Next Article in Journal
Study on the Preparation of Nd–Doped Zinc Cobalt Oxide Porous Film and Its Electrochemical Properties
Previous Article in Journal
Friction Sensitivity Test Experiment and Desensitization Mechanism of Nitrocellulose-Coated DNTF Explosive Crystals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Pomace from Oil Plants as a New Type of Raw Material for the Production of Environmentally Friendly Biocomposites

1
Institute of Wood Sciences and Furniture, Warsaw University of Life Sciences—SGGW, 159 Nowoursynowska St., 02-776 Warsaw, Poland
2
Institute of Chemical Technology and Engineering, Faculty of Chemical Technology, Poznan University of Technology, Berdychowo 4, 60-965 Poznan, Poland
3
Department of Food Engineering and Process Management, Institute of Food Sciences, Warsaw University of Life Sciences—SGGW, 159C Nowoursynowska St., 02-776 Warsaw, Poland
4
Plasma/Ion Beam Technology Division, Material Physics Department, National Centre for Nuclear Research Świerk, 7 Sołtana St., 05-400 Otwock, Poland
5
Department of Food Technology and Human Nutrition, Institute of Food Technology and Nutrition, College of Natural Sciences, University of Rzeszów, 4 Zelwerowicza St.; 35-601 Rzeszów, Poland
*
Author to whom correspondence should be addressed.
Coatings 2023, 13(10), 1722; https://doi.org/10.3390/coatings13101722
Submission received: 31 August 2023 / Revised: 29 September 2023 / Accepted: 30 September 2023 / Published: 2 October 2023

Abstract

:
The production of environmentally friendly biocomposites can be based on attractive and low-cost vegetable pomace, a waste product from oil production. In the present study, biocomposites made from HDPE and pomace from black cumin, corn, and flax seeds were subjected to evaluation of structural, morphological, and thermal parameters and susceptibility to germination by filamentous fungi. Based on the characteristics of the produced biocomposites, it should be concluded that vegetable waste from oil production, applied at a 1:1 ratio as a filler for polyethylene-based biocomposites, significantly reduces the crystallinity of the produced material and decreases its thermal stability. It should also be noted that such biocomposites are more easily overgrown by fungi, which may facilitate their biodegradation. Very poor antioxidant properties, resulting from the encapsulation of the plant fraction in polyethylene, limit the functionality of this type of material as, for example, active biomaterials to prevent free radical processes. Although the structural and physical characteristics of the produced biocomposites have been shown to be inferior to polyethylene, efforts should be made to improve these characteristics. Plant waste can be a valuable raw material for the production of materials compatible with various industries.

1. Introduction

Problems related to waste and by-products are often discussed these days. The constant development of agriculture and the food industry has its consequences in the form of large quantities of production residues [1]. Improper management of agri-food by-products can cause many serious problems, such as environmental pollution or economic losses [2]. The principles of the circular economy and the paradigm of sustainable development require and drive us not only to follow already-designed residue management solutions but also to seek new solutions [3,4]. The main strategy is to follow the principles of reduce, reuse, and recycle [5]. Reducing the number of by-products is not always possible. In this case, we should consider them not as residues but as valuable additive resources.
Routray and Orsat [6] indicate that by-products and wastes from the food industry can be divided into (1) by-products generated during agricultural harvesting; (2) post-harvest by-products; and (3) by-products and waste from food processing. In general, agricultural by-products include stems, leaves, husks, pomace, seeds, pulp, and skins of various fruits, bran, and other residues that are not of significant industrial importance for further food processing. There are many ways to reuse plant residues. They are widely regarded as a good raw material for the production of biofuels [7], organic fertilizers [8], and animal feed [9]. Another possibility for the use of plant by-products is their development as new types of raw materials for biocomposites. Plant remains can be formed into composites as a reinforcing material with many types of matrix, such as cement [10], epoxy resins [11], or thermoplastics [12,13]. It is also possible to use plant materials for the production of films [14] or to use them as a filler for polymer-based membranes [15]. In the production of composites, polymers are often combined with plant residues [16,17,18]. Depending on the intended use of the material, a suitable filler should be selected to enhance the properties that are required in this particular case. It is crucial to recognize the relationship between molecular structure and the quality parameters of the final material. The addition of plant fibers to polymers such as polylactide (PLA) or high-density polyethylene (HDPE) may affect the structural properties of the composite, which in turn determines its physical and mechanical properties.
There are numerous reports in the scientific literature on the use of plant particles in the preparation of hybrid composite materials based on polyethylene [19,20]. Fernandes et al. [21] developed biocomposites based on polyethylene reinforced with coconut and cork, while Aji et al. [22] produced a hybrid HDPE-kenaf/pineapple composite. In all cases, an increase in mechanical and physical properties was found; however, as the authors of the article suggest, these properties depend on the quality and quantity of the plant particles added. Banat et al. [23] also suggest that the thermal properties of olive pomace-reinforced polyethylene do not deteriorate, provided the correct filler ratio is chosen. It seems, therefore, that the incorporation of plant biomass into a polymer matrix opens up possibilities for new materials; however, such biomaterials need to be studied in detail to assess their suitability.
The most commonly used techniques for qualitative assessment of biocomposites are Fourier-Transform Infrared Spectroscopy (FTIR), X-ray diffraction, and microscopic visualization techniques such as SEM. These techniques make it possible to assess the chemical composition, crystallinity, and structure of biocomposites as the most important quality parameters that determine the mechanical and physical properties of materials. Gustafsson et al. [24], using electron microscopy techniques, proved that the method of preparation of the plant component affects the morphology of the biopolymer, as the chemical components of plant particles may favor the swelling process. In a study by Raschip et al. [25], it was shown that substances such as flavonoids contained in plant extracts were shown to affect the porosity and thus surface microstructure.
Morphological changes, but also changes in the degree of crystallinity of synthetic polymers caused by the addition of plant material, affect the physical, mechanical, and biological properties of biocomposites produced using them. Amorphous polymers are more susceptible to biodegradation than crystalline polymers [26]. In addition, the presence of plant particles in biocomposites, such as cellulose, hemicelluloses, or proteins, results in their faster biodegradation due to the availability of nutrients necessary for microbial growth. Biopolymers reinforced with natural fibers decompose faster than synthetic polymers, which may be due to the presence of intrachain functional groups in such biopolymers [27]. It should be noted that the most important quality parameter determining the technological properties of biopolymers is the degree of crystallinity. In polymers with a high content of plant products, the number of crystalline zones is reduced, as are the density, mechanical resistance, and hardness [28].
Another important quality parameter of biocomposites, determining their potential use, is thermal stability. The thermal stability of plant additives determines the thermal properties of the produced material. Scientific research shows that the introduction of plant components into polyethylene changes the thermal stability of this polymer [29]. Muthuraj et al. [30] showed that the addition of natural fibers changes the degradation temperature and increases the tensile and bending strengths of biopolymers. Kamaran et al. [31] performed thermogravimetry and FTIR analysis on PLA biocomposites with untreated and pretreated loofah fibers and compared the results with the ones performed for pure PLA matrix. This study’s authors found that the addition of fiber decreased thermal stability and crystallinity. In turn, Mirowski et al. [32] showed that the addition of raspberry pomace to the polyvinyl chloride composite increased its thermal stability.
Plant pomace, a waste oil production, can be an attractive and low-cost raw material for the production of environmentally friendly biocomposites. The use of plant material, on the one hand, minimizes the amount of petroleum-derived raw material used, and on the other hand, it favors the procedures for managing waste from plant production. However, the quality of the biocomposites produced is influenced by many factors, such as the chemical composition, the interactions of the individual components, and also the method of biocomposite production. In this study, we set out to create a biocomposite based on high-density polyethylene and plant particles derived from oil pressing waste. The use of plant particles, especially black cumin seeds, was interesting due to the fungicidal and antioxidant properties confirmed in the literature, which could cause the biomaterial produced to acquire biologically active properties. The biocomposite was produced by grinding, homogenizing the components, and then extruding in an extruder. The extruded biocomposite was subjected to further modifications by grinding it again to prepare the appropriate product for testing. This procedure made it possible to obtain a relatively homogeneous material. In the authors’ previous work [33], the produced biocomposites were tested to characterize their physical and mechanical properties. The susceptibility of the produced biocomposites to fungal growth was also determined. The ability of fungi to grow on the surface of composites containing plant particles may indicate that such biomaterials will be partially biodegradable. However, the demonstration of surface antioxidant properties may increase their applicability.

