Next Article in Journal
Study on the Macro-Micro Mechanical Properties of Grout Consolidated Coal under Different Loading Rates
Next Article in Special Issue
Eutectic In Situ Modification of Polyamide 12 Processed through Laser-Based Powder Bed Fusion
Previous Article in Journal
Investigation of Surface Integrity of Selective Laser Melting Additively Manufactured AlSi10Mg Alloy under Ultrasonic Elliptical Vibration-Assisted Ultra-Precision Cutting
Previous Article in Special Issue
The Influence of the Use of Technological Waste and the Simulation of Material Lifetime on the Unnotched Impact Strength of Two Different Polymer Composites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Physical Properties and Crystallization Kinetics of Biocomposite Films Based on PLLA and Spent Coffee Grounds

Faculty of Mechanical Engineering, Technical University of Liberec, Studentská 1402/2, 46117 Liberec, Czech Republic
*
Author to whom correspondence should be addressed.
Materials 2022, 15(24), 8912; https://doi.org/10.3390/ma15248912
Submission received: 21 November 2022 / Revised: 11 December 2022 / Accepted: 12 December 2022 / Published: 13 December 2022
(This article belongs to the Special Issue Application, Processing, and Testing of New Progressive Materials)

Abstract

:
In the context of today’s needs for environmental sustainability, it is important to develop new materials that are based on renewable resources and biodegrade at the end of their life. Bioplastics reinforced by agricultural waste have the potential to cause a revolution in many industrial applications. This paper reports the physical properties and crystallization kinetics of biocomposite films based on poly(L-lactic acid) (PLLA) and 10 wt.% of spent coffee grounds (SCG). To enhance adhesion between the PLLA matrix and SCG particles, a compatibilizing agent based on itaconic anhydride (IA)-grafted PLLA (PLLA-g-IA) was prepared by reactive extrusion using dicumyl peroxide (DCP). Furthermore, due to the intended application of the film in the packaging industry, the organic plasticizer acetyl tributyl citrate (ATBC) is used to improve processing and increase ductility. The crystallization behavior and thermal properties were observed by differential scanning calorimetry (DSC) and thermogravimetric analysis (TGA). Crystallinity degree increased from 3,5 (neat PLLA) up to 48% (PLLA/PLLA-g-IA/ATBC/SCG) at the highest cooling rate. The physical properties were evaluated by tensile testing and dynamic mechanical analysis (DMA). The combination of the compatibilizer, SCG, and ATBC led to a synergistic effect that positively influenced the supramolecular structure, internal damping, and overall ductility of the composite films.

1. Introduction

The invention of catalytic polymerization caused massive petroleum-based plastics production that presents a significant challenge for the environmental sustainability [1,2,3]. The bio-based materials made from renewable resources have the potential to be commercially and environmentally acceptable. Especially in Europe, attention to the production of greenhouse gas emission, renewability, and toxicity is given, and the market with these environmentally friendly materials is growing fast [4,5]. Currently, one of the most commonly used biopolymers is poly(lactic acid) (PLA) [6]. Due to its biocompatibility and bioresorbability, PLA was initially used for medical purposes, e.g., bone fixation [7]. More affordable production and the possibility of biodegradation through the industrial composting process increased PLA application potential also to packaging [8], textiles [9], molding [10], and additive technologies [11]. PLA is a linear aliphatic thermoplastic polyester with high modulus and tensile strength and excellent barrier properties [12]. By contrast, its disadvantages include brittleness [13], a low heat deflection temperature [14], crystallization [15], and a biodegradation rate [16]. Many material modifications could be made to enhance these properties. Blending with other polymers, copolymerization, and the addition of particulate fillers and fibers or plasticizers are commonly used methods [17,18,19,20,21]. For example, adding the organic plasticizers, triethyl citrate (TC) and acetyl-tributyl citrate (ATBC) are used to enhance ductility due to the increase in chain mobility while maintaining biodegradability [22]. Other PLA usage limits, such as low heat deflection temperature and mechanical properties, can be improved through the enhancement of crystallization kinetics [23,24]. One option is to increase the nucleation density by adding heterogeneous nucleation agents [25,26]. The simultaneous addition of the nucleation agents and plasticizers may results in a synergistic effect, which can positively affect the crystallinity degree of PLA [27]. In last decades, many studies have investigated the use of fiber [28] and micro or nano particulate fillers [29,30] and their potential use as nucleating agents. Currently, the natural bio-based fillers, such as hydroxyapatite [31], cellulose nanocrystals [32], biochar [33], rice husks [34], or spent coffee grounds [35], are used. Spent coffee grounds containing cellulose, hemicellulose, lignin, and coffee oil, due to its composition, preferably act as a plasticizer and filler to improve the mixing process [36,37,38]. For example, Suaduang et al. [38] found a 33% increase in MFR for PLLA containing 10% of spent coffee grounds. Another promising usage of agricultural waste is to obtain new environmental-friendly polymeric material by liquification [39]. Furthermore, due to the lack of PLA-reactive side chain groups, surface and bulk modification is challenging, which can cause poor adhesion in composite structures [40,41]. Among chemical modification possibilities, free radical grafting is arguably one of the most successful and cost-effective treatments for preparing compatibilizers and improving interface adhesion. A number of studies was published for effect assessment of silanes, styrenes, or maleic anhydride (MAH) as compatibilizing agents [41,42,43]. Using MAH in PLA composites with low filler concentrations shows also significant impact on the degree of crystallinity, especially at slow cooling rates. Běhálek et al. [44] observed an increase in the absolute degree of crystallinity (Xc) from 54.4 to 60.1% for PLA with 10 wt.% of cellulose fibers by the grafting MAH onto a PLA backbone. However, due to the poor activity of MAH toward macro-radicals, the grafting efficiency is low compared to other polyesters [45]. Additionally, the reaction of MAH-grafted polymers with proteins may result in unstable amid bonds that can easily be hydrolyzed [40]. Itaconic anhydride (IA), with its chemical similarity, represents an alternative to MAH, and moreover, it is derived from renewable resources [46]. The high degree of grafting was reached by Verbeek et al. [47] for polyethylene. Crosslinking during the grafting reactions reduced elongation at break while improving the impact strength, creep resistance and not affecting the tensile strength [48]. Ku Marsilla et al. [40] found an increase in elongation at break, tensile strength, and energy to break for PLA grafted by 5 wt.% IA. The main aim of this study was to evaluate the thermal, mechanical, and structural properties of prepared bio-composite structures, where the reactive extrusion of itaconic anhydride with PLA was applied to reduce the interfacial tension between the matrix and spent coffee grounds (SCG) filler. Due to the brittleness of the PLA and in the view of the intended application, an organic plasticizer, ATBC, was incorporated into a bio-composite. A variety of characterization methods were used, such as Fourier transform infrared spectroscopy (FT-IR), scanning electron microscopy (SEM), differential scanning calorimetry (DSC), thermogravimetric analysis (TGA), tensile test, and dynamic mechanical analysis (DMA). Furthermore, the effect of SCG, ATBC, IA grafting onto PLA chain, and cooling rate of the non-isothermal crystallization process on the crystallization kinetics was investigated in detail using a modified Avrami equation.

2. Materials and Methods

Poly(L-lactic acid) (PLLA) matrix under the trade name of Luminy L130 was supplied by Total Corbion PLA (Gorinchem, The Netherlands). It has an approximate molar mass of 170,000 g/mol, a glass transition temperature (Tg) of 60 °C, a melting point of 175 °C, and contains more than 99% of L-lactide. Particulate spent coffee ground filler (SCG) Kimbo 100% Arabica (De’Longhi S.p.A., Treviso, Italy) was ground to an average particle size lower than 20 µm. In addition to that, 95% itaconic anhydride (IA) (Sigma-Aldrich, Darmstadt, Germany) with an approximate molar mass of 112 g/mol in the form of a powder as a compatibilizer agent and 98% dicumyl peroxide (DCP) (Sigma-Aldrich, Darmstadt, Germany) with an approximate molar mass of 270 g/mol as an initiator was used. Dehydrated acetone (ACROS Organics, Geel, Belgium) was applied to dissolve DCP and IA. A liquid form of ATBC, Citroflex A-4 delivered by Vertellus Holding LLC (Indianapolis, IN, USA), was used as a plasticizer.

