Next Article in Journal
Energy-Aware Camera Location Search Algorithm for Increasing Precision of Observation in Automated Manufacturing
Previous Article in Journal
Advanced Metamaterial-Integrated Dipole Array Antenna for Enhanced Gain in 5G Millimeter-Wave Bands
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Development of a Composite Filament Based on Polypropylene and Garlic Husk Particles for 3D Printing Applications

by
Cynthia Graciela Flores-Hernández
1,
Juventino López-Barroso
1,
Claudia Esmeralda Ramos-Galván
2,
Beatriz Adriana Salazar-Cruz
2,
María Yolanda Chávez-Cinco
2 and
José Luis Rivera-Armenta
2,*
1
División de Estudios de Posgrado e Investigación, Instituto Tecnológico de Querétaro/Tecnológico Nacional de México, Av. Tecnológico S/n Esq. Gral. Mariano Escobedo, Querétaro 76000, Qro., Mexico
2
Centro de Investigación en Petroquímica, Instituto Tecnológico de Ciudad Madero/Tecnológico Nacional de México, Pról. Bahía de Aldair y Ave. de las Bahías, Parque de la Pequeña y Mediana Industria, Altamira 89603, Tams., Mexico
*
Author to whom correspondence should be addressed.
Appl. Sci. 2024, 14(19), 9139; https://doi.org/10.3390/app14199139 (registering DOI)
Submission received: 3 September 2024 / Revised: 4 October 2024 / Accepted: 7 October 2024 / Published: 9 October 2024
(This article belongs to the Special Issue Advanced Composites Processing and Manufacturing)

Abstract

:
Lignocellulosic waste materials are among the most abundant raw materials on Earth, and they have been widely studied as natural additives in materials, especially for polymer composites, with interesting results when it comes to improving physiochemical properties. The main components of these materials are cellulose, hemicellulose, and lignin, as well as small amounts of other polysaccharides, proteins, and other extractives. Several kinds of lignocellulosic materials, mainly fibers, have been evaluated in polymer matrices, and recently, the use of particles has increased due to their high surface area. Garlic is a spice seed that generates a waste husk that does not have applications, and there are no reports of industrial use of this kind of lignocellulosic material. Additive manufacturing, also known as 3D printing, is a polymer processing technique that allows for obtaining complex shapes that are hard to obtain with ordinary techniques. The use of composites based on synthetic polymers and lignocellulosic materials is a growing field of research. In the present work, the elaboration and evaluation of 3D-printed polypropylene–garlic husk particle (PP-GHP) composites are reported. First, the process of obtaining a filament by means of a single extrusion was carried out, using different GHP contents in the composites. Once the filament was obtained, it was taken to a 3D printer to obtain probes that were characterized using differential scanning calorimetry (DSC), and thermogravimetric analysis (TGA) was performed with the aim of evaluating the thermal behavior of the 3D-printed PP-GHP composites. According to the obtained results, the crystallization process and thermal stability of the PP-GHP composites were modified with the presence of GHP compared with pristine PP. Dynamic mechanical analysis (DMA) showed that the addition of GHP decreased the storage modulus of the printed composites and that the Tan δ peak width increased, which was associated with an increase in toughness and a more complex structure of the 3D-printed composites. X-ray diffraction (XRD) showed that the addition of GHP favored the presence of the β-phase of PP in the printed composites.

