Skip to main content
Erschienen in: The International Journal of Advanced Manufacturing Technology 7-8/2024

Open Access 20.02.2024 | ORIGINAL ARTICLE

Impact of process parameters on improving the performance of 3D printed recycled polylactic acid (rPLA) components

verfasst von: Mohammad Raquibul Hasan, Ian J. Davies, Alokesh Pramanik, Michele John, Wahidul K. Biswas

Erschienen in: The International Journal of Advanced Manufacturing Technology | Ausgabe 7-8/2024

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

The main goal of this research was to investigate the influence of additive manufacturing (AM) printing parameters on the mechanical properties and surface roughness of specimens fabricated using recycled polylactic acid (rPLA). In order to achieve this goal, significant printing parameters such as layer thickness, infill density, and nozzle temperature were selected based on prior research. A three-level L9 orthogonal array, based on the Taguchi method, was used in the experimental design. The mechanical properties of virgin PLA and recycled PLA printed specimens were examined and compared. To facilitate the analysis of variance (ANOVA) examination, the response data for mechanical and surface roughness parameters were transformed to signal-to-noise (S/N) ratios. The inspected responses under consideration were the surface roughness, shore D hardness, tensile strength, flexural strength, and impact strength. The main findings suggest that careful consideration of the layer height is crucial for achieving optimum mechanical properties in the recycled PLA specimens. Furthermore, the nozzle temperature also played an important factor that affected the mechanical and surface roughness properties of the 3D printed PLA specimens. Microscopic investigation demonstrated that the number and size of voids increased significantly when the layer thickness and temperature were low, namely, 0.1 mm and 195 ℃, respectively. Finally, the optimal combination of printing parameters for each performance characteristic was determined. Following this, a confirmation test was performed using the preferred combination of parameters, which indicated a strong correlation with the outcomes predicted statistically. The results obtained from this study revealed that recycled PLA exhibited mechanical properties comparable to that of virgin PLA under certain conditions. In summary, the results of this study will serve as a valuable dataset in the field of additive manufacturing, providing valuable insights for other researchers working with recycled PLA material.
Hinweise

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

1 Introduction

Technological advancements have a substantial impact on the competitiveness amongst manufacturers in various industry sectors due to the increasing demand for high-quality customised parts that are economical and have structurally sound mechanical qualities [1]. Additive manufacturing (AM) technology, also known as three-dimensional (3D) printing, has the potential to address the mentioned issues due to their design flexibility, manufacturing simplicity, lower production costs, and reduction of raw material waste [2]. These advantages make 3D printing an attractive alternative to conventional manufacturing methods [36]. There are several commercially available AM processes, such as inkjet modelling (IJM), fused deposition modelling (FDM), stereolithography (SLA), and laminated object manufacturing (LOM). However, FDM stands out as a widely utilised technology that involves the extrusion of semi-solid thermoplastic materials through a nozzle in contrast to other AM techniques that employ diverse laser systems, powders, and resins [79]. The FDM process has the capability to construct complex geometries using lightweight materials, resulting in a notable reduction in processing time [10, 11].
A range of thermoplastic polymer materials, such as polylactic acid (PLA), polyethylene terephthalate (PET), acrylonitrile–butadiene styrene (ABS), polypropylene (PP), and high-density polyethylene (HDPE), are the most extensively documented polymers employed in the FDM process [1214]. Amongst these polymers, PLA has attracted considerable interest as a bio-based and biodegradable polymer that serves as a viable substitute for petroleum-derived plastics [15]. PLA exhibits a relatively lower molecular weight and low melting point, rendering it highly compatible with a wide range of FDM equipment [14, 16]. Moreover, it has been observed that the mechanical properties of PLA exhibit a higher level of performance when compared to ABS and PETG [17]. In addition, PLA has better recyclable properties than other thermoplastic polymers [18]. By considering limited material resources and the relatively high price of virgin polymers, future recycling of PLA could become economically attractive. Therefore, recycling of PLA for 3D printing could be a feasible option, as it offers environmental benefits and comparable mechanical characteristics [19].
Previous research has indicated that the recycling process is subject to a high shear force and temperature, which facilitate chain scission processes. This causes a reduction in the molecular weight and viscosity, which ultimately reduces the mechanical strength hindering the use of recycled PLA (rPLA) [3, 20]. Furthermore, the anisotropic characteristics of FDM produce pores, which reduce its strength and other important characteristics [21]. As a consequence, whilst the utilisation of virgin PLA (vPLA) in the context of 3D printing has been extensively documented in the academic literature, the application of recycled PLA remains relatively uncommon. Hence, proper use of recycled materials is important towards the efficacy of recycled plastic materials and green manufacturing techniques.
According to existing research, the overall mechanical performance of printed specimens can be improved by modifying various printing parameters, such as the nozzle/extrusion temperature, layer thickness/height, nozzle diameter, infill percentage, infill pattern, printing speed, build orientation (flat, edge, and upright), raster/infill orientation, and top and bottom layer thicknesses [3, 17, 22]. Therefore, it is essential to optimise and control all these factors effectively within the FDM manufacturing process to fabricate components with higher dimensional precision and enhanced mechanical functionality [23]. Several studies have been conducted to improve the mechanical properties of rPLA by changing the process parameters. Dey et al. [24] altered the infill percentage, layer thickness, printing speed, nozzle temperature, build orientation, and raster angle to modify the mechanical properties of recycled PLA and observed an increase. Atakok et al. [17] conducted an experiment using a commercially recycled PLA filament to examine the impact of the layer thickness, infill percentage, and infill pattern on the mechanical properties. The findings of the investigation indicated that employing a layer thickness of 0.25 mm and an infill percentage of 70% yielded mechanical properties that closely resembled those of the vPLA. Correia et al. [25] determined that by utilising the optimal combination printing parameters, such as 0.2 mm layer thickness, 100% infill, and 40 mm/s travel speed, would effectively improve the mechanical properties of rPLA. In their study, Tan et al. [22] determined that the optimal printing parameters for enhancing the tensile strength and flexural strength of recycled polylactic acid (rPLA) include a fan cooling speed of 100%, an extruder temperature of 190 °C, and a bed temperature of 50 °C. Breški et al. [26] assessed that 30% infill and 0.1 mm layer thickness were most favourable for higher mechanical properties of rPLA. These studies showed that different infill percentages and layer thicknesses can improve the mechanical properties of rPLA 3D printed parts. Therefore, modification and optimisation of different printing parameters can improve the overall mechanical properties [27].
Numerous researchers have attempted to investigate the impact of the process parameters on the behaviour of 3D printed vPLA components. Lanzotti et al. [28] conducted a study to examine tensile properties of 3D printed PLA components by varying the layer thickness and build orientation. The findings of this study demonstrated that a decrease in layer thickness resulted in the attainment of the highest ultimate tensile strength. Eryildiz [29] experimentally determined that a flat orientation showed higher tensile properties compared to the upright build orientation due to the intralayer and interlayer phenomena. Benamira et al. [30] also printed tensile specimens in a flat position and observed improved tensile properties. Tsouknidas et al. [31] examined the influence of various printing speeds on the compressive strength of the PLA samples and concluded that by lowering the printing speeds, maximum strength could be achieved. In their study, Carneiro et al. [32] investigated parameters such as the infill density, layer thickness, and raster orientation and found that the layer thickness had the least impact on the mechanical properties. In their study, Kam et al. [33] found that the infill percentage directly influenced the tensile and Izod impact strengths of PLA parts. This study identified a significant positive correlation between strength and infill percentage. In a study conducted by Suteja and Soesanti [34], it was observed that the layer thickness had a notable impact on the tensile strength. Behzadnasab et al. [35] observed that increasing nozzle temperature from 180 to 240 °C increased mechanical strength in PLA 3D printed parts. Finally, Sun et al. [36] demonstrated that the bed temperature plays a crucial role in the bonding process, consequently leading to an enhancement in the tensile strength.
Regarding the optimization of printing parameters, researchers have utilised a range of design of experiments (DoE) methods, including full factorial design (FFD), fuzzy logic (FL), response surface methodology (RSM), central composite design (CCD), analysis of variance (ANOVA), and the Taguchi method [10, 21, 37, 38]. Several recently published studies have focused on optimising 3D printing process parameters using DoE methods. In their research, Chari et al. [39] examined the compressive strength and hardness of 3D printed PLA components by using the Taguchi method with varying nozzle temperature, infill, and layer thickness. Heidari-Rarani et al. [37] employed the Taguchi DoE method to examine the mechanical properties of PLA specimens by modifying infill density, print speed, and layer thickness. This study found that the impact of the infill density on the mechanical characteristics was significantly greater than that of the printing speed and layer thickness. To determine the impact of printing factors on the mechanical strength of FDM printed specimens, Sood et al. [40] used CCD for DoE and utilised ANOVA to optimise printing settings. The Taguchi approach was used by Alafaghani et al. [41] to investigate the impact of the infill pattern and percentage, layer thickness, and nozzle temperature on the characteristics of specimens made using PLA filaments. By utilising this approach, researchers aimed to gain insights into the optimal combination of these parameters to enhance the overall quality of the fabricated specimens. Therefore, the selection and optimisation of the FDM parameters are essential for achieving the desired technical properties of 3D printed components.
A review of the sources mentioned above indicates that recent scientific investigations have focused on analysing the influence of various factors on the mechanical strength of vPLA printed specimens, such as layer thickness, infill density, infill pattern, print speed, build orientation, and nozzle temperature. However, the mechanical strength of rPLA still remains unexplored. Additionally, no previous research has investigated the impact of layer height, infill density, and nozzle temperature on rPLA 3D printed specimens, and this is considered a novelty of the present work. Thus, it is necessary to conduct an in-depth analysis of the effects of layer height, infill density, and nozzle temperature on the performance of rPLA in order to achieve optimal outcomes and provide significant statistical information for future researchers in the field. Therefore, this study is aimed at assessing the mechanical properties of FDM manufactured samples utilising virgin and recycled PLA materials by varying the layer height, infill density, and nozzle temperature. This study will guide the researchers and encourage the manufacturing company to fabricate products for engineering applications using recycled PLA. This study employs the design of experiments (DOEs) technique to investigate the impact of selected printing parameters on the surface roughness, hardness, tensile, flexural, and impact strengths. In addition, confirmation tests were conducted to validate the optimal process parameters obtained by the DoE technique in order to propose the most appropriate printing parameter as a viable manufacturing protocol for future process optimisation.

2 Materials and methods

The main objective of this study was to investigate a preferred combination of 3D printing parameters to optimise the mechanical properties of virgin PLA and recycled PLA specimens using the Taguchi method. The proposed methodology followed in the work has been shown in Fig. 1.
As a first step, the experimental setup was planned and prepared, wherein the control parameters for 3D printing were carefully chosen. The range of the control parameters was selected, specific levels were assigned to each parameter, and then a Taguchi orthogonal array was chosen. After the initial stage, experiments were conducted using the Taguchi method. The experimental data were statistically analysed in the form of average values, S/N ratios, main effect plots, and analysis of variance (ANOVA). Finally, the optimal values for each control parameter were identified, and the optimised results were subsequently validated through confirmation experiments. A comprehensive description of the materials, printing parameter selection, testing conditions, and analysis employed in the production and characterisation of the samples will be provided in the subsequent subsections.

2.1 Materials

Polylactic acid is one of the most commonly used thermoplastics for fused deposition modelling due to its favourable printing properties and ease of use as a filament in 3D printing [17]. This biodegradable, thermoplastic material originates from cleaner sources when compared to conventional plastics, with less hazardous components, and can be printed at lower temperatures saving energy [4]. PLA has a melting point of approximately 150–160 °C, can be easily processed at 190–230 ℃, and has a high tensile strength of 50–60 MPa when compared to ABS and PET [42]. For this study, virgin PLA and recycled PLA (recycled from residual extrusion waste stream, which are subsequently recompounded and homogenised) filaments, with 1.75 mm diameter, were purchased (X3D, Perth, Western Australia) and used without further modification, for example, no additives or thermal/chemical changes. PLA filaments can absorb moisture from their surroundings, which can impair the quality of printed specimens [43]. Filaments were stored in plastic desiccators containing silica gel in order to avoid moisture adsorption prior to their use.

2.2 Selection of printing parameters

The mechanical properties of FDM printed parts are known to be influenced by several printing parameters. The literature review noted that different printing parameters had been taken into consideration by previous researchers to improve the mechanical properties of 3D printed PLA parts, including layer thickness, shell, infill percentage, infill pattern, nozzle temperature, print speed, and print orientation [13, 14, 33, 44, 45]. Therefore, based on earlier studies, parameters that directly impacted mechanical properties and surface roughness, such as infill density, layer thickness, and nozzle temperature, were chosen for this study across three levels.
The choice of the factors and their corresponding levels was motivated by the interest in maintaining the printing quality and mechanical performance of 3D printed components. The first input parameter considered was the layer height, which measures the individual layer height deposited by a nozzle tip. The printing time and printing quality of a specimen could be influenced by this printing parameter [46]. Values of 0.1 mm, 0.2 mm, and 0.3 mm were chosen as being typical values for the layer height. Second, the infill density indicates the degree to which the printed part is solid or hollow, where a 0% infill is fully hollow, and a 100% is fully solid. Printing time, part cost, and mechanical strength are known to generally increase with infill density; for this reason, three levels of infill, 60%, 80%, and 100%, were selected for analysis. The final input parameter considered was the nozzle temperature (alternatively known as the printing temperature), which denotes the temperature of the extruding nozzle. Three levels of nozzle temperatures were selected at 195 ℃, 205 ℃, and 215 ℃ to analyse the strength of the bond between the layers. It is noted that 190–220 ℃ is the temperature range recommended by the manufacturer for printing. The configurations of printing parameters have been summarised in Table 1.
Table 1
Proposed levels of 3D printing parameters for the experimental design
Factors
Level 1
Level 2
Level 3
Layer height (mm)
0.1
0.2
0.3
Infill density (%)
60
80
100
Nozzle temperature (℃)
195
205
215

2.3 Experimental design using Taguchi orthogonal approach

The number of 3D printed specimens required to assess the influence of printing parameters on the mechanical properties and microstructure needed to be chosen. Therefore, it is important to organise, conduct, evaluate, and interpret tests to assess the factors that impact the value of a parameter or set of parameters. Design of experiment is a useful tool to improve the consistency of results whilst minimising the number of experiments without loss of accuracy. The DoE methodology employs orthogonal arrays (OA) to determine the optimal number of trials and their corresponding levels [47]. Numerous types of DoE have been described in the literature, including full factorial design (FFD), central composite design (CCD), definitive screening design (DSD), Box-Behnken design (BBD), and the Taguchi method [48]. In this work, Taguchi method was selected to design the OA due to its effectiveness and ability to minimise the number of experiments performed [49]. Based on the selected parameters, an L9-OA was created using the MINITAB 21 software. The Taguchi DoE using L9-OA for the sample preparation is presented in Table 2. The array was utilised as a base for every material (vPLA and rPLA) when the test rounds were planned.
Table 2
Experimental layout based on Taguchi L9 OA design
Experimental run
Coded matrix
Un-coded matrix
A
B
C
Layer height (mm)
Infill (%)
Nozzle temperature (℃)
1
1
1
1
0.1
60
195
2
1
2
2
0.1
80
205
3
1
3
3
0.1
100
215
4
2
1
2
0.2
60
205
5
2
2
3
0.2
80
215
6
2
3
1
0.2
100
195
7
3
1
3
0.3
60
215
8
3
2
1
0.3
80
195
9
3
3
2
0.3
100
205

2.4 Fabrication of test samples by 3D printing

Once the L9 specifications of the specimens (Table 2) were determined for the different printing parameters, the specimens were printed according to these specifications. A Raise 3D E2 FDM printer with a nozzle diameter of 0.4 mm was used to manufacture the specimens according to the L9 orthogonal array. This printer uses the ideaMaker slicer software to slice the 3D model (STL file) and modifies the 3D printing process parameters, such as the infill type, percentage, layer thickness, heated bed temperature, and fan speed. After designing the test parts using Solidworks computer aided design (CAD) programme, the models were transformed into standard tessellation language (STL) files with linear tolerance of 0.05 mm and angular tolerance of 1° ASTM D638 standard [50]. These files were utilised to orient the model for the building process and mathematical slicing, which allowed for layer-by-layer printing by the FDM machine. Following this, the 3D model was converted into printing instruction code (G-code) using the ideaMaker slicer programme. This software featured various options for modifying process parameter settings, such as infill type, calculating the tool path, and regulating the heated bed temperature and fan speed. As the influencing printing parameters were selected (Sect. 2.2), therefore, considering the software settings of the 3D printer, the other printing parameters (as listed in Table 3) for the operation of the 3D printing instrument were kept constant during the printing operation. In this study, a rectilinear infill structure was selected because it presents a high yield on mechanical properties and its ability to provide strong support for solid layers with varying bottom and top thicknesses compared to other infill structures such as honeycomb and triangular [33]. The sliced 3D model was then transferred to an FDM 3D printer through its USB or SD card connector. To ensure printing accuracy and precision, and to provide a better possibility of predicting the effects of parameters [30], all samples were printed horizontally on the platform, aligned with the print head axis, whilst the vertical orientation was determined by the sample thickness. Prior to conducting every experiment, the build plate was cleaned to ensure that it was free from adhesives and that the process was unaffected [51].
Table 3
Fixed printing parameters and their corresponding values
Parameter
Specific parameter
Values (constant)
Layer
Shell thickness
2 mm
 
First layer height
0.25 mm
 
First layer flowrate
100%
 
First layer solid fill pattern type
Lines
Extruder
Extrusion width
0.4 mm
 
Retraction speed
40 mm/s
Infill
Infill pattern
Rectilinear
 
Infill angle
 − 45°/45°
 
Infill flowrate
100%
 
Infill overlap
15%
Solid fill
Bottom solid fill layers
5
 
Top solid fill layers
5
 
Bottom surface solid fill pattern type
Line
 
Top surface solid fill pattern type
Line
Speed
Printing speed
50 mm/s
 
Infill speed
50 mm/s
Support
Support structure
None
 
Platform addition
Brim
Temperature
Heat bed temperature
45 ℃

2.5 Characterisation of 3D-printed specimens

Following the 3D printing operation, all samples were kept in a sealed packet, and silica gel was used to avoid moisture absorption. For all the printed samples, tests were carried out at room temperature (typically 23 °C) according to the ASTM-D618-21 [52] standard. The mechanical properties and microstructural testing procedures followed have been described below.

2.5.1 Determination of hardness and surface roughness

The hardness of a material is a fundamental attribute that signifies its ability to withstand localised deformations, such as penetration or indentation on its surface. [53]. The hardness of the printed specimens was assessed using a specialised shore D hardness testing device (Sauter HBD 100–0) that is designed to test the hardness of rigid plastics, thermoplastics, and hard rubber materials. The ASTM-D2240-15 [54] standard, which is commonly employed for assessing the durometer hardness of polymers, was employed as the metric in this test. According to this standard, the thickness of the test specimen should be at least 6.0 mm (0.24 in). Therefore, the dimensions of the hardness test specimens were manufactured with a thickness greater than 6 mm (\(40\mathrm{ mm}\times 40\mathrm{ mm}\times 8\mathrm{ mm})\), as illustrated in Fig. 2. Five pieces of 3D printed test block were prepared for each experimental condition in an orthogonal array. Nine different surface hardness measurements with 5 mm spaces between each point were collected from each test specimen, resulting in a total of 45 hardness measurements for each experimental run.
Surface roughness (SR) is utilised as a technical criterion for mechanical components to quantify the existence of micro-irregularities on the surface texture [55]. This property is commonly employed as a technical specification for mechanical products because of its impact on aesthetic appearance and critical role in ensuring the structural precision of a component [56]. In order to evaluate the influence of printing parameters on surface roughness, a Mitutoyo surface roughness tester (model SJ-210) was utilised in this study. This compact and portable device offers a maximum resolution of 0.001 µm. The stylus moves over the sample surface at a predetermined measuring velocity and distance in order to identify any irregularities present on the surface of the workpiece [57]. Roughness was measured on the largest surface of the tensile specimens (Fig. 3). The arithmetic mean roughness (\({R}_{a}\)) and root mean square roughness (\({R}_{q}\)) were quantitatively measured at the micro-metre (μm) level.
The parameter Ra represents the mean absolute deviation of roughness irregularities from the mean line over a certain length of sampling [58]. The numerical representation of the arithmetic average height parameter is shown by Eq. 1 [58]:
$${R}_{a}=\frac{1}{n}\sum_{i=1}^{n}\left|{y}_{i}\right|$$
(1)
where \(n\) represents the total number of samples and \({y}_{i}\) represents the deviation from the sample mean line.
The root mean square roughness (\({R}_{q})\) is another significant parameter for characterising surface roughness, which indicates the standard deviation of the distribution of surface heights. The digital representation of this parameter can be mathematically expressed using Eq. 2 [58]:
$${R}_{q}=\sqrt{\frac{1}{n}\sum_{i=1}^{n}\left|{y}_{i}\right|}$$
(2)
where \(n\) represents the total number of samples and \({y}_{i}\) represents the deviation from the sample mean line.
Since waviness, peaks, and valleys affect the surface quality, five points were chosen on the top side of the tensile specimen to analyse the surface texture with a measuring speed of 0.5 mm/s (Fig. 3), and their mean values were tabulated [59, 60]. Five samples printed from each experimental run were analysed using a tracing length of 5.6 mm, an evaluation length of 4.0 mm, and a cut-off wavelength of 0.8 mm to achieve high accuracy and precision, according to ASTM-D7127-17 [61] [59]. The roughness of the printed surface was determined vertically to the feed mark [62]. The representative value of each specimen was determined by calculating the arithmetic mean of five measurements.

2.5.2 Determination of mechanical properties

The mechanical properties of the printed parts were obtained from the tensile, three-point bending (flexural), and impact tests. Following ASTM standards, the standard number of isotropic test specimens per sample was a minimum of five. Hence, for all three tests, five samples were printed for each experimental run. Therefore, 45 specimens were printed for each material (vPLA and rPLA), and in total, 90 specimens were printed for each type of mechanical test (tensile, three-point bending, and impact). The average data were reported based on the testing results of at least five samples for each material batch, along with the calculated standard deviations to evaluate the test reproducibility.
Tensile tests were conducted using dog-bone samples that were fabricated following the ASTM-D638-22 [50] Type IV specifications, with dimensions of \(115\mathrm{ mm}\times 19\mathrm{ mm}\times 3.2\mathrm{ mm}\) (as shown in Fig. 4). All tests were performed using a Shimadzu Autograph AGS-X tensile testing machine equipped with a 50 kN load cell capacity. This model has an exceptionally high level of precision, with a measurement accuracy within ± 0.5% of the indicated test force. Prior to testing the sample with the instrument, the non-shift wedge-type grips were adjusted to a range of 0 to 7 mm and calibrated according to the ASTM-E4 [63] standard. This instrument uses TRAPEZIUM-X software to manage the test process and thereby pre-process the data acquired from this machine. For the testing operation, at first, sample data such as thickness (mm), breadth (mm), and gauge length (mm) were entered into the software programme. After that, the samples were clamped in the machine and loaded at a crosshead speed of 5 mm/min until failure occurred. The computer software was used to record the force and displacement values. These values were used to generate force–displacement curves using Microsoft Excel, and the ultimate tensile strength (UTS) was evaluated from these data. The elastic modulus of each specimen was evaluated according to the initial slope of its stress/strain curve using Microsoft Excel’s ‘Trend Line’ function. For this purpose, a line diagram was plotted using strain values within the 0.0005 < ε < 0.0025 range. Several standard mechanical property evaluations were conducted, including the yield strength, ultimate tensile strength (UTS), work at UTS, and work at fracture, and the average results were tabulated to obtain the arithmetic mean.
The flexural strength (3-point bending) refers to the ability of a material to resist bending forces that act perpendicular to its longitudinal axis. The 3-point bending test specimens (strip-like samples) having dimensions of \(127\mathrm{ mm}\times 12.7\mathrm{ mm}\times 3.2\mathrm{ mm}\) (as shown in Fig. 5) were prepared and manufactured according to ASTM-D790-17 [64] standard. This test was conducted using Shimadzu Autograph AGS-X equipment with a 10 kN cell mounted with a 3-point bending apparatus, and a displacement rate of 3 mm/min and a span of 51.2 mm (span to depth ratio 16:1) were employed for the test. The instantaneous data for the load and displacement were obtained using the Trapezium-X software.
In this study, the flexural strength was determined by calculating the nominal stress in the central span, which was obtained using the maximum load value, according to Eq. 3 [64],
$$\sigma =\frac{3PL}{2b{d}^{2}}$$
(3)
where P is the load at a given point on the load–deflection curve, L is the support span length, b is the width, and d is the thickness of the specimen.
The flexural modulus was determined using Eq. 4, as specified in the ASTM-D790-17 [64] standard. This equation was derived through linear regression analysis of the load–displacement curves, considering the linear segment within the range where the correlation coefficient exceeded 95%.
$$E=\frac{{L}^{3}m}{4b{d}^{3}}$$
(4)
where m represents the slope of the tangent line to the initial linear segment of the load–deflection curve.
The Charpy impact test, alternatively referred to as the Charpy V-notch test, is a universally accepted method for evaluating the fracture behaviour of materials under high strain rates. This indicates the amount of energy absorbed by the material during the fracture [65]. Charpy impact tests were conducted to evaluate the fracture behaviour of the rPLA-printed specimens. The test specimens were manufactured using dimensions of 127 mm × 12.7 mm × 6.2 mm being notched at a 45° angle in accordance to the ASTM-D6110-18 [66] standard (Fig. 6). A Zwick 5102 (model D-7900) impact tester with a 6.5 J hammer was used to test the fabricated specimens. The impact strength was determined using the energy required to break the specimen and the depth under the notch of the specimen.

2.5.3 Determination of surface morphology

The surface morphology of polymeric materials is a significant characteristic that is originated from their chemical structure and the methods used in their production. These characteristics directly affect the final surface properties of the workpiece [67]. Optical microscopy inspects samples volumetrically and seeks to provide a preliminary perception of the structural features exhibited by a specimen [68]. The focus of this examination is to assess the surface characteristics in relation to printing quality. To examine the surface morphology of the printed samples after mechanical testing, at first, the samples were cut into sections and then embedded within acrylic resin and powder in self-curing polyacrylic cylinders with a dimension of 2 cm width and 3 cm height to expose a flat 4 × 6 mm window of surface for microstructure observations. The test parts (top surface) were prepared through a sequential polishing process utilising silicon carbide (SiC) paper with grit sizes of P400, P800, P1000, P1200, and P4000. Grit sizes determine the particle count in paper according to the Federation of European Producers of Abrasives (FEPA) system, which is the same as the ISO 6344 standard. SiC paper with a higher ‘P’ value is distinguished by the presence of smaller particles within the paper [69]. Prior to starting the manual polishing of the specimens, the surface was marked with an ink-felt tip permanent marker. The polishing process was performed until the ink was completely removed from the surface, indicating that the current grit size had replaced the previous roughness. The final polishing stage commenced with using diamond suspensions of 6 μm, followed by suspensions of 3 μm and 1 μm. Throughout the process, a continuous water flow was maintained. The specimens were adequately cleaned with water prior to use at higher grit levels. This process aided in the removal of any accumulated abrasive dust, which facilitated clear visualisation of the surface. After completion of the polishing process, the specimens were examined under an optical microscope (OM, OLYMPUS BX51M, Japan) equipped with a portable standalone colour camera. The acquired images were captured at 10 × magnifications by using magnifying lenses. For fractographic analysis, a Pro-MicroScan microscope (model DCM 900) equipped with an eyepiece camera manufactured by the Oplenic Corporation, China, was used to examine the fractured cross-sectional structure of the mechanically tested samples. This examination was primarily concerned with assessing the surface properties associated with the printing quality. A few scratches and voids were observed in the worn area, which may be due to sliding during mechanical property testing.

2.6 Assessment using the signal-to-noise ratio

Performance characteristics are fundamental quality features that determine a product’s ability to meet specified design requirements. Thus, a product of superior quality will exhibit consistent performance over its entire lifespan and under various operating conditions [70]. The Taguchi method employs a loss function to quantify the difference between the observed and desired values of the performance characteristics, which minimises performance variability. This outcome is then converted into a statistical metric called the signal-to-noise ratio (\(S/N)\) ratio to determine the optimal level for each control factor and assess the statistical significance of their impact on the response variable, in other words, to measure the process variability [10, 47]. The term ‘noise’ refers to the impact of each factor on each operation, whilst the term ‘signal’ denotes the response to changes in each operating variable [71].
The S/N ratio is a metric that considers both the average value and variability of a given quality feature. The specific formula for calculating the S/N ratio depends on the criteria used to assess the quality features that need to be investigated [48]. As a consequence, when the process is optimised in terms of S/N ratio, it ensures that the resulting optimal process conditions are robust and stable, indicating minimal process variation [72]. Hence, the S/N ratio was employed in the current analysis to determine the most favourable and optimal process conditions for the layer height, infill density, and nozzle temperature. The study of the performance characteristics involves the utilisation of three distinct types of S/N ratio loss functions: larger-the-better, nominal-the-best, and smaller-the-better. In the context of smaller-the-better conditions, these characteristics typically represent unfavourable outputs. In contrast, for the larger-the-better condition, the characteristics are generally favourable [73].

2.6.1 Processing of data for S/N value

At first, all output performance responses (data from the experiment) were analysed according to the Taguchi L9 orthogonal array-based design. In this study, the ‘larger the better’ and ‘smaller the better’ type \(S/N\) ratio responses were used to evaluate the properties of the printed specimens through the FDM process.
This study focuses on maximising the mechanical properties, including tensile strength, three-point bending, impact strength, and hardness. Therefore, the criterion ‘larger-the-better’ was chosen for these quality characteristics. The principle of ‘larger-the-better’ was employed in situations where the purpose was to maximise the values of quality characteristics [74]. The larger-the-better \(S/N\) ratio was calculated by employing Eq. 5 [75].
$$\frac{S}{N}{\text{ratio}}=(-10)\times {{\text{log}}}_{10}(\frac{1}{n}\sum_{i=1}^{n}\frac{1}{{y}_{i}^{2}})$$
(5)
where n is the number of trials, \(i\) is the experiment number, and \({y}_{i}\) is the response in each experiment.
The purpose of the smaller-the-better condition is to minimise the assessment of the quality characteristics to the smallest possible value, ideally zero, a target, or ideal value. Therefore, the principle of ‘smaller-the-better’ was employed to minimise the surface roughness of specimens produced using vPLA and rPLA. The smaller-the-better \(S/N\) ratio was computed by Eq. 6 [75].
$$\frac{S}{N}{\text{ratio}}=(-10)\times {{\text{log}}}_{10}(\frac{1}{n}\sum_{i=1}^{n}{y}_{i}^{2})$$
(6)
where n is the number of trials, \(i\) is the experiment number, and \({y}_{i}\) is the response in each experiment.
Once the \(S/N\) ratios were evaluated, the average \(S/N\) ratios of the parameters at specific levels were examined. After that, the parameter exhibiting the largest difference in arithmetic values between its highest mean \(S/N\) ratio and its lowest mean \(S/N\) ratio was determined to observe the major impact on the outcome [76]. The Taguchi approach states that when the \(S/N\) ratio is maximum, process variability is minimal. The greater the \(S/N\) ratio value, the better the outcome since it ensures the maximum quality with the least variability [73]. Therefore, the experimental values with the highest mean \(S/N\) ratio were considered as the optimal operating values for each parameter. The influence of each process parameter on the S/N ratio at different levels for each sample is presented in the results section, based on an orthogonal experimental design. Finally, the influencing parameters were determined from the calculated \(S/N\) ratios, response tables, and main effects plots.

2.7 Analysis of variance (ANOVA)

After statistically examining the \(S/N\) ratio, an analysis of variance (ANOVA) was performed to determine the accuracy of the predicted model and to investigate which parameter significantly affected the quality characteristics. ANOVA separates the variation in the dataset into two parts: between-group and within-group [77, 78]. Therefore, ANOVA primarily examines the difference between groups in relation to the variation within each group [79]. ANOVA is a statistical method used to optimise process parameters by analysing a set of experimental results and categorising them according to a common variable or parameter, as well as an objective function or response. This analysis offers valuable insights into the relevance of the observed variance as well as the connection between system parameters and their corresponding responses [76]. F-test and p-test were employed in the ANOVA to assess the extent to which each factor contributed to the overall variation [76]. The F-test was used to determine the parameters that exerted a statistically significant impact on the mechanical properties. In addition, the p value estimated the significance of the results or the extent to which a parameter affects the objective function.

2.7.1 Calculation of the F-ratio

F-ratio is applied in statistical analysis to determine the significant impact of process parameters on product characteristics, which leads to a conclusive outcome [78]. A large F-ratio indicates a higher level of variance between groups compared to within groups, thereby enhancing the likelihood of rejecting the null hypothesis, which suggests that all population means are identical. Generally, the higher the F-ratio, the greater the influence of process parameter modification on the performance characteristics [73]. For the case of F > 4, it typically indicates that modifications to the design parameters can have a significant influence on the finished product’s characteristics [78].

2.7.2 Calculation of p value

In order to determine the statistical significance of the disparity between group means, it is crucial to consider the p value, which is associated with the F-ratio [77]. The p value is a statistical measure that is used instead of critical values to determine the minimum level of significance for rejecting the null hypothesis. A smaller p value indicates a higher level of support for the alternative hypothesis [80]. The p value is commonly utilised with a threshold of 0.05, which corresponds to a significance level of 5%. For example, in a case where the calculated p value of a test statistic is less than 0.05, it would be appropriate to reject the null hypothesis [80].

2.8 DoE confirmation test

Once the optimal level of the design parameters had been selected, the final step was conducted to predict and verify the improvement of the quality characteristics using the optimal levels of the design parameters. The obtained results could subsequently be compared to the theoretically predicted optimal conditions based on the desired combination of factors and their respective levels [81]. At first, the predicted \(S/N\) ratio, which was expected for the optimal combination of parameters, was determined. After computing the predicted \(S/N\) ratio, a confirmation experiment was conducted to validate whether the optimal conditions achieved the expected response. In engineering experiments, it is expected that the level of agreement between the predicted values and confirmation results should fall within the range of ± 5%, indicating a confidence interval of greater than 95% [81]. This comparison evaluated the agreement between the obtained and expected parameters.

2.9 Statistical analysis

Microsoft Excel (Microsoft Corporation) was used for data handling. Values obtained from the material property characterisation tests are presented as mean ± standard deviation. Minitab 21 software was used to analyse the S/N ratio and ANOVA.

3 Result and discussion

The primary objective of the present investigation was to examine the most favourable printing parameters for the fabrication of 3D printed components utilising two different materials, namely, vPLA and rPLA. The present study employed a Taguchi orthogonal array experimental design to conduct the experiments. The specimens were manufactured using a fused deposition modelling 3D printer. Following the manufacturing process, a comprehensive evaluation of mechanical properties was conducted. The properties under investigation included surface roughness, hardness, tensile strength, flexural strength, and impact strength. The performance of the rPLA specimens was compared/benchmarked with 3D printed vPLA specimens using the same processing parameters. These parameters were carefully assessed to gain insights into the performance characteristics of the printed specimens. In this section, a comprehensive analysis of the results obtained from the experimental testing of the specimens has been presented with the aim of evaluating their material properties.

3.1 Surface roughness

To gain additional insight into the differences between the printed virgin and recycled materials, an analysis of the surface roughness was conducted following printing (as shown in Fig. 7). The FDM process typically results in a relatively rough surface, with the extent of this roughness being influenced by the printing quality [82]. In order to achieve a high level of surface smoothness, it is common practise to subject a product to different processing procedures [3]. Therefore, the experiment aimed to achieve maximum surface smoothness and determine whether the printing parameters under investigation have an impact on the surface roughness of the specimens.
The measurement and recording of the surface roughness for each printed part have been documented in Fig. 7. The findings presented in Fig. 7 (a) indicate that the parameter settings used in experimental run 2 (LH 0.1 mm, I 100%, and NT 205 ℃) yielded the lowest average Ra values for both vPLA and rPLA specimens, namely, 5.87 μm and 6.12 μm, respectively. This was followed by experimental run 3, which resulted in an average Ra value of 6.22 μm and 7.33 μm, and experimental run 3, which yielded average Ra values of 6.70 μm and 7.72 μm, for vPLA and rPLA, respectively. The rPLA specimens exhibited higher Ra and Rq values in comparison to the vPLA specimens, suggesting the presence of rougher surfaces. As recycled PLA is known to exhibit a lower molecular weight due to chain scission, the recycled material would be expected to exhibit a lower viscosity at the same extrusion temperature, leading to the formation of a surface with increased roughness [62]. Based on the data presented in Fig. 7, it can be observed that there is a positive correlation between the nozzle temperature and layer thickness with the surface roughness for both types of materials. Specifically, at a lower nozzle temperature (195 °C) and higher nozzle temperature (215 °C), the surface roughness exhibits an increase. The surface was roughest at 195 °C and slightly less rough at 215 °C, whereas at 205 °C, the surface roughness was lower. Maidin et al. [62] and Mani et al. [83] observed that for vPLA, the surface roughness decreased with an increase in nozzle temperature. It should be noted that the present study used different nozzle temperatures and layer thicknesses, which showed a different result but a similar trend, as in previous research on surface roughness. This may be attributed to the temperature of the nozzle, which significantly affects the raw material melting process. Incomplete melting due to lower than optimum temperature would produce a rough surface. On the other hand, the reduction in the thickness of the layers has a direct impact on the surface roughness, resulting in decreased visibility of the layers and a smoother surface texture for the printed object. Ayrilmis [84] also observed an increased roughness with increasing printing layer thickness. As the thickness of the layers increases, the surface of the printed part exhibits enhanced sharpness, thereby rendering the individual layers more noticeable [85].

3.2 Hardness test

The hardness test is a widely employed method in the field of plastics to assess the ability of a material to withstand concentrated loads that result in significant plastic deformation. This test is both rigorous and informative and provides valuable insights into the properties of the material under inspection [86]. The findings of the experiments conducted on both materials have been summarised in Fig. 8. The vPLA-printed specimens exhibited marginally higher hardness, whereas the rPLA polymers exhibited lower hardness. The highest shore D hardness was observed for experimental run 9, which indicated hardness result of 81.68 and 76.78 for vPLA and rPLA, respectively. In contrast to this, the lowest result was obtained for experimental run 1, which showed hardness of 79.53 and 73.92 for vPLA and rPLA, respectively. The findings of this study demonstrate a clear positive correlation between nozzle temperature and infill percentage and the resulting increase in hardness, which is consistent with the results evaluated by Mani et al. [83] and Maguluri et al. [87] for vPLA. According to Şirin et al. [88], an increase in the infill percentage resulted in a reduction in void formation, leading to an overall increase in hardness. Another reason is that an increase in the infill percentage leads to an expansion in the cross-sectional area of the material, thereby contributing to the development of a more rigid internal structure [89]. The results also demonstrated the influence of printing temperature on the hardness properties. A reduction in the printing temperature led to a corresponding decrease in the hardness value, the cause of which was attributed to poor layer-to-layer bonding and insufficient melting when printing at low temperatures, which resulted in increased porosity and decreased hardness. The decrease in the rate of hardness change with increasing printing temperature was attributed to the elevated temperature during the printing process and the enhanced adhesion between the fibres, resulting in a more robust material combination. Additionally, a reduction in the number of air gaps can also contribute to this phenomenon [87].

3.3 Tensile test

The output variables of elastic modulus, ultimate tensile strength, yield strength, fracture strength, work at ultimate tensile strength, and work at fracture have been observed by adjusting the input processing variables, as indicated in Table 2. Five specimens were produced for each Taguchi experiment and were chosen for tensile examination to reduce the impact of differences in the specimens on the results of the experiment. The findings from these five specimens demonstrated a consistent level of repeatability. Hence, a single stress–strain curve is illustrated, representing a set of five specimens. Figures 9 and 10 illustrate the stress–strain curves of vPLA and rPLA components manufactured through 3D printing, which included all experimental conditions. It is evident that the vPLA material exhibited higher values of tensile strength and yield strength in comparison to rPLA. The stress–strain relationships exhibited by the vPLA and rPLA specimens demonstrated a nearly linear correlation as the strain increased. Once the stress reached its maximum point, it remained relatively constant as the strain continued to rise. The examination of the outcomes revealed that both the vPLA and rPLA specimens exhibited ductile fracture behaviour. The data clearly indicate that the virgin PLA material exhibited a higher tensile strength than the recycled PLA material.
Tables 4 and 5 summarise the average values of the vPLA and rPLA tensile properties, including the elastic modulus, ultimate tensile strength, yield strength, fracture strength, and strain at fracture, for each specimen designed by the Taguchi method. The results showed that the standard deviation of all the measurements was relatively small, indicating good repeatability within the experimental data. When comparing the vPLA- and rPLA-printed specimens, it was observed that, in all of the experimental investigations, the vPLA specimens exhibited higher tensile characteristics than the rPLA-printed samples, as depicted in Figs. 9 and 10. In addition, the findings indicated that the samples exhibited better tensile properties as the layer thickness and temperature increased. In previous studies utilising recycled polylactic acid, the same declining pattern in the tensile properties was noted [3, 15, 17, 22, 90].
Table 4
Tensile properties of the vPLA printed specimens
Experimental run
Yield strength
Ultimate tensile strength (UTS)
Strain at UTS
Fracture strength
Strain at fracture
Elastic modulus
Work until UTS
Work until fracture
(MPa)
(MPa)
(%)
(MPa)
(%)
(GPa)
(kJ/m2)
(kJ/m2)
Run 1 (LH 0.1, I 60, NT 195)
33.51 ± 0.871
36.36 ± 0.289
2.27 ± 0.047
31.87 ± 0.619
3.78 ± 0.266
2.26 ± 0.296
32.08 ± 0.770
65.07 ± 1.412
Run 2 (LH 0.1, I 80, NT 205)
38.12 ± 0.931
45.30 ± 0.328
2.48 ± 0.025
42.98 ± 0.389
3.84 ± 0.229
2.71 ± 0.042
40.34 ± 0.340
85.32 ± 0.995
Run 3 (LH 0.1, I 100, NT 215)
47.97 ± 0.954
52.81 ± 0.630
2.50 ± 0.082
49.37 ± 0.829
3.70 ± 0.696
3.13 ± 0.053
43.52 ± 0.944
100.02 ± 1.607
Run 4 (LH 0.2, I 60, NT 205)
50.17 ± 1.124
53.47 ± 0.710
2.49 ± 0.066
50.53 ± 0.951
4.05 ± 0.167
3.10 ± 0.072
43.11 ± 0.632
102.04 ± 0.890
Run 5 (LH 0.2, I 80, NT 215)
50.89 ± 0.991
54.04 ± 0.429
2.51 ± 0.031
51.33 ± 0.659
5.36 ± 0.684
3.11 ± 0.036
47.56 ± 0.871
130.97 ± 1.437
Run 6 (LH 0.2, I 100, NT 195)
48.41 ± 1.202
52.68 ± 0.652
2.41 ± 0.041
47.68 ± 1.012
5.04 ± 0.688
3.08 ± 0.038
42.25 ± 0.205
120.07 ± 2.014
Run 7 (LH 0.3, I 60, NT 215)
53.56 ± 1.141
56.85 ± 0.269
2.48 ± 0.021
52.53 ± 0.475
4.74 ± 0.442
3.19 ± 0.014
45.31 ± 0.537
131.32 ± 1.018
Run 8 (LH 0.3, I 80, NT 195)
50.61 ± 1.123
51.37 ± 0.548
2.37 ± 0.044
50.23 ± 0.613
4.31 ± 0.188
3.04 ± 0.035
41.56 ± 0.759
115.90 ± 1.752
Run 9 (LH 0.3, I 100, NT 205)
51.26 ± 0.984
54.84 ± 0.698
2.45 ± 0.069
51.08 ± 0.286
5.30 ± 0.339
3.10 ± 0.038
46.73 ± 0.409
125.63 ± 1.921
*LH, layer height; I, infill percentage; NT, nozzle temperature
Table 5
Tensile properties of the rPLA printed specimens
Experimental run
Yield strength
Ultimate tensile strength (UTS)
Strain at UTS
Fracture strength
Strain at fracture
Elastic modulus
Work until UTS
Work until fracture
(MPa)
(MPa)
(%)
(MPa)
(%)
(GPa)
(kJ/m2)
(kJ/m2)
Run 1 (LH 0.1, I 60, NT 195)
22.78 ± 0.442
28.81 ± 0.227
2.06 ± 0.053
23.84 ± 0.919
5.65 ± 0.558
1.99 ± 0.031
26.77 ± 0.682
87.34 ± 0.53
Run 2 (LH 0.1, I 80, NT 205)
25.02 ± 0.738
31.20 ± 0.116
2.01 ± 0.019
25.77 ± 0.901
4.34 ± 0.230
2.19 ± 0.157
29.98 ± 0.331
74.43 ± 1.27
Run 3 (LH 0.1, I 100, NT 215)
25.84 ± 0.696
32.11 ± 0.509
2.00 ± 0.030
26.41 ± 0.863
5.15 ± 1.001
2.32 ± 0.018
24.48 ± 0.862
97.65 ± 1.75
Run 4 (LH 0.2, I 60, NT 205)
31.01 ± 0.458
35.23 ± 0.450
2.21 ± 0.031
29.93 ± 0.645
6.50 ± 1.796
2.42 ± 0.016
29.37 ± 0.774
109.63 ± 2.10
Run 5 (LH 0.2, I 80, NT 215)
32.80 ± 0.437
38.36 ± 0.429
2.17 ± 0.031
33.84 ± 0.392
6.91 ± 0.361
2.59 ± 0.007
34.63 ± 0.730
135.98 ± 1.89
Run 6 (LH 0.2, I 100, NT 195)
29.34 ± 0.436
35.86 ± 0.563
2.24 ± 0.023
30.54 ± 0.732
6.71 ± 0.792
2.38 ± 0.043
31.89 ± 0.521
120.07 ± 2.28
Run 7 (LH 0.3, I 60, NT 215)
34.12 ± 0.648
39.90 ± 0.479
2.36 ± 0.027
34.56 ± 0.992
7.10 ± 0.769
2.66 ± 0.031
40.44 ± 0.507
142.24 ± 2.29
Run 8 (LH 0.3, I 80, NT 195)
31.65 ± 0.394
37.54 ± 0.636
2.06 ± 0.031
33.16 ± 1.357
5.94 ± 1.271
2.51 ± 0.033
30.74 ± 0.461
120.70 ± 2.20
Run 9 (LH 0.3, I 100, NT 205)
33.25 ± 0.723
38.47 ± 0.338
2.16 ± 0.021
28.42 ± 0.462
6.92 ± 0.558
2.62 ± 0.019
34.66 ± 0.722
136.67 ± 2.49
*LH, layer height; I, infill percentage; NT, nozzle temperature
In the context of experimental run 7, it was observed that specimens printed with a layer height of 0.3 mm, infill density of 80%, and nozzle temperature of 215 °C exhibited the highest tensile strength. This outcome was consistent for both vPLA and rPLA materials. Based on the data presented in Table 4 and 5, it can be observed that the tensile strength of rPLA in experimental run 7 was recorded as 39.90 MPa, which is 25.22% lower than the corresponding value of 56.85 MPa for the vPLA printed specimen. The experimental results indicated that the second highest recorded value was 54.84 MPa for experimental run 9, which utilised a layer height of 0.3 mm, 100% infill, and a nozzle temperature of 205 °C. Similarly, the third highest recorded value was 54.04 MPa for experimental run 5, which employed a layer height of 0.2 mm, 100% infill, and nozzle temperature of 215 °C. In the case of the rPLA-printed specimens, the reduction in the values was 28.85% and 29.09% for the respective parameters. Conversely, a decrease in the thickness of each layer and temperature corresponded to a decline in the tensile strength. To provide an example, the results obtained from experimental run 1 revealed suboptimal results, with vPLA and rPLA demonstrating tensile strength values of 36.06 MPa and 28.81 MPa, respectively. The lower printing temperature resulted in incomplete melting, which in turn produced weak bonding between layers and loss of internal structure, resulting in easily fractured specimens [89].
A noteworthy finding was observed in relation to the strain at fracture, wherein rPLA exhibited a superior outcome compared with vPLA, with increases ranging from 14.64 to 68.83%. This ultimately indicated that rPLA demonstrated a more favourable work of fracture. In the case of both the vPLA and rPLA samples, it was observed that experimental run 7 exhibited better fracture performance, including strain at fracture, fracture strength, and work done until fracture, compared to the remaining samples. Experimental runs 5 and 9 exhibited comparable findings to experimental run 7 in terms of the work done until fracture. It is interesting to observe that both the tensile strength and fracture strain were simultaneously improved as the nozzle temperature increased. This phenomenon can be attributed to the improved bond quality between the layers deposited at the interface [10].
Previous research indicated that reducing the infill density leads to a decrease in the contact and adhesion between layers, resulting in a reduction in the mechanical properties [89]. However, the findings of the present study indicate that the infill density did not have a significant impact when the nozzle temperature and layer thickness were suitably high, as observed in experimental run 7. The rationale behind this phenomenon may lie in the reduction of the number of layers during printing with a higher layer thickness, resulting in a decrease in the distortion effect, consequently leading to an increase in strength. Moreover, an increase in the thickness of each layer facilitates the movement of the partially melted substance within the expanded gaps between the printed lines, consequently leading to a notable enhancement in the structural integrity [91]. An additional factor that could contribute to this phenomenon is the utilisation of a higher printing temperature, which increases the fluidity of the material and subsequently enhances the interfacial bond strength between successive layers. Furthermore, the reduction in interchain interaction and decrease in the number of air gaps would be expected to contribute to an increase in the tensile strength of the material, as demonstrated by Hsueh et al. [89].

3.4 Flexural test

The experimental values of flexural properties have been presented in Table 6 and 7. These tables provide a comprehensive overview of the flexural performances of the tested specimens. In addition, the stress vs. strain diagrams for the flexural tests have been depicted in Figs. 11 and 12 and visually illustrate the relationship between stress and strain, thereby offering valuable insight into the mechanical behaviour of the specimens under flexural loading. The flexural stress–strain diagrams were determined using Eqs. 4 and 5, respectively. Work until flexural strength was determined from the area under the bending load–displacement curve until the point of maximum stress. During loading, the fracture strength of the flexural tests could not be determined, because the samples did not completely fracture up to the testing limit. All tests were terminated before the specimens reached their fracture point because they came into contact with the grips prior to fracture. This was primarily attributed to the excessive ductility of the polymers; therefore, the measurement of strain at fracture was not achieved [3, 92]. The findings presented in this study showed comparatively poorer mechanical properties exhibited by the specimens printed using rPLA in comparison to those printed using vPLA. In previous investigations employing rPLA, a comparable finding related to a decreasing trend in flexural characteristics was documented [3, 17, 22]. This phenomenon can be attributed to the reduced mechanical properties of recycled PLA owing to its low molecular weight [93].
Table 6
Test results for vPLA flexural specimens
Experimental run
Yield strength
Flexural strength
Flexural strain
Elastic modulus
Work until flexural strength
MPa
MPa
%
GPa
kJ/m2
Run 1 (LH 0.1, I 60, NT 195)
60.25 ± 0.25
64.78 ± 2.21
4.78 ± 0.21
2.31 ± 1.20
28.74 ± 1.41
Run 2 (LH 0.1, I 80, NT 205)
66.44 ± 0.58
71.15 ± 1.21
5.11 ± 0.85
2.56 ± 1.24
31.11 ± 2.20
Run 3 (LH 0.1, I 100, NT 215)
73.85 ± 1.20
79.98 ± 1.48
5.23 ± 0.98
2.94 ± 1.24
35.86 ± 2.01
Run 4 (LH 0.2, I 60, NT 205)
69.45 ± 1.07
82.75 ± 2.01
6.60 ± 0.67
2.91 ± 1.83
39.47 ± 1.91
Run 5 (LH 0.2, I 80, NT 215)
74.11 ± 0.87
87.39 ± 1.24
6.92 ± 0.70
2.91 ± 1.51
40.87 ± 1.01
Run 6 (LH 0.2, I 100, NT 195)
70.72 ± 1.23
82.53 ± 1.09
7.90 ± 1.24
2.89 ± 1.81
45.83 ± 1.46
Run 7 (LH 0.3, I 60, NT 215)
77.34 ± 0.89
92.44 ± 2.01
6.85 ± 1.20
3.01 ± 0.95
49.96 ± 1.47
Run 8 (LH 0.3, I 80, NT 195)
71.48 ± 1.87
86.11 ± 2.42
8.11 ± 0.51
2.85 ± 0.89
46.07 ± 2.03
Run 9 (LH 0.3, I 100, NT 205)
76.60 ± 1.74
90.26 ± 1.74
6.50 ± 0.81
2.95 ± 1.01
46.35 ± 1.58
*LH, layer height; I, infill percentage; NT, nozzle temperature
Table 7
Test results for rPLA flexural specimens
Experimental run
Yield strength
Flexural strength
Flexural strain
Elastic modulus
Work until flexural strength
MPa
MPa
%
GPa
kJ/m2
Run 1 (LH 0.1, I 60, NT 195)
41.50 ± 1.23
46.78 ± 1.83
8.75 ± 2.21
1.86 ± 2.41
71.93 ± 1.10
Run 2 (LH 0.1, I 80, NT 205)
43.15 ± 1.74
52.31 ± 1.94
8.95 ± 2.85
2.05 ± 1.08
73.44 ± 0.95
Run 3 (LH 0.1, I 100, NT 215)
53.02 ± 0.84
65.54 ± 0.89
9.20 ± 1.27
2.13 ± 1.30
79.01 ± 2.01
Run 4 (LH 0.2, I 60, NT 205)
63.40 ± 1.01
75.84 ± 1.38
7.71 ± 1.38
2.35 ± 1.49
84.48 ± 1.28
Run 5 (LH 0.2, I 80, NT 215)
68.35 ± 0.51
82.40 ± 2.01
8.55 ± 1.78
2.40 ± 1.74
92.66 ± 1.87
Run 6 (LH 0.2, I 100, NT 195)
60.33 ± 1.21
74.22 ± 1.89
8.50 ± 2.12
2.19 ± 2.41
82.50 ± 1.91
Run 7 (LH 0.3, I 60, NT 215)
72.06 ± 1.87
84.74 ± 2.24
9.90 ± 1.45
2.52 ± 1.34
93.45 ± 2.41
Run 8 (LH 0.3, I 80, NT 195)
66.34 ± 0.98
79.70 ± 1.56
8.70 ± 1.81
2.36 ± 2.29
85.94 ± 1.01
Run 9 (LH 0.3, I 100, NT 205)
67.33 ± 1.23
81.44 ± 2.03
8.77 ± 1.78
2.48 ± 2.10
86.36 ± 1.58
*LH, layer height; I, infill percentage; NT, nozzle temperature
According to the findings presented in Table 6 and 7, it is evident that experimental run 7 yielded the highest value for all the flexural properties. Specifically, this condition demonstrated a maximum flexural strength of 92.44 MPa for vPLA and 84.74 MPa for rPLA. This result may be attributed to the fact that experimental run 7 was printed with a high nozzle temperature and high layer height, which resulted in improved layer adhesion and minimal void formation; as a consequence, the specimens exhibited excellent flexural properties in comparison to other conditions [94]. In contrast to this, the findings of experimental run 1 yielded unsatisfactory outcomes for both vPLA and rPLA with respective flexural strengths of 64.78 MPa and 46.78 MPa. The stress and strain curves for vPLA (Fig. 11) showed that experimental run 7 possessed maximum strength, but that the strain capacity was not maximum. Experimental run 8 exhibited lower flexural strength in comparison with experimental run 7, but it demonstrated greater ductility because of its better strain-absorbing capacity. This was attributed to the domination of additional porosity within the sample due to the lower melting temperature, which resulted in increased deflection [1]. The stress–strain curves, as illustrated in Fig. 12, provide valuable insights into the mechanical properties of rPLA. Notably, analysis of the stress and strain curves revealed that experimental run 7 exhibited the highest levels of flexural stress and strain amongst all of the samples tested. In comparing the two curves, it became apparent that the strain displayed by rPLA is greater in magnitude when compared to that of vPLA. The rationale behind this phenomenon was that elongation of thermoplastic polymers is directly influenced by the recycling process. As previously mentioned, the thermal degradation process during recycling induces chain scissoring, which subsequently causes a reduction in the strength of the intermolecular bonds within the polymers. Hence, the polymers demonstrated a tendency to undergo elongation [95].

3.5 Impact strength

In relation to the influence of the printing parameters on the impact strength, Fig. 13 presents the average Charpy impact values, which exhibit a direct correlation with the infill percentage and layer thickness. The study conducted by Kamaal et al. [96] found that there was no significant difference in impact strength at low infill percentages. However, this study observed that when the infill percentage reached 100%, and the layer height was 0.3 mm, there was an increase in impact strength. This can be attributed to the reduction in voids within the printed material, as well as the increased amount of material available to absorb the impact energy [97]. The outcome of the impact tests indicated that rPLA exhibited an impact strength that is comparable to that of vPLA. The impact strength of the components made from recycled PLA exhibited only a marginally reduced value (ranging from 4.06 to 11.08%) when compared to the components made from virgin PLA. Experimental run 9 exhibited the highest average impact value for both the vPLA and rPLA specimens. In contrast to this, the sample with the lowest average value was observed in experimental run 1 when the layer thickness was reduced, with this being consistent with Atakok et al. [17], who found a decrease in impact strength as the layer height decreased, with this being attributed to the reduced resistance to crack propagation.

3.6 Fractography studies

The experimental tests revealed that the nature of fracture for the vPLA and rPLA specimens was ductile, which was due to stretching and reorientation of the material that resulted in deformation. The fracture surfaces of the 3D printed PLA samples exhibited a ductile type of fracture. The fracture surfaces were analysed using an optical microscope with results being presented in Figs. 14 and 15. It can be seen from Fig. 14 that a large number of pores were present when the layer thickness was low, even though the infill amount and temperature were high (run 1) as shown in Fig. 14a, which resulted in poor bonding of the material. When the layer thickness and nozzle temperature were high (run 7), a smaller number of pores were present, as shown in Fig. 14b, with this being mainly attributed to the increase in partially melted polymer within the expanded gaps between the printed lines, consequently leading to a notable enhancement in structural integrity. However, when the layer thickness and infill were high, but the nozzle temperature was low (run 9), the presence of pores increased compared to run 7, but it was still very small compared to run 1 (shown in Fig. 14c). In contrast to this, it was observed from Fig. 14 that, although the infill and layer thickness increased, if the nozzle temperature was low then the presence of large voids was noted (Fig. 14d). This was attributed to the low fluidity of the material, which ultimately increased the size and quantity of pores. For the case of rPLA, the same type of observations was made. In this case, large voids were observed for the same experimental conditions as had been noted for vPLA, i.e., Fig. 15a–d. An in-depth investigation utilising microscopy tools was conducted to identify the underlying factors contributing to the observed reduction in mechanical strength caused by variations in the nozzle temperature. The findings indicate that the inadequate flow of material, stemming from a lower nozzle temperature during the deposition process, leads to the formation of gaps and voids within the layers. Although the infill density, as specified in the slicing software, was set to 100%, the presence of voids was observed [98]. Figure 16a–c shows that with decreasing nozzle temperature, the voids between layers increased owing to the poor adhesion between layers, which ultimately decreased the strength of the rPLA samples. In conclusion, this analysis showed a positive correlation between layer thickness and nozzle temperature, i.e., even for the same nominal layer thickness, increasing the nozzle temperature led to a larger actual layer thickness. Thus, the nozzle temperature and layer thickness must be selected appropriately to improve the microstructure of the specimen. Overall, a good correlation was noted between the optical microscopic studies and the resulting experimental data.

3.7 Optimisation of process parameters

The optimisation of process parameters within the FDM 3D printer with respect to hardness, tensile strength, and surface roughness was carried out using Taguchi analysis and the Minitab 21 software. The S/N ratios for tensile, flexural, impact strength, hardness, and surface roughness were evaluated with ‘larger the better’ criteria for tensile properties, flexural properties, impact strength and hardness, and ‘lower the better’ criteria for surface roughness. The optimal levels were determined by utilising the average S/N ratios for each response at each level. A higher value of S/N ratio indicates superior quality characteristics [23]. Therefore, to maximise the mechanical strength, the optimal parameters were selected by choosing the highest S/N ratio for each factor.
Figure 17 shows a comparison of the main effects plot, illustrating the S/N ratios for both the vPLA and rPLA tensile properties, as well as the FDM printing parameters. The figure reveals that layer thickness was the most significant factor affecting tensile properties, and the temperature of the nozzle head also had a substantial impact on strain at UTS. It is anticipated that the melt viscosity of PLA is highly influenced by temperature, and as the nozzle temperature increases during the printing process, the melt viscosity of PLA decreases, thereby leading to improved interlayer adhesion [99]. As shown in Fig. 17, the optimum combination of parameters was found to be A3B3C3, with a layer thickness of 0.3 mm, infill of 100%, and a nozzle temperature of 215 °C.
Figure 18 illustrates the S/N ratio and main effects plot for flexural strength (FS) for the larger-the-better criteria. The FS exhibited an almost identical parametric response to that of the tensile strength. Here, it was determined that the flexural characteristics exhibited maximum values when the layer thickness, infill density, and nozzle temperature were adjusted to high levels. Therefore, for the maximum flexural strength response, 0.3 mm layer thickness, 100% infill, and 215 °C nozzle temperature were determined to be the optimal combination of parameters. The reasoning behind this combination was the same as already discussed for the case of tensile strength.
From a dynamic application standpoint, maximising the required impact strength would be important [21]. In the current study, the resulting Charpy impact strengths were examined under the condition of larger the better, as illustrated in Fig. 19. Analysis of the S/N ratio response plots revealed that the layer height and infill significantly affected the impact strength of the vPLA and rPLA-printed specimens. However, it was observed that nozzle temperature did not significantly influence the impact properties, such as had been the case for TS and FS. Therefore, 0.3 mm layer thickness, 100% infill, and 205 °C nozzle temperature were determined to be the optimal combination to achieve high impact strength. In particular, the interlayer bonding exhibited poor performance at 195 °C but improved when the temperature was raised to 205 °C. However, a further increase in the temperature resulted in increased fluidity of the molten plastic, leading to a decrease in viscosity and the formation of voids. Consequently, this had a slight negative impact on the overall performance in terms of impact strength [97].
The experimental findings were examined in order to determine the optimal printing parameters for minimising surface roughness. The analysis was conducted based on the S/N ratios, employing the ‘smaller the better’ criteria and presented in Fig. 20. The investigation revealed that the primary factor contributing to the reduction in the surface roughness of the specimens was the reduction in layer height. Furthermore, the main effects plot demonstrated that low density had a statistically significant impact on surface roughness. Therefore, the most favourable arrangement of processing parameters for minimising surface roughness was determined as follows: a layer thickness of 0.1 mm, an infill percentage of 60%, and a nozzle temperature of 205 °C for vPLA; a layer thickness of 0.1 mm, an infill percentage of 60%, and a nozzle temperature of 205 °C for rPLA.
Figure 21 presents the S/N ratios and mean plots (using the ‘larger the better’ criterion) for shore D hardness of the 3D printed specimens. Overall, it was found that the specimen hardness increased by increasing the infill density, layer height, and nozzle temperature for vPLA. The reason behind this trend was ascribed to the reduction in porosity, which would be expected to improve the hardness [21]. However, a layer thickness of 0.2 mm exhibited the highest hardness for rPLA and, therefore, the optimum combinations were found to be layer height of 0.3 mm, infill of 100%, and nozzle temperature of 215 ℃ for vPLA and 0.2 mm layer height, 100% infill, and 215 ℃ nozzle temperature for rPLA.

3.8 ANOVA analysis

An ANOVA study was conducted with a confidence level of 95% to examine the relative contribution of the parameters on the individual responses of the tensile strength, flexural strength, impact strength, shore D hardness, and surface roughness properties of both virgin PLA and recycled PLA materials. Table 8 shows the ANOVA results, where the rows labelled as ‘Residual Error’ refer to the errors that arise from uncontrollable factors, specifically noise, which are not accounted for in the experiment and are thus considered as experimental errors [100]. A low residual error was observed for each property that exhibited good control over the processing of components. The main purpose of the percentage contribution is to assess the extent to which different parameters influence mechanical properties. In other words, a higher percentage value indicates a stronger influence of a particular parameter on the response variable [101]. The findings indicated that the layer thickness had a higher percentage contribution to the parameters examined (as shown in Table 8) compared to the infill and nozzle temperatures. Therefore, the major determinant for the properties investigated in this work was found to be the layer thickness. The second factor that exerted an influence on the system was the nozzle temperature. In contrast to this, the influence of the infill percentage on the outcome was found to be negligible (at least within the range of 60–100% investigated in this work). From the ANOVA results, it was determined that overall optimum tensile and flexural strength could be achieved by selecting high layer thickness, high nozzle temperature, and low infill density.
Table 8
ANOVA responses for S/N ratios of various properties examined in this work
Response
Source
DF
Seq SS
Adj SS
Adj MS
F
p
% contribution
Ultimate tensile strength (vPLA)
LH
2
3.09
3.09
1.54
130.35
0.008
68.04
Infill
2
0.19
0.19
0.09
8.17
0.109
4.27
NT
2
1.23
1.23
0.61
52.05
0.019
27.17
Residual error
2
0.023
0.02
0.01
  
0.52
Total
8
4.55
    
100.00
Ultimate tensile strength (rPLA)
LH
2
2.94
2.94
1.47
273.29
0.004
73.48
Infill
2
0.14
0.14
0.07
13.04
0.071
3.51
NT
2
0.91
0.91
0.45
84.58
0.012
22.74
Residual error
2
0.01
0.01
0.00
  
0.27
Total
8
4.01
    
100.00
Flexural strength (vPLA)
LH
2
5.97
5.97
2.98
39.68
0.025
74.55
Infill
2
0.43
0.43
0.22
2.86
0.259
5.38
NT
2
1.46
1.46
0.73
9.68
0.094
18.19
Residual error
2
0.15
0.15
0.08
  
1.88
Total
8
8.00
    
100.00
Flexural strength (rPLA)
LH
2
23.50
23.50
11.75
23.61
0.041
80.39
Infill
2
1.42
1.42
0.71
1.43
0.412
4.86
NT
2
3.31
3.31
1.65
3.33
0.231
11.35
Residual error
2
0.99
0.99
0.49
  
3.40
Total
8
29.23
    
100.00
Impact strength (vPLA)
LH
2
2.35
2.35
1.17
10.88
0.084
47.72
Infill
2
1.72
1.72
0.86
7.96
0.112
34.91
NT
2
0.64
0.64
0.32
2.96
0.252
12.99
Residual error
2
0.22
0.22
0.11
  
4.38
Total
8
4.92
    
100.00
Impact strength (rPLA)
LH
2
1.44
1.44
0.72
4.16
0.194
35.98
Infill
2
1.62
1.62
0.81
4.68
0.176
40.50
NT
2
0.60
0.60
0.30
1.72
0.368
14.86
Residual error
2
0.35
0.35
0.17
  
8.65
Total
8
4.01
    
100.00
Shore D hardness (vPLA)
LH
2
0.003
0.003
0.001763
2.46
0.289
32.11
Infill
2
0.004
0.004
0.0021
2.93
0.254
38.24
NT
2
0.001
0.001
0.000911
1.27
0.44
16.59
Residual error
2
0.001
0.001
0.000716
  
13.05
Total
8
0.011
    
100.00
Shore D hardness (rPLA)
LH
2
0.028
0.028
0.014
28.77
0.034
68.94
Infill
2
0.008
0.008
0.004
8.13
0.11
19.49
NT
2
0.003
0.003
0.001
3.83
0.207
9.17
Residual error
2
0.001
0.001
0.0004
  
2.40
Total
8
0.041
    
100.00
Surface roughness Ra (vPLA)
LH
2
47.42
47.42
23.71
85.37
0.012
82.09
Infill
2
1.63
1.63
0.81
2.95
0.253
2.83
NT
2
8.15
8.15
4.07
14.68
0.064
14.11
Residual error
2
0.55
0.55
0.27
  
0.96
Total
8
57.76
    
100.00
Surface roughness Ra (rPLA)
LH
2
43.64
43.64
21.82
44.58
0.022
75.99
Infill
2
0.85
0.85
0.42
0.87
0.536
1.48
NT
2
11.96
11.96
5.98
12.22
0.076
20.83
Residual error
2
0.98
0.98
0.49
  
1.70
Total
8
57.43
    
100.00
*DF represents the degree of freedom; Seq SS represents the sequential sum of squares; Adj SS represents the adjusted sum of squares; Adj MS represents the adjusted mean square; F represents the ratio of explained variance to unexplained variance; p represents the probability of obtaining F value

3.9 DoE confirmation test

The optimal parameters resulting in the most favourable mechanical properties for the vPLA and rPLA specimens have been summarised in Table 9. A confirmation experiment was conducted to validate the potential of the optimised parameter conditions determined by Taguchi analysis to improve the mechanical properties. The test was necessary to validate the conditions suggested by the Taguchi analysis, which resulted in the optimal properties. The predicted results for the optimal conditions were determined solely by considering the optimum performance factors. Confirmation test results for several mechanical properties, such as tensile strength, flexural strength, impact strength, shore D hardness, and surface roughness, were obtained using the optimal combination of factors, as listed in Table 9. The obtained results for these properties were found to be 61.25 MPa and 45.97 MPa for tensile strength, 104.05 MPa and 89.87 MPa for flexural strength, 5.61 and 5.37 kJ/m2 for impact strength, 82.30 and 79.60 for shore D hardness, and 1.97 μm and 2.28 μm for surface roughness of vPLA and rPLA specimens, respectively. The results demonstrated that the samples produced using the optimal combination of factors exhibited higher levels of accuracy than those generated by the L9 orthogonal array. Therefore, the S/N ratio of the selected combination of factors was employed for S/Nconfirmation.
Table 9
Summary of the optimum parameters required to attain the highest possible mechanical characteristics
Factors
Tensile Strength
Flexural Strength
Impact Strength
vPLA
rPLA
vPLA
rPLA
vPLA
rPLA
Level
Description
Level
Description
Level
Description
Level
Description
Level
Description
Level
Description
Layer Height (mm)
3
0.3
3
0.3
3
0.3
3
0.3
3
0.3
3
0.3
Infill (%)
3
100
3
100
3
100
3
100
3
100
3
100
Nozzle Temperature (°C)
3
215
3
215
3
215
3
215
2
205
3
215
Factors
Shore D hardness
Surface Roughness
vPLA
rPLA
vPLA
rPLA
Level
Description
Level
Description
Level
Description
Level
Description
Layer Height (mm)
3
0.3
3
0.3
1
0.1
1
0.1
Infill (%)
3
100
3
100
1
60
2
60
Nozzle Temperature (°C)
3
215
3
215
2
205
2
205
The optimal combination of factors was determined by calculating the predicted values. The comparison between the confirmation test data and the predicted results is presented in Table 10. It is widely recognised in the engineering community that achieving a 95% confidence interval is a fundamental principle in engineering tests in order to ensure high-quality products and testing validity [102]. Therefore, the CI for both the confirmed and predicted outcomes were determined. For all the attributes listed in Table 10, the findings showed a significant degree of overlap between the CI obtained from the confirmation tests and predicted outcomes. A significant observation was made for the surface roughness property, as the experimental value showed a comparatively better result than the predicted value. The primary reason for this was the improvement in interlayer adhesion when the rPLA samples were printed using an optimal combination of printing parameters [17, 97]. This result suggests that the selected combination of factors in terms of material preparation and processing factors was acceptable and could be considered as optimal conditions during manufacturing. The ANOVA results suggested that the significant factors could be accurately predicted and combined to achieve good reproducibility.
Table 10
Confirmation test summary based on optimum factors
Output response
Prediction
Confirmation
\({S/N}_{{\text{predicted}}}\)
\({C.I}_{{\text{predicted}}}\)
\({S/N}_{{\text{confirmation}}}\)
\({C.I}_{{\text{confirmation}}}\)
Tensile strength
vPLA
35.80
 ± 2.54
35.81
 ± 0.21
rPLA
32.21
 ± 2.07
33.11
 ± 1.21
Flexural strength
vPLA
39.74
 ± 6.48
40.34
 ± 1.45
rPLA
39.43
 ± 8.44
39.07
 ± 1.47
Impact strength
vPLA
14.75
 ± 0.31
14.97
 ± 0.12
rPLA
14.12
 ± 0.26
14.60
 ± 0.13
Shore D hardness
vPLA
38.21
 ± 0.96
38.23
 ± 1.41
rPLA
37.62
 ± 1.14
38.07
 ± 1.12
Surface roughness
vPLA
 − 14.11
 ± 2.12
 − 5.90
 ± 0.17
rPLA
 − 15.35
 ± 2.37
 − 7.19
 ± 0.14

4 Conclusions

The present study employed vPLA and rPLA filaments to investigate the influence of the layer thickness, infill percentage, and nozzle temperature on 3D printed specimen surface roughness, hardness, tensile, flexural, and Charpy impact properties. This study used the Taguchi L9 OA method to design the experiments, whilst the obtained results were additionally analysed by ANOVA to determine the optimal printing parameters for maximising the mechanical strength of recycled PLA specimens. The experimental results indicated that vPLA possessed overall stronger mechanical properties compared to rPLA at all the selected printing factors. The rPLA specimen properties exhibited lower overall results than the vPLA for the same layer thickness and nozzle temperature, but the former with a high layer thickness and high nozzle temperature could attain almost the same property as the latter. The main effect plots show that the tensile and flexural strengths exhibited the highest values when the layer thickness and nozzle temperature were high. ANOVA analysis also indicated that layer thickness had the highest significance amongst the input variables, with p values below 0.05, for all responses. In addition, the temperature of the nozzle also played a significant influence in enhancing the tensile characteristics. In contrast, the infill density had an insignificant influence on the range of values encountered in this study (60–100%). The optimum printing parameters for FDM using vPLA and rPLA filament based on ANOVA results for tensile, flexural, impact, and shore D hardness tests were 0.3 mm layer thickness, 215 ℃ nozzle temperature, and 100% infill density with the values for surface roughness being 0.1 mm, 205 ℃, and 80%, respectively. The results revealed that for a high layer height, infill amount, and nozzle temperature, the bonding between layers increased, which also had a beneficial effect on the mechanical strength. Microscopic investigations showed that a high nozzle temperature and layer height resulted in a stronger microstructure. In contrast to this, the existence of voids was observed when the temperature was low (195 ℃), causing a reduction in mechanical properties. Finally, the predicted responses for individual parameters were verified through experimental confirmation. These additional experiments showed that using the optimally determined printing parameters, rPLA achieved a high tensile strength of 45.97 MPa, which was comparable to the tensile strength of vPLA when printed with 0.1 mm layer thickness, 215 °C nozzle temperature, and 80% infill density. Therefore, the experimental investigation proved that rPLA could be a feasible environmentally friendly alternative material for virgin PLA owing to the similarity of its mechanical properties under optimum processing conditions. Since this study evaluated the optimal printing parameters for the layer thickness, infill, and nozzle temperature of recycled PLA specimens, future research could focus on choosing other important printing parameters, such as printing speed, infill angle for the evaluation of mechanical properties, and determining the fatigue properties of vPLA and rPLA specimens. The investigation and minimisation of the residual stress could be an interesting field for future research. In addition, investigating the use of blends of recycled and virgin PLA may be useful in 3D printing in order to evaluate the resulting mechanical properties and attendant sustainability benefits.

Acknowledgements

The authors are grateful to the technical staff of the School of Civil and Mechanical Engineering, Curtin University, Australia, for their valuable technical support.

Declarations

The article followed the guidelines of the Committee on Publication Ethics (COPE) and involved no studies on human or animal subjects.
All authors read and approved the final manuscript.

Competing interests

The authors declare no competing interests.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Literatur
6.
Zurück zum Zitat Pulok MKH, Rahman MS, Chakravarty UK (2021) Crack Propagation and Fracture Toughness of Additively Manufactured Polymers. Proceedings of the ASME 2021 International Mechanical Engineering Congress and Exposition Volume 4: Advances in Aerospace Technology. Virtual, Online. November 1-5, 2021. ASME. https://doi.org/10.1115/imece2021-72061 Pulok MKH, Rahman MS, Chakravarty UK (2021) Crack Propagation and Fracture Toughness of Additively Manufactured Polymers. Proceedings of the ASME 2021 International Mechanical Engineering Congress and Exposition Volume 4: Advances in Aerospace Technology. Virtual, Online. November 1-5, 2021. ASME. https://​doi.​org/​10.​1115/​imece2021-72061
9.
Zurück zum Zitat Uddin MS, Sidek MFR, Faizal MA, Ghomashchi R, Pramanik A (2017) Evaluating mechanical properties and failure mechanisms of fused deposition modeling acrylonitrile butadiene styrene parts. J Manuf Sci Eng 139. https://doi.org/10.1115/1.4036713 Uddin MS, Sidek MFR, Faizal MA, Ghomashchi R, Pramanik A (2017) Evaluating mechanical properties and failure mechanisms of fused deposition modeling acrylonitrile butadiene styrene parts. J Manuf Sci Eng 139. https://​doi.​org/​10.​1115/​1.​4036713
22.
23.
Zurück zum Zitat Ahmad MN, Ishak MR, Mohammad Taha M, Mustapha F, Leman Z, Anak Lukista DD, Irianto Ghazali I (2022) Application of Taguchi method to optimize the parameter of fused deposition modeling (FDM) using oil palm fiber reinforced thermoplastic composites. Polymers (Basel) 14. https://doi.org/10.3390/polym14112140 Ahmad MN, Ishak MR, Mohammad Taha M, Mustapha F, Leman Z, Anak Lukista DD, Irianto Ghazali I (2022) Application of Taguchi method to optimize the parameter of fused deposition modeling (FDM) using oil palm fiber reinforced thermoplastic composites. Polymers (Basel) 14. https://​doi.​org/​10.​3390/​polym14112140
38.
Zurück zum Zitat Maazinejad B, Mohammadnia O, Ali GAM, Makhlouf ASH, Nadagouda MN, Sillanpää M, Asiri AM, Agarwal S, Gupta VK, Sadegh H (2020) Taguchi L9 (34) orthogonal array study based on methylene blue removal by single-walled carbon nanotubes-amine: adsorption optimization using the experimental design method, kinetics, equilibrium and thermodynamics. J Mol Liq 298. https://doi.org/10.1016/j.molliq.2019.112001 Maazinejad B, Mohammadnia O, Ali GAM, Makhlouf ASH, Nadagouda MN, Sillanpää M, Asiri AM, Agarwal S, Gupta VK, Sadegh H (2020) Taguchi L9 (34) orthogonal array study based on methylene blue removal by single-walled carbon nanotubes-amine: adsorption optimization using the experimental design method, kinetics, equilibrium and thermodynamics. J Mol Liq 298. https://​doi.​org/​10.​1016/​j.​molliq.​2019.​112001
42.
Zurück zum Zitat Freeland B, McCarthy E, Balakrishnan R, Fahy S, Boland A, Rochfort KD, Dabros M, Marti R, Kelleher SM, Gaughran J (2022) A review of polylactic acid as a replacement material for single-use laboratory components. Materials 15. https://doi.org/10.3390/ma15092989 Freeland B, McCarthy E, Balakrishnan R, Fahy S, Boland A, Rochfort KD, Dabros M, Marti R, Kelleher SM, Gaughran J (2022) A review of polylactic acid as a replacement material for single-use laboratory components. Materials 15. https://​doi.​org/​10.​3390/​ma15092989
47.
Zurück zum Zitat Oloyede CT, Jekayinfa SO, Alade AO, Ogunkunle O, Laseinde OT, Adebayo AO, Abdulkareem AI, Smaisim GF, Fattah IMR (2023) Synthesis of biobased composite heterogeneous catalyst for biodiesel production using simplex lattice design mixture: optimization process by Taguchi method. Energies 16. https://doi.org/10.3390/en16052197 Oloyede CT, Jekayinfa SO, Alade AO, Ogunkunle O, Laseinde OT, Adebayo AO, Abdulkareem AI, Smaisim GF, Fattah IMR (2023) Synthesis of biobased composite heterogeneous catalyst for biodiesel production using simplex lattice design mixture: optimization process by Taguchi method. Energies 16. https://​doi.​org/​10.​3390/​en16052197
51.
Zurück zum Zitat Ramesh M, Panneerselvam K (2020) PLA-Based Material Design and Investigation of Its Properties by FDM. In: Shunmugam MS, Kanthababu M (eds) Advances in Additive Manufacturing and Joining. Lecture Notes on Multidisciplinary Industrial Engineering. Springer, Singapore. pp. 229–241. https://doi.org/10.1007/978-981-32-9433-2_20 Ramesh M, Panneerselvam K (2020) PLA-Based Material Design and Investigation of Its Properties by FDM. In: Shunmugam MS, Kanthababu M (eds) Advances in Additive Manufacturing and Joining. Lecture Notes on Multidisciplinary Industrial Engineering. Springer, Singapore. pp. 229–241. https://​doi.​org/​10.​1007/​978-981-32-9433-2_​20
57.
Zurück zum Zitat Mishra P, Sood S, Bharadwaj V, Aggarwal A, Khanna P (2023) Parametric modeling and optimization of dimensional error and surface roughness of fused deposition modeling printed polyethylene terephthalate glycol parts. Polymers (Basel) 15. https://doi.org/10.3390/polym15030546 Mishra P, Sood S, Bharadwaj V, Aggarwal A, Khanna P (2023) Parametric modeling and optimization of dimensional error and surface roughness of fused deposition modeling printed polyethylene terephthalate glycol parts. Polymers (Basel) 15. https://​doi.​org/​10.​3390/​polym15030546
61.
Zurück zum Zitat ASTM-D7127-17 (2017) Standard test method for measurement of surface roughness of abrasive blast cleaned metal surfaces using a portable stylus instrument. ASTM International. West Conshohocken, PA, United States. https://doi.org/10.1520/d7127-17 ASTM-D7127-17 (2017) Standard test method for measurement of surface roughness of abrasive blast cleaned metal surfaces using a portable stylus instrument. ASTM International. West Conshohocken, PA, United States. https://​doi.​org/​10.​1520/​d7127-17
64.
Zurück zum Zitat ASTM-D790-17 (2017) Standard test methods for flexural properties of unreinforced and reinforced plastics and electrical insulating materials. ASTM International. West Conshohocken, PA, United States. https://doi.org/10.1520/D0790-17 ASTM-D790-17 (2017) Standard test methods for flexural properties of unreinforced and reinforced plastics and electrical insulating materials. ASTM International. West Conshohocken, PA, United States. https://​doi.​org/​10.​1520/​D0790-17
68.
Zurück zum Zitat Vidakis N, David C, Petousis M, Sagris D, Mountakis N (2023) Optimization of key quality indicators in material extrusion 3D printing of acrylonitrile butadiene styrene: the impact of critical process control parameters on the surface roughness, dimensional accuracy, and porosity. Mater Today Commun 34. https://doi.org/10.1016/j.mtcomm.2022.105171 Vidakis N, David C, Petousis M, Sagris D, Mountakis N (2023) Optimization of key quality indicators in material extrusion 3D printing of acrylonitrile butadiene styrene: the impact of critical process control parameters on the surface roughness, dimensional accuracy, and porosity. Mater Today Commun 34. https://​doi.​org/​10.​1016/​j.​mtcomm.​2022.​105171
69.
Zurück zum Zitat Kim YJA (2014) The Effect of Surface Roughness on Biofilm Adhesion to Clear Heat-cured Poly (methyl Methacrylate) used for Ocular Prostheses. Master of Health Sciences, University of Otago. [Online]. Available: http://hdl.handle.net/10523/4792 Kim YJA (2014) The Effect of Surface Roughness on Biofilm Adhesion to Clear Heat-cured Poly (methyl Methacrylate) used for Ocular Prostheses. Master of Health Sciences, University of Otago. [Online]. Available: http://​hdl.​handle.​net/​10523/​4792
70.
Zurück zum Zitat Kackar RN (1989) Taguchi’s quality philosophy: analysis and commentary. In: Dehnad K (ed) Quality Control, Robust Design, and the Taguchi Method. Springer, US Boston, MA, pp 3–21CrossRef Kackar RN (1989) Taguchi’s quality philosophy: analysis and commentary. In: Dehnad K (ed) Quality Control, Robust Design, and the Taguchi Method. Springer, US Boston, MA, pp 3–21CrossRef
71.
Zurück zum Zitat Peace GS (1993) Taguchi methods: a hands-on approach. Addison Wesley Publishing Company Peace GS (1993) Taguchi methods: a hands-on approach. Addison Wesley Publishing Company
72.
Zurück zum Zitat Phadke MS, Kackar RN, Speeney DV, Grieco MJ (1989) Off-line quality control in integrated circuit fabrication using experimental design. In: Dehnad K (ed) Quality Control, Robust Design, and the Taguchi Method. Springer, US Boston, MA, pp 99–141CrossRef Phadke MS, Kackar RN, Speeney DV, Grieco MJ (1989) Off-line quality control in integrated circuit fabrication using experimental design. In: Dehnad K (ed) Quality Control, Robust Design, and the Taguchi Method. Springer, US Boston, MA, pp 99–141CrossRef
73.
Zurück zum Zitat Dehnad K (2012) Quality control, robust design, and the Taguchi method Springer New York, NY Dehnad K (2012) Quality control, robust design, and the Taguchi method Springer New York, NY
75.
Zurück zum Zitat Ross PJ (1988) Taguchi techniques for quality engineering: loss function, orthogonal experiments, parameter and tolerance design. McGraw-Hill Ross PJ (1988) Taguchi techniques for quality engineering: loss function, orthogonal experiments, parameter and tolerance design. McGraw-Hill
76.
Zurück zum Zitat Chen W-H, Carrera Uribe M, Kwon EE, Lin K-YA, Park Y-K, Ding L, Saw LH (2022) A comprehensive review of thermoelectric generation optimization by statistical approach: Taguchi method, analysis of variance (ANOVA), and response surface methodology (RSM). Renew Sustain Energy Rev 169. https://doi.org/10.1016/j.rser.2022.112917 Chen W-H, Carrera Uribe M, Kwon EE, Lin K-YA, Park Y-K, Ding L, Saw LH (2022) A comprehensive review of thermoelectric generation optimization by statistical approach: Taguchi method, analysis of variance (ANOVA), and response surface methodology (RSM). Renew Sustain Energy Rev 169. https://​doi.​org/​10.​1016/​j.​rser.​2022.​112917
81.
Zurück zum Zitat Belavendram N (1995) Quality by design: taguchi techniques for industrial experimentation. Prentice Hall Belavendram N (1995) Quality by design: taguchi techniques for industrial experimentation. Prentice Hall
89.
Zurück zum Zitat Hsueh MH, Lai CJ, Liu KY, Chung CF, Wang SH, Pan CY, Huang WC, Hsieh CH, Zeng YS (2021) Effects of printing temperature and filling percentage on the mechanical behavior of fused deposition molding technology components for 3d printing. Polymers (Basel) 13. https://doi.org/10.3390/polym13172910 Hsueh MH, Lai CJ, Liu KY, Chung CF, Wang SH, Pan CY, Huang WC, Hsieh CH, Zeng YS (2021) Effects of printing temperature and filling percentage on the mechanical behavior of fused deposition molding technology components for 3d printing. Polymers (Basel) 13. https://​doi.​org/​10.​3390/​polym13172910
93.
Zurück zum Zitat Puchalski M, Kwolek S, Szparaga G, Chrzanowski M, Krucinska I (2017) Investigation of the influence of PLA molecular structure on the crystalline forms (alpha’ and alpha) and mechanical properties of wet spinning fibres. Polymers (Basel) 9. https://doi.org/10.3390/polym9010018 Puchalski M, Kwolek S, Szparaga G, Chrzanowski M, Krucinska I (2017) Investigation of the influence of PLA molecular structure on the crystalline forms (alpha’ and alpha) and mechanical properties of wet spinning fibres. Polymers (Basel) 9. https://​doi.​org/​10.​3390/​polym9010018
99.
Zurück zum Zitat Akhoundi B, Nabipour M, Hajami F, Shakoori D (2020) An experimental study of nozzle temperature and heat treatment (annealing) effects on mechanical properties of high-temperature polylactic acid in fused deposition modeling. Polym Eng Sci 60:979–987. https://doi.org/10.1002/pen.25353CrossRef Akhoundi B, Nabipour M, Hajami F, Shakoori D (2020) An experimental study of nozzle temperature and heat treatment (annealing) effects on mechanical properties of high-temperature polylactic acid in fused deposition modeling. Polym Eng Sci 60:979–987. https://​doi.​org/​10.​1002/​pen.​25353CrossRef
Metadaten
Titel
Impact of process parameters on improving the performance of 3D printed recycled polylactic acid (rPLA) components
verfasst von
Mohammad Raquibul Hasan
Ian J. Davies
Alokesh Pramanik
Michele John
Wahidul K. Biswas
Publikationsdatum
20.02.2024
Verlag
Springer London
Erschienen in
The International Journal of Advanced Manufacturing Technology / Ausgabe 7-8/2024
Print ISSN: 0268-3768
Elektronische ISSN: 1433-3015
DOI
https://doi.org/10.1007/s00170-024-13140-7

Weitere Artikel der Ausgabe 7-8/2024

The International Journal of Advanced Manufacturing Technology 7-8/2024 Zur Ausgabe

    Marktübersichten

    Die im Laufe eines Jahres in der „adhäsion“ veröffentlichten Marktübersichten helfen Anwendern verschiedenster Branchen, sich einen gezielten Überblick über Lieferantenangebote zu verschaffen.