2.1. Masterbatch Preparation and Film Processing

For the bio-composite film extrusion, at first, it was necessary to prepare masterbatches (MB) (Table 1), which were combined with PLLA pellets to reach the targeted compositions (Table 2). All additives, as well as the PLLA, were dried in vacuum oven VD53 (Binder GmbH, Tuttlingen, Germany) to remove moisture for 24 h at 60 °C. For the MB1 preparation, IA and DCP were dissolved in 50 mL of dehydrated acetone and added to PLLA. Further, the material was dried again and reactive extrusion with a subsequent granulation process was carried out. As a processing device, the laboratory micro-compounder MC 15 HT (Xplore, Sittard, The Netherlands) with conical screws, a constant temperature profile of 180 °C and a screw speed of 60 rpm was used. In this study, the IA and DCP concentration was fixed to 6 and 0.5% according to conclusions of Ku Marsilla et al. [40]. The dispersion and distribution were controlled by the level of torque. Consequently, the dried bio-composite pellets were processed on the twin-screw extruder MC 15 HT with a flat film die (210 °C, 0.2 mm gap size) at a melt temperature of 200 °C and 60 rpm screw speed. The extruded films were drawn by air-cooled rolls.

2.2. Spectral Analysis

To study the grafting of IA onto the PLLA, the Fourier transform infrared spectroscopy (FT-IR) using a Nicolet iS10 (Thermo Fisher Scientific, Waltham, MA, USA) spectrometer was performed. Measurements were taken with single-reflection attenuated total reflectance (ATR) system equipped with a diamond crystal. The specimens were dried in a vacuum oven VD53 at 60 °C for 24 h before the characterization. Each measurement consisted of 32 scans with 4 cm−1 resolution in the 400 to 4000 cm−1 range.

2.3. Scanning Electron Microscopy (SEM)

The morphology of fractured surfaces was examined by field emission scanning electron microscopy (FE-SEM) using the TESCAN MIRA 3 (Tescan, Brno, Czech Republic) instrument with an accelerated voltage of 5 kV. The film specimens were frozen in liquid nitrogen (−196 °C) and crushed. The fractured surfaces were coated with 5 nm of gold using a sputter coater LEICA EM ACE 200 (Leica Microsystems, Wetzlar, Germany).

2.4. Crystallization Kinetics

The non-isothermal study of crystallization kinetics was conducted by differential scanning calorimetry (DSC) analysis using a DSC 1/700 (Mettler Toledo, Greifensee, Switzerland) at different cooling rates at 5 to 40 °C min−1. The facility was calibrated via zinc and indium standards. For each measurement, it was prepared about 4–8 mg of the film sample, and an empty pan was used as a reference. The samples were heated from 0 to 200 °C with a heating rate of 10 °C min−1 and performed under a constant nitrogen flow of 50 mL min−1. The following thermal properties were observed for each composite film: the enthalpy of melting (ΔHm), the enthalpy of cold crystallization (ΔHcc), and the enthalpy of pre-melt crystallization (ΔHpc). In addition to that, the degree of crystallinity (Xc) was determined according to the following equation:
X c   = H m H cc H pc H m 0 · w m
where (ΔHm0) is the melting enthalpy of 100% crystalline PLLA (106 J g−1) [49] and wm is the mass content of PLLA in the composites.
Further, the melting temperature (Tm), cold crystallization temperature (Tcc), pre-melt crystallization temperature (Tpc), melt crystallization temperature (Tc), and enthalpy of melt crystallization (ΔHc) were tracked. Relative crystallinity (XT) was calculated according to the following equation:
X T   = H T H C = T 0 T dH c dT dT T 0 T dH c dT dT
where (ΔHT) is the heat of melt crystallization, (T1) initial temperature, (T) temperature at a given moment, and (T) final temperature. The melt crystallization time (tc) was determined by the following equation:
t c   = T 0 T CR
where (CR) is the cooling rate.
Jeziorny-modified model of the logarithmic transformation of Avrami Equation (4) was used to consider the temperature changes during the crystallization process [50,51]. The growth rate parameter (Zt) is corrected for CR, giving a modified growth rate constant (Zc), as follows:
log ln 1 X T = log Z t n log t
log Z c = log Z t CR

2.5. Thermal Degradation

The temperature of thermal decomposition of each additive and matrix was set by thermogravimetric analysis (TGA) using TGA2 instrument (Mettler Toledo, Greifensee, Switzerland). The samples were heated from 50 to 600 °C under N2 atmosphere at the heating ramp of 10 °C min−1. The initial decomposition temperature was determined at 5% mass loss. Prior to the measurements, the temperature sensor was calibrated using the Curie temperature of ferromagnetic materials based on isotherm, trafoperm, and nickel alloy.

2.6. Mechanical Measurement

Measurement of mechanical properties was performed by using a TIRA test 2300 (Tira, Schalke, Germany) testing machine equipped with a load cell of 1 kN and extensometer MFX 500-B (Mess & Feinwerktechnik GmbH, Velbert, Germany). Measurements have been performed according to the ISO 527 standard. Young’s modulus measurements were performed at a crosshead speed of 1 mm min−1 and tensile strength with elongation at break at a crosshead speed of 5 mm min−1. According to ISO 291 standard, all the specimens were conditioned in a KSP climatic chamber (Teseco, Kostelec nad Orlici, Czech Republic) at 23 °C and 50% relative humidity for one week.

2.7. Dynamic Mechanical Analysis (DMA)

The measurement of dynamic mechanical properties was carried out on the DMA DX04T (RMI s.r.o., Lazne Bohdanec, Czech Republic) equipment with the constant sinusoidal strain under the temperature of 0 to 80 °C, frequency of 1 Hz, and deformation amplitude of 0.2 mm. The square specimens (10 × 10 mm) were obtained from the films and tensile tested. The heating rate was 2 °C min−1, and the storage modulus (E′), loss modulus (E″), complex modulus (E*), and loss factor (tan δ) were evaluated according to Equations (6)–(9). The values of these parameters were determined at temperatures of 10, 23, 40, and 50 °C.
E = E   .   cos δ
E = E   .   sin δ
tan δ = E E
E = E +   i   . E

3. Results

3.1. Spectral Analysis

The presence of IA on the PLLA chain was identified through FT-IR measurements, which were compared to neat PLLA and IA (Figure 1). Full spectra of each component are shown in Figure 1a. As shown in Figure 1b,c, new absorption bands appear at 1649 cm−1 and 928 cm−1 in the spectra of PLLA-g-IA. A new peak can be observed at 1649 cm−1 for PLLA-g-IA, which is not noticeable for a neat PLLA sample (Figure 1b). This absorption was ascribed to the stretching vibrations of the cyclic anhydride functional group (C=C), which prove the mechanism of IA grafting onto PLA by Ku Marsilla [40]. Absorption band at around 2920 cm−1 can be assigned to CH2 functional groups, which could be considered as proof of grafted IA since CH2 groups are included in its molecule [52]. A peak reflecting C=O stretching vibration of an anhydride ring was probably not observed due to the low concentration of grafted IA [46]. Absorption band at 928 cm−1 can be ascribed to the symmetric stretching vibration of the C-O-C functional group (Figure 1c). This conclusion confirms successful grafting of IA monomers onto the PLLA backbone.

