Indranil
Dey‡
,
Debashrita
Kundu‡
,
Sayon
Ghosh
,
Samir
Mandal
,
Ketaki
Samanta
* and
Suryasarathi
Bose
*
Department of Materials Engineering, Indian Institute of Science, Bengaluru – 560012, India. E-mail: ketakisamanta123@gmail.com; sbose@iisc.ac.in
First published on 13th March 2025
Post-consumer recycled polypropylene (PCR PP) is promising for sustainable applications, yet its limitations in electrical conductivity and mechanical properties require modifications. This study develops a vitrimer nanocomposite by modifying PCR PP via styrene-assisted maleic anhydride grafting and incorporating a molecule containing multiple epoxide groups facilitating effective crosslinking. Graphene oxide (GO) is added as a nanofiller, improving rheological, thermal, electrical and infrared thermal properties. Characterization techniques confirm structural enhancements, while tensile testing shows significant gains in strength and modulus. The vitrimer nanocomposite demonstrates recyclability and high performance, offering a sustainable path for advanced engineering applications within a circular economy framework.
Incorporating dynamic crosslinkers into waste plastics presents a viable solution to this problem. This entails converting waste plastics into covalent adaptive networks (CANs), where the crosslinks can interchange dynamically in response to external triggers.18 The strong mechanical characteristics of conventional thermosets and the malleability or reprocessability of thermoplastics are combined in an ideal CAN.19–25 Bulk reprocessing using injection molding, compression molding, melt blowing, or twin-screw extrusion has been shown to be effective in CANs that are based on dynamic linkages such as acetals,26,27 vinylogous ureas/urethanes,28,29 Diels–Alder structures,30 hydrazines,31 dioxaborolanes,32 boroxines,33,34 olefins,35,36 silyl ether,37 disulfides,38,39 siloxanes,40 diketoenamines,41 and so forth. Even with these advancements, most CANs still rely on dynamic linkages that are typically absent from well-established commercial thermosets, necessitating the creation, synthesis, and development of novel materials. The integration of various polar structures through the use of CANs is anticipated to potentially expand the variety and value of PP's uses in addition to addressing the material's notable loss of mechanical qualities after repeated recycling.42 Yet, because of the strict synthetic requirements and the non-reactive PP backbones, the production of CANs using PP requires adherence to more specific concepts than when using many other polymers. To meet industrial demands, dynamic chemistry should be incorporated into polypropylene (PP) in a way that is scalable, features a universal structure, avoids costly additives, and remains resistant during post-processing.43 The potential to directly convert standard polymers into CANs without additional modification or optimization of current facilities makes the implantation of dynamic links via post polymerization functionalization attractive.25 Conversely, the effective synthesis of PP-CANs is contingent upon a restricted set of dynamic reactions because of their chemical and thermal stability within the matrix.44
Over the past ten years, transesterification has emerged as a widely renowned technique for creating CANs, thanks to the groundbreaking research conducted on vitrimers in 2011.45 Therefore, adding maleic anhydride (MA) side groups to PP, the most thoroughly researched commercial product among all post-functionalized PP materials, presents a workable platform for transesterification.46–48 To accomplish transesterification, such reliable dynamic reactions typically require the presence of external catalysts.49–52
We developed a straightforward, scalable method for upcycling PCR PP into vitrimers via transesterification using zinc acetate as an external catalyst. The process involved a two-step reaction: first, grafting maleic anhydride (MA) onto PCR PP with styrene as a co-agent and dicumyl peroxide (DCP) as an initiator, followed by crosslinking with TGDDM epoxy in a twin-screw extruder (Fig. 1). Post evaluating the vitrimer thoroughly, graphene oxide (GO) was incorporated to design vitrimer nanocomposites, allowing us to study the effects of GO on the nanocomposite's thermomechanical, rheological, electrical, and infrared thermal properties, both pre- and post-recycling. There are plenty of literature studies on polypropylene nanocomposites,53–58 however, currently, there is no literature exploring the impact of adding nanofillers to polyolefin vitrimer nanocomposites or recycling them. Therefore, understanding the role of nanofillers in polyolefin vitrimer systems is essential. This study investigates the thermomechanical, rheological, electrical, and infrared thermal properties of PCR PP vitrimer nanocomposites by varying the amount of graphene oxide (GO). Additionally, the performance of these properties was carefully evaluated after recycling to assess the nanocomposites' durability and potential for sustainable reuse.
![]() | ||
Fig. 1 Schematic diagram for the preparation of the PCR PP vitrimer nanocomposite and reprocessing via dynamic exchange. |
In the following step, PCR PP vitrimer batches were fabricated by crosslinking the prepared m′-PCR PP batches with TGDDM in the presence of zinc acetate as a transesterification catalyst60 by subjecting them to extrusion at 180 °C and 150 rpm speed for 2 min residence time. PCR PP vitrimer batches were fabricated at different concentrations of TGDDM crosslinker (Table S2†). The nanocomposites of the most optimum crosslinked sample were formulated with the GO nanofiller at varying concentrations (Table S3†). The conditions for preparing the nanocomposites were kept consistent with those for the vitrimer samples. After extruding each batch, tensile samples were prepared through injection molding, with the barrel temperature set at 180 °C and the mold kept at ambient temperature. Four dog bone-shaped specimens were molded for each sample, using an injection pressure of 14 bar for mechanical testing.
Gel fraction (%) = W1/W0 × 100 | (1) |
The crystallization temperature (Tc) was determined from the exothermic peak observed during cooling to 40 °C. Subsequently, a second heating cycle was conducted to measure the melting temperature (Tm) from the endothermic peak. The heating and cooling rates were set at 10 °C min−1 under a nitrogen flow of 50 mL min−1.
The fraction of crystallinity (% Xc) of each vitrimer was determined by comparing its melting enthalpy (ΔHm) with the theoretical enthalpy of 100% crystalline isotactic polypropylene (i-PP), ΔHm0, which is 207.0 J g−1, where61
% Xc = ΔHm/ΔHm0 × 100 | (2) |
![]() | ||
Fig. 2 (a) FT-IR spectrum of m′-PCR PP and vitrimers with different concentrations of TGDDM. (b) Gel fraction of PCR PP and PCR PP vitrimers. |
Graphene oxide (GO) was specifically chosen as the nanofiller for the vitrimer nanocomposites due to its unique surface chemistry, which includes a high density of oxygen-containing functional groups such as hydroxyl, epoxy, and carboxyl groups. These functional groups facilitate strong interfacial interactions with the epoxide groups of TGDDM (the crosslinker used in our study) and maleic anhydride. FTIR spectra of 0.5 GO and 1 GO samples were analysed and compared with that of the 15 TGDDM sample (see Fig. S1†). It is evident that the intensity of the anhydride peak at 1782 cm−1 decreases from the 15 TGDDM to the 1 GO sample and the intensity of the ester peak at 1730 cm−1 increases. This confirms that GO reacted with maleic anhydride and formed an ester bond. This results in improved interfacial adhesion between GO and the vitrimer matrix.
The degree of crosslinking in PCR PP and the corresponding vitrimer samples was assessed through gel fraction studies by subjecting the samples to reflux conditions in xylene at 120 °C overnight. The analysis shows that the level of maleic anhydride groups grafted onto PCR PP is crucial for promoting crosslinking.16,43 Functionalized PCR PP (m′-PCR PP) systems present active functional sites capable of reacting with the TGDDM molecule, which promotes the crosslinking of the polymer and facilitates the formation of dynamic covalent adaptable networks. This leads to the conclusion that the degree of crosslinking is intrinsically linked to the maleation efficiency. As depicted in Fig. 2b, pristine PCR PP has a 0% gel fraction, implying its non-crosslinked network which accounts for its complete dissolution in hot xylene during the experiment. In contrast, the corresponding vitrimers of PCR PP with TGDDM did not dissolve, corroborating the formation of a crosslinked network that does not dissolve in xylene. The gel fraction progressively increased from 44% to 74% with TGDDM concentrations ranging from 5 wt% to 20 wt%, corroborating the enhancement in crosslinked networks in the system owing to availability of more TGDDM moieties to attach to the MA groups on the PCR PP backbone. Interestingly, the gel fraction saturated at 74% for both the samples 15 TGDDM and 20 TGDDM, suggesting that the crosslinking density of the PCR PP vitrimer approaches a saturation threshold at 15 wt% of TGDDM beyond which additional crosslinkers do not significantly enhance dynamic network formation and may be left dangling. Additional gel fraction analyses of GO incorporated vitrimer nanocomposites reveals that the addition of GO did not significantly alter the crosslink density (Fig. S2†). This result suggests that GO has minimal effect on the crosslinking of PCR PP.