2. Materials and Methods

2.1. Preparation of Biocomposites

The material for the production of biocomposites was high-density polyethylene-HDPE (Hostalen GD 7255, Basell Orlen Polyolefins Sp. z o.o., Płock, Poland) and nigella, maize, and flax seeds, which were waste in the form of pomace after the oil pressing process. The biocomposite production method was carried out according to the procedure described by Betlej et al. [33]. Plant pomace ground to a dusty fraction using a knife mill (OB-RPPD Sp. z o.o., Czarna Woda, Poland) was combined with thermoplastic particles with the assumed mass fraction of 50% and then subjected to homogenization using a high-speed mixer (KMOD SGGW, Warsaw, Poland). From the obtained mixture, biocomposites were produced using an extruder (Leistritz Extrusionstechnik GmbH, Nürnberg, Germany), which was then comminuted using a knife mill (OB-RPPD sp. z o. o., Czarna Woda, Poland) to a fine fraction, which was subjected to further processing. From the ground fraction, particles < 1.5 mm in size were obtained, which were used to prepare biocomposites with a thickness of 0.31–0.42 mm. The graphic way of preparing biocomposites is shown in Figure 1 and Figure 2.
The obtained biocomposites were marked with the following codes: PE-N (HDPE + black cumin—nigella—seed pomace), PE-M (HDPE + maize seed pomace), PE-F (HDPE + flax seed pomace), PE—pure high-density polyethylene.

2.2. Characterization of Quality Parameters

2.2.1. X-ray Diffraction Analysis (XRD)

To evaluate the supermolecular structures, the polyethylene matrix and composite materials were subjected to X-ray diffraction analysis using a RIGAKU diffractometer (Japan). The operating parameters for the lamp with Cu Kα radiation (wavelength 1.5418 A°) were: 30 kV and 25 mA anode excitation. The analysis was performed in the angle range 2θ = 10–30° with a count step of 0.04. The deconvolution of peaks was carried out by the method proposed by Hindeleh and Johnson [34]. After the separation of X-ray diffraction lines, the degree of crystallinity (Xc) of polyethylene and composites was calculated by comparing the areas under crystalline peaks and amorphous curves.

2.2.2. Thermal Properties of Film

The thermal properties were evaluated on the basis of differential scanning calorimetry (DSC) measurements as well as thermogravimetry analysis (TGA).
For differential scanning calorimetry (DSC) measurements, a DSC 3+ differential calorimeter (Mettler Toledo, Greifensee, Switzerland) was used. Temperatures and enthalpies of the melting process of the components of the analyzed materials were measured. Prior to analysis, the material was dried at 30 °C for 48 g in a vacuum oven (103 Pa) and stored over P2O5 until analysis. Approximately 3–5 mg of the sample was weighed into 40 µL aluminum pans, and small holes were made in the lids. The samples were heated in the temperature range of 20–250 °C at a rate of 5 °C/min, and a nitrogen flow of 50 mL/min was used. The DSC was calibrated using pure indium (Tm = 156.6 °C, ΔHm = 28.6 J/g). Curves were analyzed using STAR software version 16.10.
For the Thermogravimetry Analysis (TGA) measurements, a TGA/DSC 3+ thermogravimeter (Mettler Toledo, Greifensee, Switzerland) was used. About 5 mg of the ground material was weighed into 70 µL of aluminum crucibles. Mass changes were recorded in the range of 30–600 °C at a rate of 10 °C/min in a nitrogen atmosphere (50 mL/min). The resulting TGA and DTG curves were analyzed using STAR software version 16.10.

2.2.3. Fourier-Transform Infrared Spectroscopy (FTIR)

Infrared spectra were obtained using a Cary 630 spectrometer (Agilent Technologies Inc., Santa Clara, CA, USA) with a diamond crystal ATR reflection attachment. The analysis was carried out in the wavelength range of 650–4000 cm−1, with a resolution of 4 cm−1. Each sample was scanned 32 times and the background spectrum was collected after each measurement. The obtained spectra were analyzed using MicroLab PC software (Agilent Technologies Inc., Santa Clara, CA, USA).

2.2.4. SEM Analysis

The surface observations of the non-modified polyethylene, polyethylene-maize, polyethylene-nigella, and polyethylene-flax materials were performed using a Zeiss EVO® (Oberkochen, Germany) MA10 scanning electron microscope. The observations were performed for magnifications of, e.g., 50, 100, 200, 500, 1000, and 2000×, using the secondary electron (marked as SE or SED) detector. The acceleration voltage of the electrons was 20 kV. The range of electrons in pure polyethylene (33 at.% of C and 67 at.% of H), with a density of about 0.92 g/cm3, determined by the Suspre code (https://uknibc.co.uk/SUSPRE, accessed on 31 August 2023), was about 11 µm. The samples were covered by a 5 nm gold layer before the observations using a Q150R ES rotary pumped coating system by Quorum Technologies Ltd. (Laughton, UK), due to their poor electrical conductivity.