3.2. Scanning Electron Microscopy (SEM)

Structural analysis of the fractured surfaces of the composite films using electron microscopy was performed to assess the effect of the compatibilizing agent (IA) and plasticizer (ATBC) on the interfacial adhesion between the SCG and the PLLA matrix. The brittle fracture of all studied films could be observed in the microscopic images. In this case, Figure 2b shows that the adhesion between the hydrophobic matrix and the hydrophilic natural filler is sufficient even without the use of a compatibilizing agent. However, with the addition of PLLA-g-IA, SCG particles were coated and bonded to the matrix better than in the PLLA/SCG composite (Figure 2d). The reduction in the interfacial tension between the filler and matrix should be mainly reflected in the mechanical properties of the films. For the sample PLLA/ATBC/SCG, the interfacial adhesion is also at a very good level, which may be caused by higher chain mobility due to the addition of a plasticizer that contributed to better envelopment of the SCG filler (Figure 3b). The same results were observed for PLLA/PLLA-g-IA/ATBC/SCG composite films (Figure 3d).

3.3. Crystallization Kinetics

The extruded film samples were studied using DSC analysis according to the ISO 11,357 standard to specify morphology and thermal properties. In this section, the results obtained from the melt cooling run and the second heating run, which determine the crystallization capability of the materials, are discussed. In terms of the transition temperatures, the results of the DSC analysis show that the melting temperature of the composite structures is mainly determined by the polymer matrix in range of 172 ± 2 °C, see Table 3. In the presence of ATBC, the melting temperature slightly decreases to values of 168 ± 2 °C. As expected, the effect of ATBC is particularly pronounced in changing the Tg, which is in range of 45 ± 5 while for PLLA material structures without ATBC in range of 59 ± 2 °C. However, the determination of the Tg by DSC is affected by a methodological error, which is particularly visible at lower cooling rates, when the primary crystallization of the polymer system is greatest, increasing its degree of crystallinity and resulting in difficulty in determining the glass transition region. The study of crystallization kinetics has shown, in addition to well-known facts, such as the effect of CR on the Xc and a positive effect of SCG on the formation of the PLLA crystalline structure, see Figure 4. In contrast, no positive effect of IA on the Xc of the PLLA/SCG composite structure was demonstrated, although grafting IA onto PLLA chain alone resulted in an increase in Xc. Comparing the PLLA and PLLA/PLLA-g-IA matrices, an increase in crystallinity of 2.6 times can be found at CR of 5 °C/min. In contrast, the comparison of the PLLA/SCG and PLLA/PLLA-g-IA/SCG composites shows a 5% decrease in crystallinity, which could be observed at all investigated CRs. Thus, it can be concluded that the grafting of IA onto the PLLA chain does not achieve a synergistic effect in the crystallization of SCG composite structures, whereas in the case of the PLLA/ATBC/SCG, it is noticeable, and an increase in the Xc was observed at all CRs.
A positive finding is that the increase in Xc is particularly evident even at high CRs, which otherwise generally hinder polymer crystallization. At slow melt CR (5 °C/min), the PLLA/PLLA-g-IA/ATBC/SCG achieved the highest value of Xc = 48%, showing an increase of 18% compared to PLLA/PLLA-g-IA/ATBC and a fourfold increase compared to neat PLLA matrix. A similar result can be observed for the PLLA/SCG structure. At higher CRs (40 °C/min), the crystallization ability of PLLA is limited. The values of Xc show that a higher change in the macromolecule arrangement can be achieved at higher CRs due to the SCG addition; for example, a more than threefold Xc increase was observed for the PLLA/SCG structure, while PLLA/PLLA-g-IA/SCG increased 2.4 times (only 1.4-fold at 5 °C/min). The highest values of Xc at CR = 40 °C/min were achieved by the PLLA/ATBC/SCG (Xc = 20%), where an increase of about 2 times was observed compared to the PLLA/ATBC matrix, while at CR = 5 °C/min this increase was only 10%. Overall, the organic plasticizer based on ATBC can be attributed to an increase in the conformational processes of macromolecules promoting the crystallization behavior of the polymer. Its addition to PLLA results in an increase in Xc at all CRs. The positive effect of the SCG can be clearly seen in the increase in enthalpy changes in the primary crystallization and the suppression of the secondary crystallization. The temperature of secondary crystallization (Tsc) shifts to significantly lower values (Table 3), indicating that SCG composites require lower kinetic energy for the internal rotations of macromolecules, similar to the PLLA/ATBC structures.
Such as cellulose nanocrystals [26], lignin nanoparticles [53] or cellulose fibers, and rice husks [44], SCG act as a nucleating agent to promote overall crystallization of PLLA. This is due to the higher onset of the crystallization temperature and the peak of the primary crystallization temperature (Tc), respectively, see Table 4. For example, at CR of 5 °C/min, Tc = 101.7 °C for PLLA and 108.2 °C for PLLA/SCG. Similarly, this phenomenon can be observed for the PLLA/ATBC/SCG structure, where at the same CR, the Tc increased from 93.8 to 101.6 °C, see Table 4. The relative crystallization (XT) determined at Tc and halftime of the crystallization process (t1/2) corresponding to XT = 50% were also monitored to assess the crystallization kinetics. As was mentioned above, the presence of SCG increases the primary crystallization temperature, whereas ATBC decreases it. Thus, the PLLA/SCG crystallinity half-time is smaller than PLLA/ATBC/SCG at all studied CRs. The effect of SCG presence in the PLLA, PLLA/PLLA-g-IA, PLLA/ATBC, and PLLA/PLLA-g-IA/ATBC is shown in Figure 5 and Figure 6.
An increase in the crystallization halftime also occurs when IA is grafted onto the PLLA chain. At a CR = 5 °C/min, in all composite structures, the t1/2 decreased due to the presence of SCG. At higher melt CRs, this trend persisted especially for the PLLA/SCG and PLLA/ATBC/SCG. For the PLLA/PLLA-g-IA/SCG and PLLA/PLLA-g-IA/ATBC/SCG, there is a slight increase in the t1/2. Thus, it can be assumed that grafting of IA onto the PLLA backbone suppresses the rate of crystallization. The results of the change in XT clearly indicate a higher rate of the PLLA crystallization process in the presence of SCG and ATBC. Especially at lower CRs, the synergistic effect of SCG with ATBC in the PLLA matrix was confirmed (XT has grown up to 52% compared to neat PLLA). These conclusions are further supported by the Jeziorny modification of Avrami crystallization kinetics model. The statistical parameters obtained by the linear fitting of the dependence of log[-ln(1-XT)] and log t are listed in Table 5.
The Avrami exponent n as a measure of uniform crystal growth increases by the addition of SCG from 3.25 to 4.58 at CR = 5 °C/min. The values of n could be concluded to be a mixed-type growth with variable nucleation. With the increase in the CR, the values of Zc increase and t1/2 decreases, which indicates the higher crystallization rate with the increase in CR. Additionally, in the presence of SCG, the values of Zc and t1/2 changed from 0.00069 and 3.23 min at CR = 5 °C/min to 0.037839 and 0.34 min at CR = 40 °C/min. Therefore, based on the results of t1/2 and Zc, we could draw the conclusion that the presence of SCG has a positive effect on PLLA crystallization kinetics, especially at higher CRs. The same conclusion was reached for PLLA/PLLA-g-IA and PLLA/ATBC matrix. The value of n is depending on the nature of the nucleation and growth geometry of crystals [54]. This indicates that the presence of SCG has a great influence on the nucleation and growth geometry of PLLA. The highest effect could be observed for PLLA/PLLA-g-IA/SCG and PLLA/ATBC/SCG, where the value of n was 2.63 and 2.96 at CR = 40 °C/min, while for the other composite structures, there was not a noticeable area of primary crystallization.