The mechanical properties of post-consumer recycled polypropylene (PCR PP), along with vitrimers containing varying concentrations of TGDDM and GO-incorporated vitrimer-nanocomposites, were examined through tensile testing at room temperature. The mechanical properties and corresponding stress–strain curves are presented in Fig. 3a and S3† respectively and listed in Table S4.† All samples exhibited significantly higher yield compared to their thermoplastic precursor, PCR PP, prior to failure.
A notable observation was the steady increase in Young's modulus (YM) and yield stress (YS) with increasing TGDDM concentrations, from 5% to 20%. This trend highlights the substantial impact of the degree of crosslinking and crystallinity (Xc) on the mechanical properties of these semi-crystalline polymers. The observed behaviour can be explained by the preferential reaction of styrene with macroradicals of PP, leading to the formation of more stable styryl macroradicals. These radicals copolymerize with MA in the modified PCR PP (m′-PCR PP), forming branches. As a result, the addition of styrene mitigates PP chain scission while simultaneously enhancing the grafting of MA, thus improving the overall mechanical performance of the vitrimer system.63,64 The crosslinking of maleic anhydride (MA) and TGDDM, resulting in the formation of a covalent adaptable network (CAN) within the PCR PP matrix results in this upcycling of PCR PP. Given that the transesterification reaction responsible for topological rearrangement of CANs is slow at room temperature, the PCR PP vitrimer samples behave similarly to non-dynamically crosslinked PCR PP. This creates a significant barrier to crystallite rearrangement, which explains the higher Young's modulus (YM) and yield stress (YS) compared to those of the precursor material. The CANs thus demonstrate their potential for upcycling PCR PP, as evidenced by the increase in YS from 28 MPa in PCR PP to 33 MPa for the 20% TGDDM system.65,66 All vitrimers and the vitrimer-nanocomposite showed a similar elongation at yield (EY) as that of PCR PP. At 20% TGDDM content, the vitrimer reached a saturation point in both Young's modulus (YM) and yield stress (YS), showing similar values to that of the 15% TGDDM system. This suggests that further increases in crosslinker concentration may lead to crosslinker aggregation, negatively affecting mechanical properties. As a result, the 15 TGDDM system was selected for further studies due to its optimum mechanical properties. To explain these observations, it is proposed that at lower gel content, the increased crosslink density enhanced the interactions between molecular chains, improving the mechanical properties of m′-PCR PP TGDDM vitrimers. However, at higher gel content, the denser crosslinked network restricted the mobility and elongation of the molecular chains, causing a decline in EB as the TGDDM content increased from 5% to 20%.67 The underlying reason lies in the fact that the mechanical strength of semi-crystalline polymers like polypropylene (PP) is largely derived from their crystalline morphology. When crosslinked with active agents, the mechanical strength initially increases as crosslinker concentration increases. However, at higher concentrations, the crystalline structure is disrupted, resulting in reduced elongation, toughness, and overall mechanical performance.68,69 Interestingly, the percentage crystallinity showed only a slight decrease from PCR PP to the 20% TGDDM vitrimer (which is further discussed in the later sections), indicating that TGDDM is primarily concentrated in the amorphous regions of PP and has minimal impact on the crystalline regions. Similarly, previous work by Kar et al. reported the use of a di-epoxy crosslinker, DGEBA, to upcycle recycled PP and PE, demonstrating a comparable approach to enhance the mechanical properties of recycled polymers.43 However, TGDDM, with its four epoxide functional groups, exhibits higher reactivity, potentially promoting twice as many crosslinking reactions during epoxy-anhydride curing compared to the former epoxy crosslinker. This enhanced reactivity occurs through interactions with the pendant maleic anhydride groups grafted onto the PCR PP chain via the styrene-maleic anhydride (SMA) charge transfer complex (CTC). As a result, the mechanical properties of PCR PP are significantly enhanced, facilitating more effective upcycling of the material post vitrimer formation.