2.2.5. Growth of Fungi

Discs 20 mm in diameter were cut from HDPE-pomace plant-based biocomposites, which were then sterilized under UV light (Bionovo, Legnica, Poland) for 30 min. Sterile discs were placed on maltose-agar medium containing 2.5% maltose extract (Linegal Chemicals sp. z o.o., Blizne Łaszczyńskiego, Poland) and 2.5% agar (Polaura, Morąg, Poland). Plastic sterile pads separated the biocomposites from direct contact with the moist medium. The mold inoculum was placed at four equal intervals around the disks. Each inoculum was placed approximately 10 mm from the edge of the samples. Four species of fungi were used: Aspergillus niger Tiegh, strain ATCC 1688; Chaetomium globosum Kunze, strain A-141 (ATCC 6205); Penicillium chrysogenum Thom, strain A-130 (832); and Trichoderma viride Pers., strain A-102, from the pure culture collection of the Department of Wood Science and Wood Preservation of the Warsaw University of Life Sciences. The cultures were grown for 14 days in a Thermolyne Type 42,000 model incubator (ThermoFisher Scientific, Waltham, MA, USA) under temperature and relative humidity conditions of 26 ± 2 °C and 63 ± 2%, respectively. Measurements were taken at 24 h intervals by taking high-resolution images on a laboratory photo-taking station. The evaluation of the degree of fungus coverage on the surface of the samples was determined based on the guidelines of the methodology described by Betlej et al. [35]. The percentage of fungus coverage of the biocomposite surface was calculated relative to the total biocomposite surface using ImageJ2 image analysis software (Fiji v1.52i). The percentage of film surface fouling was determined with an accuracy of 5%. The test was performed in triplicate for each variant.

2.2.6. Antioxidant Properties

The research material was cut into fragments with an area of 10 cm2, which were flooded with 20 cm3 of water with pH 4 for 12 h at 4 °C, after which the biocomposites were removed and the solution was tested for antioxidant activity using the DPPH method. The ability to reduce free radicals of the simulant obtained after 12 h of incubation at 4 °C was measured on the basis of a modified method [36] using a stable free radical DPPH at a maximum absorbance of 515 nm. The radical solution was prepared by dissolving 2.4 mg of DPPH in 100 mL of methanol. The test solution (5 μL) was added to 3.995 mL of methanolic DPPH. After shaking, the mixture was kept in the dark for 20 min. The absorbance of the solution was measured spectrophotometrically at a wavelength of 515 nm. The reference point (the blank test) was the absorption of pure water with a pH of 4. The tests were repeated three times. The reported concentrations are based on a previously generated curve of various concentrations of Trolox. The final results are expressed in µm Trolox per 100 cm3.

2.3. Test Standards

The statistical analysis of the results was carried out with the application Statistica version 13 (TIBCO Software Inc., CA, USA). An analysis of variance (ANOVA) was used to test (α = 0.05) for significant differences between factors. A comparison of the mean values was performed using Tukey’s test, with α = 0.05.

3. Results and Discussion

Figure 3a shows X-ray diffraction patterns of polyethylene matrix and composites with various lignocellulosic fillers. Characteristic diffraction peaks are identified at angles of 2Θ = 21.5 and 23.7, which are consistent with those found in the literature for originating polyethylene. These peaks came from the lattice planes (110) and (200), respectively, and indicate that the HDPE used in this investigation shows an orthorhombic structure [37,38]. Moreover, regardless of the vegetable raw material used, the position of the peaks in each sample remains unchanged.
From the diffraction curves obtained, some differences in the intensity of the maxima can be observed. The highest intensity was found for the unfilled polyethylene matrix. The addition of fillers results in a decrease in intensity, which is particularly noticeable in composites with nigella filler (PE-N). Note that the intensity of the diffraction peak is directly proportional to the crystallinity of the polymer. Therefore, in the next step, the degree of crystallinity was determined for all the tested systems, which are listed in Table 1.
The degree of crystallinity for pure polyethylene is 52%. The introduction of each type of filler into the polymer matrix causes a decrease in Xc. For composites containing corn (PE-M) and flax fillers (PE-F), the crystallinity was about 40%. It is noteworthy that the greatest decrease in the degree of crystallinity was observed for composites containing black seed pomace waste (about 33%). Such a significant reduction in crystallinity in the case of the PE-N composite means that the black seed particle filler has reduced nucleation activity. The reasons for this phenomenon can be attributed to the anatomical structure of plant particles and the differences in chemical composition between plant particles and the polymer [39,40]. As suggested by Bajsic et al. [41], plant particles as components of a polyethylene-based biocomposite affect the crystallinity of polyethylene and may act as modulating agents for crystallization nucleation. If the crystallinity of plant particles is high, they may favor processes that lead to an increase in the level of crystallinity of the biocomposites produced with them. Other researchers argue that the reason for this phenomenon should be sought in the quality of the polymer used and in the method of producing the biocomposite [42]. The crystallization of the polymer matrix in the presence of plant fibers may also be determined by interactions at the polymer-fiber interface, which may affect the formation of the supermolecular structure. The renewable fillers used have different anatomical structures and thermal expansion, which can affect the value of the shear stresses at the interface and, consequently, the structure of the composites, as extensively described by Garbarczyk et al. [43]. An additional factor influencing the crystallization process of the polymer matrix is the particle size distribution and average pore size of the fillers [44].
Infrared spectra of all analyzed samples, such as polyethylene and polyethylene with the addition of plant material (linen, corn, and black cumin), contain vibrations of functional groups characteristic of polyethylene (Figure 3b). The peak of 2919 cm−1 corresponds to asymmetric vibrations, and 2851 cm−1 to symmetric vibrations of the methylene group (CH2). At the wavelength of 1470 cm−1, deformation vibrations of the C-H bond in the CH2 methylene group appear, and at the wavelength of 719 cm−1, rocking vibrations of the CH2 group appear [45,46]. Only in the HDPE with the addition of flax seed pomace was the appearance of the C-H group in the 3009 cm−1 wavelength observed [47]. Furthermore, the material containing the addition of black cumin seeds and the flax was characterized by vibrations at the wavelengths of 1707 and 1748 cm−1, respectively. These correspond to the stretching vibrations of the C=O group coming from ketones [48]. The greatest differences were observed in the material with the addition of linen. This spectrum has peaks in the 3000–3630 cm−1 range, characteristic of natural cellulose fibers belonging to the OH group. In the range of 1250–100 cm−1 there are strong C–O–C and C–O stretching vibrations [49]. Peak 1606 cm−1 corresponds to the C-C aromatic stretching. The signals present in the samples of polyethylene with additives come from their main components: lignins, hemicelluloses, pectins, and waxes [50].
Figure 4 shows the DSC curves for biocomposites based on plant particles and PE. The total energy consumption for the melting process of pure polyethylene was 472.93 mJ, while the addition of plant particles reduces the amount of heat necessary to raise the temperature, which means that plant particles have a lower specific heat. At the same time, it can be seen that the melting point (Tm) of pure polyethylene and biocomposites is similar, although the addition of plant particles causes a slight decrease in the melting point of the crystalline phase (Table 2). The lowering of the melting point may be due to the lower crystallites in biocomposites confirmed in earlier studies. Similar observations were made by Suárez et al. [51], who found that the addition of plant particles did not modify polyethylene, and lower energy values may be associated with a lower crystallinity of the biocomposite. The research by Andrzejewski et al. [52] also confirms these suppositions. The decrease in enthalpy in biocomposites, as opposed to pure polyethylene, may be related to the increased presence of the amorphous phase [53].
In the study of the effect of plant particles on the thermal stability of biocomposites, it was found that plant additives reduce thermal stability. The initial value of the thermal degradation of biocomposites was lower than that of pure PE. In the conducted research, it is noteworthy that the degradation of the biocomposites with the addition of black cumin begins already at a temperature of about 200 °C, which indicates that the addition of this plant material causes a significant weakening of thermal stability compared to other analyzed plant particles (Figure 5). Biocomposites based on PE and black cumin seed pomace showed three phases of thermal decomposition, and the observed first phase is related to the decomposition processes of particles other than lignin and cellulose contained in plants. Biocomposites based on PE, flax, and maize seeds turn out to be more thermally stable. Table 3 presents the temperature values and mass losses of biocomposites depending on the thermal degradation phase. According to the degradation temperatures of polyethylene and lignocellulosic material, it can be concluded that the decomposition of the main polymers of the biocomposites took place in phase III of the process, although the thermal processes observed in phase II could be related to the decomposition of cellulose and hemicelluloses. As indicated in the literature, the first DTG peak for biocomposites based on natural fibers and synthetic polymers begins at a temperature of about 250 °C and is a complex process depending on the stability of the fibers against the matrix [54]. Similar conclusions from research are presented by da Costa Borges et al. [55]. Their studies indicate that the observed degradation processes of biofilms produced based on plant particles, starting in the temperature range of 120–165 °C, are attributed to the loss of water and volatile compounds. Whereas the second phase of decomposition observed at temperatures up to 250 °C concerns the decomposition mainly of hemicellulose, which is a less stable organic substance.
Figure 6 shows the results of the SEM observation of the surface of the typical regions of non-modified polyethylene, polyethylene-maize, polyethylene-nigella, and polyethylene-flax materials for the observation magnification of 500×. The scale bar of 100 µm refers to all photos. In the biocomposites with the addition of black cumin, dark streaks are visible, which are probably the remains of the oily fraction enclosed in the structure of the material, which could have gotten into the composite from seed particles during the pressing process. Despite the fractions of non-structural substances visible in the biocomposite with the addition of black cumin particles, the expected high antioxidant properties of biocomposites for this plant were not observed (Table 4). In the case of other biocomposites, slight surface antioxidant properties were also identified. It seems that the cause of this phenomenon is illustrated by the morphological structure of biocomposites shown in SEM images. Both plant particles and non-structural fractions of plant origin are surrounded by polyethylene; therefore, potential antioxidant substances cannot move to the outer surface of the biocomposites. It can be assumed that these low antioxidant values of biocomposites result from the extraction of compounds from the pomace from the surface that has been opened at the cutting site.
In the studies on the degree of fouling of the biopolymer surface by fungi, a variable ability of selected fungal species to infect the surface of biocomposites was observed. Pure polyethylene was not susceptible to fungal fouling. Chaetomium globosum and Trichoderma viride, typical representatives of cellulolytic fungi, completely overgrew the samples containing corn particles. It is clearly seen that the addition of black cumin inhibited the growth of these fungi. Penicillium notatum also showed variable growth activity towards polymers containing plant particles (Figure 7). While the fungus completely overgrew the surfaces of biocomposites containing corn particles, in the case of additives in the form of black cumin and flax, a clear growth-inhibiting effect was observed. The only species insensitive to the antifungal effect of plant particles was Aspergillus niger, which completely overgrown the surface of biocomposites in a very short time. It can therefore be assumed that the metabolic activity of the fungi could have led to microdamage to the biocomposites, from which plant components were released, which on the one hand could have been a growth-stimulating nutrient, and on the other hand, as in the case of the biocomposite with black cumin, the components contained in it could lead to inhibition of sensitive fungal species. In addition, as reported in their research by Abdel-Latif et al. [56], oil obtained from black cumin seeds (Nigella sativa) has good fungicidal properties.
Of the factors analyzed influencing fungal overgrowth in biocomposites (Table 5), the highest percentage influence (48.8%) was shown by the type of biocomposite. On the other hand, the fungus species had an impact of 23.7%. It is worth mentioning that the influence of factors not included in this study (error) was only 11.9%. Analyzing the degree of fungal growth on the biopolymer surface after 12 days, it can be seen that, irrespective of the fungal species, the sample made of pure PE (homogeneous group A) was the least grown. Regardless of the type of additive (PE-M, PE-N, or PE-F), the highest degree of overgrowth was observed for the fungus Aspergillus niger (homogeneous group C). With regard to biocomposites, the PE-F variant proved to be the most susceptible to the effects of fungi (Table 6).
The use of plant raw materials for the production of biocomposites and biopolymers has become a frequent subject of scientific research in recent years. In addition, oilseed crops appear to be important because of their potential for use in the production of polyethylene-like materials [57]. FTIR spectrum analysis confirmed that the analyzed biocomposites contain active substances in the polyethylene matrix. The migration of active substances from the plant particles into the polyethylene matrix is also visible in the SEM images. These results correspond with the conclusions of other research authors [58,59]. The inclusion of plant particles or their extracts imparts new properties to the manufactured materials, which are referred to as active properties. Ordon et al. [60] showed that the addition of rosemary, raspberry, and pomegranate extracts to polyethylene imparts antimicrobial and antiviral properties to the biopolymer. Similar results were obtained by Solano et al. [61] by saturating a polyethylene matrix with essential oils. The results obtained indicate that the presence of lignocellulosic material can stimulate fungi to grow on the surface of the biocomposite; however, the non-structural substances present in the plant particles, especially in the nigella seeds, show an inhibitory effect on fungal growth. Undoubtedly, a valuable functionality of plant-particle-based biocomposites is their antioxidant properties, which are attracting interest in the area of food products. Wang et al. [62] pointed out that the antioxidant properties of polymers, however, depend on a number of factors; however, the types of extracts and the way the biomaterial is formed are key in this respect. In the present study, no significant antioxidant properties were obtained, which may suggest that the substances responsible for these functions do not diffuse to the surface of the biocomposite. Souza et al. [63] have indicated that the chemical environment of antioxidant substances determines their ability to migrate to the surface of biopolymers. The authors of this study indicate that certain chemicals can entrap antioxidant compounds in the polymer matrix.

4. Conclusions

Polymer biocomposites have been an important point in the development of sustainable materials science for several years. Restrictions on the use of environmentally unfriendly conventional plastics have increased industry interest in composites based on plant-based raw materials. The management of plant raw materials, especially waste materials, will, on the one hand, promote sustainable agriculture and, on the other hand, lead to the development of an industry producing environmentally friendly, biodegradable, and easy-to-manufacture biocomposites. The properties of such biocomposites depend on many factors, such as the chemical composition of plant particles, their volume fraction, geometry and orientation in the matrix, and the interaction between plant particles and the polymer. On the basis of the characteristics of the biocomposites produced, it should be concluded that the plant waste generated in the production of oil, used in the amount of 50% as a filler for biocomposites based on polyethylene, significantly reduces the crystallinity of the produced material and reduces its thermal stability. Despite its reduced thermal stability, the biocomposite can be used wherever the ambient temperature does not exceed 100 °C.
At the same time, it has been noted that molds are more likely to overgrow the surface of biocomposts due to the presence of lignocellulosic particles as a source of assimilable carbon. On the other hand, the presence of non-structural substances of plant origin, especially from black cumin waste confirmed by FTIR and SEM results, may confer fungal growth-inhibiting properties to biocomposites. Very poor antioxidant properties, resulting from the closure of the plant fraction in polyethylene, limit the functionality of this type of material as, for example, active biomaterials that prevent free radical processes. Although the structural and physical characteristics of the produced biocomposites were found to be inferior to those of polyethylene, efforts should be made to improve these characteristics. It seems that it is particularly important to choose the right proportions of plant particles, but also to determine their physico-chemical properties, as these will determine the properties of the entire biocomposite, and this will translate into potential application properties. The research presented here appears to be of interest because it allows the suitability of low-cost waste plant biomass to be assessed for the production of more sustainable biocomposites. However, in order to fully indicate the performance characteristics of such biocomposites, the authors of this paper will focus on the physico-chemical analysis of the raw plant material and its appropriate selection during subsequent experiments in order to obtain a biocomposite with properties no worse than those of pure polyethylene.

Author Contributions

Conceptualization, I.B. and P.B. (Piotr Borysiuk); methodology, I.B., K.L., P.B. (Piotr Borysiuk), K.R., M.N., S.B., B.A., M.B., K.K. and T.C.; software, I.B. and K.L; validation, I.B.; formal analysis, I.B. and M.N.; investigation, I.B. and K.L.; resources, I.B. and K.L.; data curation, I.B.; writing—original draft preparation, I.B.; writing—review and editing, M.N., K.L. and P.B. (Piotr Boruszewski); visualization, I.B.; supervision, I.B. and P.B. (Piotr Boruszewski); project administration, I.B., K.L. and P.B. (Piotr Boruszewski); funding acquisition, P.B. (Piotr Boruszewski). All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors of the publication would like to thank Leszek Danecki from Research & Development Center for Wood-Based Panels Sp. z o. o. in Czarna Woda for the preparation of biocomposites and the Institute of Wood Sciences and Furniture Warsaw University of Life Sciences for the support in financing research from the science development fund.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Food and Agriculture Organization of United Nations (FAO). The Future of Food and Agriculture—Trends and Challenges; FAO: Rome, Italy, 2017; ISBN 978-92-5-109551-5. Available online: https://www.fao.org/3/i6583e/i6583e.pdf (accessed on 25 August 2023).
  2. Babu, S.; Singh Rathore, S.; Singh, R.; Kumar, S.; Singh, V.K.; Yadav, S.K.; Yadav, V.; Raj, R.; Yadav, D.; Shekhawat, K.; et al. Exploring Agricultural Waste Biomass for Energy, Food and Feed Production and Pollution Mitigation: A Review. Bioresour. Technol. 2022, 360, 127566. [Google Scholar] [CrossRef]
  3. Winans, K.; Kendall, A.; Deng, H. The History and Current Applications of the Circular Economy Concept. Renew. Sustain. Energy Rev. 2017, 68, 825–833. [Google Scholar] [CrossRef]
  4. Koul, B.; Yakoob, M.; Shah, M.P. Agricultural Waste Management Strategies for Environmental Sustainability. Environ. Res. 2022, 206, 112285. [Google Scholar] [CrossRef] [PubMed]
  5. Ari, I.; Yikmaz, R.F. Chapter 4—Greening of Industry in a Resource- and Environment-Constrained World. In Handbook of Green Economics; Acar, S., Yeldan, E., Eds.; Academic Press: Cambridge, MA, USA, 2019; pp. 53–68. [Google Scholar] [CrossRef]
  6. Routray, W.; Orsat, V. Chapter 8—Plant By-Products and Food Industry Waste: A Source of Nutraceuticals and Biopolymers. In Handbook of Food Bioengineering, Food Bioconversion; Grumezescu, A.M., Holban, A.M., Eds.; Academic Press: Cambridge, MA, USA, 2017; pp. 279–315. [Google Scholar] [CrossRef]
  7. Gupte, A.P.; Basaglia, M.; Casella, S.; Favaro, L. Rice Waste Streams as a Promising Source of Biofuels: Feedstocks, Biotechnologies and Future Perspectives. Renew. Sustain. Energy Rev. 2022, 167, 112673. [Google Scholar] [CrossRef]
  8. Krasowska, M.; Kowalczyk-Sadowy, M. Evaluation of the possibility of using apple pomace for fertilizing purposes. J. Res. Appl. Agric. Eng. 2018, 63, 89–93. [Google Scholar]
  9. Galal, T.M.; Ali, E.F.; Eid, E.M.; Al-Yasi, H.M.; Magrashi, A.; Althobaiti, F.; Farahat, E.A. Evaluating the Nutrient Contents and Nutritive Value of Taif’s Rose (Rosa damascena Mill var. trigintipetala) Waste to Be Used as Animal Forage or Soil Organic Fertilizers. Agriculture 2022, 12, 1481. [Google Scholar] [CrossRef]
  10. Juárez-Alvarado, C.A.; Magniont, C.; Escadeillas, G.; Terán-Torres, B.T.; Rosas-Diaz, F.; Valdez-Tamez, P.L. Sustainable Proposal for Plant-Based Cementitious Composites, Evaluation of Their Mechanical, Durability and Comfort Properties. Sustainability 2022, 14, 14397. [Google Scholar] [CrossRef]
  11. Shih, Y.-F.; Cai, J.-X.; Kuan, C.-S.; Hsieh, C.-F. Plant Fibers and Wasted Fiber/Epoxy Green Composites. Compos. Part B Eng. 2012, 43, 2817–2821. [Google Scholar] [CrossRef]
  12. Elgamsy, R.; Allah Abo Elmagd, A.; Elrahman Mokhtar, A.M.; Khalid, I.; Taha, N.; Sadek, S.; Tawfic, M.L.; Attia, T.; Elsabbagh, A. Developing Fire Retardant Composites of Biodegradable Polyethylene Reinforced with Agricultural Wastes. Ain Shams Eng. J. 2022, 13, 101768. [Google Scholar] [CrossRef]
  13. Ashori, A. 2—Hybrid thermoplastic composites using nonwood plant fibers. Hybrid Polym. Compos. Mater. Prop. Characteris. 2017, 39–56. [Google Scholar] [CrossRef]
  14. Singh, A.; Gu, Y.; Castellarin, S.D.; Kitts, D.D.; Pratap-Singh, A. Development and Characterization of the Edible Packaging Films Incorporated with Blueberry Pomace. Foods 2020, 9, 1599. [Google Scholar] [CrossRef] [PubMed]
  15. Ballinas-Casarrubias, L.; Camacho-Davila, A.; Gutierrez Méndez, N.; Ramos-Sánchez, V.H.; Chávez-Flores, D.; Manjarrez-Nevárez, L.; Zaragoza-Galán, G.; González-Sanchez, G. Biopolymers from Waste Biomass—Extraction, Modification and Ulterior Uses. In Recent Advances in Biopolymers; InTech: Rijeku, Croatia, 2016. [Google Scholar] [CrossRef]
  16. Asim, M.; Saba, N.; Jawaid, M.; Nasir, M. 12—Potential of Natural Fiber/Biomass Filler-Reinforced Polymer Composites in Aerospace Applications. In Woodhead Publishing Series in Composites Science and Engineering; Jawaid, M., Thariq, M., Eds.; Woodhead Publishing Series in Composites Science and Engineering: Duxford, UK, 2018; pp. 253–268. [Google Scholar] [CrossRef]
  17. Arrakhiz, F.Z.; El Achaby, M.; Malha, M.; Bensalah, M.O.; Fassi-Fehri, O.; Bouhfid, R.; Benmoussa, K.; Qaiss, A. Mechanical and Thermal Properties of Natural Fibers Reinforced Polymer Composites: Doum/Low Density Polyethylene. Mater. Des. 2013, 43, 200–205. [Google Scholar] [CrossRef]
  18. Sullins, T.; Pillay, S.; Komus, A.; Ning, H. Hemp Fiber Reinforced Polypropylene Composites: The Effects of Material Treatments. Compos. Part B Eng. 2017, 114, 15–22. [Google Scholar] [CrossRef]
  19. Barczewski, M.; Aniśko, J.; Hejna, A.; Mysiukiewicz, O.; Kosmela, P.; Sałasińska, K.; Boczkowska, A.; Przybylska-Balcerek, A.; Stuper-Szablewska, K. Ground lemon and stevia leaves as renewable functional fillers with antioxidant activity for high-density polyethylene composites. Clean Technol. Environ. Policy 2023. [Google Scholar] [CrossRef]
  20. Jayaraman, K.; Bhattacharyya, D. Mechanical performance of woodfibre- waste plastic composite materials. Resour. Conserv. Recycl. 2004, 41, 307–319. [Google Scholar] [CrossRef]
  21. Fernandes, E.M.; Correlo, V.M.; Mano, J.F.; Reis, R.L. Novel cork polymer composites reinforced with short natural coconut fibres: E_ect of fibre loading and coupling agent addition. Compos. Sci. Technol. 2013, 78, 56–62. [Google Scholar] [CrossRef]
  22. Aji, I.S.; Zainudin, E.S.; Khalina, A.; Sapuan, S.M.; Khairul, M.D. Studying the e_ect of fiber size and fiber loading on the mechanical properties of hybridized kenaf/ PALF reinforced HDPE composite. J. Reinf. Plast. Compos. 2011, 30, 546553. [Google Scholar] [CrossRef]
  23. Banat, R.; Al-Rawashdeh, M.; Aljnaid, M. Thermal Properties of Olive Pomace Filled Linear Low Density Polyethylene Bio-Composites. J. Macromol. Sci. Part B 2021, 60, 957–969. [Google Scholar] [CrossRef]
  24. Gustafsson, J.; Landberg, M.; Bátori, V.; Åkesson, D.; Taherzadeh, M.J.; Zamani, A. Development of Bio-Based Films and 3D Objects from Apple Pomace. Polymers 2019, 11, 289. [Google Scholar] [CrossRef]
  25. Raschip, I.E.; Fifere, N.; Dinu, M.V. A Comparative Analysis on the Effect of Variety of Grape Pomace Extracts on the Ice-Templated 3D Cryogel Features. Gels 2021, 7, 76. [Google Scholar] [CrossRef]
  26. Kumar, A.; Weig, A.R.; Agarwal, S. Balancing Degradability and Physical Properties of Amorphous Poly(d,l-lactide) by Making Blends. Macromol. Mater. Eng. 2022, 307, 2100602. [Google Scholar] [CrossRef]
  27. Mecking, S. Polyethylene-like materials from plant oils. Philos. Trans. R. Soc. A 2020, 378, 20190266. [Google Scholar] [CrossRef] [PubMed]
  28. Samir, A.; Ashour, F.H.; Hakim, A.A.A.; Bassyouni, M. Recent advances in biodegradable polymers for sustainable applications. NPJ Mater. Degrad. 2022, 6, 68. [Google Scholar] [CrossRef]
  29. Prochon, M.; Dzeikala, O. Biopolymer Composites as an Alternative to Materials for the Production of Ecological Packaging. Polymers 2021, 13, 592. [Google Scholar] [CrossRef] [PubMed]
  30. Muthuraj, R.; Lacoste, C.; Lacroix, P.; Bergeret, A. Sustainable thermal insulation biocomposites from rice husk, wheat husk, wood fibers and textile waste fibers: Elaboration and performances evaluation. Ind. Crops Prod. 2019, 135, 238–245. [Google Scholar] [CrossRef]
  31. Kamran, M.J.; Jayamani, E.; Heng, S.K.; Wong, Y.C.; Rahman, M.R.; Al-Bogami, A.S.; Huda, D.; Bakri, M.K.B.; Rahman, M.M. Characterization and Comparative Study on Chemically Treated Luffa Fiber as Reinforcement for Polylactic Acid Bio-Composites. BioResources 2022, 17, 2576–2597. [Google Scholar] [CrossRef]
  32. Mirowski, J.; Oliwa, R.; Oleksy, M.; Tomaszewska, J.; Ryszkowska, J.; Budzik, G. Poly(vinyl chloride) Composites with Raspberry Pomace Filler. Polymers 2021, 13, 1079. [Google Scholar] [CrossRef]
  33. Betlej, I.; Salerno-Kochan, R.; Borysiuk, P.; Boruszewski, P.; Monder, S.; Krajewski, K.; Andres, B.; Krochmal-Marczak, B.; Pisulewska, E.; Danecki, L.; et al. Quality Parameters of PE–Pomace Based Membranes. Membranes 2022, 12, 1086. [Google Scholar] [CrossRef]
  34. Hindeleh, A.M.; Johnson, D.J. The resolution of multipeak data in fibre science. J. Phys. Appl. Phys. 1971, 4, 259–263. [Google Scholar] [CrossRef]
  35. Betlej, I.; Krajewski, K.; Borysiuk, P. An assessment of the susceptibility of bacterial cellulose films to fouling by mold fungi. Ann. WULS-SGGW For. Wood Technol. 2020, 110, 103–109. [Google Scholar] [CrossRef]
  36. Nenadis, N.; Tsimidou, M.Z. DPPH (2,2-di(4-tert-octylphenyl)-1-picrylhydrazyl) radical scavenging mixed-mode colorimetric assay(s). In Measurement of Antioxidant Activity & Capacity: Recent Trends and Applications; Apak, R., Capanoglu, E., Eds.; Wiley: Hoboken, NJ, USA, 2018; pp. 141–161. [Google Scholar] [CrossRef]
  37. Yijian, L.; Weichuan, D.; Demin, T.; Wei, Z.; Qiangguo, D. Space chargé distribution and crystalline structure in low density polyethylene (LDPE) blended with high density polyethylene (HDPE). Polym. Int. 2005, 54, 465–470. [Google Scholar]
  38. Liao, C.Z.; Tjong, S.C. Mechanical and thermal performance of high-density polyethylene/alumina nanocomposites. J. Macromol. Sci. B. 2012, 52, 812–825. [Google Scholar] [CrossRef]
  39. de Souza Coelho, C.C.; Soares Silva, R.B.; Carvalho, C.W.P.; Rossi, A.L.; Teixeira, J.A.; Freitas-Silva, O.; Correa Cabral, L.M. Cellulose nanocrystals from grape pomace and their use for the development of starch-based nanocomposite films. Int. J. Biol. Macromol. 2020, 159, 1048–1061. [Google Scholar] [CrossRef] [PubMed]
  40. Dintcheva, N.T.; Infurna, G.; Baiamonte, M.; D’Anna, F. Natural Compounds as Sustainable Additives for Biopolymers. Polymers 2020, 12, 732. [Google Scholar] [CrossRef]
  41. Bajsic, E.G.; Persic, A.; Jemric, T.; Buhin, J.; Kucic Grgic, D.; Zdraveva, E.; Zizek, K.; Holjevac Grguric, T. Preparation and Characterization of Polyethylene Biocomposites Reinforced by Rice Husk: Application as Potential Packaging Material. Chemistry 2021, 3, 1344–1362. [Google Scholar] [CrossRef]
  42. Cisneros-López, E.O.; Pérez-Fonseca, A.A.; González-García, Y.; Ramírez-Arreola, D.E.; González-Núñez, R.; Rodrigue, D.; Robledo-Ortíz, J.R. Polylactic acid–agave fiber biocomposites produced by rotational molding: A comparative study with compression molding. Adv. Polym. Technol. 2018, 37, 2528–2540. [Google Scholar] [CrossRef]
  43. Garbarczyk, J.; Borysiak, S. Polypropylene—Cellulose fibres composites. Part I. Influence of conditions of extrusion and injection processes on the structure of polypropylene matrix. Polimery 2004, 49, 541–546. [Google Scholar] [CrossRef]
  44. Grząbka-Zasadzińska, A.; Klapiszewski, Ł.; Borysiak, S.; Jesionowski, T. Thermal and Mechanical Properties of Silica–Lignin/Polylactide Composites Subjected to Biodegradation. Materials 2018, 11, 2257. [Google Scholar] [CrossRef]
  45. Smith, B. The infrared spectra of polymers II: Polyethylene. Spectroscopy 2021, 36, 22–25. [Google Scholar] [CrossRef]
  46. Kovács, R.L.; Csontos, M.; Gyöngyösi, S.; Elek, J.; Parditka, B.; Deák, G.; Kuki, Á.; Kéki, S.; Erdélyi, Z. Surface characterization of plasma-modified low density polyethylene by attenuated total reflectance fourier-transform infrared (ATR-FTIR) spectroscopy combined with chemometrics. Polym. Test. 2021, 96, 107080. [Google Scholar] [CrossRef]
  47. Rohman, A.; Sudjadi, D.W.; Lukitaningsih, E.; Rosman, A.S. Use of Fourier Transform Infrared Spectroscopy in Combination with Partial Least Square for Authentication of Black Seed Oil. Int. J. Food Prop. 2015, 18, 775–784. [Google Scholar] [CrossRef]
  48. Mohammed, S.J.; Amin, H.H.; Aziz, S.B.; Sha, A.M.; Hassan, S.; Abdul Aziz, J.M.; Rahman, H.S. Structural characterization, antimicrobial activity, and in vitro cytotoxicity effect of black seed oil. Evid Based Complement. Altern. Med. 2019, 2019, 6515671. [Google Scholar] [CrossRef]
  49. Alix, S.; Philippe, E.; Bessadok, A.; Lebrun, L.; Morvan, C.; Marais, S. Effect of chemical treatments on water sorption and mechanical properties of flax fibres. Bioresour. Technol. 2009, 100, 4742–4749. [Google Scholar] [CrossRef] [PubMed]
  50. Geminiani, L.; Campione, F.P.; Corti, C.; Luraschi, M.; Motella, S.; Recchia, S.; Rampazzi, L. Differentiating between Natural and Modified Cellulosic Fibres Using ATR-FTIR Spectroscopy. Heritage 2022, 5, 4114–4139. [Google Scholar] [CrossRef]
  51. Suárez, L.; Ortega, Z.; Romero, F.; Paz, R.; Marrero, M. Influence of Giant Reed Fibers on Mechanical, Thermal, and Disintegration Behavior of Rotomolded PLA and PE Composites. J. Polym. Environ. 2022, 30, 4848–4862. [Google Scholar] [CrossRef]
  52. Andrzejewski, J.; Krawczak, A.; Wesoły, K.; Szostak, M. Rotational molding of biocomposites with addition of buckwheat husk filler. Structure-property correlation assessment for materials based on polyethylene (PE) and poly(lactic acid) PLA. Compos. Part B 2020, 202, 108410. [Google Scholar] [CrossRef]
  53. Korol, J.; Lenża, J.; Formela, K. Manufacture and research of TPS/PE biocomposites properties. Compos. Part B 2015, 68, 310–316. [Google Scholar] [CrossRef]
  54. Monteiro, S.N.; Calado, V.; Rodriguez, R.J.S.; Margem, F.M. Thermogravimetric behavior of natural fibers reinforced polymer composites—An overview. Mater. Sci. Eng. A 2012, 557, 17–28. [Google Scholar] [CrossRef]
  55. da Costa Borges, M.A.; Sorigotti, A.R.; Paschoalin, R.T.; Júnior, J.A.P.; da Silva, L.H.D.; Dias, D.S.; Ribeiro, C.A.; de Araújo, E.S.; Resende, F.A.; da Silva Barud, H. Self-Supported Biopolymeric Films Based on Onion Bulb (Allium cepa L.): Gamma-Radiation Effects in Sterilizing Doses. Polymers 2023, 15, 914. [Google Scholar] [CrossRef]
  56. Abdel-Latif, E.F.; Abbas, K.A.; Abdelmontaleb, H.S.; Hamdy, S.M. Nigella sativa oil: A promising prospective antifungal agent in the manufacture of low-salt soft cheese. Ital. J. Food Saf. 2021, 10, 9862. [Google Scholar] [CrossRef]
  57. Alix, S.; Marais, S.; Morvan, C.; Lebrun, L. Biocomposite materials from flax plants: Preparation and properties. Compos. Part A Appl. Sci. Manuf. 2008, 39, 1793–1801. [Google Scholar] [CrossRef]
  58. Canales, D.; Montoille, L.; Rivas, L.M.; Ortiz, J.A.; Yañez-S, M.; Rabagliati, F.M.; Ulloa, M.T.; Alvarez, E.; Zapata, P.A. Fungicides Films of Low-Density Polyethylene (LDPE)/Inclusion Complexes (Carvacrol and Cinnamaldehyde) against Botrytis cinereal. Coatings 2019, 9, 795. [Google Scholar] [CrossRef]
  59. Al-Maqtari, Q.A.; Ali Mahdi, A.; Al-Ansi, W.; Mohammed, J.K.; Wei, M.; Yao, W. Evaluation of bioactive compounds and antibacterial activity of Pulicaria jaubertii extract obtained by supercritical and conventional methods. J. Food Meas. Charact. 2021, 15, 449–456. [Google Scholar] [CrossRef]
  60. Ordon, M.; Nawrotek, P.; Stachurska, X.; Mizielińska, M. Polyethylene Films Coated with Antibacterial and Antiviral Layers Based on CO2 Extracts of Raspberry Seeds, of Pomegranate Seeds and of Rosemary. Coatings 2021, 11, 1179. [Google Scholar] [CrossRef]
  61. Solano, A.C.V.; de Rojas Gante, C. Two different processes to obtain antimicrobial packaging containing natural oils. Food. Bioprocess Technol. 2012, 5, 2522–2528. [Google Scholar] [CrossRef]
  62. Wang, S.; Marcone, M.F.; Barbut, S.; Lim, L.-T. Fortification of dietary biopolymers-based packaging material with bioactive plant extracts. Food Res. Int. 2012, 49, 80–91. [Google Scholar] [CrossRef]
  63. Souza, V.G.L.; Rodrigues, P.F.; Duarte, M.P.; Fernando, A.L. Antioxidant Migration Studies in Chitosan Films Incorporated with Plant Extracts. J. Renew. Mater. 2018, 6, 548–558. [Google Scholar]
Figure 1. Biocomposites production methods.
Figure 1. Biocomposites production methods.
Coatings 13 01722 g001
Figure 2. Biocomposites: PE-N (HDPE + black cumin—nigella—seed pomace), PE-M (HDPE + maize seed pomace), PE-F (HDPE + flax seed pomace), PE—pure high-density polyethylene.
Figure 2. Biocomposites: PE-N (HDPE + black cumin—nigella—seed pomace), PE-M (HDPE + maize seed pomace), PE-F (HDPE + flax seed pomace), PE—pure high-density polyethylene.
Coatings 13 01722 g002
Figure 3. The XRD (a) and FTIR (b) curves of biocomposites, PE—pure polyethylene, PE—M (HDPE + maize seed pomace), PE-N (HDPE + black cumin-nigella-seed pomace), PE-F (HDPE + flax seed pomace).
Figure 3. The XRD (a) and FTIR (b) curves of biocomposites, PE—pure polyethylene, PE—M (HDPE + maize seed pomace), PE-N (HDPE + black cumin-nigella-seed pomace), PE-F (HDPE + flax seed pomace).
Coatings 13 01722 g003
Figure 4. DSC curves of tested iocomposites.
Figure 4. DSC curves of tested iocomposites.
Coatings 13 01722 g004
Figure 5. Thermogravimetric curves (a) and their derivatives (b) of the biocomposites.
Figure 5. Thermogravimetric curves (a) and their derivatives (b) of the biocomposites.
Coatings 13 01722 g005
Figure 6. The SEM results of the surface observation of non-modified polyethylene, polyethylene-maize, polyethylene-nigelle, and polyethylene-flax materials.
Figure 6. The SEM results of the surface observation of non-modified polyethylene, polyethylene-maize, polyethylene-nigelle, and polyethylene-flax materials.
Coatings 13 01722 g006
Figure 7. Results of the growth of biocomposites on surfaces by mold fungi.
Figure 7. Results of the growth of biocomposites on surfaces by mold fungi.
Coatings 13 01722 g007
Table 1. Degree of crystallinity of the biocomposite materials produced.
Table 1. Degree of crystallinity of the biocomposite materials produced.
Name *Degree of Crystallinity Xc (%)
PE52
PE-M40
PE-N33
PE-F42
* The names of materials are defined in the text.
Table 2. Thermodynamic parameters of the melting characteristics of the biocomposites produced.
Table 2. Thermodynamic parameters of the melting characteristics of the biocomposites produced.
NameTm (°C)Total Energy Consumption (mJ)ΔH (J/g)
PE109.94472.94122.21
PE-M107.36332.7582.77
PE-N106.24261.0265.58
PE-F106.66234.7765.95
Tm—melting temperature, ΔH—melting enthalpy.
Table 3. Thermal degradation phases and related temperatures and mass losses.
Table 3. Thermal degradation phases and related temperatures and mass losses.
NamePhase IPhase IIPhase III
Temperature (°C)Loss of Weight (%)Temperature (°C)Loss of Weight (%)Temperature (°C)Loss of Weight (%)
PE----465.3998.24
PE-M--300.1814.63466.2676.56
PE-N207.9714.80301.2839.17467.3720.43
PE-F----466.4184.16
Table 4. Antioxidant activity of biocomposites.
Table 4. Antioxidant activity of biocomposites.
NameDPPH Radical Scavenging Activity
(µmol Trolox/cm3)
PE0.02 A
PE-M0.43 B
PE-N0.64 C
PE-F0.51 D
A,B,C,D—homogeneous groups in Tukey’s test with α = 0.05.
Table 5. ANOVA for selected factors influencing the fouling by fungi.
Table 5. ANOVA for selected factors influencing the fouling by fungi.
Source of VariationSum of SquaresMean Sum of SquaresFisher’s F-TestSignificance LevelPercentage of Contribution
SSMSFpP [%]
type of biocomposite52,149.417,383.147.66690.00000048.8
species of fungi25,298.212,649.134.68560.00000023.7
type of biocomposite x specie of fungi16,707.32784.67.63560.00002715.6
error12,763.8364.7 11.9
Table 6. Comparison of fungal growth of biocomposite surfaces.
Table 6. Comparison of fungal growth of biocomposite surfaces.
Type of BiocompositesSpecie of FungiDegree of Fouling of the Biocomposite Surface by Fungi after 12 Days (%)Homogeneous Groups in Tukey’s Test with α = 0.05
PEAspergillus niger0.14A
PETrichoderma viride0.00A
PEPenicillium notatum0.00A
PEChaetomium globosum0.00A
PE-MAspergillus niger100.00D
PE-MTrichoderma viride1.18A
PE-MPenicillium notatum51.76B, C
PE-MChaetomium globosum100.00D
PE-NAspergillus niger100.00D
PE-NTrichoderma viride0.00A
PE-NPenicillium notatum83.98C, D
PE-NChaetomium globosum31.33A, B
PE-FAspergillus niger100.00D
PE-FTrichoderma viride100.00D
PE-FPenicillium notatum100.00D
PE-FChaetomium globosum100.00D
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Betlej, I.; Borysiuk, P.; Borysiak, S.; Rybak, K.; Nowacka, M.; Barlak, M.; Andres, B.; Krajewski, K.; Lipska, K.; Cebulak, T.; et al. Pomace from Oil Plants as a New Type of Raw Material for the Production of Environmentally Friendly Biocomposites. Coatings 2023, 13, 1722. https://doi.org/10.3390/coatings13101722

AMA Style

Betlej I, Borysiuk P, Borysiak S, Rybak K, Nowacka M, Barlak M, Andres B, Krajewski K, Lipska K, Cebulak T, et al. Pomace from Oil Plants as a New Type of Raw Material for the Production of Environmentally Friendly Biocomposites. Coatings. 2023; 13(10):1722. https://doi.org/10.3390/coatings13101722

Chicago/Turabian Style

Betlej, Izabela, Piotr Borysiuk, Sławomir Borysiak, Katarzyna Rybak, Małgorzata Nowacka, Marek Barlak, Bogusław Andres, Krzysztof Krajewski, Karolina Lipska, Tomasz Cebulak, and et al. 2023. "Pomace from Oil Plants as a New Type of Raw Material for the Production of Environmentally Friendly Biocomposites" Coatings 13, no. 10: 1722. https://doi.org/10.3390/coatings13101722

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