1. Introduction

The use of lignocellulosic materials (fibers, cellulose, and nanofibers) as reinforcements in the preparation of biocomposites has been widely studied; interesting results have been reported, and the physical–mechanical properties of polymer matrices have been improved in the existing literature. On the other hand, there are few reports on the use of these kinds of materials as additives applied in 3D printing, which generally provide greater rigidity and mechanical resistance [1]. The development of sustainable and renewable materials from biomass materials for the 3D printing process has become an interesting option [2]. Cellulose, cellulose nanofibers, microcrystalline cellulose, and cellulose nanocrystals are some of the most studied biomass materials in 3D printing [2,3]. In 3D printing, the use of thermoplastic biocomposite filaments is generating interest because of their promise for decreasing the cost of materials, decreasing environmental impacts, improving processing conditions, and improving mechanical properties [4]. There have been studies on the production of filaments using PP composites with plant fibers such as bamboo, ramie [5], wool [6], and coconut fiber [7] to obtain biocomposites that are applicable in 3D printing processes due to their viscosity and good behavior in the printing process, as well as their thinning, resolution, and mechanical resistance [8,9]. Three-dimensional printing or additive manufacturing is a polymer processing technology that helps to obtain complex shape pieces without it being necessary to use mold tools, unlike in extrusion, injection, or molding processes, and there are low costs for manufacturers and less impact on the environment [10,11]. This process can be applied in areas such as the biomedical, aerospace, and automotive industries, among others [12,13]. One of the first application areas was in weaving technology due to the fact that the textile industry plays an important role in producing many necessary ordinary items, and the development of 3D fiber-reinforced composites has since become a new technology to supplement business strategies [14].
Usually, 3D printing involves processing simple materials. The most common kinds of polymers that are processed by 3D printing are thermoplastics, which can be processed through melt mixing and extrusion due to their simple preparation process [15]. However, due to the need to improve the properties of filaments, the development of composite materials for filaments processed via 3D printing has gained interest, presenting an opportunity to improve the available commercial filaments. The most important challenge in the development of composite materials for 3D printing is our lack of knowledge regarding the behavior of polymers during the process and how the addition of a reinforcer or filler modifies the polymer [16].
PP is a thermoplastic polymer with a wide range of applications and is used in 3D printing as a filament with different matrix properties. For instance, a lower melt flow index (MFI) is often used due to its high molecular weight, thus benefitting the 3D printing process.
There are works that report the use of fibers designed to reinforce recycled PP (r-PP) for 3D printing filaments. Fiberglass has been reported as a sustainable material; in addition to improving the thermal stability of rPP, it has tensile properties that have been associated with increases in Young’s modulus of up to 40% [17]. Another work reported that the addition of glass fibers to PP filaments can be an option for parts used in the automotive industry, indicating that they can significantly improve the energy absorption capacity, which is attributed to the textile architecture that is formed and allows it to withstand higher loads in different directions [16]. Another type of fiber that has been studied as a reinforcement in thermoplastic polymers is carbon fiber; it was found that 3D-printed parts with carbon fiber showed greater resistance to failure due to a good interaction with the polymer matrix [18]. Rice husk is a waste product that has been used as a reinforcement in polymers for filaments for the 3D printing process; research has found that the tensile strength of this r-PP is lower, a behavior that can be attributed to a weak interfacial attraction between the polymer and the particles [19]. Another type of lignocellulosic material that has been studied as a reinforcement in PP is cocoa shell, which was found to decrease the thermal stability of PP; in addition, the crystallization and fusion processes were affected by the addition of this type of particle, but the mechanical properties increased depending on the printing conditions, such as the loading direction and raster angle, among other conditions [20]. Inorganic particles such as CaCO3 have also been added to PP to obtain PP composite filaments [21].
Some commercial filament polymers, such as buta-1,3-diene;prop-2-enenitrile;styrene (ABS), poly (vinyl acetate) (PVA), polycaprolatone (PCL), and nylon, are available [22]. One of the most studied polymers for 3D printing is poly (lactic acid) (PLA) [23], and some other composites have been reported. PLA-based composites with almond shell particles were shown to have improvements in mechanical properties, such as shore D hardness, Young’s modulus, and compression, which indicated that these properties depend on the printing pattern [24]. Among the most commonly reported polymer matrices in composites for 3D printing, PP is a widely used polymer due to its properties such as easy processing and low density [25]. Several works report the reinforcement of PP with natural fibers, such as wheat straw, bagasse, sunflower stems, and corn husks [26,27,28].
Garlic husk (GH) is a lignocellulosic waste material obtained from garlic seeds and is widely used in the food industry as a spice, as a vegetable coating, for the extraction of antioxidants, and for the extraction of contaminants in water [29,30,31]. However, research on the use of this waste material as a polymer additive is lacking. Our group reported the use of garlic husk particles (GHPs) in thermoplastic composites with interesting results, mainly in relation to the thermal properties, which were improved in comparison with those of pristine PP [32,33].
Considering the growing interest in the use of lignocellulosic waste as an additive in thermoplastic polymers and the fact that the use of PP filled with garlic husk particles (GHPs) in 3D printing is unexplored, the present work reports the use of GHPs as a reinforcement in a PP matrix to obtain composite PP–GHP filaments with different contents, all of which were developed using a single extrusion. The filaments were used to obtain 3D-printed composite specimens, whose thermal properties were evaluated using thermogravimetric analysis (TGA), differential scanning calorimetry (DSC), and dynamic mechanical analysis (DMA). X-ray diffraction (XRD) was also carried out to evaluate the crystallinity of the 3D-printed specimens. The aim was to obtain a sustainable material and find industrial applications for lignocellulosic waste materials such as GHPs. The results obtained herein for the thermal stability, thermal behavior, and viscoelastic properties are discussed.

2. Materials and Methods

2.1. Materials

A melt flow index of 20 g/10 min (ASTM-D1238 [34]) and a density of 0.9 g/cm3 (ASTM D792A [35]) were used for PP. This material was provided by Indelpro S.A. de C.V., Altamira, Tams., Mexico. Mexican garlic husk particles (GHPs) were collected domestically and received the following treatment: washing with distilled water and drying in an oven at 40 °C; later, they were milled to reduce the particle size until a homogeneous particle size of 53 μm (mesh 270 ASTM) was obtained. The reason for using small particles was because small particles have a higher surface area than that of fibers, generating a better interaction between the matrix and reinforcer particles. When particles are used in composite filaments for 3D printing, the printed pieces have less porosity than that of composite filaments filled with fibers [18,36].

2.2. Composite Filament Preparation

The filaments were obtained using a Beutelspacher single extruder (Beutelspacher, Mexico City, Mexico) and a processing temperature profile of 160, 170, 175, and 180 °C and a screw speed of 30 rpm. The GHP contents added to the composites were 1, 3, and 5 phr (parts per hundred of resin). Table 1 reports the formulations that were prepared. The content of additives was varied considering the convention for establishing the concentration of additives in composite formulations based on phr. The filaments were extruded while keeping a stable diameter of 1.8 mm; then, they were taken to a Flashforge 3D printer (model: Creator Pro 2 (Hangzhou, China)). The printing conditions were a temperature of 240 °C in the nozzle and 100 °C as the bed temperature, a raster angle of 90°, a filling density of 100%, and a printing speed of 60 mm/s. Figure 1 presents the process of obtaining the filaments, the printing process, and the obtained specimens.

2.3. 3D-Printed Composite Characterization

The obtained pieces were analyzed in a Perkin Elmer DSC (model: DSC8000, Waltham, MA, USA), where a heating–quenching–heating cycle was carried out. The first heating step was used to delete the thermal history of the material, and heated the material from −30 to 230 °C, with a heating rate of 20 °C/min; then, it was kept for 5 min at 230 °C, quenched with a heating rate of 20 °C/min until the temperature reached −30 °C, and kept for 5 min at this temperature. The second heating cycle ranged from −30 to 230 °C with a heating rate of 10 °C/min, and the quenching and second heating cycle were used to calculate the crystallization temperature (Tc), crystallization enthalpy (∆Hc), melting temperature (Tm), and melting enthalpy (∆Hm) to evaluate the crystallization process. In addition, the degree of crystallinity (Xc) was calculated using the following formula:
Xc = (∆Hm/∆) × 100
where ∆H° is the melting enthalpy per mass unit of the 100% crystalline PP, with a value of 207 J/g [37] and ∆Hm corresponds to the melting enthalpy. Each sample was tested in duplicate.
The thermogravimetric behavior was evaluated using a TA Instruments Q600 Simultaneous TGA/DSC (SDT) (New Castle, DE, USA) in a temperature range from 30 to 600 °C, with a heating rate of 10 °C/min in a N2 atmosphere. The dynamic mechanical analysis (DMA) was carried out using a TA Instruments Q800 (New Castle, DE, USA) with a temperature range from −40 to 150 °C and a heating rate of 5 °C/min in the multifrequency mode with a frequence of 1 Hz and a dual-cantilever clamp. The X-ray diffraction (XRD) analysis was performed using an Empyrean Malvern Panalytical diffractometer (Malvern, UK), with 45 kV and 40 mA, a Bragg–Bretano configuration, CuKα1 (λ = 1.5406 Å) radiation, 0.0263° 2θ, and an interval from 4 to 60° 2θ.

3. Results

3.1. Results of the DSC Characterization

The crystallization and melting behaviors of the PP and PP–GHP composites during 3D printing were studied using DSC. Figure 2 shows that the typical DSC thermograms for the 3D-printed PP and PP-GHP composites presented a similar path; the quenching cycle and second heating cycle were used to calculate Tc, ∆Hc, and Tm, and those data are reported in Table 2. It was possible to observe that the Tc value of the 3D-printed PP–GHP composites was not affected by the presence of GHPs, suggesting that the crystallization rate was not affected. However, ∆Hc showed an increase for the 3D-printed PP–GHP composite compared with that of the 3D-printed PP, but with a higher content, and the enthalpy decreased to a value similar to that of pristine PP. This behavior can be associated with the threshold percolation involved in the generation of a modification of this property. Regarding the melting process, the presence of GHPs caused a decrease in Tm compared with the value in pristine PP; a lower GHP content resulted in a greater effect (8 °C), which was associated with the fact that the presence of particles advanced the fusion process, and the crystallinity content increased when using a content of 3 phr. ∆Hm increased with the presence of GHPs when the content was 1 or 3 phr, and with 5 phr, it decreased to a value similar to that of pristine PP.
On the other hand, the percentage crystallinity increased compared with that of pristine PP for the 3D-printed PP–1GHP and PP–3GHP composites, showing an increase of up to 15% for the PP–3GHP composite, but it decreased for the 3D-printed PP–5GHP composite, to a values similar to that of pristine PP, meaning that GHPs generated a nucleating effect in the PP matrix. This was associated with an improvement in the interfacial interaction between the PP matrix and GHPs. The later increase in the GHP content and decrease in crystallinity could have been due to a threshold percolation limit, meaning that after a certain particle content, the matrix was not positively affected. One of the main reasons to provide a matrix modifi-cation it the adhesion between particles and matrix, when this adhesion is low, void formation increases, and the presence of crack nucleation leads to smaller property modifications [15]. A similar behavior was already reported by our group in a previous work in compounds prepared from PP and GHPs through compression molding. However, the highest values reported were with a higher GHP content (8 phr) [32]. This difference could be associated with the processing conditions used to obtain the filaments; when they flowed through the printer nozzle, a higher shear rate was generated, meaning that the particles could be better oriented in the PP matrix to favor the generation of a greater number of crystalline zones. However, this behavior has not been reported in works that obtained PP composite filaments reinforced with waste lignocellulosic particles. Previous work reported that the addition of untreated and chemically treated bamboo fibers to PP and PLA had a nucleating effect, accelerating the crystallization speed and increasing the crystallinity of the PP/PLA bamboo fiber filaments for 3D printing.
The crystallinity of 3D-printed composites is an important property, since it can be associated with the shrinkage capacity of a material when deposited on a printing bed. Morales et al. (2021) [19] reported that a decrease in the crystallization temperature of composite PP–rice husk filaments reduced the presence of shrinkage during the printing process in comparison with that of r-PP. Other reports indicated that the presence of amorphous compounds, such as hemicellulose and lignin, produced a decrease in the content of crystalline zones in PP–cocoa shell filaments for 3D printing [20].

3.2. Results of TGA Characterization

TGA was used in the present work to evaluate the thermal stability of materials and the effect of the presence of GHPs in the thermal decomposition of 3D-printed composites. In the TGA thermogram (Figure 3), it is possible to observe that the addition of a low content of GHPs did not have a significant effect on the thermal stability of the pristine PP matrix because the 3D-printed PP–1GHP composite showed a similar decomposition temperature to that of pristine PP, but the 3D-printed PP–3GHP and PP–5GHP composites showed an increase in the decomposition temperature, which could indicate an improvement in the thermal stability of the 3D-printed composites. This behavior can be attributed to the residue from char formation in the material from the GHPs, which acted as a barrier that inhibited free heat flow to slow the thermal decomposition of the material. A similar behavior was reported with other kinds of lignocellulosic materials in 3D-printed PP composites [38].
On the other hand, Figure 4 shows a derivative curve from the TGA, where the peak represents the medium decomposition temperature of the material, which is associated with the decomposition susceptibility or the speed of variation of weight with respect to the temperature of the material. It is possible to observe that the temperature peaks in the DTG curves were similar for the 3D-printed PP and PP–GHP composites. The main difference was that the 3D-printed PP–3GHP composite showed a higher quotient value at its peak, which indicated that this material presented the highest susceptibility to weight loss (it decomposed faster) in comparison with the 3D-printed PP and other PP–GHP composites. Previous works reported the opposite behavior, with the presence of lignocellulosic materials such as bamboo fibers in a 3D-printed PP matrix resulting in a lower decomposition temperature than that of 3D-printed lignocellulosic composites [24].

3.3. Results of DMA Characterization

DMA is a powerful tool for evaluating the properties of composites. It consists of applying an oscillating force to a sample and evaluating the material’s response. The capacity to absorb or dissipate energy is reported. Furthermore, the damping factor or Tan δ that is obtained from the loss ratio and the storage modulus can help in the evaluation of the dispersion level when a filler or reinforcer is added to a polymer matrix. Figure 5 depicts the storage modulus versus temperature for the 3D-printed PP and PP–GHP composites, and it can be clearly observed that the 3D-printed PP–1GHP composite showed a similar behavior to that of the 3D-printed pristine PP, but when the GHP content was increased, there was a significant decrease in the modulus, indicating that there is a loss of capacity to support stress. It is possible that there was an increase in the sliding or flow to the polymer chains and particles, which was reflected in the decrease in the storage modulus. A diminishing storage modulus in polymer composite materials was previously reported, and it was found that, with an increase in temperature, the composites exhibited more ductility and became softer than the bulk matrix, reflecting lower storage modulus values [39]. Another observation is that the storage modulus was higher when the temperature was increased for all 3D-printed composites compared with the pristine PP; this could be attributed to the motions of the polymer matrix being restricted due to the presence of GHPs with the increase in temperature.
In composites, there is a damping factor due to the nature of the matrix and filler materials, the friction generated due to slipping in the resin/fiber interface, energy dissipation at cracks, and delaminations produced at damaged locations, in addition to viscoplastic and thermoelastic damping. Figure 6 depicts the Tan δ curves for the 3D-printed PP and PP–GHP composites, and it is possible to observe two transitions. The first transition occurred around 10 °C and was associated with the glass transition of the polymer matrix; this did not show a significant change when GHPs were added, but the peak height was lower for the 3D-printed pristine PP compared with that of the 3D-printed composites. This was because the presence of GHPs restricted the mobility of segments of the amorphous phase of PP. The height of the Tan δ peak also varied when GHPs were added. The higher value of the 3D-printed PP–3GHP composite indicated that the particles increased the mobility of the chain segments in the amorphous phase, and the composites had a nonelastic strain component. It has been reported that a higher Tan δ peak value is associated with a less elastic than viscous nature compared with that of pristine PP [40]. Another interesting observation was that the width of the Tan δ peak increased for the 3D-printed PP–GHP composites, which was indicative of a more complex structure and suggested that there were molecular relaxations in the 3D-printed composites that were not present in pristine PP because the peak became broader with the increase in GHPs. The second transition was observed around 80 °C; this was associated with the crystalline structure of PP and was related to the laminar and rotational sliding of ordered crystalline structures. The Tan δ value increased for the 3D-printed composites, indicating that there was a more ordered structure in the filaments. This was in accordance with the DSC results, indicated that the crystallinity of the 3D-printed PP–GHP composites had a high percentage in comparison with that of the 3D-printed pristine PP. Similar results were reported before, attributing this to the relaxation of restricted amorphous PP macromolecules in the crystalline phase and to mechanisms of lamellar slip and rotation in the crystalline phase [41].

3.4. Results of XRD Characterization

Figure 7 presents the X-ray diffraction pattern for the 3D-printed PP and PP–GHP composites, in which typical signals associated with the crystalline structure of PP can be identified. The diffraction peaks located at ~13.8, 16.6, 18.3, 20.87, 21.58, 25.21, and 28.20° for 2θ were associated with the monoclinic α phase of PP and the (110), (040), (130), and (060) planes, while the peak of diffraction located at ~15.8° for 2θ was related to the β-hexagonal phase [42,43,44]. The presence of the β phase in the PP filament generally occurs under special processing conditions or due to the addition of a nucleating agent [45]. Therefore, processing in the extruder and subsequent processing in the 3D printer meant that this type of structure was favored, and the presence of GHPs could also generate this effect, as has been previously reported. It is important to mention that this crystalline phase was maintained for the 3D-printed PP–GHP composites. There was a decrease in the intensities of the peaks in the 3D-printed composites compared with that in the 3D-printed pristine PP, which represented a change in the crystallite size. On the other hand, all 3D-printed PP–GHP composites showed a shift at smaller angles, which suggested that there was an expansion of the basal space between the crystalline PP structures, which was attributed to the penetration of the GHPs into the structure. This behavior is consistent with what has been previously reported [46].

4. Conclusions

According to the obtained results, the following conclusions were found:
-
It is possible to obtain composite PP–GHP filaments through extrusion and use a 3D printing process to obtain probes with interesting properties in comparison with those of pristine PP.
-
The DSC results showed that the degree of crystallinity increased for PP–GHP composites compared with those of pristine PP. There was a higher Xc value for the 3D-printed PP–3GHP composites, and an increase in the content of GHPs generated a decrease in Xc.
-
The DTG curves showed that the thermal stability was affected by the addition of GHPs, and it was higher for the 3D-printed PP–3GHP composites.
-
The DMA results indicated that the storage modulus decreased with the addition of GHPs to PP, indicating that the stiffness decreased, and the Tan δ curve helped demonstrate that the 3D-printed composites had a more complex structure because the peak width increased for the 3D-printed composites in comparison with that of the 3D-printed pristine PP. This was related to molecular relaxations in the 3D composites that were not present in PP.
-
The XRD results corroborated the presence of the monoclinic α phase of PP and the β-hexagonal phase, and there was a slight displacement of the signals of the PP filament to lower angles on the 2θ scale when GHPs are added to PP, which was related to changes in the structure of the 3D-printed composites.

Author Contributions

Conceptualization, C.G.F.-H., J.L.-B. and J.L.R.-A.; methodology, C.G.F.-H., J.L.-B., B.A.S.-C., C.E.R.-G., M.Y.C.-C. and J.L.R.-A.; software, B.A.S.-C., C.E.R.-G. and M.Y.C.-C.; validation, B.A.S.-C., C.E.R.-G. and M.Y.C.-C.; formal analysis, B.A.S.-C., C.E.R.-G. and M.Y.C.-C.; investigation, J.L.-B., C.G.F.-H., B.A.S.-C., M.Y.C.-C. and C.E.R.-G.; resources, J.L.-B., C.G.F.-H. and J.L.R.-A.; data curation, J.L.-B., B.A.S.-C. and M.Y.C.-C.; writing—original draft preparation, C.G.F.-H., J.L.-B., B.A.S.-C., C.E.R.-G., M.Y.C.-C. and J.L.R.-A.; writing—review and editing, C.G.F.-H. and J.L.R.-A.; visualization, J.L.-B. and C.E.R.-G.; supervision, C.E.R.-G., M.Y.C.-C. and B.A.S.-C.; project administration, C.G.F.-H. and J.L.R.-A.; funding acquisition, J.L.R.-A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Tecnológico Nacional de México (TECNM), grant number 16612.23-P.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Acknowledgments

The authors wish to thank Sarahi Gallardo Ponce and Ana Cristal del. Valle Gonzalez for technical support in the characterization of the filaments and 3D printing process.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zarna, C.; Opedal, M.T.; Echtermeyer, A.T.; Chinga-Carrasco, G. Reinforcement ability of lignocellulosic components in biocomposites and their 3D printed applications–a review. Compos. Part C Open Access 2021, 6, 100171. [Google Scholar] [CrossRef]
  2. Ji, A.; Zhang, S.; Bhagia, S.; Yoo, C.G.; Ragauskas, A.J. 3D printing of biomass-derived composites: Application and characterization approaches. RSC Adv. 2020, 10, 21698–21723. [Google Scholar] [CrossRef]
  3. Singh, B.; Kumar, R.; Chohan, J.S. Polymer matrix composites in 3D printing: A state of art review. Mater. Today Proc. 2020, 33, 1562–1567. [Google Scholar] [CrossRef]
  4. Mazzanti, V.; Malagutti, L.; Mollica, F. FDM 3D printing of polymers containing natural fillers: A review of their mechanical properties. Polymers 2019, 11, 1094. [Google Scholar] [CrossRef] [PubMed]
  5. Cai, R.; Wen, W.; Wang, K.; Peng, Y.; Ahzi, S.; Chinesta, F. Tailoring interfacial properties of 3D-printed continuous natural fiber reinforced polypropylene composites through parameter optimization using machine learning methods. Mater. Today Commun. 2022, 32, 103985. [Google Scholar] [CrossRef]
  6. Röhl, V.; Müssig, J. Wool fiber-reinforced thermoplastic polymers for injection molding and 3D-printing. In Wool Fiber Reinforced Polymer Composites; Woodhead Publishing: Sawston, UK, 2022; pp. 351–386. [Google Scholar]
  7. Gama, N.; Magina, S.; Barros-Timmons, A.; Ferreira, A. Enhanced compatibility between coconut fibers/PP via chemical modification for 3D printing. Prog. Addit. Manuf. 2022, 7, 213–223. [Google Scholar] [CrossRef]
  8. Bi, X.; Huang, R. 3D printing of natural fiber and composites: A state-of-the-art review. Mater. Des. 2022, 222, 111065. [Google Scholar] [CrossRef]
  9. Le Duigou, A.; Correa, D.; Ueda, M.; Matsuzaki, R.; Castro, M. A review of 3D and 4D printing of natural fibre biocomposites. Mater. Des. 2020, 194, 108911. [Google Scholar] [CrossRef]
  10. Wittbrodt, B.T.; Glover, A.G.; Laureto, J.; Anzalone, G.C.; Oppliger, D.; Irwin, J.L.; Pearce, J.M. Life-cycle economic analysis of distributed manufacturing with open-source 3-D printers. Mechatronics 2013, 23, 713–726. [Google Scholar] [CrossRef]
  11. Kreiger, M.; Pearce, J.M. Environmental Impacts of Distributed Manufacturing from 3-DPrinting of Polymer Components and Products, in Symposium D/G Materials for Sustainable Development—Challenges and Opportunities. MRS Online Proc. Libr. 2013, 1492, 85–90. [Google Scholar] [CrossRef]
  12. Decuir, F.; Phelan, K.; Hollins, B.C. Mechanical strength of 3-D printed filaments. In Proceedings of the 2016 32nd Southern Biomedical Engineering Conference (SBEC), Shreveport, LA, USA, 11–13 March 2016; pp. 47–48. [Google Scholar]
  13. Guvendiren, M.; Molde, J.; Soares, R.M.; Kohn, J. Designing biomaterials for 3D printing. ACS Biomater. Sci. Eng. 2016, 2, 1679–1693. [Google Scholar] [CrossRef] [PubMed]
  14. Golra, O.A.; Tariq, J.; Ehsan, N.; Mirza, E. Strategy for introducing 3D fiber reinforced composites weaving technology. Procedia Technol. 2012, 1, 211–216. [Google Scholar] [CrossRef]
  15. Valino, A.D.; Dizon, J.R.C.; Espera Jr, A.H.; Chen, Q.; Messman, J.; Advincula, R.C. Advances in 3D printing of thermoplastic polymer composites and nanocomposites. Prog. Poly Sci. 2019, 98, 101162. [Google Scholar] [CrossRef]
  16. Hufenbach, W.; Böhm, R.; Thieme, M.; Winkler, A.; Mäder, E.; Rausch, J.; Schade, M. Polypropylene/glass fibre 3D-textile reinforced composites for automotive applications. Mater. Des. 2011, 32, 1468–1476. [Google Scholar] [CrossRef]
  17. Kristiawan, R.B.; Rusdyanto, B.; Imaduddin, F.; Ariawan, D. Glass powder additive on recycled polypropylene filaments: A sustainable material in 3D printing. Polymers 2021, 14, 5. [Google Scholar] [CrossRef]
  18. Kong, X.; Luo, J.; Luo, Q.; Li, Q.; Sun, G. Experimental study on interface failure behavior of 3D printed continuous fiber reinforced composites. Addit. Manuf. 2022, 59, 103077. [Google Scholar] [CrossRef]
  19. Morales, M.A.; Atencio Martinez, C.L.; Maranon, A.; Hernandez, C.; Michaud, V.; Porras, A. Development and characterization of rice husk and recycled polypropylene composite filaments for 3D printing. Polymers 2021, 13, 1067. [Google Scholar] [CrossRef]
  20. Morales, M.A.; Maranon, A.; Hernandez, C.; Porras, A. Development and characterization of a 3D printed cocoa bean shell filled recycled polypropylene for sustainable composites. Polymers 2021, 13, 3162. [Google Scholar] [CrossRef]
  21. Kumar, S.; Singh, R. On Preparation of CaCO3 Reinforced Polypropylene Composite With 3D Printing for Biomedical Application. Encycl. Mater. Plast. Polym. 2022, 1, 378–385. [Google Scholar]
  22. Tian, X.; Liu, T.; Yang, C.; Wang, Q.; Li, D. Interface and performance of 3D printed continuous carbon fiber reinforced PLA composites. Compos. Part A Appl. Sci. Manufact. 2016, 88, 198–205. [Google Scholar] [CrossRef]
  23. Long, H.; Wu, Z.; Dong, Q.; Shen, Y.; Zhou, W.; Luo, Y.; Zhang, C.; Dong, X. Mechanical and thermal properties of bamboo fiber reinforced polypropylene/polylactic acid composites for 3D printing. Polym. Eng. Sci. 2019, 59 (Suppl. S2), E247–E260. [Google Scholar] [CrossRef]
  24. Singh, R.; Kumar, R.; Singh, M.; Preet, P. On compressive and morphological features of 3D printed almond skin powder reinforced PLA matrix. Mater. Res. Express 2020, 7, 025311. [Google Scholar] [CrossRef]
  25. Ashori, A.; Nourbakhsh, A. Bio-based composites from waste agricultural residues. Waste Manag. 2010, 30, 680–684. [Google Scholar] [CrossRef] [PubMed]
  26. Hansen, H.K.; Arancibia, F.; Gutiérrez, C. Adsorption of copper onto agriculture waste materials. J. Hazard. Mater. 2010, 180, 442–448. [Google Scholar] [CrossRef] [PubMed]
  27. Hernández-Varela, J.D.; Chanona-Pérez, J.J.; Benavides, H.A.C.; Sodi, F.C.; Vicente-Flores, M. Effect of ball milling on cellulose nanoparticles structure obtained from garlic and agave waste. Carbohydr. Polym. 2021, 255, 117347. [Google Scholar] [CrossRef] [PubMed]
  28. Kallel, F.; Bettaieb, F.; Khiari, R.; García, A.; Bras, J.; Chaabouni, S.E. Isolation and structural characterization of cellulose nanocrystals extracted from garlic straw residues. Ind. Crop. Prod. 2016, 87, 287–296. [Google Scholar] [CrossRef]
  29. Chen, X.; Yin, L.; Zhou, H.; Liu, J.; Li, X.; Ai, X.; Huang, K. Efficient removal of lead from washing effluent of lead-contaminated soil with garlic peel. Chem. Res. Chin. Univ. 2018, 34, 1020–1027. [Google Scholar] [CrossRef]
  30. Hawal, L.; Al-Sulttani, A.O.H.; Kariem, N.O. Chromium elimination from contaminated soil by electro-kinetic remediation, using garlic peels powder. J. Ecol. Eng. 2021, 22, 252–259. [Google Scholar] [CrossRef]
  31. Liu, W.; Liu, Y.; Tao, Y.; Yu, Y.; Jiang, H.; Lian, H. Comparative study of adsorption of Pb (II) on native garlic peel and mercerized garlic peel. Env. Sci. Pollut. Res. 2014, 21, 2054–2063. [Google Scholar] [CrossRef]
  32. Villalobos-Neri, E.E.; Macchlesh Del Pino-Perez, L.A.; Velasco-Ocejo, H.A.; Rivera-Armenta, J.L.; Espindola-Flores, A.C. Preparation and Evaluation of Thermal Properties of Composites Based on Polypropylene Reinforced with Garlic Husk Particles (GHP). J. Nat. Fibers 2023, 20, 2145409. [Google Scholar] [CrossRef]
  33. Flores-Hernández, C.G.; López-Barroso, J.; Salazar-Cruz, B.A.; Saucedo-Rivalcoba, V.; Almendarez-Camarillo, A.; Rivera-Armenta, J.L. Evaluation of Starch–Garlic Husk Polymeric Composites through Mechanical, Thermal, and Thermo-Mechanical Tests. Polymers 2024, 16, 289. [Google Scholar] [CrossRef] [PubMed]
  34. ASTM-D1238; Standard Test Method for Melt Flow Rates of Thermoplastics by Extrusion Plastometer. American Society for Testing and Materials: West Conshohocken, PA, USA, 2023.
  35. ASTM D792A; Standard Test Methods for Density and Specific Gravity (Relative Density) of Plastics by Displacement. American Society for Testing and Materials: West Conshohocken, PA, USA, 2020.
  36. Xanthos, M. (Ed.) Functional Fillers for Plastics; John Wiley & Sons: Hoboken, NJ, USA, 2010. [Google Scholar]
  37. Bazan, P.; Salasińska, K.; Kuciel, S. Flame retardant polypropylene reinforced with natural additives. Ind. Crop. Prod. 2021, 164, 113356. [Google Scholar] [CrossRef]
  38. Tan, Y.A.; Chan, M.Y.; Koay, S.C.; Ong, T.K. 3D polymer composite filament development from post-consumer polypropylene and disposable chopstick fillers. J. Vinyl Addit. Tech. 2023, 29, 909–922. [Google Scholar] [CrossRef]
  39. Hristov, V.; Vasileva, S. Dynamic mechanical and thermal properties of modified poly (propylene) wood fiber composites. Macromol. Mater. Eng. 2003, 288, 798–806. [Google Scholar] [CrossRef]
  40. Tajvidi, M.; Falk, R.H.; Hermanson, J.C. Effect of natural fibers on thermal and mechanical properties of natural fiber polypropylene composites studied by dynamic mechanical analysis. J. Appl. Polym. Sci. 2006, 101, 4341–4349. [Google Scholar] [CrossRef]
  41. Amash, A.; Zugenmaier, P. Thermal and dynamic mechanical investigations on fiber-reinforced polypropylene composites. J. Appl. Polym. Sci. 1997, 63, 1143–1154. [Google Scholar] [CrossRef]
  42. Wang, S.; Ajji, A.; Guo, S.; Xiong, C. Preparation of microporous polypropylene/titanium dioxide composite membranes with enhanced electrolyte uptake capability via melt extruding and stretching. Polymers 2017, 9, 110. [Google Scholar] [CrossRef]
  43. Chrissopoulou, K.; Anastasiadis, S.H. Polyolefin/layered silicate nanocomposites with functional compatibilizers. Eur. Polym. J. 2011, 47, 600–613. [Google Scholar] [CrossRef]
  44. Niu, P.; Liu, B.; Wei, X.; Wang, X.; Yang, J. Study on mechanical properties and thermal stability of polypropylene/hemp fiber composites. J. Reinf. Plast. Compos. 2011, 30, 36–44. [Google Scholar]
  45. Klapiszewski, Ł.; Grząbka-Zasadzińska, A.; Borysiak, S.; Jesionowski, T. Preparation and characterization of polypropylene composites reinforced by functional ZnO/lignin hybrid materials. Polym. Test. 2019, 79, 106058. [Google Scholar] [CrossRef]
  46. Wang, D.; Wang, J.; He, S.; Yan, Y.; Zhang, J.; Dong, J. Efficient approach to produce functional polypropylene via solvent assisted solid-phase free radical grafting of multi-monomers. Appl. Petrochem. Res. 2021, 11, 99–111. [Google Scholar] [CrossRef]
Figure 1. Scheme of process for obtaining 3D-printed specimens: (a) extruder with temperatures, (b) filaments obtained by extrusion, (c) 3D printing, (d) obtained 3D-printed composites.
Figure 1. Scheme of process for obtaining 3D-printed specimens: (a) extruder with temperatures, (b) filaments obtained by extrusion, (c) 3D printing, (d) obtained 3D-printed composites.
Applsci 14 09139 g001
Figure 2. DSC thermogram for 3D-printed PP, 1st heating step (red), cooling step (green), and 2nd heating step (blue).
Figure 2. DSC thermogram for 3D-printed PP, 1st heating step (red), cooling step (green), and 2nd heating step (blue).
Applsci 14 09139 g002
Figure 3. TGA thermogram for 3D-printed PP and PP-GHP composites.
Figure 3. TGA thermogram for 3D-printed PP and PP-GHP composites.
Applsci 14 09139 g003
Figure 4. DTGA thermogram curves for 3D-printed PP and PP-GHP composites.
Figure 4. DTGA thermogram curves for 3D-printed PP and PP-GHP composites.
Applsci 14 09139 g004
Figure 5. DMA thermogram of storage modulus versus temperature for 3D-printed PP and PP-GHP composites.
Figure 5. DMA thermogram of storage modulus versus temperature for 3D-printed PP and PP-GHP composites.
Applsci 14 09139 g005
Figure 6. DMA thermogram of Tan δ versus temperature for 3D-printed PP-GHP composites.
Figure 6. DMA thermogram of Tan δ versus temperature for 3D-printed PP-GHP composites.
Applsci 14 09139 g006
Figure 7. XRD diffractogram for PP 3D-printed and PP-GHP 3D-printed composites.
Figure 7. XRD diffractogram for PP 3D-printed and PP-GHP 3D-printed composites.
Applsci 14 09139 g007
Table 1. Formulations and nomenclature for the 3D-printed PP–GHP composites.
Table 1. Formulations and nomenclature for the 3D-printed PP–GHP composites.
Composite FilamentGHP Content [phr]Code
PP0PP
PP 1 phr GHP1PP-1 GHP
PP 3 phr GHP3PP-3 GHP
PP 5 phr GHP5PP-5 GHP
phr: parts per hundred of resin.
Table 2. DSC results for 3D-printed PP-GHP composites.
Table 2. DSC results for 3D-printed PP-GHP composites.
Composite FilamentΔHc [J/g]Tc [°C]ΔHm [J/g]Tm [°C]Xc [%]
PP92.78511564.10917030.97
PP-1 GHP91.10911268.63116233.15
PP-3 GHP100.10511379.4716638.39
PP-5 GHP90.57311465.5616631.67
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

Flores-Hernández, C.G.; López-Barroso, J.; Ramos-Galván, C.E.; Salazar-Cruz, B.A.; Chávez-Cinco, M.Y.; Rivera-Armenta, J.L. Development of a Composite Filament Based on Polypropylene and Garlic Husk Particles for 3D Printing Applications. Appl. Sci. 2024, 14, 9139. https://doi.org/10.3390/app14199139

AMA Style

Flores-Hernández CG, López-Barroso J, Ramos-Galván CE, Salazar-Cruz BA, Chávez-Cinco MY, Rivera-Armenta JL. Development of a Composite Filament Based on Polypropylene and Garlic Husk Particles for 3D Printing Applications. Applied Sciences. 2024; 14(19):9139. https://doi.org/10.3390/app14199139

Chicago/Turabian Style

Flores-Hernández, Cynthia Graciela, Juventino López-Barroso, Claudia Esmeralda Ramos-Galván, Beatriz Adriana Salazar-Cruz, María Yolanda Chávez-Cinco, and José Luis Rivera-Armenta. 2024. "Development of a Composite Filament Based on Polypropylene and Garlic Husk Particles for 3D Printing Applications" Applied Sciences 14, no. 19: 9139. https://doi.org/10.3390/app14199139

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop