched impact strength of the specimens was measured on a JJ-20 instrumented impact machine at 20 °C according to the ASTM D256-04. At least five specimens were tested for each sample to obtain an average value.

Scanning electron microscopy (SEM) experiments were performed using a XL30 ESEM FEG (FEI Co.) instrument with an acceleration voltage of 8 kV. The fracture surfaces of Izod notched impact test were coated with gold to increase the contrast.

Differential scanning calorimetry (DSC) analysis was performed on a Perkin-Elmer DSC-7 instrument under N2 atmosphere. The sample was first heated from −50 °C to 50 °C at 10 °C min−1 and then rapidly quenched to −50 °C, followed by a second heating process to obtain the glass transition temperature (Tg).

Rheological measurements were performed at 140 °C using an AR 2000 (TA, USA) rheometer with a parallel plate geometry, and the diameters of plates was 25 mm. Dynamic amplitude sweeps from 0.1% to 100% strain at 1 rad s−1 frequency were executed to determine the linear viscoelasticity range. Then, dynamic frequency sweeps were performed from 0.01 to 100 rad s−1 at a small strain of 1% within the linear viscoelastic zone.

The TEM images were recorded on a transmission electron microscope (Tecnai G2 F20 S-TWIN) with an acceleration voltage of 200 kV. The samples for TEM analysis were prepared by microtoming films with 50–100 nm thickness from the blends with an ultramicrotome (LEICA ULTR CUTR ME1-057) and stained with 1% aqueous phosphotungstic acid. The rubber particle size was calculated by TEM using Nanomeasure software. The number and weight average rubber particle sizes were calculated using the following equations.

image file: c5ra07142a-t1.tif

image file: c5ra07142a-t2.tif

3. Results and discussion

3.1 Tensile properties and impact strength

The stress–strain curves of PPC/PCO2PU blends are plotted in Fig. 1, and the corresponding parameters are listed in Table 1. Neat PPC displayed brittle fracture with elongation at break of 9.83%. This value increased to 28.74% when the loading of PCO2PU1 reached 5 wt%. It was noteworthy that both the tensile strength and the modulus of the blend with 5 wt% PCO2PU1 increased by 10.21 MPa and 89.48 MPa, respectively compared with those of neat PPC. Further increasing the PCO2PU1 loading to 20 wt% increased the elongation at break to 49.83%, whereas the tensile strength remained at 47.66 MPa. When the loading of PCO2PU1 was increased to 30 wt%, the elongation at break reached 320.76% with the tensile strength decreasing to 27.36 MPa. Similar toughening effect can be observed in PPC/PCO2PU2 blends. When the PCO2PU2 loading was 20 wt%, the elongation at break increased to 33.06% while the tensile strength was still 4.32 MPa higher than that of the neat PPC. Therefore, a simultaneous increase of tensile strength and elongation at break was realized in this polyblending system, especially at relatively low PCO2PU loading. This may be attributed to the intermolecular hydrogen bonding between the C[double bond, length as m-dash]O of PPC and the NH of PCO2PU, as reported for the PLA/hyperbranched polyamide blend15 and PPC/low-molecular weight urethane blend.8 It is common that the addition of rubbery particles toughens the matrix while decreasing the tensile strength. However, the existence of hydrogen bonds can enhance interfacial adhesion and improve tensile strength.9 Therefore, the tensile strength of the blend increased with PCO2PU1 loading below 15 wt%. When the PCO2PU1 loading was above 15 wt%, the steric hindrance and dilution effect became dominant, suppressing the reinforcing effect of PCO2PU. Compared with PCO2PU1, PCO2PU2, with a higher carbonate content, showed a more significant reinforcing effect, which may be related to the increased content of hydrogen bonded C[double bond, length as m-dash]O groups in PPC. Moreover, it appears that PCO2PU1 is more effective than PCO2PU2 in toughening PPC. Therefore, an impact test was conducted to confirm the toughening effect.
image file: c5ra07142a-f1.tif
Fig. 1 Stress–strain curves of PPC/PCO2PU1 (a) and PPC/PCO2PU2 (b).
Table 1 Main mechanical properties of PPC/PCO2PU with various PCO2PU loading
Sample PCO2PU ratio (wt%) Young's modulus (MPa) Tensile strength (MPa) Elongation at break (%)
PPC/PCO2PU1 0 1392.91 ± 60.84 53.47 ± 2.48 9.83 ± 5.28
PPC/PCO2PU1 2.5 1407.07 ± 51.74 59.83 ± 3.13 24.60 ± 9.32
PPC/PCO2PU1 5 1482.39 ± 67.85 63.68 ± 3.31 28.74 ± 5.07
PPC/PCO2PU1 10 1420.6 ± 134.29 60.69 ± 4.85 28.71 ± 4.17
PPC/PCO2PU1 15 1144.66 ± 79.13 49.01 ± 3.04 21.02 ± 5.28
PPC/PCO2PU1 20 1086.76 ± 126.49 47.66 ± 1.90 49.48 ± 7.53
PPC/PCO2PU1 30 680.05 ± 109.49 27.36 ± 1.40 320.76 ± 51.44
PPC/PCO2PU2 2.5 1424.79 ± 72.52 61.99 ± 3.39 21.97 ± 4.96
PPC/PCO2PU2 5 1449.79 ± 67.56 63.28 ± 2.89 22.37 ± 4.96
PPC/PCO2PU2 10 1474.11 ± 25.50 66.23 ± 1.79 20.22 ± 4.59
PPC/PCO2PU2 15 1323.54 ± 48.24 63.78 ± 2.72 33.74 ± 6.31
PPC/PCO2PU2 20 1164.90 ± 40.76 57.79 ± 2.96 33.06 ± 5.15
PPC/PCO2PU2 30 730.59 ± 109.01 28.03 ± 1.44 233.01 ± 49.70


The notched impact strengths of PPC/PCO2PU blends are displayed in Table 2 and Fig. 2. When the PCO2PU1 loading was below 20 wt%, the notched impact strength increased very little. It increased suddenly by 33.4 J m−1 from 20.8 J m−1 to 54.2 J m−1 when PCO2PU1 loading was increased to 20 wt%. Further increasing the PCO2PU1 loading to 30 wt% increased the notched impact strength to 228.3 J m−1. PCO2PU2 showed nearly no toughening effect in the range of compositions studied. PCO2PU1 was more effective in toughening PPC, which was more obvious in the notched impact test.

Table 2 Notched impact strengths of PPC/PCO2PUs with various PCO2PU loadings
Sample 0 2.5 wt% 5 wt% 10 wt% 15 wt% 20 wt% 30 wt%
PPC/PCO2PU1 (J m−1) 20.8 ± 1.4 24.7 ± 3.2 24.2 ± 2.9 22.2 ± 3.2 20.5 ± 1.3 54.2 ± 5.1 228.3 ± 46.3
PPC/PCO2PU2 (J m−1) 20.8 ± 1.4 21.3 ± 1.6 22.7 ± 2.1 24.7 ± 3.1 20.8 ± 3.1 22.8 ± 2.1 25.5 ± 8.6



image file: c5ra07142a-f2.tif
Fig. 2 Notched impact strengths of PPC/PCO2PUs with various PCO2PU loadings.

It is reasonable that the elongation at break of PPC/PCO2PU2 blends increased significantly, while the notched impact strength was improved little because notched impact strength was more accurate. As far as the PPC/PCO2PU1 blend was concerned, PCO2PU1 effectively toughened PPC at 20 wt% loading and the notched impact strength was comparable with that of neat PA6.16 When loading was increased to 30 wt%, although the tensile strength decreased to 27.36 MPa, the notched impact strength increased to 10.9 fold that of neat PPC, which was even higher than the bisphenol A polycarbonate.17

The improved notched impact strength may indicate the occurrence of a brittle–ductile transition. This transition may be accompanied by the formation of microstructures such as shear yielding, crazes and microvoids. To confirm the microstructure formed during the impact test and understand the toughening mechanism, the fracture surfaces of neat PPC and PPC/PCO2PU blends under 20 wt% and 30 wt% loading were studied.

3.2 Toughening mechanism

SEM images obtained on the fracture surface of PPC and the polyblends are shown in Fig. 3. In Fig. 3a, the fracture surface of neat PPC had a typical brittle appearance; it was relatively smooth with elliptical marks, indicating that the secondary-crack-front velocity was bigger than that of main-crack-front.18 When PPC was blended with 20 wt% PCO2PU1, the corresponding fracture surface showed no elliptical mark and became coarser while matrix shear yielding appeared (Fig. 3c). Moreover, a few cavitations could be found in the magnified image (Fig. 3d). At 30 wt% PCO2PU1 loading, a large area of shear yielding with cavitation was observed and the fracture surface showed a typical ductile appearance (Fig. 3e). Compared with PCO2PU1, the addition of 20 wt% and 30 wt% PCO2PU2 did not considerably change the fracture surface of PPC. No cavitation and massive shear yielding can be observed, corresponding to the limited increase in the impact strength of PPC/PCO2PU2 (Fig. 3g and i).
image file: c5ra07142a-f3.tif
Fig. 3 SEM images of the fracture surface of PPC/PCO2PU impact sample: (a and b) PPC, (c and d) PPC/PCO2PU1 (80/20), (e and f) PPC/PCO2PU1 (80/30), (g and h) PPC/PCO2PU2 (80/20), (i and j) PPC/PCO2PU2 (80/30).

It has been reported that for polyblends showing massive shear yielding in the matrix upon impacting, most energy will be dissipated from matrix yielding.19 Though the formation of microvoids is the secondary factor contributing to toughening, it is necessary in the formation of massive shear yielding. The cavitation reduces the critical stress at the onset of shear yielding to avoid obtaining the critical strain energy release rate first.12b In the PPC/PCO2PU1 blends with 20 wt% PCO2PU1, cavitation as well as shear yielding was observed, indicating that a transition from brittle to ductile occurred and the impact strength was increased. However, the absence of massive cavitation determined that shear yielding was not as massive as that observed in PLA blend20 or in PP blend.21 At 30 wt% PCO2PU1 loading, the cavitation as well as the shear yielding was expanded to the whole range and the impact strength was improved largely. Therefore, the cavitation and shear yielding resulted in the increase of notched impact strength. In addition, the occurrence of cavitation and shear yielding in the PPC/PCO2PU1 blends at the loading of 20 wt% may be related to the morphology of the blend. Therefore, the miscibility and morphology were studied.

3.3 DSC analysis of various PPC/PCO2PU blends

To study the miscibility between PPC and two types of PCO2PU, the DSC curves of PPC/PCO2PU1 and PPC/PCO2PU2 with various compositions were recorded in Fig. 4. There was only a single glass transition temperature at the low loading of PCO2PU for all the blends, suggesting good miscibility between PPC and PCO2PU at the loading below 5 wt%. However, another glass transition temperature appeared when PCO2PU loading was above 5 wt%, indicating the occurrence of obvious phase separation. The glass transition temperature of the miscible blend was between that of neat PPC and PU, which was a little higher than the value calculated by the Fox equation. In immiscible blends composed of PPC and PCO2PU1, the two glass transition temperatures, displayed in all the blends, were also between those of the two components due to the diffusion of a mutually soluble Tg-reducing component from one phase to another. But interestingly, the lower Tgs in the PPC/PCO2PU2 blends, corresponding to rubbery particle rich phase, were lower than that of the PCO2PU2. As reported in the PS/PE blends, the phenomenon may be due to the interactions between the two components.22
image file: c5ra07142a-f4.tif
Fig. 4 DSC curves of PPC/PCO2PUs with various PCO2PU loadings: (a) PPC/PCO2PU1 and (b) PPC/PCO2PU2.

3.4 Rheology analysis of various PPC/PCO2PU blends

To confirm the miscibility of PPC and PCO2PUs, rheology analysis has been conducted. Fig. S1 and S2 show the dynamic frequency sweeps of PPC and PPC/PCO2PU blends with various compositions. All the blends exhibited a decrease in viscosity with an increase in frequency, indicating an obvious shear thinning behavior and pseudoplastic characteristic. Moreover, the addition of PCO2PU decreased the viscosity of PPC, which may favor processability.

Due to the sensitivity of dynamic rheology to changes in morphology, it is generally used to study the miscibility of blends. For homopolymer or miscible blends, when the frequency is small enough, the slope of log[thin space (1/6-em)]Gvs. log[thin space (1/6-em)]ω is 2 and the slope of log[thin space (1/6-em)]G′′ vs. log[thin space (1/6-em)]ω is 1.23 When phase separation occurs, the slope becomes smaller due to the contribution of the elasticity of the interface and the time–temperature superposition fails. The plots of log[thin space (1/6-em)]Gvs. log[thin space (1/6-em)]ω for the PPC/PCO2PU blends with various compositions are displayed in Fig. 5. As the content of PCO2PU increased, the slope of log[thin space (1/6-em)]Gvs. log[thin space (1/6-em)]ω decreased gradually. When the loadings of two types of PCO2PUs were 5 wt%, the slope was very close to 2, indicating the miscibility of the blends. With the loading increased to 10 wt%, the slope decreased to 1.78 and 1.68 for PPC/PCO2PU1 and PPC/PCO2PU2, respectively. The slope further decreased when PCO2PU loading was 30 wt%. Therefore, the blends became immiscible with PCO2PUs loading above 10 wt%, which was in accordance with the DSC analysis. The different toughing effects of PPC/PCO2PU blends may be related with phase morphology.


image file: c5ra07142a-f5.tif
Fig. 5 G′ of PPC/PCO2PUs with various compositions at 1% strains. (a) 5 wt%, (b) 10 wt%, (c) 20 wt%, (d) 30 wt%.

3.5 Phase morphology

Phase morphology is of vital importance for the toughening effect. The size,24 distribution, and even the shape of the distribution phase25 affect the microstructure formed during the impact process and thus influence the value of impact strength. For cavitation observed initially at 20 wt% PCO2PU1 loading, the phase morphologies of PPC/PCO2PU blends under 20 wt% PCO2PU loadings were studied and the diameters of the dispersed phase are displayed in Table S1 and Fig. 6. As shown in Fig. 6, PCO2PU was dispersed uniformly as spherical particles in the PPC matrix in all the blends. When the content of PCO2PU1 increased from 15 wt% to 20 wt%, the weight average diameter increased from 0.09 μm to 0.20 μm. For the blend with PCO2PU2 content of 20 wt%, the weight average diameter of the dispersed phase was 0.11 μm, which indicated that PCO2PU2 had better miscibility with PPC because of larger carbonate content. However, it appeared that the impact strength could be increased when the diameter of the dispersed phase was as large as 0.20 μm. When the dispersed phase had a diameter around 0.10 μm, there was nearly no contribution to the impact strength. A similar phenomenon was observed in PLA blends, where optimum particle size for increasing toughness was between 0.5 and 0.9 μm.26 In the PLA/P(CL-co-LA) blend, the observed particle size was 0.70 μm, and the impact strength increased to 2 times of neat PLA.25a According to the theory developed by Paul,12b the shear yielding of the matrix occurs when the stress reaches the critical value, which is influenced by the cavitation of the dispersed phase. When the diameter of the rubber phase is too small to form the microvoids, the critical stress at the onset of shear yielding is so large that the strain energy release rate is sufficient to initiate crazing and crack growth in the plane strain region. As a result, the development of shear yielding is limited and fracture occurs first. It is noteworthy that the resistance of cavitation is related to the diameter of rubber phase, i.e., the critical volume strain at cavitation increases as the diameter decreases. Therefore, there is a critical diameter beyond which the cavitation occurs before shear yielding. The cavitation favors the shear yielding by reducing the critical value at which shear yielding occurs. As a result, energy dissipation is promoted and the impact strength increases.
image file: c5ra07142a-f6.tif
Fig. 6 Phase morphologies of PPC/PCO2PUs with various compositions: (A) PPC/PCO2PU1 (85/15), (B) PPC/PCO2PU1 (80/20), (C) PPC/PCO2PU2 (80/20).

In the PPC/PCO2PU blends, the diameter of the dispersed phase is believed to be a determining factor. The cavitation was found in the PPC/PCO2PU1 (80/20) blend, in which the diameter of the dispersed phase reached 0.20 μm, and shear yielding occurred with energy dissipation during the process of impact. However, in other blends, the small diameter of the dispersed phase led to the absence of microvoids, and unstable crazes propagated before the shear yielding occurred. As a result, the notched impact strength increased little. The smaller toughening efficiency of PCO2PU2 can be efficiently explained by the smaller diameter of dispersed phase, which results from better miscibility of PCO2PU2 with PPC; this is unfavorable for the occurrence of cavitation.

4. Conclusion

Polyurethane containing carbonate units was synthesized and used to toughen PPC taking advantage of their potential miscibility. Both the elongation at break and the tensile strength were improved and simultaneous reinforcement and toughening were realized. The notched impact strength of PPC was improved to 228.3 J m−1 for the first time, which was comparable to traditional bisphenol A polycarbonate. It was found that the increase in impact strength was related to the diameter of dispersed phase. When the diameter of dispersed phase reached 0.20 μm, cavitation and matrix yielding occurred; this led to the dissipation of energy and ductile fracture. Our investigations were helpful to understand the relationship between morphology and toughness in addition to determining the range of efficient diameters of the rubber phase for toughening PPC.

Acknowledgements

The work was financially supported by the National Natural Science Foundation of China (Grant no. 51321062 and no. 21134002).

Notes and references

  1. (a) D. J. Darensbourg, Chem. Rev., 2007, 107, 2388 CrossRef CAS PubMed; (b) S. Klaus, M. W. Lehenmeier, C. E. Anderson and B. Rieger, Coord. Chem. Rev., 2011, 255, 1460 CrossRef CAS PubMed; (c) T. Sakakura, J.-C. Choi and H. Yasuda, Chem. Rev., 2007, 107, 2365 CrossRef CAS PubMed.
  2. (a) N. Nagiah, G. Ramanathan, T. S. Uma, L. Madhavi, R. Anitha and T. S. Natarajan, Adv. Polym. Technol., 2013, 32, 21370 CrossRef PubMed; (b) A. Welle, M. Kroeger, M. Doering, K. Niederer, E. Pindel and I. S. Chronakis, Biomaterials, 2007, 28, 2211 CrossRef CAS PubMed.
  3. Q. Lu, Y. Gao, Q. Zhao, J. Li, X. Wang and F. Wang, J. Power Sources, 2013, 242, 677 CrossRef CAS PubMed.
  4. F. Gao, Q. Zhou, Y. Dong, Y. Qin, X. Wang, X. Zhao and F. Wang, J. Polym. Res., 2012, 19, 9877 CrossRef.
  5. (a) Z. Bartczak, A. S. Argon, R. E. Cohen and M. Weinberg, Polymer, 1999, 40, 2331 CrossRef CAS; (b) Q. Yuan, W. Jiang, H. Zhang, J. Yin, L. An and R. K. Y. Li, J. Polym. Sci., Part B: Polym. Phys., 2001, 39, 1855 CrossRef CAS PubMed.
  6. (a) G. Gao, J. Zhang, H. Yang, C. Zhou and H. Zhang, Polym. Int., 2006, 55, 1215 CrossRef CAS PubMed; (b) L. D. Zhu, H. Y. Yang, G. Di Cai, C. Zhou, G. F. Wu, M. Y. Zhang, G. H. Gao and H. X. Zhang, J. Appl. Polym. Sci., 2013, 129, 224 CrossRef CAS PubMed.
  7. Q. Zhou, F. Gao, X. Wang, X. Zhao and F. Wang, Acta Polym. Sin., 2009, 227 CrossRef CAS.
  8. L. Chen, Y. Qin, X. Wang, X. Zhao and F. Wang, Polymer, 2011, 52, 4873 CrossRef CAS PubMed.
  9. L. Chen, Y. Qin, X. Wang, Y. Li, X. Zhao and F. Wang, Polym. Int., 2011, 60, 1697 CrossRef CAS PubMed.
  10. (a) J. J. Huang, H. Keskkula and D. R. Paul, Polymer, 2004, 45, 4203 CrossRef CAS PubMed; (b) S. H. Wu, Polymer, 1985, 26, 1855 CrossRef CAS; (c) S. H. Wu, J. Appl. Polym. Sci., 1988, 35, 549 CrossRef CAS PubMed; (d) Z. Z. Yu, Y. C. Ke, Y. C. Ou and G. H. Hu, J. Appl. Polym. Sci., 2000, 76, 1285 CrossRef CAS; (e) B. Majumdar, H. Keskkula and D. R. Paul, J. Polym. Sci., Part B: Polym. Phys., 1994, 32, 2127 CrossRef CAS PubMed.
  11. (a) A. M. Donald and E. J. Kramer, J. Appl. Polym. Sci., 1982, 27, 3729 CrossRef CAS PubMed; (b) D. G. Gilbert and A. M. Donald, J. Mater. Sci., 1986, 21, 1819 CrossRef CAS; (c) M. Matsuo, T. T. Wang and T. K. Kwei, J. Polym. Sci., Part A-2, 1972, 10, 1085 CrossRef CAS PubMed; (d) J. C. Stendahl, E. R. Zubarev, M. S. Arnold, M. C. Hersam, H. J. Sue and S. I. Stupp, Adv. Funct. Mater., 2005, 15, 487 CrossRef CAS PubMed.
  12. (a) C. B. Bucknall and D. R. Paul, Polymer, 2013, 54, 320 CrossRef CAS PubMed; (b) C. B. Bucknall and D. R. Paul, Polymer, 2009, 50, 5539 CrossRef CAS PubMed.
  13. Y. Gao, L. Gu, Y. Qin, X. Wang and F. Wang, J. Polym. Sci., Part A: Polym. Chem., 2012, 50, 5177 CrossRef CAS PubMed.
  14. (a) I. Kim, M. J. Yi, K. J. Lee, D. W. Park, B. U. Kim and C. S. Ha, Catal. Today, 2006, 111, 292 CrossRef CAS PubMed; (b) B. Y. Liu, X. J. Zhao, X. H. Wang and F. S. Wang, Polymer, 2003, 44, 1803 CrossRef CAS.
  15. Y. Lin, K.-Y. Zhang, Z.-M. Dong, L.-S. Dong and Y.-S. Li, Macromolecules, 2007, 40, 6257 CrossRef CAS.
  16. J. J. Huang, H. Keskkula and D. R. Paul, Polymer, 2006, 47, 639 CrossRef CAS PubMed.
  17. (a) S. Balakrishnan and N. R. Neelakantan, Polym. Int., 1998, 45, 347 CrossRef CAS; (b) X. Zhi, H.-B. Zhang, Y.-F. Liao, Q.-H. Hu, C.-X. Gui and Z.-Z. Yu, Carbon, 2015, 82, 195 CrossRef CAS PubMed.
  18. S. B. Newman and I. Wolock, Fracture Processes in Polymeric Solids: Phenomena and theory, John Wiley & Sons Inc, New York, 1964 Search PubMed.
  19. (a) G. M. Kim and G. H. Michler, Polymer, 1998, 39, 5689 CrossRef CAS; (b) G. M. Kim and G. H. Michler, Polymer, 1998, 39, 5699 CrossRef CAS.
  20. H. Kang, B. Qiao, R. Wang, Z. Wang, L. Zhang, J. Ma and P. Coates, Polymer, 2013, 54, 2450 CrossRef CAS PubMed.
  21. C. Geng, J. Su, S. Han, K. Wang and Q. Fu, Polymer, 2013, 54, 3392 CrossRef CAS PubMed.
  22. V. Thirtha, R. Lehman and T. Nosker, Polymer, 2006, 47, 5392 CrossRef CAS PubMed.
  23. J. D. Ferry, Viscoelastic properties of polymers, Wiley, New York, 1980 Search PubMed.
  24. Y.-S. He, J.-B. Zeng, G.-C. Liu, Q.-T. Li and Y.-Z. Wang, RSC Adv., 2014, 4, 12857 RSC.
  25. (a) J. Odent, P. Leclere, J.-M. Raquez and P. Dubois, Eur. Polym. J., 2013, 49, 914 CrossRef CAS PubMed; (b) H. Xiu, C. Huang, H. Bai, J. Jiang, F. Chen, H. Deng, K. Wang, Q. Zhang and Q. Fu, Polymer, 2014, 55, 1593 CrossRef CAS PubMed.
  26. (a) G. C. Liu, Y. S. He, J. B. Zeng, Y. Xu and Y. Z. Wang, Polym. Chem., 2014, 5, 2530 RSC; (b) W. M. Gramlich, M. L. Robertson and M. A. Hillmyer, Macromolecules, 2010, 43, 2313 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra07142a

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.