In the case of graphene oxide (GO), especially in the 15% TGDDM vitrimer a significant increase in Young's modulus was observed while maintaining comparable tensile strength and elongation at yield as listed in Table S4.† This improvement can be attributed to the oxygen-containing functional groups on the GO surface, which likely formed covalent bonds with the epoxide groups of unreacted TGDDM. As a nanofiller, GO possesses a large surface area, facilitating strong interfacial adhesion with the matrix. Additionally, its surface chemistry fosters strong interfacial interactions with the epoxide groups, resulting in a marked enhancement of the composite's mechanical properties. The robust interfacial adhesion between GO and the vitrimer matrix translates to higher energy absorption at failure, as observed in the composites with as low as 0.5% GO in the vitrimer.
However, when the GO content was increased to 1.0%, the system exhibited a reduction in tensile strength, Young's modulus, and elongation at yield, leading to diminished overall toughness in the sample 1 GO. This decline is likely due to the short mixing time (2 minutes) during extrusion, which may have been insufficient to achieve proper dispersion of 1% GO within the vitrimer matrix. As a result, GO agglomerated creating stress concentration points that led to premature failure of the 1 GO vitrimer nanocomposite.70 Thus, the 0.5 GO vitrimer nanocomposite system was selected for further experimental work, as it demonstrated the most optimal mechanical performance.
Thermograms reveal a less intense peak at around 125 °C in heating scans and 114 °C in cooling scans of PCR PP, attributed to a minor concentration of polyethylene in the waste PCR PP stream, which potentially entered during collection, sorting, etc. As crosslinking density increased, Tc exhibited an upward trend, reflecting that less undercooling was needed for crystallization due to heterogeneous microdomains acting as nucleation sites in PCR PP. For the 15 TGDDM vitrimer with a gel fraction of 74%, Tc increased marginally as compared to neat PCR PP (Tc = 125 °C), and further addition of TGDDM did not elevate Tc beyond that of 15 TGDDM (Tc = 132 °C), indicating that non-bonded esters aggregated in the amorphous regions, as corroborated by a consistent gel fraction of 74% between 15% and 20% TGDDM.
Moreover, the peak of the crystallization exotherm decreased and broadened with increasing TGDDM concentration, signifying a rise in matrix heterogeneity. This trend was also observed for Tm, where reduced crystalline domains required less energy to melt. Crosslinking constraints significantly hindered the segmental alignment of polyolefin chains, creating a physical barrier to chain packing and reducing crystallinity. At lower TGDDM concentrations, nucleation and crosslinking were balanced, while higher TGDDM content led to dominant crosslinking as evidenced by the 20 TGDDM vitrimer, where crystallinity is lower compared to that of the 5 TGDDM system, although the crystallization temperature (Tc) is the highest as depicted in Table S5.†43 The incorporation of styrene during the melt mixing of PCR PP and maleic anhydride (MA) reduces PP chain scission. It activates MA double bonds via charge transfer complex (CTC) formation, enhancing grafting and reactivity with TGDDM.59,72 This results in increased gel content and decreased crystallinity. Despite the high TGDDM concentration, the vitrimer maintains a notable semi-crystalline structure with only a slight reduction in crystallinity compared to its thermoplastic precursor, PCR PP, thereby preserving its mechanical properties.
The addition of 0.5% GO to the vitrimer results in a minor baseline shift in the DSC thermogram (Fig. 3b), indicating a transition towards a composite material. Given the low concentration of 0.5% GO, it does not significantly influence the crystallinity, since their number and size are insufficient compared to the crosslinking junctions, which are the primary nucleation sites. Consequently, the Tc and crystallization exotherm of the 0.5 GO vitrimer nanocomposite system (ΔHc = 49.6 J g−1) remain comparable to those of the 15 TGDDM vitrimer system (ΔHc = 50.19 J g−1). Similarly, the melting endotherm around Tm exhibits little change, reflecting the minimal impact on crystallinity and thermal transitions.
In contrast, the 1% GO system displays a pronounced baseline shift in the DSC thermogram, suggesting that it has evolved into a distinct vitrimer-nanocomposite system. Although 1% GO content is too low to act as major nucleation sites, the presence of GO domains impedes the smooth packing and folding of the PP matrix, reducing crystallinity (Xc) to 16% while keeping Tc relatively unchanged at 131 °C. This disruption leads to a lower crystallization exotherm of 35.81 J g−1. The diminished crystalline domains require less energy to melt and induce flow, resulting in a reduced melting endotherm of 33.87 J g−1, even though Tm remains intact. Thus, the presence of GO in the 1 GO sample creates barriers that hinder crystalline structure formation, affecting the thermal and mechanical properties of the vitrimer nanocomposite.
TGA thermograms of PCR PP, 15 TGDDM, and 0.5 GO are presented in Fig. S5† which illustrates the differences in thermal stability during the transition from thermoplastic to vitrimer to vitrimer-composite. Notably, the Td5 of both 15 TGDDM and 0.5 GO is observed at approximately 328 °C, which is lower than that of PCR PP (372 °C). However, the Td95 of 0.5 GO exceeds that of both PCR PP and 15 TGDDM, reaching 450 °C. The identical Td5 at 328 °C for both 15 TGDDM and 0.5 GO is presumably due to the volatiles originating from the onset of TGDDM decomposition. According to Lee,73 it was suggested that the thermal breakdown process of cured epoxy resins occurs through the decomposition of the glycidyl ether unit within the network. The TGA analysis demonstrated a significant change in slope at around 373 °C, indicating the decomposition of a substantial amount of TGDDM into glycidyl ether units.74 The higher Td95 of 15 TGDDM compared to PCR PP, along with a residual weight loss of 5.55% at 800 °C, can be attributed to the strong dynamic networks of the β-hydroxy ester in the vitrimer system undergoing thermally activated bond exchange even at elevated temperatures imparting enhanced thermal stability compared to neat PCR PP. Conversely, the Td95 of 0.5 GO, not even identified within the temperature scan range, with a residual weight loss of 7% at 800 °C, indicates that the decomposition temperature is elevated by the addition of GO nanofillers within the 15 TGDDM vitrimer matrix. The reason behind this originates from the phenomenon that the nanofiller GO shows spatial interference which enforces the development of the highly crosslinked molecular structure of 0.5 GO vitrimer-nanocomposites.75
Rheological experiments were conducted to better understand the flow behaviour and viscoelastic nature of vitrimers at increased temperatures, which has a direct impact on their processibility and end use. To investigate the viscoelastic properties of vitrimers, small amplitude oscillatory sweeps (SAOS) were used to assess flow characteristics at different frequencies or shear rates. The frequency sweeps at 180 °C showed that both the vitrimer and the vitrimer-composite exhibited power-law behaviour (G′ ∼ G′′ ∼ ωn), as shown in Fig. 4a. This common power-law behaviour has been regularly observed in various earlier vitrimer investigations.76,77 The classic behaviour of a viscoelastic polymer melt with a significant frequency dependency is evident in the case of PCR PP, where G′ and G′′ increase with an increase in frequency, probably resulting in difficulty in relaxation at higher frequencies. Interestingly, PCR PP exhibited no crossover points between its storage (G′) and loss (G′′) moduli. Furthermore, Fig. 4b shows that PCR PP shows more viscous character (G′′ > G′) than elastic across the full frequency range investigated. Dynamic crosslinking with TGDDM resulted in vitrimers showing elastic solid behaviour with a frequency-independent G′ at lower frequencies, virtually reaching a plateau, and a significantly lower G′′ than G′, as is typical of crosslinked materials.78 Furthermore, the G′ and G′′ trends of 15 TGDDM vitrimer samples in Fig. 4a demonstrated no crossover point in the examined frequency range.
![]() | ||
Fig. 4 (a) Storage modulus (G′) and loss modulus (G′′) versus angular frequency and (b) complex viscosity versus angular frequency for PCR PP, vitrimer and the vitrimer nanocomposite at 180 °C. |
Interestingly, the addition of 0.5% GO in 15 TGDDM did not change its viscoelastic characteristics significantly, only enhancing both G′ and G′′ marginally. The minimal increase in G′ is presumably due to the reaction of GO with the unreacted TGDDM leading to the formation of covalent bonds, and therefore a better filler–matrix interface, and improved load transfer characteristics.70,79,80 The incorporated GO in the vitrimer matrix might bridge the separated polymer chains in the vitrimer-nanocomposite matrix reducing the slippage of the chains during viscous flow and therefore increasing the G′′.81
Fig. 4b shows that PCR PP attained zero-shear viscosity at roughly 103 Pa s, and that of 15 TGDDM continued to increase dramatically, nearly two orders of magnitude higher, with no evidence of zero-shear viscosity within the studied frequency range. This result represents a significant increase in melt strength due to the creation of a crosslinked network in the vitrimers. The 0.5 GO vitrimer-nanocomposite system behaves similarly to the 15 TGDDM vitrimer system in terms of shear thinning. The marginal increase in complex viscosity can be attributed to the formation of a GO nanoparticle network in the vitrimer system corroborated by the literature reported on graphite oxide suspensions in PDMS.81
The surface morphology of tensile fractured PCR PP is reported as shear bands and microfibres in our prior work.16 However, after transforming PCR PP into a vitrimer, the tensile fracture morphology remained unchanged and exhibited the same fibrous texture with no phase separation, as depicted in Fig. 5a. A similar fibrous morphology was observed following the addition of 0.5 wt% GO to the PCR PP vitrimer (Fig. 5b).
![]() | ||
Fig. 5 SEM micrograph of tensile fractured surfaces of (a) the PCR PP vitrimer (15 TGDDM) and (b) its nanocomposites (0.5 GO). |
The electrical conductivity of PCR PP and its vitrimer 15 TGDDM is very low due to the insulating nature of the polymer. However, the incorporation of 0.5 wt% GO into the vitrimer increased AC conductivity to 10−10 S cm−1 (Fig. 7). This enhancement is likely due to the formation of a network structure within the polymer matrix, which facilitates charge transport. Chammingkwan et al. observed a similar range of AC conductivity by the addition of 1 wt% GO into PP.84 Therefore, lightweight conducting composites derived from PCR PP and GO can potentially serve as antistatic packaging for electronic components, automotive parts for static dissipation, and EMI shielding materials for sensitive electronic equipment.85
![]() | ||
Fig. 7 AC electrical conductivity of PCR PP, the PCR PP vitrimer (15 TGDDM), and its nanocomposite (0.5 GO). |
The thermally activated bond exchange induced by transesterification in the vitrimer enables the vitrimer-nanocomposite to be reprocessable. The initially extruded strands of 0.5 GO nanocomposite samples were cut into small pieces, re-extruded, and then injection molded under the same processing conditions to produce thermo-mechanically reprocessed vitrimer-nanocomposite samples (0.5 GO-r). The TGA thermogram (Fig. S7†) of the recycled 0.5 GO-r closely matches with that of the original 0.5 GO vitrimer nanocomposite sample showing that no change in the weight loss pattern has occurred, proving that no early degradation occurred due to recycling. Mechanical data (Table S6†) showed that the yield strength (YS), elongation at yield (EY), and Young's modulus (YM) remained almost unchanged after recycling, as evidenced in the stress–strain graph (Fig. S8†) for the 0.5 GO samples. This is further supported by the recovery rates of 100% for both TS and EY including 90% for YM, as depicted in the recovery rate (RR) bar plot (Fig. 8b). The fracture morphology of 0.5 GO-r was analyzed after tensile testing which revealed that the fracture surface exhibited characteristics similar to those of the parent material, including shear bands and microfibers aligned with the load direction, with no apparent phase separation (Fig. S9†). However, during the recycling of 0.5 GO, the AC conductivity slightly decreased to 10−11 S cm−1 due to a minor breakdown of the network during extrusion (Fig. 8c).86 Despite this reduction being minimal, it can be concluded that the AC conductivity is largely preserved after recycling. Furthermore, in the infrared thermal study of 0.5 GO-r, the temperature increase was up to 104 °C at 300 s of laser exposure time and there was an instantaneous fall in temperature (up to 32 °C) within 120 s of turning off the laser beam (Fig. 8d). This suggests the excellent heat dissipation capability of the prepared 0.5 GO vitrimer nanocomposite due to the high thermal conductivity of GO even after vigorous recycling. Thus, a thermally conductive vitrimer-nanocomposite produced from PCR PP helped to synthesise a recyclable nanocomposite for reprocessable CANs that fully recover their crosslink densities, associated thermomechanical properties and AC conductivity.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4na00904e |
‡ Equal contribution. |
This journal is © The Royal Society of Chemistry 2025 |