3.4. Thermal Degradation

The thermodegradation temperatures are summarized in Table 6. Thermal stability of neat PLLA, SCG, and ATBC reaches its limit at 328, 253, and 222 °C, respectively. Thus, the addition of ATBC into PLLA films caused a significant decrease in the initial decomposition temperature. Similar results were observed by Maiza et al. [22] and others [55,56]. Grafting IA onto the PLLA backbone slightly increased thermal stability, which is ascribed to the new C=C bond. The addition of SCG reduced the decomposition temperature to approximately 310 °C, which is related to the degradation of low molecular weight components, such as hemicellulose, that is present in the lignocellulosic residue [57]. A similar reduction in thermal stability was observed by Terroba-Delicado et al. [25] or Mendes et al. [58] when studying PE/SCG thermal properties. Nekhamanurak et al. [59] observed PLLA/CaCO3 fatty acid-treated composites and concluded the reaction of the fatty acids and oils with ester linkage of PLLA decreases its thermal stability. The incorporation of SCG into PLLA/PLLA-g-IA/ATBC matrix points to the synergistic effect between SCG and ATBC in the presence of PLLA-g-IA.

3.5. Mechanical Measurements

The results from tensile tests of neat PLLA, modified PLLA, and SCG-based composites prepared by extrusion are shown in Table 7. The measured values show that the Young’s modulus depending on the presence of SCG does not change in practice, which is positive since the moisture content of the SCG composite film 0.7–0.9% is higher than that of the neat PLLA 0.3–0.4%. The same conclusion was reached for the composite with grafted itaconic anhydride onto the PLLA chain. Significant changes are observed for the PLLA/ATBC/SCG film, where the presence of coffee grounds in the PLLA/ATBC matrix causes a decrease in tensile modulus of about 28% (from 2834 MPa to 2049 MPa) and an increase in the elongation more than five times (from 7 to 36%), although the degree of crystallinity increased by 63%. These results indicate the synergistic effect of ATBC and SCG containing fatty acids and other components reducing the stiffness and increasing the ductility of the material, which was not achieved by PLLA containing SCG and ATBC separately. This synergistic effect was also observed for the composite films PLLA/PLLA-g-IA/ATBC and PLLA/PLLA-g-IA/ATBC/SCG, respectively. The presence of SCG led to a 12% decrease in the average modulus and in an increase in ductility by approximately 90%. A comparison of the tensile test results shows that the addition of SCG to PLLA, PLLA/PLLA-g-IA, and PLLA/ATBC matrix reduces the tensile strength as expected. This effect is also supported by the increase in the elongation at break, which confirms the SCG plasticizing ability. Similar results were observed by Suaduang et al. [35] on the bio-composite PLA/SCG blown films. It was found that a greater content of SCG significantly decreases the tensile strength, and the results of elongation at break showed the reduction to brittleness compared to neat PLA. This brittleness reduction is supported by ATBC content. The ATBC with abundant polar groups easily enters the gap between PLA chains, which weakens the binding and entanglement force between PLA chains and increases their mobility [60]. Tian et al. observed that 20 wt.% of ATBC resulted in the great increase in the elongation at break while the tensile strength and tensile modulus reduce observably when compared with the 15 wt.% ATBC modified PLA, which is attributed to the reduction in intermolecular forces [60,61].

3.6. Dynamic Mechanical Analysis (DMA)

The experimental research has investigated the temperature dependence of the dynamic modulus of elasticity (E*), including its real (E′) and imaginary (E″) components, and the loss factor (tan δ) characterizing the internal damping of the material. The parameters determined by DMA at temperatures of 10, 23, 40, and 50 °C, respectively, are shown in Table 8. The temperature of 10 °C represents cooler conditions, the temperature of 23 °C corresponds to standard environmental conditions, the temperature of 40 °C corresponds to the mean glass transition temperature of PLLA/ATBC-based films, and the temperature of 50 °C corresponds to the glass transition of other studied composite films. With the addition of SCG, the results of the DMA show an increase in the imaginary component of the dynamic modulus, which is characterizing the damping properties of the material. The change in tan δ is insignificant as the imaginary component of the dynamic modulus increases along with its real component. This is not the case for the PLLA/ATBC/SCG film due to the content of a plasticizer, which reduces the elastic modulus. In contrast to the static tensile test, a positive effect of SCG on the elastic modulus and damping properties was observed. An increase in these properties may be promoted by the presence of fatty acids in SCG, which is in line with the conclusions of Coelho et al. [62]. The differences in the values of the elastic modulus found in static and dynamic loading are caused by the nature of the test and the range of deformation regions. The greatest effect of SCG onto damping properties was clearly observed for the PLLA/ATBC/SCG film, which corresponds to the change in the nominal relative elongation at break, see Table 7. The SCG in the given composite system increased the imaginary elastic modulus and the loss factor up to 2.9 and 3.8 times, respectively. The deviations observed for the PLLA/PLLA-g-IA/ATBC/SCG composite structure may be due not only to the frequency of the measurements but also to the frequency of the recorded data and their approximation.

4. Conclusions

In this paper, the itaconic anhydride (IA) was successfully grafted onto PLLA backbone by a reactive extrusion process, and consequently, acetyl tributyl citrate (ATBC) and spent coffee grounds (SCG) were incorporated. To increase the application potential, the effect of the organic plasticizer on the supramolecular structure, thermal, mechanical, and dynamic-mechanical properties of the PLLA/SCG bio-composite were investigated. The presence of SCG and ATBC caused not only an increase in elongation at break of PLLA composites but also enhanced crystallization kinetics. The positive effect of SCG was observed as the nucleation barrier in the primary crystallization region decreased. A synergistic effect occurring especially at higher cooling rates (CR) was achieved for the PLLA/ATBC/SCG composite structure. At lower CRs, PLLA/PLLA-g-IA/ATBC/SCG reached the highest values of the degree of crystallinity (Xc). Thus, the study of crystallization kinetics showed that SCG significantly increased the crystallization rate in all studied materials. In terms of thermal stability, SCG reduced the initial temperature of degradation. However, the chemical reaction of the fatty acids contained in SCG with the ester groups of PLLA was partially prevented by grafting IA onto its backbone. It also had a positive effect on the formation of a supramolecular structure by increasing the Xc. The same effect was achieved by the presence of an ATBC plasticizer. The combination of the abovementioned modifications resulted into highest Xc values while maintaining the desirable properties of the extruded films, such as high elongation at break. A positive effect of SCG on the overall ductility and the damping properties was also demonstrated by DMA analysis and an increase in the imaginary component of the dynamic modulus of elasticity.

Author Contributions

Conceptualization, L.B., J.N., M.B. and P.L.; methodology, L.B.; validation, P.L. and M.B.; formal analysis, L.B., J.N. and M.B.; resources, J.N. and M.B.; writing—original draft preparation, J.N.; writing—review and editing, M.B.; visualization, J.N., M.B. and L.B.; supervision, M.B. and P.L., project administration, J.N.; funding acquisition, J.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Student Grand Competition of the Technical University of Liberec under the project No. SGS-2022-5011 and the Institutional Endowment for the Long Term Conceptual Development of Research Institutes, as provided by the Ministry of Education, Youth and Sports of the Czech Republic.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mülhaupt, R. Catalytic Polymerization and Post Polymerization Catalysis Fifty Years After the Discovery of Ziegler’s Catalysts. Macromol. Chem. Phys. 2003, 204, 289–327. [Google Scholar] [CrossRef]
  2. Philp, J.C.; Ritchie, R.J.; Guy, K. Biobased Plastics in a Bioeconomy. Trends Biotechnol. 2013, 31, 65–67. [Google Scholar] [CrossRef] [PubMed]
  3. Barker, T.; Bashmakov, I.; Bernstein, L.; Bogner, J.; Bosch, P.; Dave, R.; Davidson, O.; Fisher, B.; Grubb, M.; Gupta, S. Technical Summary. In Climate Change 2007: Mitigation. In Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change; Cambridge University Press: Cambridge, UK, 2007. [Google Scholar]
  4. La Mantia, F.; Morreale, M. Green Composites: A Brief Review. Compos. Part A Appl. Sci. Manuf. 2011, 42, 579–588. [Google Scholar] [CrossRef]
  5. Moshood, T.D.; Nawanir, G.; Mahmud, F.; Mohamad, F.; Ahmad, M.H.; Abdul Ghani, A. Expanding Policy for Biodegradable Plastic Products and Market Dynamics of Bio-Based Plastics: Challenges and Opportunities. Sustainability 2021, 13, 6170. [Google Scholar] [CrossRef]
  6. EUBIO_Admin Market. European Bioplastics e.V.
  7. Farah, S.; Anderson, D.G.; Langer, R. Physical and Mechanical Properties of PLA, and Their Functions in Widespread Applications—A Comprehensive Review. Adv. Drug Deliv. Rev. 2016, 107, 367–392. [Google Scholar] [CrossRef] [Green Version]
  8. Auras, R.; Harte, B.; Selke, S. An Overview of Polylactides as Packaging Materials. Macromol. Biosci. 2004, 4, 835–864. [Google Scholar] [CrossRef]
  9. Mochizuki, M. Textile Applications. Poly (Lactic Acid) Synthesis, Structures, Properties, Processing, Applications, and End of Life; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2022; pp. 619–629. [Google Scholar]
  10. Muller, J.; González-Martínez, C.; Chiralt, A. Poly (Lactic) Acid (PLA) and Starch Bilayer Films, Containing Cinnamaldehyde, Obtained by Compression Moulding. Eur. Polym. J. 2017, 95, 56–70. [Google Scholar] [CrossRef]
  11. Rasselet, D.; Caro-Bretelle, A.-S.; Taguet, A.; Lopez-Cuesta, J.-M. Reactive Compatibilization of PLA/PA11 Blends and Their Application in Additive Manufacturing. Materials 2019, 12, 485. [Google Scholar] [CrossRef] [Green Version]
  12. Sin, L.T. Polylactic Acid: PLA Biopolymer Technology and Applications; William Andrew: Norwich, NY, USA, 2012; ISBN 1-4377-4460-5. [Google Scholar]
  13. Zengwen, C.; Pan, H.; Chen, Y.; Bian, J.; Han, L.; Zhang, H.; Dong, L.; Yang, Y. Transform Poly (Lactic Acid) Packaging Film from Brittleness to Toughness Using Traditional Industrial Equipments. Polymer 2019, 180, 121728. [Google Scholar] [CrossRef]
  14. Tábi, T.; Ageyeva, T.; Kovács, J.G. Improving the Ductility and Heat Deflection Temperature of Injection Molded Poly (Lactic Acid) Products: A Comprehensive Review. Polym. Test. 2021, 101, 107282. [Google Scholar] [CrossRef]
  15. Salmerón Sánchez, M.; Mathot, V.B.F.; Vanden Poel, G.; Gómez Ribelles, J.L. Effect of the Cooling Rate on the Nucleation Kinetics of Poly(l-Lactic Acid) and Its Influence on Morphology. Macromolecules 2007, 40, 7989–7997. [Google Scholar] [CrossRef]
  16. Aitor, L.; Erlantz, L. Review on the Thermomechanical Properties and Biodegradation Behaviour of Polyester. Eur. Polym. J 2019, 121, 1–31. [Google Scholar]
  17. Yang, Y.; Zhang, L.; Xiong, Z.; Tang, Z.; Zhang, R.; Zhu, J. Research Progress in the Heat Resistance, Toughening and Filling Modification of PLA. Sci. China Chem. 2016, 59, 1355–1368. [Google Scholar] [CrossRef]
  18. Ramot, Y.; Haim-Zada, M.; Domb, A.J.; Nyska, A. Biocompatibility and Safety of PLA and Its Copolymers. Adv. Drug Deliv. Rev. 2016, 107, 153–162. [Google Scholar] [CrossRef]
  19. Fortelny, I.; Ujcic, A.; Fambri, L.; Slouf, M. Phase Structure, Compatibility, and Toughness of PLA/PCL Blends: A Review. Front. Mater. 2019, 6, 206. [Google Scholar] [CrossRef] [Green Version]
  20. Choi, K.; Choi, M.-C.; Han, D.-H.; Park, T.-S.; Ha, C.-S. Plasticization of Poly (Lactic Acid)(PLA) through Chemical Grafting of Poly (Ethylene Glycol)(PEG) via in Situ Reactive Blending. Eur. Polym. J. 2013, 49, 2356–2364. [Google Scholar] [CrossRef]
  21. Murariu, M.; Dubois, P. PLA Composites: From Production to Properties. Adv. Drug Deliv. Rev. 2016, 107, 17–46. [Google Scholar] [CrossRef]
  22. Maiza, M.; Benaniba, M.T.; Quintard, G.; Massardier-Nageotte, V. Biobased Additive Plasticizing Polylactic Acid (PLA). Polimeros 2015, 25, 581–590. [Google Scholar] [CrossRef]
  23. Dartora, P.C.; da Rosa Loureiro, M.; de Camargo Forte, M.M. Crystallization Kinetics and Morphology of Poly (Lactic Acid) with Polysaccharide as Nucleating Agent. J. Therm. Anal. Calorim. 2018, 134, 1705–1713. [Google Scholar] [CrossRef]
  24. Battegazzore, D.; Bocchini, S.; Frache, A. Crystallization Kinetics of Poly (Lactic Acid)-Talc Composites. Express Polym. Lett. 2011, 5, 849–858. [Google Scholar] [CrossRef] [Green Version]
  25. Terroba-Delicado, E.; Fiori, S.; Torres-Giner, S.; Gomez-Caturla, J.; Montanes, N.; Sanchez-Nacher, L. Improving the Mechanical Ductility and Toughness of Injection-Molded Polylactide Pieces by the Dual Incorporation of Liquor Waste Derived Spent Coffee Grounds and Oligomers of Lactic Acid. 2 November 2021. Preprint (Version 1). Available online: https://assets.researchsquare.com/files/rs-1035447/v1/97b0a683-ae91-4208-92fc-0b971337b616.pdf?c=1648436444 (accessed on 20 November 2022). [CrossRef]
  26. Gupta, A.; Simmons, W.; Schueneman, G.T.; Mintz, E.A. Lignin-Coated Cellulose Nanocrystals as Promising Nucleating Agent for Poly (Lactic Acid). J. Therm. Anal. Calorim. 2016, 126, 1243–1251. [Google Scholar] [CrossRef]
  27. Saeidlou, S.; Huneault, M.A.; Li, H.; Park, C.B. Poly (Lactic Acid) Crystallization. Prog. Polym. Sci. 2012, 37, 1657–1677. [Google Scholar] [CrossRef]
  28. Běhálek, L.; Seidl, M.; Dobránsky, J. Crystallization of Polylactic Acid Composites with Banana and Hemp Fibres by Means of DSC and XRD Methods. Appl. Mech. Mater. 2014, 616, 325–332. [Google Scholar] [CrossRef]
  29. Brzeziński, M.; Biela, T. Polylactide Nanocomposites with Functionalized Carbon Nanotubes and Their Stereocomplexes: A Focused Review. Mater. Lett. 2014, 121, 244–250. [Google Scholar] [CrossRef]
  30. Murariu, M.; Laoutid, F.; Dubois, P.; Fontaine, G.; Bourbigot, S.; Devaux, E.; Campagne, C.; Ferreira, M.; Solarski, S. Pathways to Biodegradable Flame Retardant Polymer (Nano) Composites. In Polymer Green Flame Retardants; Elsevier: Amsterdam, The Netherlands, 2014; pp. 709–773. [Google Scholar]
  31. Akindoyo, J.O.; Beg, M.D.; Ghazali, S.; Heim, H.P.; Feldmann, M. Impact Modified PLA-Hydroxyapatite Composites–Thermo-Mechanical Properties. Compos. Part A Appl. Sci. Manuf. 2018, 107, 326–333. [Google Scholar] [CrossRef]
  32. Herrera, N.; Salaberria, A.M.; Mathew, A.P.; Oksman, K. Plasticized Polylactic Acid Nanocomposite Films with Cellulose and Chitin Nanocrystals Prepared Using Extrusion and Compression Molding with Two Cooling Rates: Effects on Mechanical, Thermal and Optical Properties. Compos. Part A Appl. Sci. Manuf. 2016, 83, 89–97. [Google Scholar] [CrossRef] [Green Version]
  33. Arrigo, R.; Bartoli, M.; Malucelli, G. Poly (Lactic Acid)–Biochar Biocomposites: Effect of Processing and Filler Content on Rheological, Thermal, and Mechanical Properties. Polymers 2020, 12, 892. [Google Scholar] [CrossRef] [Green Version]
  34. Battegazzore, D.; Bocchini, S.; Alongi, J.; Frache, A. Rice Husk as Bio-Source of Silica: Preparation and Characterization of PLA–Silica Bio-Composites. RSC Adv. 2014, 4, 54703–54712. [Google Scholar] [CrossRef] [Green Version]
  35. Suaduang, N.; Ross, S.; Ross, G.; Pratumshat, S.; Mahasaranon, S. Effect of Spent Coffee Grounds Filler on the Physical and Mechanical Properties of Poly (Lactic Acid) Bio-Composite Films. Mater. Today Proc. 2019, 17, 2104–2110. [Google Scholar] [CrossRef]
  36. Ballesteros, L.F.; Teixeira, J.A.; Mussatto, S.I. Chemical, Functional, and Structural Properties of Spent Coffee Grounds and Coffee Silverskin. Food Bioprocess Technol. 2014, 7, 3493–3503. [Google Scholar] [CrossRef]
  37. Mussatto, S.I.; Machado, E.; Martins, S.; Teixeira, J.A. Production, Composition, and Application of Coffee and Its Industrial Residues. Food Bioprocess Technol. 2011, 4, 661–672. [Google Scholar] [CrossRef] [Green Version]
  38. Suaduang, N.; Ross, S.; Ross, G.M.; Wangsoub, S.; Mahasaranon, S. The Physical and Mechanical Properties of Biocomposite Films Composed of Poly (Lactic Acid) with Spent Coffee Grounds.; Trans Tech Publications Ltd.: Bäch, Switzerland, 2019; Volume 824, pp. 87–93. [Google Scholar]
  39. Naguib, H.M.; Hou, G. Exploitation of Natural and Recycled Biomass Resources to Get Eco-Friendly Polymer. J. Polym. Environ. 2022. [Google Scholar] [CrossRef]
  40. Ku Marsilla, K.I.; Verbeek, C.J.R. Modification of Poly(Lactic Acid) Using Itaconic Anhydride by Reactive Extrusion. Eur. Polym. J. 2015, 67, 213–223. [Google Scholar] [CrossRef]
  41. Essabir, H.; Raji, M.; Laaziz, S.A.; Rodrique, D.; Bouhfid, R. Thermo-Mechanical Performances of Polypropylene Biocomposites Based on Untreated, Treated and Compatibilized Spent Coffee Grounds. Compos. Part B Eng. 2018, 149, 1–11. [Google Scholar] [CrossRef]
  42. Jo, M.Y.; Ryu, Y.J.; Ko, J.H.; Yoon, J. Effects of Compatibilizers on the Mechanical Properties of ABS/PLA Composites. J. Appl. Polym. Sci. 2012, 125, E231–E238. [Google Scholar] [CrossRef]
  43. Du, J.; Wang, Y.; Xie, X.; Xu, M.; Song, Y. Styrene-Assisted Maleic Anhydride Grafted Poly (Lactic Acid) as an Effective Compatibilizer for Wood Flour/Poly (Lactic Acid) Bio-Composites. Polymers 2017, 9, 623. [Google Scholar] [CrossRef] [Green Version]
  44. Běhálek, L.; Borůvka, M.; Brdlík, P.; Habr, J.; Lenfeld, P.; Kroisová, D.; Veselka, F.; Novák, J. Thermal Properties and Non-Isothermal Crystallization Kinetics of Biocomposites Based on Poly(Lactic Acid), Rice Husks and Cellulose Fibres. J. Therm. Anal. Calorim. 2020, 142, 629–649. [Google Scholar] [CrossRef]
  45. Ma, P.; Jiang, L.; Ye, T.; Dong, W.; Chen, M. Melt Free-Radical Grafting of Maleic Anhydride onto Biodegradable Poly(Lactic Acid) by Using Styrene as A Comonomer. Polymers 2014, 6, 1528–1543. [Google Scholar] [CrossRef] [Green Version]
  46. Petruš, J.; Kučera, F.; Petrůj, J. Post-Polymerization Modification of Poly (Lactic Acid) via Radical Grafting with Itaconic Anhydride. Eur. Polym. J. 2016, 77, 16–30. [Google Scholar] [CrossRef]
  47. Verbeek, C.J.R.; Hanipah, S.H. Grafting Itaconic Anhydride onto Polyethylene Using Extrusion. J. Appl. Polym. Sci. 2010, 116, 3118–3126. [Google Scholar] [CrossRef]
  48. Tamboli, S.M.; Mhaske, S.T.; Kale, D.D. Crosslinked Polyethylene. Indian J. Chem. Technol. 2004, 11, 853–864. [Google Scholar]
  49. Sarasua, J.-R.; Prud’homme, R.E.; Wisniewski, M.; Le Borgne, A.; Spassky, N. Crystallization and Melting Behavior of Polylactides. Macromolecules 1998, 31, 3895–3905. [Google Scholar] [CrossRef]
  50. Jeziorny, A. Parameters Characterizing the Kinetics of the Non-Isothermal Crystallization of Poly (Ethylene Terephthalate) Determined by DSC. Polymer 1978, 19, 1142–1144. [Google Scholar] [CrossRef]
  51. Coburn, N.; Douglas, P.; Kaya, D.; Gupta, J.; McNally, T. Isothermal and Non-Isothermal Crystallization Kinetics of Composites of Poly (Propylene) and MWCNTs. Adv. Ind. Eng. Polym. Res. 2018, 1, 99–110. [Google Scholar] [CrossRef]
  52. Muenprasat, D.; Suttireungwong, S.; Tongpin, C. Functionalization of Poly (Lactic Acid) with Maleic Anhydride for Biomedical Application. J. Met. Mater. Miner. 2010, 20, 189–192. [Google Scholar]
  53. Yang, W.; Fortunati, E.; Dominici, F.; Giovanale, G.; Mazzaglia, A.; Balestra, G.M.; Kenny, J.; Puglia, D. Synergic Effect of Cellulose and Lignin Nanostructures in PLA Based Systems for Food Antibacterial Packaging. Eur. Polym. J. 2016, 79, 1–12. [Google Scholar] [CrossRef]
  54. Qiu, Z.; Fujinami, S.; Komura, M.; Nakajima, K.; Ikehara, T.; Nishi, T. Nonisothermal Crystallization Kinetics of Poly (Butylene Succinate) and Poly (Ethylene Succinate). Polym. J. 2004, 36, 642–646. [Google Scholar] [CrossRef] [Green Version]
  55. Brdlík, P.; Borůvka, M.; Běhálek, L.; Lenfeld, P. Biodegradation of Poly (Lactic Acid) Biocomposites under Controlled Composting Conditions and Freshwater Biotope. Polymers 2021, 13, 594. [Google Scholar] [CrossRef]
  56. Arrieta, M.P.; Peponi, L.; López, D.; Fernández-García, M. Recovery of Yerba Mate (Ilex Paraguariensis) Residue for the Development of PLA-Based Bionanocomposite Films. Ind. Crops Prod. 2018, 111, 317–328. [Google Scholar] [CrossRef]
  57. Morrell, J.J. Degradation of Lignocellulosic Materials and Its Prevention. Jom 2014, 66, 580–587. [Google Scholar] [CrossRef]
  58. Mendes, J.F.; Martins, J.T.; Manrich, A.; Luchesi, B.R.; Dantas, A.P.S.; Vanderlei, R.M.; Claro, P.C.; Neto, A.R.D.S.; Mattoso, L.H.C.; Martins, M.A. Thermo-Physical and Mechanical Characteristics of Composites Based on High-Density Polyethylene (HDPE) e Spent Coffee Grounds (SCG). J. Polym. Environ. 2021, 29, 2888–2900. [Google Scholar] [CrossRef]
  59. Nekhamanurak, B.; Patanathabutr, P.; Hongsriphan, N. The Influence of Micro-/Nano-CaCO3 on Thermal Stability and Melt Rheology Behavior of Poly (Lactic Acid). Energy Procedia 2014, 56, 118–128. [Google Scholar] [CrossRef]
  60. Tian, J.; Cao, Z.; Qian, S.; Xia, Y.; Zhang, J.; Kong, Y.; Sheng, K.; Zhang, Y.; Wan, Y.; Takahashi, J. Improving Tensile Strength and Impact Toughness of Plasticized Poly(Lactic Acid) Biocomposites by Incorporating Nanofibrillated Cellulose. Nanotechnology 2022, 11, 2469–2482. [Google Scholar] [CrossRef]
  61. Gálvez, J.; Correa Aguirre, J.P.; Hidalgo Salazar, M.A.; Vera Mondragón, B.; Wagner, E.; Caicedo, C. Effect of Extrusion Screw Speed and Plasticizer Proportions on the Rheological, Thermal, Mechanical, Morphological and Superficial Properties of PLA. Polymers 2020, 12, 2111. [Google Scholar] [CrossRef]
  62. Coelho, G.O.; Batista, M.J.; Ávila, A.F.; Franca, A.S.; Oliveira, L.S. Development and Characterization of Biopolymeric Films of Galactomannans Recovered from Spent Coffee Grounds. J. Food Eng. 2021, 289, 110083. [Google Scholar] [CrossRef]
Figure 1. FT-IR spectra of PLLA Luminy L130, IA, and PLLA-g-IA: (a) full FT-IR spectrum from 4000 cm−1 to 400 cm−1, (b) absorbance of the C=C stretching at 1649 cm−1, and (c) absorbance of the C-O-C stretching at 928 cm−1.
Figure 1. FT-IR spectra of PLLA Luminy L130, IA, and PLLA-g-IA: (a) full FT-IR spectrum from 4000 cm−1 to 400 cm−1, (b) absorbance of the C=C stretching at 1649 cm−1, and (c) absorbance of the C-O-C stretching at 928 cm−1.
Materials 15 08912 g001
Figure 2. SEM analysis of fractured surfaces: (a) PLLA, (b) PLLA/SCG, (c) PLLA/PLLA-g-IA, and (d) PLLA/PLLA-g-IA/SCG.
Figure 2. SEM analysis of fractured surfaces: (a) PLLA, (b) PLLA/SCG, (c) PLLA/PLLA-g-IA, and (d) PLLA/PLLA-g-IA/SCG.
Materials 15 08912 g002
Figure 3. SEM analysis of fractured surfaces: (a) PLLA/ATBC, (b) PLLA/ATBC/SCG, (c) PLLA/PLLA-g-IA/ATBC, and (d) PLLA/PLLA-g-IA/ATBC/SCG.
Figure 3. SEM analysis of fractured surfaces: (a) PLLA/ATBC, (b) PLLA/ATBC/SCG, (c) PLLA/PLLA-g-IA/ATBC, and (d) PLLA/PLLA-g-IA/ATBC/SCG.
Materials 15 08912 g003
Figure 4. Crystallinity degree of biocomposite structures at different CRs.
Figure 4. Crystallinity degree of biocomposite structures at different CRs.
Materials 15 08912 g004
Figure 5. Variation in XT during the cooling rates 5, 10, 20, and 40 °C/min for (a) neat PLLA, (b) PLLA/SCG, (c) PLLA/PLLA-g-IA, (d) PLLA/PLLA-g-IA/SCG, (e) PLLA/ATBC, (f) PLLA/ATBC/SCG, (g) PLLA/PLLA-g-IA/ATBC, and (h) PLLA/PLLA-g-IA/ATBC/SCG.
Figure 5. Variation in XT during the cooling rates 5, 10, 20, and 40 °C/min for (a) neat PLLA, (b) PLLA/SCG, (c) PLLA/PLLA-g-IA, (d) PLLA/PLLA-g-IA/SCG, (e) PLLA/ATBC, (f) PLLA/ATBC/SCG, (g) PLLA/PLLA-g-IA/ATBC, and (h) PLLA/PLLA-g-IA/ATBC/SCG.
Materials 15 08912 g005
Figure 6. Plots of XT as a function of time during the cooling rates 5, 10, 20, and 40 °C/min for (a) neat PLLA, (b) PLLA/SCG, (c) PLLA/PLLA-g-IA, (d) PLLA/PLLA-g-IA/SCG, (e) PLLA/ATBC, (f) PLLA/ATBC/SCG, (g) PLLA/PLLA-g-IA/ATBC, and (h) PLLA/PLLA-g-IA/ATBC/SCG.
Figure 6. Plots of XT as a function of time during the cooling rates 5, 10, 20, and 40 °C/min for (a) neat PLLA, (b) PLLA/SCG, (c) PLLA/PLLA-g-IA, (d) PLLA/PLLA-g-IA/SCG, (e) PLLA/ATBC, (f) PLLA/ATBC/SCG, (g) PLLA/PLLA-g-IA/ATBC, and (h) PLLA/PLLA-g-IA/ATBC/SCG.
Materials 15 08912 g006
Table 1. Masterbatch compositions.
Table 1. Masterbatch compositions.
SamplesPLLA
(g)
DCP
(g)
IA
(g)
ATBC
(g)
SCG
(g)
MB150.000.8510.2000
MB250.000.8510.20017.00
MB374.950017.000
MB460.0000017.00
Table 2. Sample compositions.
Table 2. Sample compositions.
SamplesPLLA
(wt.%)
IA
(wt.%)
ATBC
(wt.%)
SCG
(wt.%)
PLLA100000
PLLA/SCG900010
PLLA/PLLA-g-IA94600
PLLA/PLLA-g-IA/SCG846010
PLLA/ATBC900100
PLLA/ATBC/SCG8001010
PLLA/PLLA-g-IA/ATBC846100
PLLA/PLLA-g-IA/ATBC/SCG7461010
Table 3. The results from the second heating run of DSC analysis.
Table 3. The results from the second heating run of DSC analysis.
SamplesCR
(°C/min)
Tg
(°C)
Tm
(°C)
Tpc
(°C)
Tsc
(°C)
ΔHm
(J/g)
ΔHpc
(J/g)
ΔHsc
(J/g)
Xc
(%)
PLLA560.9175.5160.9113.040.21.425.912.2
1060.3174.7160.4111.939.10.532.26.0
2060.9175.5160.9113.038.20.333.34.4
4060.9175.7156.1113.038.40.334.43.5
PLLA/SCG5-173.9--43.40.00.045.5
10-173.6160.3-41.21.80.041.3
2059.0172.5155.394.146.56.615.825.4
4058.8172.4155.397.245.96.726.713.1
PLLA/PLLA-g-IA559.6172.4156.1105.144.64.58.631.7
1058.9172.7157.2108.241.61.828.711.1
2059.4173.1158.6109.241.21.033.76.4
4059.3173.1158.6109.440.41.035.04.4
PLLA/PLLA-g-IA/SCG5-172.6159.6-39.71.30.043.1
10-172.0155.594.843.15.04.138.2
2059.1171.7154.999.243.75.724.914.6
4059.0171.7155.199.243.35.528.310.6
PLLA/ATBC5-169.2153.1-40.41.50.040.8
10-169.1147.286.642.55.210.228.4
2040.8168.9148.992.742.34.424.214.3
4040.1168.9149.393.042.24.127.511.2
PLLA/ATBC/SCG5-169.8--38.50.00.045.4
10-169.3154.9-38.50.80.044.5
20-169.0145.777.142.05.85.636.1
4040.9168.7143.982.542.36.518.620.4
PLLA/PLLA-g-IA/ATBC5-170.0153.1-37.51.40.040.6
10-169.1146.790.140.54.912.026.5
2043.9169.5148.994.539.24.123.712.8
4043.6169.4149.194.439.13.726.111.0
PLLA/PLLA-g-IA/ATBC/SCG5-171.5--37.70.00.048.0
10-170.2152.091.839.93.90.945.2
2049.1170.2149.391.342.55.919.222.2
4049.1170.2148.991.442.35.923.816.1
Table 4. The results from the cooling run of DSC analysis.
Table 4. The results from the cooling run of DSC analysis.
SamplesCR
(°C/min)
Tc
(°C)
ΔHc
(J/g)
ΔHTc
(J/g)
XT
(%)
PLLA5101.75.432.4044.2
10100.21.180.4840.7
20----
40----
PLLA/SCG5108.235.4520.1356.8
10105.733.3517.2251.6
2096.212.205.9748.9
4093.70.810.4454.3
PLLA/PLLA-g-IA5100.921.949.6043.8
1096.63.291.6148.9
20----
40----
PLLA/PLLA-g-IA/SCG5105.933.3717.8753.6
1099.323.4810.9146.5
2094.52.671.4654.7
4091.70.100.0660.0
PLLA/ATBC593.831.4817.9557.0
1086.616.978.2248.4
2087.51.370.5842.3
40----
PLLA/ATBC/SCG5101.633.1722.3367.3
1094.229.3518.1461.8
2086.917.348.2147.3
4083.70.990.5353.5
PLLA/PLLA-g-IA/ATBC595.328.8616.0255.5
1088.611.925.5546.6
2088.11.240.5040.7
40----
PLLA/PLLA-g-IA/ATBC/SCG5101.931.8418.2757.4
1095.025.4712.3648.5
2090.94.211.9145.4
40----
Table 5. The modified Avrami parameters for the non-isothermal crystallization of PLLA, PLLA/PLLA-g-IA, PLLA/ATBC, and PLLA/PLLA-g-IA/ATBC composites.
Table 5. The modified Avrami parameters for the non-isothermal crystallization of PLLA, PLLA/PLLA-g-IA, PLLA/ATBC, and PLLA/PLLA-g-IA/ATBC composites.
SamplesCR
(°C/min)
t1/2
(min)
log ZtZcnr2
PLLA53.87−2.100.001593.250.99
102.35−1.700.002013.980.99
20-----
40-----
PLLA/SCG53.23−2.460.000694.580.98
102.03−1.570.002674.750.98
201.25−0.580.013153.860.99
400.341.180.378392.710.98
PLLA/PLLA-g-IA54.43−2.720.000383.840.99
102.50−1.830.001483.710.97
20-----
40-----
PLLA/PLLA-g-IA/SCG54.42−3.070.000174.550.96
102.97−2.030.000943.810.96
201.70−1.080.004164.050.98
400.351.190.387202.630.96
PLLA/ATBC55.32−2.920.000243.830.99
103.79−2.060.000883.500.99
201.18−0.540.014424.060.99
40-----
PLLA/ATBC/SCG54.65−3.130.000154.560.95
102.54−1.640.002284.180.96
201.75−1.050.004463.220.95
400.401.050.280502.960.99
PLLA/PLLA-g-IA/ATBC54.54−2.710.000393.930.99
103.10−1.860.001383.570.99
201.11-0.380.020843.640.99
40-----
PLLA/PLLA-g-IA/ATBC/SCG54.37−2.880.000274.250.97
103.18−2.230.000594.010.96
201.34−0.670.010693.790.99
40-----
Table 6. Thermodegradation temperatures.
Table 6. Thermodegradation temperatures.
SamplesTd,5 (°C)
PLLA328
PLLA/SCG310
PLLA/PLLA-g-IA337
PLLA/PLLA-g-IA/SCG316
PLLA/ATBC272
PLLA/ATBC/SCG272
PLLA/PLLA-g-IA/ATBC272
PLLA/PLLA-g-IA/ATBC/SCG279
Table 7. Tensile modulus, strength, and elongation at break of PLLA-based films.
Table 7. Tensile modulus, strength, and elongation at break of PLLA-based films.
SamplesEt (MPa)σm (MPa)εtb (%)
x ¯ sCI(95) x ¯ sCI(95) x ¯ sCI(95)
PLLA35301931205032811
PLLA/SCG359922213836211942
PLLA/PLLA-g-IA36031731074621411
PLLA/PLLA-g-IA/SCG338622113733211142
PLLA/ATBC28343001863921732
PLLA/ATBC/SCG2049193120251136117
PLLA/PLLA-g-IA/ATBC14751358425211278855
PLLA/PLLA-g-IA/ATBC/SCG129326716523322413522
Table 8. Values of storage modulus, loss modulus, complex modulus, and loss factor.
Table 8. Values of storage modulus, loss modulus, complex modulus, and loss factor.
SamplesT
(°C)
E′
(MPa)
E″
(MPa)
tan δ
(-)
E*
(MPa)
PLLA102276370.0162277
232222420.0192223
402140510.0242141
502088650.0312090
PLLA/SCG102632430.0172633
232596440.0172597
402465560.0232466
502354760.0322356
PLLA/PLLA-g-IA102449350.0152449
232374450.0192375
402260580.0262261
502147800.0382150
PLLA/PLLA-g-IA/SCG103074450.0153074
233012480.0153013
402833690.0242834
5026171390.0472623
PLLA/ATBC102282650.0282284
232251630.0282253
402017900.0442021
50----
PLLA/ATBC/SCG1020091850.0932027
2319071810.0961925
4014122380.1691453
50----
PLLA/PLLA-g-IA/ATBC1025061330.0532513
2323611440.0612370
4010073840.0381131
50----
PLLA/PLLA-g-IA/ATBC/SCG1027331220.0452738
2325161330.0532523
4011313760.0331241
50----
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Novák, J.; Běhálek, L.; Borůvka, M.; Lenfeld, P. The Physical Properties and Crystallization Kinetics of Biocomposite Films Based on PLLA and Spent Coffee Grounds. Materials 2022, 15, 8912. https://doi.org/10.3390/ma15248912

AMA Style

Novák J, Běhálek L, Borůvka M, Lenfeld P. The Physical Properties and Crystallization Kinetics of Biocomposite Films Based on PLLA and Spent Coffee Grounds. Materials. 2022; 15(24):8912. https://doi.org/10.3390/ma15248912

Chicago/Turabian Style

Novák, Jan, Luboš Běhálek, Martin Borůvka, and Petr Lenfeld. 2022. "The Physical Properties and Crystallization Kinetics of Biocomposite Films Based on PLLA and Spent Coffee Grounds" Materials 15, no. 24: 8912. https://doi.org/10.3390/ma15248912

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop