Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

3D printing of biomass-derived composites: application and characterization approaches

Anqi Ji a, Shuyang Zhangb, Samarthya Bhagiac, Chang Geun Yoo*a and Arthur J. Ragauskas*bcd
aDepartment of Chemical Engineering, State University of New York College of Environmental Science and Forestry, Syracuse, NY 13210, USA. E-mail: cyoo05@esf.edu; Tel: +1-315-470-6516
bDepartment of Chemical and Biomolecular Engineering, University of Tennessee, Knoxville, TN 37996, USA. E-mail: aragausk@utk.edu; Tel: +1-865-974-2042
cBiosciences Division, Oak Ridge National Laboratory, Oak Ridge, TN 37831, USA
dDepartment of Forestry, Wildlife, and Fisheries, Center for Renewable Carbon, The University of Tennessee Institute of Agriculture, Knoxville, TN 37996, USA

Received 22nd April 2020 , Accepted 29th May 2020

First published on 8th June 2020


Abstract

Three-dimensional (3D) printing is an additive manufacturing technique with a wide range of 3D structure fabrication and minimal waste generation. Recently, lignocellulosic biomass and its derivatives have been used in 3D printing due to their renewable nature and sustainability. This review provides a summary of the development of different types of biomass and its components such as cellulose and lignin in 3D printing, brief data analysis and introduction to characterization methods of the 3D printed composites. Mechanical properties such as tensile properties, Izod impact properties, and flexural properties, thermal properties and morphological properties of 3D-printed composites are discussed. In addition, other available characterization methods of 3D-printed composites are reported. The future direction of biomass and its derivatives in the field of 3D printing is also discussed.


1. Introduction

Three-dimensional (3D) printing, also known as additive manufacturing (AM), is a process that makes physical components from 3D model data, by building the components layer by layer.1–3 It can fabricate self-supporting 3D objects without molds with single or multiple materials in a short period of time.4 Early in the 1980s, Kodama reported a method to fabricate a 3D plastic model by layer-by-layer stacking with masks to form each photosynthesized layer, which was considered as the earliest reported 3D printing method and the prototype of stereolithography apparatus.5 During the last few decades, 3D printing has evolved into various types to accommodate the printing of different species of materials. Recent advances in computer technology have made 3D printing user-friendly.6 Versatile properties of the printed structure are available based on the structure–processing–property relationship with diverse materials.2 These advantages make 3D printing applicable in many fields such as aerospace, automotive, medical, architecture, and construction.

In recent decades, the demands for depleting petroleum resources and environmental concerns have focused attention on eco-friendly materials, including the field of 3D printing. Lignocellulosic biomass is one of the most abundant raw materials on earth and under intensive focus. It is mainly composed of cellulose, hemicellulose, and lignin, as well as a small amount of proteins and other extractives.7 It has been utilized widely because of its sustainability, non-toxicity, and biocompatibility in several applications, such as papermaking, biofuels and biocomposite materials. Biomass is biodegraded to small molecules8 (e.g., monomeric sugars, carbon dioxide, methane, water) by microorganisms,9,10 which participate in the carbon cycle that is eco-friendly compared with petroleum-based materials. This characteristic of biomass also allows many biomass-derived products to have a certain degree of biodegradability. Recent scientific and technological progress, which helped in a broader and deeper understanding of biomass characteristics,11–16 has enabled an expanded utilization. The development of biomass-derived materials using 3D printing technology as an alternative to fossil oil-based plastics will provide an opportunity to achieve sustainable and renewable bioeconomy.17,18 With the increase of demands for renewable materials and the advancement of 3D printing technology, the use of biomass-derived materials for 3D printing has been widely studied. As shown in the Fig. 1, the number of patents for 3D printing using biomass and its components revealed an increasing tendency. Among them, the most researched is to use cellulose for 3D printing. In the past five years, the number of patents for 3D printing with cellulose has reached about 5100, almost double the number before 2000. This trend implies that the application of biomass and its components in 3D printing has become a hot topic and cellulose for 3D printing has been widely used.


image file: d0ra03620j-f1.tif
Fig. 1 Number of patents for cellulose and biomass-derived in 3D printing.

In this review, biomass-derived 3D printing materials, including major biomass components such as cellulose and lignin, whole-cell wall, and other potential biomass components, are discussed. Recent 3D printing technologies and their characterization approaches for the printing composites are also reviewed.

2. 3D printing techniques for biomass-derived materials

In 3D printing, target objects are built layer by layer by the printer according to the codes that can be executed by 3D printing software. Noteworthily, the 3D techniques are mainly distinguished by the way how each layer is made and attached to its contiguous layers.4 3D printing techniques applied to biomass-derived materials are introduced in this section.

2.1 Fused deposition modeling (FDM)

Fused deposition modeling (FDM), also named as fused filament fabrication (FFF), is a material extrusion method to print 3D structures.3 As shown in Fig. 2a, the prefabricated filament is mechanically fed into a liquefier system by a pair of gears. The feeding rate is controlled by the printing software (e.g., Repetier). The diameter of the filament is decided according to the default gap between the two gears to ensure proper friction between the gears and the filament. When the filament is continuously fed into the liquefier, the subsequent cold end will push the preceding melt end through a fixed-diameter nozzle onto a preheated plate or the previous layer with continuous lines. The target pattern is then printed and accumulated on the Z-axis. FDM is widely applied for thermoplastic polymers, such as acrylonitrile butadiene styrene (ABS),19–21 polylactic acid (PLA),19,20,22 polyamide (PA),23 and polycarbonate (PC).24 In recent reports, biomass was added as additional components or reinforcements to extend the functionality of the 3D printed products.2 The addition of biomass components affect the polymer matrix in processing behaviors, with respect to rheology properties and the thermal stability. A “printing zone” was introduced to unify the important factors such as temperature, extruding rates and the viscosity for different materials and give proper processing conditions for the FDM printing (Fig. 2b).25
image file: d0ra03620j-f2.tif
Fig. 2 (a) Mechanism of FDM/FFF. (b) “Printing zone” defined in an FDM printing (ABS, HIPS and NBR41–HW represent acrylonitrile–butadiene–styrene, high impact polystyrene and acrylonitrile butadiene rubber with 41 mol% of nitrile contents, respectively). Reprinted with copyright permission from ref. 25. Copyright 2018 Science Advances.

2.2 Direct ink writing (DIW)

DIW is another extrusion-based 3D printing method.3 As presented in Fig. 3, the ink in DIW is usually extruded by a pneumatic pressure system through a nozzle to a platform on an XY plane.26,27 Unlike FDM the filament is melted to the target state with proper flowability and then solidified by cooling, the state of the ink in most DIW is not changed; therefore, the properties of the inks play a pivotal role in the processing. The formulation of the ink needs a proper shear-thinning property to ensure a low resistance in the extrusion, as well as enough yield stress and fast elastic recovery after the extrusion to prevent any collapse of the printed objects.27–30 Thus the ratio of solvent, biomass-derived materials and other components need to be formulated and mixed carefully. To formulate the inks, the biomass-derived materials were first dispersed in the solvent with other components by stirring.28 For the homogeneous suspension of the materials in the solvent, further mixing was performed with homogenizers,31,32 speed mixer,29,30 and/or sonication.26,33 Prior to printing, the ink was degassed to avoid bubbles that might be flaws in the printed structures and ensure the inks are extruded smoothly.30,34 For aqueous-based inks, direct cryo writing (DCW), a new printing method derived from DIW has been developed.35,36 The ink can be printed under relatively cold conditions, thereby fixing the structures immediately. Cellulose was widely used in DIW because of its hydrophilicity, and the resultant hydrogels have been utilized in health-related fields.37,38
image file: d0ra03620j-f3.tif
Fig. 3 Mechanism of DIW.

2.3 Stereolithography (SLA)/digital light processing (DLP)

Stereolithography (SLA) is one of the earliest developed 3D printing techniques. The idea to form each 2D pattern of SLA is to synthesize the photoreactive resin by UV light. Initially, in SLA printing, the plate is immersed in a resin tank filled with curable resins. During the printing, the distance from the plate or the precured layer to the bottom of the tank is controllable as the layer thickness, and the UV light rapidly scans the 2D pattern point by point to initiate the photopolymerization of the resin at the selected positions. After each layer of printing, the distance of the plate to the bottom increases with one-layer height for the next 2D pattern (Fig. 4). DLP runs in a similar way to that of SLA, but the photopolymerization of each target pattern is initiated by the projection of the same UV pattern to the target XY plane instead of points scanning, as shown in Fig. 4. Therefore, the DLP process prints faster than SLA. Recent reports showed that the biomass materials could be utilized in SLA/DLP, which will extend their applications in 3D printing.39–41
image file: d0ra03620j-f4.tif
Fig. 4 Mechanism of SLA/DLP. Redrawn based on ref. 86. Copyright 2019 ACS Omega.

2.4 Binder jetting

Unlike other techniques above, binder jetting deposits the liquid binding agents on the raw materials' powder to form each layer. As shown in Fig. 5, the powder is uniformly spread on the platform or the previous layer during the printing, and then the binder adhesive is selectively deposited to the powder, forming the target pattern. After one layer is formed, the platform drops a layer height followed by spreading another layer of powder by a leveling roller for the next cycle of binder deposition. This technique has shown potential in fabricating drug delivery tablets with cellulose-based materials.42
image file: d0ra03620j-f5.tif
Fig. 5 Mechanism of binder jetting.

3. Biomass-derived 3D printing materials

3.1 Cellulose

Cellulose, widely distributed in plant cell walls, has the highest annual production among natural polymers and has been intensively studied.43 Cellulose forms a plant cell wall with other components such as hemicellulose and lignin (Fig. 6a).44 Depending on the isolation methods and feedstock sources, cellulose has different sizes and shapes such as cellulose nanofibrils (CNF, Fig. 6b), cellulose nanocrystals (CNC, Fig. 6c), and bacterial cellulose (BC, Fig. 6d).45
image file: d0ra03620j-f6.tif
Fig. 6 (a) Schematic of the tree hierarchical structure illustrating the role of cellulose. Reprinted with permission from ref. 44. Copyright 2011 Chemical Society Reviews. (b) SEM image of the CNF, scale bar 6 μm. Reprinted with permission from ref. 30. Copyright 2019 Advanced Functional Materials. (c) TEM image of CNCs, scale bar 100 nm. Reprinted with permission from ref. 30. Copyright 2019 Advanced Functional Materials. (d) SEM image of BC produced by Komagataeibacter xylinus. Scale bar 5 μm. Reprinted with permission from ref. 46. Copyright 2017 RSC Advances.

Recent cellulose-based 3D printing studies are summarized in Table 1 with various applications regarding cellulose's different forms. With the inherent hydrophilic property of cellulose arosing from the abundant hydroxyl groups on its surface, it can be well-dispersed in water as a stable suspension in various forms. This makes cellulose a good ink candidate for DIW 3D printing to form hydrogels for many applications. In the DIW 3D printing method, the cellulose-based suspension can be printed to hydrogel structures directly, where shear-thinning performance, sufficient yield stress to prevent collapse, a finite elastic modulus, and fast elastic recovery are key rheological properties.27–30 Shao et al. utilized the shear thinning behavior of MFC suspension on hydrogel printing at various concentrations employing DIW.28 A 2 wt% MFC suspension was printed to a cube structure with high fidelity but with high shrinkage rates after air drying. Lignosulfonate (LS) can significantly prevent this shrinkage, making MFC/LS a good candidate in DIW. CNF hydrogels can also form effective networks that own shear-thinning properties as well as maintain the printed shapes due to its relatively high aspect ratio. Kuzmenko et al. reported that DIW printed CNF hydrogels as biocompatible matrices for neural tissue engineering in which carbon nanotubes (CNT) were blended as the conductive materials.47 In this report, the interaction between CNF and CNT can be tailored using NaOH with different pH values. To improve the interaction between cellulose fibers in the hydrogel, Sanandiya et al. proposed fungus-like structures utilizing in which small amounts of chitin to fill the gaps between the cellulose fibers, which provides good mechanical performance for the printed structures.48 This method was also provided the capability of printing large items like wind turbine blades. Li et al. introduced 2,2,6,6-tetramethyl-1-piperidinyloxy (TEMPO) mediated oxidation to the surface of CNF (T-CNF) for a better entangled CNF network.32 Furthermore, crosslinking with polyamide–epichlorohydrin enhanced the mechanical performance of the composite. Functional materials based on T-CNF/CNT hydrogels was reported by Li et al. The T-CNF/CNT hydrogels were printed directly into an ethanol solution by a DIW method followed by an in situ solvent-exchange.49 This method fixed and densified the structure at the same time. It resulted in excellent mechanical performance as well as yielding a percolation network of CNTs. Cao et al. also applied the in-site solvent change to their hydrogels made of Ti3C2 transition metal carbides/nitrides (MXene) with T-CNF.50 With the addition of 30% of Ti3C2, the printed hybrid fiber shows a tensile strength of 136.5 ± 21.5 MPa and Young's modulus of 9.3 ± 2.4 GPa. The printed fiber also showed responsive behaviors to multiple external stimuli, which can be further applied as smart textiles.

Table 1 3D printing of cellulose composites
Species Printing methods Contents and sizes of the biomass Form of printed samples Applications Other materials Ref.
Microfibrillated cellulose (MFC)/lignosulfonate (LS) DIW MFC: 0.5 wt%–11.4 wt%; Hydrogel Further carbonization for conductive materials   28
LS: 20 wt%–50 wt%
CNF 1.6 wt% Hydrogel Neural tissue engineering (NTE) 0.4 wt% SWCNT (acidified) 47
  Hydrogel scaffold Cell scaffolds Waterborne PU 55
11.25 wt% Aerogel Triboelectric nanogenerator PDMS, Ag paste 33
CNF 1 wt% Hydrogel Bio-medical field Galactoglucomannan methacrylates (GGMMAs) 58
GGMMAs 1, 2, 3 wt%
2 wt% Hydrogel Conductive hydrogel CNTs 60
CNF and xylan-tyramine (XT) CNF < 1 wt% Hydrogel 4D printing H2O2, horseradish peroxidase 26
CNF/XT (xylan modified with tyramine) <3wt% Hydrogel and aerogel Wound dressings, smart textiles, packaging, or soft robots   61
CNF     Sensors Nisopropylacrylamide 62
Cellulose fiber 50 wt%–89 wt% (<200 μm) Bulk Structural material Chitosan (75–85% deacetylated) 48
Bacterial cellulose nanofibrils (BCNF) <1.4 wt% Hydrogel Tissue engineering Silk fibroin, gelatin and genipin 57
TEMPO–CNF 2.8 wt% Hydrogel Oil/water separation, and electronic related applications Kymene (0.06 wt%) 32
TEMPO–CNF   Aerogel Conductive material CNT 49
TEMPO–CNF   Hybrid fiber Smart textiles MXene 50
CNC <10 wt% Hydrogel Scaffold Oxidized dextran (OD)/gelatin (GEL) 52
3, 5, 10, 20 wt% Hydrogel Scaffold Gelatin 37
6 wt% Hydrogel scaffolds Bio- or medical application Sodium alginate (SA), gelatin 54
20 wt% Hydrogel Rheology study   29
CNC + CNF     Ion sensors Ag nanowhisker 59
Dialdehyde cellulose nanocrystals (DAC) 5, 10, 20 wt% Hydrogel Tissue engineering Gelatin 53
Bacterial cellulose (BC) 0–2.25 wt% Hydrogel Tissue engineering and regenerative medicine Cu2+, alginate, 51
CNC and BAPO modified CNC DLP <6.14 wt% Hydrogel and aerogel High mechanical performance aerogel Pegmem, BAPO–ONa, 34
CNC + TEMPO–CNF DIW >27.5 wt% total cellulose Hydrogel and aerogel High mechanical performance aerogel   30
CNC DCW Overall solid 4 wt% in dispersion Aerogel   Xyloglucan 35
Various Aerogel Green materials Xyloglucan, wood flour 36
Microcrystalline cellulose (MCC) FDM 1, 3, 5 wt% Bulk   PLA 65
CNF 30 wt% Bulk   PLA 64
CNF 30 wt% Bulk   PP block copolymer 63
CNC 0.5, 1 wt% Bulk   ABS 66
CNC <1 wt% Bulk   ABS, methacrylate-based resin 67
CNF ROP-grafted with PLA 0, 1, 3 wt% Bulk   PLA 69
CNC-modified by silica sol 1 wt% Bulk   ABS, KH550 et al. 68
CNF grafted with PLA CNF-g-PLA 1, 3, 5 wt% Bulk   PLA 70
TEMPO–BC 1, 1.5, 2, 2.5 wt% Bulk   PLA 71
Lignin-coated CNC (L-CNC) SLA <1 wt% Bulk   MA 74
CNC 0.5, 1, 2, 4 wt% Bulk   MA 73
CNC <1.2 wt% Bulk Medical industry PEGDA, photoinitiator 39
MMA-modified CNC 0.5, 1, 2, 4 wt%     MMA 75
CNC DLP 0.2, 0.5, 1, 2, 5 wt% Bulk Biomedical application PEGDA, DiGlyDA, photoinitiator 40
MCC Cement printing 0.5, 1, 1.5 wt% Bulk Cement materials Cement 72
Methyl cellulose (MC) Ceramic printing 0.25 wt% Bulk MC-assisted ceramic printing Magnesium aluminate spinel 76
Hydroxypropyl cellulose (HPC) FDM 45 wt% Bulk (low resolution) Drug delivery Theophylline, triacetin 79
Hydroxypropyl cellulose (HPC) Binder jetting 10, 30, 50 wt% Bulk Drug delivery Caffeine (medicine); magnesium stearate and colloidal silicone dioxide; 70 v/v% ethanol 42
Ethyl cellulose (EC) FDM 50–80 wt% Bulk Drug delivery Ibuprofen (medicine), release modifier (hydroxypropyl methylcellulose, sodium alginate, xanthan gum, polyvinyl alcohol) 80
Hydroxyethyl cellulose (HEC) DIW 0.5–2.5 wt% Bulk   Lignin, microcrystalline cellulose, citric acid 81
CA 25–35 wt%   Antimicrobial Acetone 38
CA (to cellulose) <22 wt% Hydrogel Oil/water separation Ethyl acetate 77
Carboxymethylated hydrophilic CNF (Hphil-CNF) + methyltrimethoxysilaned hydrophobic CNF (Hphob-CNF)   Hydrogel Bio-medical   83
Carboxymethyl cellulose (CMC)   Paste Battery Silver nano whisker 82
Cellulose fiber (CF) and CMC CF: 15–45 wt% Between bulk and hydrogel     84
CMC: 5–20 wt%


In DIW printing, CNC hydrogels require a higher concentration to reach the rheology requirement (20 wt%)27,31 to print stable structures compared to printing with CNF or MFC (2 wt%).28 To reach the rheological requirement of the printing as well as the structural stability of the hydrogels after printing, blending polymers that own thixotropy behaviors with cellulose materials showed potential in composites fabrication in DIW. Gutierrez et al. utilized alginate as the matrix to fabricate the hydrogels with bacterial cellulose. The 3D printed hydrogel can be used for antimicrobial applications when loaded with Cu2+.51 Jiang et al. reported a series of results utilizing gelatin/cellulose systems to 3D-print hydrogels for tissue repair.37,52,53 Sultan et al. combined sodium alginate (SA) and gelatin, as a polymer matrix in hydrogels for DIW.54 The results showed that the addition of CNC particles significantly increased the viscosity of the hydrogels, and thereby the SA/gelatin/CNC hydrogels can be printed to fine-structure grids without crosslinking. Chen et al. synthesized a waterborne polyurethane (PU), which was compounded with CNF suspensions.55 The hybrid ink could be printed to stable grids as a scaffold for cell proliferation. The interaction between PU nanoparticles and CNF was reported to be the key to improve the viscosity of the ink and the structural stability of the PU/CNF hydrogel.

To obtain high mechanical performance for cellulose hydrogels, crosslinking is a direct method that has been widely applied in cellulose hydrogels.56 Ions (e.g., Ca2+ in CaCl2) that can introduce reversible crosslinking in the cellulose network hydrogels51 and some other non-ion agents, like genipin that can react with the hydroxyl groups on the surface of cellulose, were studied to strengthen the network in DIW printed cellulose products.37,52,53,57 Another promising approach for crosslinking is to utilize modified biomass derivates as the crosslinking agents. Markstedt et al. added tyramine-substituted xylan (XT) in printable CNF hydrogels, and the XT can form an irreversible crosslinked structure after printing.26 The ion-crosslinked CNF hydrogels were confined by the XT network, which could be swelled and deswelled. By controlling the crosslink density of XT, the hydrogel strength, as well as the swelling behavior, can be adjusted, which will extend the application of the hydrogel in the field of 4D printing where humidity change can induce movement. Xu et al. blended UV-curable galactoglucomannan methacrylates (GGMMAs, a hemicellulose derivative), which own a similar repeating unit to that of CNF, with TEMPO-oxidized CNFs for a DIW ink formulation.58 The GGMMAs were polymerized with TEMPO-oxidized CNF by a UV post-cure and increased the compressive Young's modulus by 10-fold compared to that of CNF hydrogel printed samples. The resultant products were all-biocompatible polymers that can be used in tissue engineering. Modification of cellulose particles is another approach to introduce crosslinking. Siqueira et al. modified CNC particles into methacrylate-based (MA-based) CNCs for its crosslinking under UV exposure with a photoinitiation system.27 However, the modified CNC favored a hydrophobic environment to react with some other MA monomers. Wang et al. reported a method to synthesize bis(acyl)phosphane oxide (BAPO-) modified CNC.34 The CNC–BAPO can be printed with other components (poly(ethylene glycol) methyl ether methacrylate and BAPO–ONa) in an aqueous solution and has up to 1100% of swelling, which showed an approach to fabricate cellulose composites with superior swelling capacity and improved mechanical properties.

Of the many applications, cellulose-based aerogels after selected drying processes of the printed hydrogels is an important research area. The prepared aerogels showed potential applications in value-added areas, like triboelectric nanogenerator33 and ion sensors.59 However, during the drying process, shrinkage may cause the collapse of the structure, which can hinder the transition from hydrogels to aerogels. To study the influence of different drying processes, Hakansson et al. dried CaCl2-crosslinked CNF hydrogels by air drying, surfactant treatment followed by air drying, solvent exchange followed by air drying or freeze drying.60 The results showed that the freeze drying could mostly retain the geometry while the other air-drying methods led to significant shrinkage. In this report, it concluded that controlling the shrinkage can make it possible to obtain structures with higher resolution (comparing to the nozzle size) and mechanical performance (comparing to the freeze-dried samples), where a higher density of the dry mass was claimed to contribute to the higher tensile strength. A similar conclusion was made by Hausmann et al., and the idea of increasing the density of dry mass was applied to obtain high-mechanical aerogels.30 For this approach, water in the printed cellulose (CNC and TEMPO–CNF) hydrogels can be replaced with ethanol, acetone or acetonitrile. After a drying process to eliminate these non-aqueous solvents, a dense cellulose-based aerogel was fabricated. With tailoring the orientation of the fibrils by shear-inducing during the printing process, the resultant cellulose structure showed anisotropy on the mechanical performance with high tensile strength. A 6.1 g hook made of the aerogel can hold a load of up to 4.5 kg (Fig. 7a). This is also proved in Markstedt et al.'s report, that the orientation of the anisotropic fillers can significantly impact the tensile strength of the products.61 Jiang et al. took advantage of the anisotropy of CNF and glass fibers in swelling behaviors.62 The DIW printed self-actuating structures can respond to different environmental stimuli to realize simple logic judgements, with controlled timing of actuation in response. In another report from Hausmann et al., the alignment of the anisotropic particles in extrusion was studied with concentrated CNC hydrogels.29 The polarized-light effect of CNC was utilized to trace the alignment of CNC particles in the hydrogels by in-site polarized light imaging within a pure shearing field. The results indicate that the particles will be aligned when a shear beyond yield stress was applied. The quantitative prediction of the rheology behavior regarding the nozzle dimensions, printing conditions, and rheological properties for the hydrogels was studied and used to tune the printing behavior for cellulose hydrogels in DIW. Recently, a new method derived from DIW, called direct cryo writing (DCW), was developed.35,36 In this method, the hydrogels were extruded to a cooling environment (a cool liquid, plate or chamber) to freeze the printed structures immediately. Kam et al. utilized the DCW method to print XG/CNC hydrogels as well as a composite consisting of CNC, XG and wood flour to a bolt and a nut (Fig. 7b).36


image file: d0ra03620j-f7.tif
Fig. 7 Various studies on 3D printing of cellulose and cellulose derivates. (a) The alignment of cellulose nanofibers (CNF) and nanocrystals (CNC) controlled by the flow in a DIW printing (left), leading to a strong aerogel hook (right). Reprinted with permission from ref. 30. Copyright 2019 Advanced Functional Materials. (b) DCW printing of cellulose composites with other biomasses. Reprinted with permission from ref. 36. Copyright 2019 Advanced Materials Technologies. (c) Wood cell mimicking structure combined FDM printing structure with UV-cured resin and CNC. Reprinted with permission from ref. 67. Copyright 2018 Materials & Design. (d) SLA printing of CNC reinforced structure that can be used in medical fields. Reprinted with permission from ref. 39. Copyright 2017 ACS Applied Materials & Interfaces. (e) MC-assisted ceramic slurry showed unique rheology behavior on printing. Below two images showed the prototype (left) and the sintered counterpart (right). Reprinted with permission from ref. 76. Copyright 2019 Journal of Alloys and Compounds. (f) CA-based oil/water separation mesh and its anti-oil-fouling property. Reprinted with permission from ref. 77. Copyright 2019 ACS Applied Materials & Interfaces.

Due to the high crystallinity of cellulose, its particles have been applied to improve the mechanical properties of 3D printing applications.63–65 In FDM printed samples, the gaps between the printed layers/lines and the interface between the fillers and the matrix can affect their mechanical performance. Various methods have been utilized to solve these challenges with FDM printed cellulose-based composite. Shariatnia et al. proposed a scalable blending method by spraying the CNC aqueous suspensions to each of the printing layers for the gaps of the printed layers/lines, thus the dispersed CNC cluster was included in the printed ABS/CNC composites.66 The CNC agglomerates were observed as “nano-stitches” in SEM images, and the tensile performance was reported to increase with the addition of 0.5–1.0 wt% of CNC. Feng et al. utilized CNC-included UV-curable resins to fill in the gaps/voids in the ABS printed patterns (Fig. 7c).67 Using the UV post-curing, the gaps in the samples were significantly reduced, and the elastic modulus and hardness were improved. In terms of the interface enhancement between cellulose particles and polymer matrix, introducing coupling reagents to the surface of the cellulose increased the bonding on the interface that reduces the warpage and shrinkage after printing.68 Other studies synthesized the same polymer chains of the matrix from the surface of cellulose particles and the printed composites with cellulose-grafted-polymers showed enhanced tensile strength.69,70 Li et al. reported a “Pickering emulsion approach” to promote the dispersity of TEMPO treated CNF in the PLA matrix with a slight improvement of the mechanical performance of resultant products.71 The mechanical properties of PLA were enhanced by the nucleating effect of the hybrid cellulose fillers.69–71 Recently, a report on cement printing with MCC expanded the application of cellulose in the non-polymer-based matrix to reduce its carbon print.72

In the regards of SLA printing, photoreactive resins physically blended with cellulose particles also attracted increasing attention.39,40,73,74 By replacing the commercial methacrylate-based resin with a biocompatible photosensitive resin system, the printed structure can be considered as candidates in the medical field. Palaganas et al. proposed an ink formulation based on poly(ethylene glycol) diacrylate and CNC (PEGDA/CNC) that can print complex structures with good fidelity, potentially applicable in the tissue engineering and reconstructive surgery (Fig. 7d).39 The addition of CNC in ink improved the tensile behavior of the printed products. Li et al. reported that the tensile strength of the products improved from 6 MPa to 8 MPa, and their Young's modulus increased from 52 MPa to 80 MPa with 1.0 wt% of CNC addition.40 The compatibility of PEGDA/CNC can also be enhanced by adding 1,3-diglycerolate diacrylate (DiGlyDA). The dispersity of CNC in a photoreactive was improved by modification of the CNC surface.75

Cellulose derivatives having different physical and chemical properties from cellulose were also utilized in 3D printing. Methyl cellulose (MC) solutions at certain concentrations have a special complex viscosity transition by increasing the temperature. Biswas et al. blended MC in a ceramic-based slurry and found that the ceramic prototypes can be printed and then solidified by increasing the temperature to >40 °C that the sharply increased viscosity of MC can help stabilize the printed structure (Fig. 7e).76 Hydroxypropyl cellulose (HPC), is soluble in water and has been applied in medical applications (e.g., eye drops).78 The thixotropy behavior of the HPC molecular chains also gives it good printability. Arafat et al. FDM-printed tablets with HPC as the matrix for drug delivery and the release behavior of the target medicine (i.e., theophylline) for the tablets were studied by varying structural parameters.79 Infanger et al. also used HPC as a binder combined with 70 v/v% ethanol solution to fabricate a 3D printed drug delivery system with a binder jetting printer.42 Because of the high solubility of HPC in water, the drug release time became less than 120 min in the aforementioned systems.42,79 Yang et al. replaced the polymer matrix from hydrophilic HPC to hydrophobic ethyl cellulose (EC) that lead to a sustained 24 h release behavior for the EC-based drug system.80 In this system, HPC was also applied to adjust the release behavior for its water solubility. Hydroxyethyl cellulose (HEC), widely used in the food and cosmetics industry, was recently applied to fabricate functionally graded materials.81 The gelation process and the rheology behavior of the HEC solution, combined with the embed gradient information in models design, lead to DIW printing of continuous, high-contrast, and multidirectional stiffness gradients materials. Carboxymethyl cellulose (CMC) is another cellulose derivate that showed good rheology properties for DIW printing. A silver nanowhisker conductive network was formed with the assist of the CMC solution and applied in an assembled battery.82 Shin et al. proposed a method to combine hydrophilic carboxymethylated CNF (CM-CNF) and methyltrimethoxysilane-modified CM-CNF to DIW print a cell culture platform which showed the potential in drug delivery.83 Thibaut et al. also introduced CMC as the main component with CNF to obtain distortion-free structures with high resolution in DIW printing. Cellulose acetate (CA) can be dissolved in acetone thus used to be fabricated to gels.84 Unlike cellulose-based hydrogels, solvent (acetone) in CA/acetone gels can be easily removed. Pattinson and Hart took advantage of this feature to print CA/acetone gels where acetone was evaporated soon after DIW printing.38 The obtained material was built with a dense cellulose structure that showed good tensile strength (∼40 MPa). Koh et al. printed meshes with high fidelity by CA concentrated solutions (ethyl acetate as the solvent) and regenerated the CA-based meshes into cellulose-based meshes by NaOH/CH3OH treatment.77 The resultant meshes showed a greater than 95% separation efficiency for water/oil mixture with high flux and anti-oil fouling/self-cleaning ability. This performance was attributed to the abundant hydrogen bonds on the cellulose that form hydration layer/shell, which can prevent oil penetration through the meshes (Fig. 7f).

3.2 Lignin

Lignin, the second most abundant terrestrial biopolymer after cellulose, has been under-utilized and, to date, is mostly used for direct combustion.85 Therefore, the valorization of lignin has drawn great attention in the current biorefinery process. Given that lignin contributes to the hydrophobicity, antimicrobial, and antioxidant activities of the plant cell wall, it can be a reinforcing agent in 3D printing composites. The application of lignin in 3D printing is summarized in Table 2 and recent applications of lignin were mostly focused on mechanical reinforcement in composites. Though these studies for 3D printing of lignin are still at an early stage for further application, the versatile properties of lignin including antiaging, flame retardant, and UV absorption86 can provide various potential functions to the printed products. Spruce lignin was extracted by Tanase-Opedal et al. and applied in FDM 3D printing with polylactide (PLA) at different printing temperatures from 205 °C to 230 °C.87 The stronger layer bonding of the printed composite with lignin resulted in competitive tensile performance. In addition, antioxidant properties were significantly improved by adding lignin. Dominguez-Robles et al. reported that lignin improved the antioxidant capability for an FDM printed PLA/lignin/tetracycline drug delivery material.88 Vaidya et al. also reported the feasibility of lignin in FDM with PHB.89 The composite showed weaker shear-thinning and less shrinkage/thermal contraction with lignin loading, thereby easing the warpage of the printed samples. Liu et al. observed enhanced tensile strength from FDM printed materials with 15% of lignin addition to PLA compared to that of pure PLA and other PLA/biomass composites.90 Mimini et al. investigated the chemical structures, thermal properties, printability with FDM, and physical properties of PLA-based composites blended respectively with pine kraft lignin, beech organosolv lignin, and beech LS.91 All the three materials had slightly lower thermal stability and impact strength when those lignin contents increased (up to 15 wt%). Nguyen et al. considered that the poor mechanical behavior of FDM printed composites was due to poor adhesion between layers.92 Thus, they proposed a method to improve the mechanical behaviors for lignin-based composites by introducing a component (acrylonitrile butadiene rubber 41, NBR41) that can form hydrogen bonds between the polymer (acrylonitrile–butadiene–styrene, ABS) and lignin in the composites. The ABS/NBR41/lignin (5[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]4) printed sample showed almost twice the tensile strength and the elongation at break compared to that of ABS/lignin (6[thin space (1/6-em)]:[thin space (1/6-em)]4), implying that NBR41 played a vital role in the mechanical performance. To further improve the mechanical performance of the ABS/NBR41/lignin (5[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]4), 10 wt% of carbon fibers were added to the printed sample, and the composite showed higher tensile strength and Young's modulus than that of pristine ABS. For a deeper understanding of the structure–processing–property relationship of lignin-based FDM printing composite, Nguyen et al. studied the correlations between chemical structures and rheological behaviors of kraft softwood (SW) and organosolv hardwood (HW) lignin.93 The results showed kraft SW lignin exhibited higher complex viscosity which was attributed to the steric hindrance from the molecular structure than that of organosolv HW lignin. HW lignin was mixed with different polymer (Nylon 12 and ABS, with NBR41) with the presence of carbon fibers and up to 60 wt% of lignin can be added in the composite (Fig. 8a) with good mechanical performances.
Table 2 3D printing of lignin composites
Biomass species Printing methods Highest contents of the biomass Other materials Ref.
Lignin from spruce FDM 40 wt% PLA 87
Softwood kraft lignin 3 wt% PLA (matrix), TC (medicine) 88
Lignin (from Pinus radiata wood chips) 50 wt% PHB 89
Lignin 20 wt% PLA 90
Kraft lignin, beech organosolv lignin and beech lignosulfonate 15 wt% PLA 91
Organosolv hardwood lignin 40 wt% ABS, NBR41, carbon fiber 92
Softwood kraft lignin; organosolv hardwood lignin 60 wt% ABS, NBR41, Nylon 12, carbon fiber 93
Softwood kraft lignin SLA 1 wt% Commercial methacrylate resin 86
Organosolv lignin 3 wt% Polyurethane acrylate/morpholine/tripropylene glycol diacrylate 95
Lignin modified by MA 15 wt% Ethoxylated pentaerythritol tetraacrylate/aliphatic urethane acrylate/urethane acrylate 41



image file: d0ra03620j-f8.tif
Fig. 8 Studies on lignin 3D printing. (a) FDM printing process of lignin-included composite that owns the highest reported lignin contents (60 wt%) and the printed oak leaf. Reprinted with permission from ref. 93 Copyright 2018 Science Advances. (b) SLA printing of lignin-included resin that showed an improvement of the tensile strength. Reprinted with permission from ref. 86. https://pubs.acs.org/doi/abs/10.1021/acsomega.9b02455, Copyright 2019 ACS Omega. Further permissions related to the material excerpted should be directed to the ACS. (c) Modified lignin in SLA printing can be printed with the highest concentration of 15 wt%. Reprinted with permission from ref. 41. Copyright 2018 ACS Applied Materials & Interfaces. Further permissions related to the material excerpted should be directed to the ACS.

Zhang et al. reported that a small amount of lignin (0.2 wt%) enhanced the tensile properties of the printed kraft SW lignin/polymethacrylate composites, extending the application of lignin in SLA as a reinforcement agent (Fig. 8b).86 However, the UV-absorbance of lignin94 hindered the UV-initiated photopolymerization, thus the composite could not be printed with more than 1 wt% loading of lignin. Ibrahim et al. used up to 3 wt% of organosolv lignin (from oil palm empty fruit bunches fibers) in a commercial photocurable resin (polyurethane acrylate/morpholine/tripropylene glycol diacrylate).95 The highest tensile performance (tensile strength and Young's modulus) was obtained with a relatively low lignin content of 0.6 wt%, and the lowest was with 3 wt% lignin. To obtain a high lignin loading composite with modest mechanical properties, Sutton et al. modified organosolv lignin (isolated from Populus trichocarpa and Populus deltoides) into a UV-curable reactant with methacrylic anhydride.41 Up to 15 wt% of the modified lignin was loaded in the methacrylate-based resin and was successfully printed by SLA. Though the Young's modulus of the printed sample decreased, the ultra-tensile strength and the elongation at break increased as the modified lignin was added (Fig. 8c).

3.3 Whole biomass

Unlike cellulose and lignin, whole biomass has been utilized as a feedstock by simple size revision without complex physical and chemical processes. For example, “mud–straw walls” have been widely used as a building material. A similar idea can be applied in terms of whole biomass 3D printing. Among many 3D printing biomass species, wood has been one of the most popular biomasses in 3D printing applications (Table 3). Wood was first introduced by Henke and Treml into Liquid Deposition Modeling (LDM) to fabricate the composite.96 With wood chips/sawdust as the matrix and other materials as binders, printed samples can be used as non-structural materials. Rosenthal et al. proposed an air-dry sawdust/methylcellulose/water system, which can be printed with LDM using an 8 mm nozzle.97 A nozzle with a smaller diameter was expected to increase the resolution of the LDM printed samples. The resolution for this method can be further increased by applying a 3 mm nozzle diameter, reported by Kariz et al., with polyvinyl acetate (PVAc) or urea-formaldehyde (UF) as the binder in the LDM materials system.98 The average bending strength and modulus of the elasticity were compared with PVAc- and UF-based composites, and UF favored the mechanical performance.
Table 3 3D printing of wood composites
Species Printing methods Size of the biomass Contents Other materials Ref.
Wood chips/sawdust LDM 0.8–2 mm   Gypsum, methyl cellulose, sodium silicate and different types of cement 96
Air-dry sawdust from beech and methylcellulose (MC)   ∼90 wt% wood MC (as binder and lubrication) 97
Beech wood powder <0.237 mm 13 wt%–25 wt% PVAc and urea-formaldehyde 98
Wood flour (poplar) FDM Sieve into 140–160 mesh 30 wt% Three types of plasticizer 112
Wood-filled PLA   30 wt% PLA 108
Wood fibre-filled PLA   40 wt% PLA 111
Wood flour   14 wt% PLA 113
Beech wood   10, 20, 30, 40, 50 wt% PLA 102
Beech wood   10, 20, 30, 40, 50 wt% ABS and PLA 99
Beech wood   10, 20, 30, 40, 50 wt% PLA 100
Recycled pine wood   30 wt% PLA/PHA 107
Wood-filled PLA   30 wt% PLA 109
Wood flour   30 wt% PLA 101
Wood-filled filament (commercial)   30–40 wt% PLA 110
Pine powder, bleach pulp, mechanical pulp, newspaper pulp, eucalyptus powder   4, 6, 10, 15, 20 wt% PLA 90
Wastepaper, cardboard, wood flour   <20 wt% Recycled polypropylene, commercial PP 63
Wood   40 wt% PLA, ceramic, metal, carbon fiber 114
Recycled wood fiber   15 wt% PHA and PLA 103
Commercial wood filament     Polymer resin 104
Wood-filled PLA   30 wt% PLA 105
Wood   5 wt% PLA 106


Nowadays, FDM 3D printing is being actively investigated in many wood composite applications. The properties of the resultant products can be affected by various factors. As a filler, wood particles affect the rheology and mechanical performance according to Kariz et al.'s report.99 They found that the trend of storage modulus in PLA/wood composites changed before and after the testing temperature reached the glass transition temperature of the PLA matrix. The tensile strength of beech wood/PLA composites by FDM increased with up to 10 wt% of biomass loading, while it decreased beyond this concentration. Ayrilmis et al. reported that as the beechwood content was increased, the surface roughness and the water contact angle of the composites increased.100 The ascending water contact angle implied the decreasing wettability as more wood was incorporated in the composites.

Given that wood materials consist of various hydrophilic materials, the water/moisture sensitivity of the wood/PLA composites was also a factor that will affect the final properties. Ecker et al. immersed the PLA/wood flour composites in water for 7 days and then tested the impact strength.101 The “softening effect” by water that caused a strong reduction of the mechanical performance of the immersed composites for both PLA and PLA/wood composites, and the softening effect had a stronger influence on the 3D printed samples with higher wood contents. Kariz et al. placed the wood/PLA composites in three climates with various relative humidity (RH, 33%, 65%, and 87%) for several days.102 The results indicated that the wood contents dominated the mechanical behaviors, while the moisture variation did not significantly affect the performance. The water/moisture sensitivity of wood in its composites was utilized by Le Duigou et al.103 The wood fiber oriented in the FDM samples was swelled by water absorption, thereby causing the anisotropic strain of the printed sample. The moisture-driven strain can be a promising property in the moisture-actuation system. Similarly, Correa et al. developed a method for programable hygroscopic wood material.104 These reports utilized the anisotropy as well as the hydrophily of wood in smart materials by mimicking the behaviors of plants. Wood/PLA composites were also used as the matrix on the application of antenna105 and artificial hygromorphs.106

Besides the wood contents, other factors can also affect the properties of resultant FDM printed materials. The setting of printing parameters played important roles in FDM printing. Various printing temperature (210–250 °C) was used by Guessasma et al. to print polymer (PLA and PHA)/recycled pine wood.107 To balance the mechanical performance and the thermal stability of wood contents, 220 °C was shown to be preferred temperature in the printing process. Layer thickness in FDM was also an important parameter. Ayrilmis reported that a thicker layer increased the surface roughness and water contact angles and also brought bigger gaps in the printed samples, resulting in lower tensile performance and higher water absorption capacity.108,109 Vigneshwaran et al. studied the influence of layer thickness with infill density and pattern in FDM printing to evaluate the mechanical performance of the PLA/wood composites.110 Dong et al. compared the mechanical performance under various conditions (infill density, layer thickness and the number of shells and their combinations) and reported that the number of shells in an FDM printing of wood-filled PLA filaments dominated the mechanical performance.111 Other factors like the addition of plasticizer112 and UV posttreatment113 had influences in final properties. Some other biomass-derived materials were also utilized in FDM and compared with wood. Zander et al. reported the mixing of polypropylene (commercial PP and recycled PP) and wastepaper, cardboard, and wood flour, respectively, by solid-state shear pulverization (SSSP) and then printed by FDM.63 The tensile strength results showed that the composites had lower tensile strength compared to that of the pure PP. Liu et al. compared to the filling of wood, carbon fiber, ceramic, aluminum, and copper in FDM PLA composites.114 The mechanical performance was lower with the loading of wood and carbon fiber and higher with loading inorganic fillers.

Table 4 summarizes the 3D printing with other biomass-derived materials. Starch, another type of glucose polymer linked through α-1,4- and α-1,6-glycosidic bonds, is an important component in the human diet.115 The desire for personalized food fabrication drove the research on the 3D printing of starch.116–118 An et al. tried to combine potato starch with Nostoc sphaeroides in 3D food printing.117 The rheological behaviors of the mixture were characterized by various blend ratios and processing conditions in detail. Liu et al. reported a method to adjust the rheological behaviors of mashed potatoes by potato starch for the 3D food printing.119 To understand the influence of the origins of starch on 3D food printing, Zheng et al. studied the difference of corn, potato and wheat starch in the same food printing process in 20 wt% starch hydrogels.120 Wheat starch samples showed the least shrinkages, which own the best fidelity of the final products. For food applications, starch was blended with materials from diverse food resources in 3D food printing. Lille et al. optimized mixing ratio and processing conditions for healthy and customized food products with starch, milk powder, cellulose nanofiber, rye bran, oat protein concentrate, and faba bean protein.121 The pharmaceutical industry is also interested in starch 3D printing for individualized medicine products. Binder jetting printing was utilized for a starch-based drug delivery system, and the obtained drug meet the requirement of the mechanical and disintegration properties.122

Table 4 3D printing of other biomass composites
Species Printing methods Contents Form of products Applications Other materials Ref.
Starch DIW (with mixing channel before the nozzle) 7 wt% Hydrogel Customized healthy food Water, bovine gelatin, sucrose, egg white protein 116
Potato starch and Nostoc sphaeroides 3D food printing (DIW) Potato starch (<0.48 wt%), Nostoc sphaeroides and the derivative (4.8 wt%) Hydrogel Food printing Water 117
Starches (potato, rice, and corn) Hot-extrusion 3D printing <30 wt% Hydrogel Food printing   118
Mashed potatoes and potato starch 3D food printing Mashed potato < 85 wt%; potato starch 15 wt%   Food printing Trehalose 119
Potato, wheat and corn starch High-temperature food printing 20 wt%   Food printing Water 120
Potato starch and pea protein 3D food printing Potato starch > 92 wt%; pea protein < 8 wt%   Food printing Water 127
Pregelatinized starch, microcrystalline cellulose (MCC)   Starch 46 wt%; MCC 5 wt% Bulk Drug and medicine Warfarin sodium, D-sucrose, povidone K30, silicon dioxide 122
Starch DIW Starch 71 wt%–75 wt%   Supporting materials in ceramic 3D printing Polyvinylpyrrolidone 123
Thermal plastic starch FDM 30 wt% Bulk   Plasticizers; compatibilizer, impact modifier 124
Spirulina platensis, Tetraselmis suecica and lignin DIW <3 wt% Bulk Cementitious construction Metakaolin, alkaline activator, bentonite 126
Corn starch/CA (SCA) FDM 50[thin space (1/6-em)]:[thin space (1/6-em)]50 Bulk Medical devices   125


Apart from applications in edible products, starch is also widely used in other fields in terms of additive manufacturing. Yang et al. used a bi-nozzle extrusion system to print the starch slurry with ceramic slurry in which the starch slurry acted as the supporting material.123 In the post sintering process, a complex ceramic structure was obtained, while the starch slurry was thermally decomposed into gases. As a renewable polymer, starch was also researched for replacing petrol-based materials in FDM. To improve the compatibility of starch and petrol-based polymer, Kuo et al. blended debranched starch with ABS with SMA (compatibilizer) and MBS (plasticizer).124 The so-called thermal plastic starch (TPS) was successfully printed by FDM with ABS and TiO2, and carbon black can be used to change the color, which can extend the application of the printed products. Paggi et al. fabricated a filament based on starch and CA with a weight ratio of 50[thin space (1/6-em)]:[thin space (1/6-em)]50, and the physical properties (porosity and mechanical properties) were optimized by varying the processing temperature and extrusion rates.125 Recently, some other biomass such as Spirulina platensis and Tetraselmis suecica, which were acquired from wastewater treatment plants, were also applied in 3D printing.126 The addition of the microalgal biomass wastes eased the extrusion process by decrease the yield stress in an LDM process, which also expanded the possible applications for biomass 3D printing.

4. Characterization methods for biomass-derived 3D printed products

The application of biomass as a feedstock of 3D printing can contribute to carbon-neutrality by replacing the current petroleum-based materials. As discussed in the previous sections, characteristics of 3D printed products can be influenced by the composition of resins, the properties of biomass feedstock and commercial resins, processing methods and others. Proper evaluation of these characteristics and understanding of the correlations of these properties with feedstock and processing factors are crucial to make competitive and commercially feasible 3D printed products. This section will introduce important mechanical, thermal, and other properties of 3D printed products with possible analysis methods and examples from previous studies.

4.1 Mechanical properties

Mechanical properties are basic and essential information of most material products. These characteristics including tensile, Izod impact resistance, and flexural properties have been used to evaluate the performance of biomass-derived 3D printed products. Especially in regards to the alternation of commercial plastic resins with biomass-derived materials, significant changes in mechanical properties have been reported.64,93,124,128,129 Recent advancements in biomass modification and printing techniques allow for the enhancement of those performances.2,130 Depending on the substrates and printing technologies, the mechanical properties of the products vary, affecting the application scopes. In the following context, the definitions of these mechanical properties, their analytical methods, and data interpretation with biomass-derived 3D printed products in the previous studies are summarized.
4.1.1 Tensile properties. The tensile properties such as Young's modulus, ultimate tensile strength, yield stress, elongation at break are typically measured by Universal Testing Machine.131–133 Young's modulus (E), also known as tensile modulus, describes the deformation-stiffness resistance of the material. According to Hooke's law which is the law of elasticity, when the material behaves elastic and within the elastic limit of the material, the stress becomes proportional to the strain, expressed as eqn (1),134 Young's modulus is the slope of the linear relationship of the elastic region.134
 
image file: d0ra03620j-t1.tif(1)
where E is the Young's modulus, MPa; σ is the stress, MPa; ε is the strain, % or unitless.

Stress is defined as the ratio of applied force to cross-sectional area and expressed as:

 
image file: d0ra03620j-t2.tif(2)
where σ is the stress, MPa; F is the applied force, N; A is the cross-sectional area, mm2

Strain is defined as the ratio of length variation to original length:

 
image file: d0ra03620j-t3.tif(3)
where ε is the strain, % or unitless; Δl is the length variation, mm; lo is the initial length, mm.

Several properties such as tensile strength and tensile elongation have been used to analyze the mechanical properties of the printed products. Tensile strength, also known as ultimate tensile strength or ultimate strength, is the maximum stress of the specimen before breaking/failure.135 Similarly, tensile elongation, also termed as tensile elongation at break and fracture stain, is the ratio of the tensile length to the original length at the breakage of the specimen.

Table 5 presents the tensile properties of biomass-derived 3D printed products in previous studies. Young's modulus, tensile strength and tensile elongation of 3D printed products composed of commercial plastics and several types of lignins41,87,93,138,140 (softwood lignin,87 hardwood lignin,41 hydrothermally extracted lignin,93 kraft lignin140) with different weight fractions (0.2–40%) have been tested. Depending on the fraction of lignin and/or type of bio-materials, the values of Young's modulus (375–6600 MPa) and tensile strength (23–52 MPa) varied in previous studies. A few studies reported tensile elongation in a range of 1.5–3.3% with different lignin fractions. Nguyen et al. printed blends of Nylon 12 and organosolv hardwood (HW) lignin (60[thin space (1/6-em)]:[thin space (1/6-em)]40 wt%) with additional 4–16% carbon fibers through the FDM method.93 They found that with the increase of carbon fibers, Young's modulus increased dramatically to ∼7500 MPa compared to ∼1770 MPa for pristine Nylon 12. They concluded that the addition of 4 to 16 wt% of carbon fibers enhanced mechanical stiffness and printing speed.

Table 5 Mechanical properties of biomass-derived 3D printed products
Materials Young's modulus (MPa) Tensile strength (MPa) Tensile elongation (%) Ref.
PLA + softwood lignin (20–40%) 1746–2843 27–46 1.5–2.0 87
Photo-curable polyurethane + organosolv lignin (0.2–3.0%) 4–12 8–25 95
Photo-curable polyurethane + organosolv lignin/graphene (0.2–3.0%) 4.5–13 8–28
Tetra-acrylate oligomer (33–38%) + aliphatic urethane acrylate (33–38%) + monofunctional urethane acrylate (16–19%) + hardwood lignin (5–15%) 3500–6600 41
Alkali-treated bamboo fiber (ABF)/polypropylene (PP)/PLA 21–38 10.6–14.2 136
Thermoplastic polyurethane elastomer (TPU)/wood flour (80[thin space (1/6-em)]:[thin space (1/6-em)]20 wt%) mixture + EPDM-g-MAH (0–10%) 13–17 205.2–591.2 137
Poly(L-lactic acid) + lignin (10–40%) 1275–1888 23–47 2.2–3.3 138
Methacrylate (MA) + lignin-coated cellulose nanocrystals (L-CNC) 610–1230 32–69 2.1–11.0 74
Thermoplastic starches (TPS)/acrylonitrile–butadiene–styrene (ABS) (30[thin space (1/6-em)]:[thin space (1/6-em)]70 wt%) mixture + styrene maleic anhydride (0–1%) + methyl-methacrylate butadiene styrene (0–2%) + TiO2 (0–5%) + carbon black (0–5%) 34–49 124
PLA + bamboo fiber (20%) 51 2.2 139
PLA + alkali-treated bamboo fiber (10–30%) 57–66 6.0–8.5
PLA + lignin (5–15%) 2280–2470 40–52 1.5–2.9 140
PLA + cellulose nanofibrils (10–50%) 3000–9500 50–105 0.5–5.8 64
Cellulose acetate (CA) 2000–2400 42–47 38
Nylon 12/hot water lignin (60[thin space (1/6-em)]:[thin space (1/6-em)]40 wt%) mixture + carbon fibers (0–16%) 2000–7500 40–100 93


The mechanical properties of the printed products with different types of cellulose such as cellulose nanofibrils64 and cellulose acetate38 were also reported as presented in Table 5. Up to 9500 MPa of Young's modulus, 100 MPa of tensile strength, 0.5–5.8% of tensile elongation was observed. Instead of using the cellulose blended with plastic resin,64 Pattinson and Hart used cellulose acetate (CA) powder as a single component in a DIW printing, and their Young's modulus (E) and strength (σ) were 2.2 ± 0.1 GPa and 45.0 ± 1.9 MPa for extrusion-parallel-load and 2.2 ± 0.2 GPa and 44.7 ± 2.2 MPa for extrusion perpendicular-load, respectively, indicating no anisotropy in mechanical properties.38

The mechanical characteristics of 3D printed products with whole biomass (e.g., wood flour, bamboo fiber) were investigated.136,137,139 The tensile strength was changed when bamboo fibers were printed with PLA and polypropylene (PP)/PLA.136,139 Also, it was reported that a significant improvement for the elongation at break was observed from this type of the printed products.137 Bi et al. reported that a thermoplastic polyurethane elastomer (TPU)/wood flour composite showed significant improvement of the elongation at break from 205.26% to 591.17% by adding 4 wt% of ethylene-propylene-diene-monomer grafted maleic anhydride (EPDM-g-MAH) with wood flour.137 They discussed that the increase of elongation at break was possibly due to the esterification between EPDM-g-MAH and wood flour, which provided high interfacial adhesion. In addition, the physical crosslinking of TPU and EPDM can also promote the interface interaction of the composite, which also improves the tensile properties. Long et al. reported that the tensile strength and tensile elongation of alkali-treated bamboo fiber (ABF)/polypropylene (PP)/PLA composites through different methods were 33.730 MPa and 10.80% by injection molding and 21.832 MPa and 14.20% by FDM printing, respectively, with various blending ratios.136 The tensile strength and tensile elongation of silane coupling agent treated bamboo fiber/PP/PLA composites by injection molding and FDM printing were 37.014 MPa and 28.100 MPa, 10.66% and 11.80%, respectively. For the same composite, products obtained by injection molding show better tensile strength than that obtained by FDM processes while the products from FDM methods have a higher elongation at break.

The Young's modulus, tensile strength, and tensile elongation of commercial resins blended with biomass-derived mixture74,95 through 3D printing have also been tested. Ibrahim et al. reported the Young's modulus (4.5–13 MPa) and tensile strength (8–28 MPa) of photo-curable PU blended with 0.2% to 3.0% organosolv lignin/graphene through SLA and concluded that photo-curable PU-0.6% lignin/graphene had high tensile strength and resistance to deformation.95

4.1.2 Other mechanical properties. Izod impact resistance and flexural properties have also been measured for some 3D printed products.91,124,136,139 The Izod impact resistance can be measured on an Izod141,142 or Charpy143,144 impact test machine for the toughness of a material. For the impact testing, a notched sample is subjected to an impact, and the impact strength, which represents the absorbed energy per unit area, is measured. Flexural properties can be obtained from the bending test.91,124,129,136,139 Flexural modulus, also known as the elasticity bending modulus, is the ratio of stress to strain in the elastic limit. Flexural strength is the maximum flexural stress in a bending test.145 The bending test is carried out by fixing the specimen on the support from both ends and pressed by applied force until the specimen break/failure. These properties can be tested by crosshead position indicator or deflection indicator.145

Long et al. tested the flexural strength (4000–6250 MPa), flexural modulus (87.5–97.5 MPa) and impact strength (2.40–3.60 kJ m−2) of untreated bamboo fiber (BF)/PLA composites and 10–30% alkali-treated bamboo fiber (ABF)/PLA composites.139 With the increase of ABF content, the flexural strength of composite material first increased and then decreased. They believed that the increase in strength of the composite was due to the increase in crystallinity of the material, while the decrease in mechanical properties with the increase in ABF content was due to the poor interface compatibility and poor interface adhesion caused by excessive ABF. The impact strength of the composites increased after adding ABF but decreased when the content of ABF exceeded 10%. The reason is that excessive ABF can form noteworthy agglomeration readily, resulting in poor dispersion and poor interface adhesion, which reduces the impact energy absorption of the material. Mimini et al. tested the flexural strength and impact strength of PLA with different kinds (kraft lignin, organosolv lignin, and LS) and contents of the lignins.91 They found that various concentrations of different types of lignin did not improve the flexural strength of PLA. In addition, the impact strength of PLA/lignin composites was reduced due to the introduction of flaws such as voids and air pockets in the composites.

In general, mechanical property testing can be different depending on materials (plastics,133,146,147 vulcanized rubber,148 thermoplastic elastomers,148 polymer matrix composite149,150) and shapes of products (e.g., thin plastic sheeting150,151). For instance, measurement method of tensile test and formula for calculating tensile parameters are determined by these factors. Similarly, impact resistance142,152–155 and flexural properties145,156 have different standards depending on the types of testing materials. Biomass-derived 3D printed products are typically formed with biomass and other materials such as plastic. Due to the variety of biomass and/or co-materials, these properties result in a broad range of values, and universal standards for these mechanical properties are not established yet. The necessity of proper standards for these materials will increase due to the advances in 3D printing techniques and increase of biomass-derived material applications.

4.2 Thermal properties

The addition of biomass into commercial resins affects their thermal properties.87,128 Depending on the end-use of the product, the required properties vary. Therefore, measuring and analyzing thermal properties are important in the field of 3D printing.4 Thermal properties including glass transition temperature, melting temperature, crystallization temperature, cold crystallization temperature, and thermostability can be characterized by thermo-analytical techniques such as thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC).
4.2.1 Thermogravimetric analysis (TGA). TGA is a technique for the thermal stability measurement of materials by monitoring the mass of substance as a function of temperature and time.157 A thermogravimetric analyzer is composed of a thermobalance, a temperature controller, a data collection device, containers that are inert to the specimen and a gas flow control device.158 By analyzing the thermogravimetry curve, composition, thermal stability, thermal decomposition and products of the multicomponent system can be measured.159

Fig. 9 shows an example of the TGA plot of lignin, PLA and PLA/lignin biocomposites.87 The TGA curve presents the weight change of the biocomposites with temperature. The decrease of TGA thermal curve indicates weight loss. Onset temperature was determined at the starting point of the decreasing, implying the beginning of decomposition. From Fig. 9, lignin and lignin-added biocomposites show lower thermal stability than that of PLA because of their lower onset temperatures.


image file: d0ra03620j-f9.tif
Fig. 9 TGA plots of lignin, PLA and PLA/lignin biocomposites. Reprinted with permission from ref. 87. Copyright 2019 Materials.

In addition, the gas products from TGA testing can be further analyzed by multiple techniques, thermogravimetric analyzer equipped with Fourier transform infrared spectroscopy (TGA/FTIR) was used to analyze the degradation pathways of polymers and copolymers.160 By analyzing the degradation products at various temperatures, the overall structure of polymers and the changes of polymer-based products can be understood. The thermogravimetric analyzer with a mass spectrometer (TGA/MS) can be used to detect low levels of components in the printed products.161 Thermogravimetric analyzer connected with gas chromatography mass spectrometer (TGA/GC-MS) could get a more detailed characterization compared with TGA/MS. The gaseous compounds released during thermal decomposition are first separated by GC, then identified and quantified by MS,162 which could separate overlapping events and detect very low levels of components in complex matrices.

4.2.2 Differential scanning calorimetry. Differential scanning calorimetry (DSC) is another type of thermal analysis technology, which shows the change of material in regards to the rate of flow of heat under the programmed temperature.163 Typical DSC curve represents the endothermic and exothermic performance of the sample with changing temperature and heat flow rate (dQ/dt). Two types of DSC, heat flux DSC and power composition DSC are commonly used.164

Gkartzou et al. experimentally studied the PLA/lignin bulk composites, as Fig. 10a presents, they predicted that the increase of lignin content promotes the PLA's double melting behavior, while no significant changes were observed in the glass transition temperature and the corresponding heat capacity before and after the glass transition.140 From Fig. 10b, a step can be observed in the heating curve indicating that the material is undergoing glass transition. For polymers, glass transition refers to the segmental motion of amorphous polymer chains start to be frozen or unfrozen,165 where the temperature for glass transition is called glass transition temperature (Tg). Polymer changes from a hard and fragile state to viscous or rubber state with the increase of temperature around Tg. Determination of a representative Tg, properties of materials and methods such as half-step-height method, inflection method, equal-areas method and conditions are considered.166 In the DSC cooling curve, exothermic peaks representing the changes of the specimen from a liquid state to a fully or partially solid state can be observed. This process is called crystallization. As the temperature rises, an endothermic behavior occurs since the specimen undergoes the process of changing from a completely or partially solid state to a liquid state, which is melting.167 It is worth noticing that sometimes there is cold crystallization, which is the process when the crystallization occurs during the heating process and above the glass transition because the molecular chains reobtain the mobility which is severely limited below the Tg.168,169 This thermal behavior occurs because of the specimen previously cooled very quickly and has no time to crystallize.169 In Fig. 10b, at about 113.6 °C, there is cold crystallization occurring, which appears as an exothermic peak.


image file: d0ra03620j-f10.tif
Fig. 10 (a) DSC thermographs of PLA/lignin bulk composites. (b) DSC thermograph of the sample with 5 wt% lignin. Quantities for characterization of the glass transition of the sample containing 5 wt% lignin: extrapolated onset temperature (Tge), half-step temperature (Tg1/2), change of the normalized heat capacity during the transition (ΔCp), initial (Tgi) and final (Tgf) temperatures of the glass transition. Reprinted with permission from ref. 140. Copyright 2017 Manufacturing Review.

Table 6 summarizes the thermal properties of biomass-derived 3D printed products.

Table 6 Thermal properties of biomass-derived 3D printed products
Materials Glass transition temperature Tg (°C) Melting temperature Tm (°C) Crystallization temperature Tc (°C) Cold crystallization temperature Tcc (°C) Ref.
PLA + softwood lignin (20–40%) 59–71 179.2–186.6 87
PLA + microcrystalline cellulose (MCC) (1–5%) ∼60 159–161 112–119 65
PLA + titanate coupling agent modified MCC (1–5%) ∼59 159–160 107–115
PLA + lignin (5–15%) 56–61 147–152 111–117 91 and 140
Polypropylene (22.5–56%) + PLA (22.5–52.5%) + maleated polypropylene (0–5%) + untreated bamboo fiber (20%) 159 124 136
Polypropylene (22.5–56%) + PLA (22.5–52.5%) + maleated polypropylene (0–5%) + alkali-treated bamboo fiber (20%) 157–167 121–127
Bamboo fiber/Polypropylene/PLA + alkali-treated bamboo fiber (10–30%) 161–162 115–119
PLA + wood flour (5%) 60 167 97 170
Poly (L-lactic acid) + lignin (10–40%) 57–64 161–170 138
Methacrylate (MA) + lignin-coated cellulose nanocrystals (L-CNC) 82–101 74
Polypropylene (PP)/n-octyltriethoxysilane (OTS)/aminopropyltriethoxysilane (APTES)/perfluorooctyltriethoxysilane (PFOS) + microcrystalline cellulose (MCC) 147–151 109–113 171
PLA + cellulose nanofibrils (10–50%) 58–60 64
Methacrylate (MA) + cellulose nanocrystal (0.5–4%) 57.6–74.5 73


The glass transition temperature (56–71 °C), melting temperature (147–186.6 °C), cold crystallization temperature (111–117 °C) of 5–40% lignin blend with PLA87,91,140 and poly(L-lactic acid)138 were reported in previous studies. Tanase-Opedal et al. reported that Tg decreased for the PLA/lignin as the lignin content increased, which can be explained by different molecular factors172 such as crosslinking density, interchain hydrogen bonding, molecular mass, and rigid phenyl groups.87

The thermal characteristics of 3D printed products with different types of cellulose such as microcrystalline cellulose (MCC),65,171 cellulose nanocrystal,73 cellulose nanofibrils64 have been characterized. The glass transition temperature, melting temperature, crystallization temperature, and cold crystallization temperature were 57.6–74.5 °C, 147–161 °C, 109–113 °C, 107–115 °C, respectively. The similar thermal properties such as melting temperature (157–167 °C) and the crystallization temperature (115–127 °C) were observed from the printed products with whole biomass (bamboo fiber136 and wood flour170). Yang et al. prepared methacrylate (MA) nanocomposite blended with cellulose nanocrystal (CNC) (0.5–4%) through SLA.73 After printing, the sample was postcured under different temperatures (room temperature, 120 °C, 140 °C and 160 °C). After postcure process, a favorable thermal property was obtained at 160 °C while comparable thermal properties were also obtained at 120 °C and 140 °C. However, compared to 140 °C, the mechanical property of the printed composites decreased at 160 °C. In addition, as the CNC exceeded 1%, tensile strength decreased at 120 °C, 140 °C, and 160 °C. Therefore, they suggested MA/CNC (0.5–1%) and postcure temperature between 120 °C and 140 °C for 3D printed CNC/MA nanocomposites.

The glass transition temperature of commercial resins blend with biomass-derived mixture74 through 3D printing has also been tested. Feng et al. tested the thermal properties of the samples printed through SLA with methacrylate (MA) and 0.1% to 1% lignin-coated cellulose nanocrystals (L-CNC).74 They found that as the L-CNC content increase from 0.1% to 1%, the Tg didn't change significantly. Compared to the neat MA, the glass transition temperature decreased. The reasons may be the increased fluidity of the polymer chain or the addition of L-CNC induced gaps between L-CNC and the MA matrix which increased the free volume around the MA chain.

4.3 Morphological properties

Morphological analysis has been used in characterizing the biomass-derived 3D printing materials intuitively. By investigating the influence of raw materials and printing technologies on the shape and structure of the products, important information on enhancing the quality of the 3D printed products can be obtained.173
4.3.1 Scanning electron microscope. A scanning electron microscope (SEM) is a type of electron microscope to observe the surface morphology of samples by secondary electron signal imaging. For characterization of the biomass-derived 3D printed products, SEM can be used to observe the binding property of printing layers,74,87,136,137,171 the microcrystalline domains of the raw materials,65 the morphology of the biocomposites,38,41,64,65,74,87,89,91,93,95,129,136,137,139,140,170,171,174,175 interfacial adhesion and composite compatibility,64,74,93,95,139,170 fracture surface of the tensile specimen,64,87,140 and inner-bead porosity,64 etc.

The fracture surface of PLA/lignin composites blended with 20% lignin, or 40% lignin printed at 205 °C, 215 °C, 230 °C with FDM are shown in Fig. 11. Tanase-Opedal et al. found that printing at 215 °C is the most suitable temperature according to their SEM images.87 The combination of the printing layers at 215 °C was improved, which was beneficial to the mechanical properties of the samples.


image file: d0ra03620j-f11.tif
Fig. 11 Scanning electron microscopy (SEM) analysis of the fracture surface of tensile tested dogbones. Reprinted with permission from ref. 87. Copyright 2019 Materials.

Ibrahim et al. used organosolv lignin and graphene nanoplatelets as filler to reinforce photo-curable PU through SLA to get the final product.95 PU showed a wide range of slight cracks and a relatively flat surface (Fig. 12a). With the addition of graphene nanoplatelets, the fillers were observed on the surface of PU–G (Fig. 12b). After adding 0.6% of lignin–graphene, graphene was reported to be evenly distributed in the polymer matrix, as shown in Fig. 12c. The top surface of PU-0.6% lignin/G exposed to the UV light indicated that with the assistance of lignin as the filler, the graphene nanoplatelets in photo-curable PU became homogenous (Fig. 12d). The authors believed that the composition of lignin-graphene promoted the uniform distribution of biocomposites in the polymer matrix and significantly improved the mechanical properties of the 3D printed composites.


image file: d0ra03620j-f12.tif
Fig. 12 Micrograph images of the fracture surface after tensile testing (a) photo-curable PU, (b) PU–graphene (PU–G), (c) PU-0.6% lignin/G, and (d) top surface of PU-0.6% lignin/G. Reprinted with permission from ref. 95. Copyright 2019 Polymers.
4.3.2 Optical microscope. Optical microscope, also called light microscope, uses optical principles to create images that magnify the vision of small objects. An optical microscope was used to determine the surface roughness,140 dispersion of biomass-derived components in the surrounding matrix,140 shrinkages of static filament after solvent evaporation,38 and crystalline morphology (by polarized optical microscope)175 for biocomposites.

Gkartzou et al. used an optical microscope to observe the morphology of single fibers extruded at different printing speeds.140 Compared with pure PLA fiber (Fig. 13a), lignin caused a significant increase in fibers' surface roughness (Fig. 13b–d). As the diameter of the nozzle decreased to the size that was comparable to the lignin aggregate, the diameter inconsistencies became more prominent at higher speeds, the aggregation of lignin resulted in the increase of local diameter up to 30% at 60 mm s−1 printing speed (Fig. 13c and d). Speed of 20 mm s−1 as the general printing speed of all nozzles was selected.


image file: d0ra03620j-f13.tif
Fig. 13 Light micrographs of (a) individual PLA fibers extruded from 0.2 mm nozzle 20 mm s−1 printing speed; (b) individual PLA blends with 5 wt% lignin fibers extruded from 0.2 mm nozzle 20 mm s−1 printing speed; (c and d) individual PLA blends with 5 wt% lignin fibers extruded from 0.2 mm nozzle 60 mm s−1 printing speed. Reprinted with permission from ref. 140. Copyright 2017 Manufacturing Review.

4.4 Other approaches

4.4.1 Contact angle measurement. Contact angle measurement is a common and effective method to describe the relative hydrophobic behavior of the composites by using water as the liquid.89,176,177 By measuring the wettability of the material surface.177,178 Wetting can be judged by measuring the contact angle between the liquid and solid at the interface of gas, liquid and target materials. If the wetting tendency increase, the contact angle or surface tension will decrease.178 It is worth noting that the contact angle is affected by surface porosity, roughness, non-uniformity, and other factors.177 Lower than 90° of the contact angle value between the liquid and the solid indicates a high affinity between the liquid drop and the solid. If the contact angle is higher than 90° with water, the solid is considered as a hydrophobic material.177,179

Vaidya et al. blended PHB with 10% to 50% biorefinery lignin and printed by FDM.89 The water contact angle of the composites increased with the increasing content of biorefinery lignin, indicating that the relative higher hydrophobicity of the composite surface increased with more lignin.

4.4.2 Shape memory effect. Shape memory polymer (SMP) can be fixed to a temporary shape and restore its original shape after external stimulation (such as heating or lighting).137,180 Therefore, sometimes it is necessary to measure the shape memory effect of the biocomposites.

Bi et al. tested the shape memory behavior of wood flour/thermoplastic polyurethane (TPU) mixture blend with different EPDM-g-MAH contents at different temperatures (room temperature or 60° in the oven).137 The characterization process is heating the sample with the original shape and angle (180°) to the temperature above the soft segments melting temperature of TPU for 2 h, then bend the sample to a fixed angle and place it in a temperature which is below glass transition temperature for 12 h to fix the temporary shape, and finally placing the sample at the corresponding temperature to determine the shape recovery state and shape recovery rate. The shape recovery rate (Rr) can be obtained by the following formula:

 
image file: d0ra03620j-t4.tif(4)
where Ar is the recovery angle; At is the fixed angle.

Based on the final changing state images of the samples and Rr% data, they found that the Rr% of the unmodified and 4% EPDM-g-MAH–TPU/WF (4% TPU/WF) composites were both higher than the Rr% at room temperature, and the shape memory performance of the 4% TPU/WF composites was slightly higher than that of the unmodified TPU/WF composites. Therefore, they concluded that temperature was a key factor in thermally induced shape memory effect and 4% TPU/WF composites had better shape memory performance. In addition, by studying the heat-induced shape memory effect of TPU/WF composites at 60 °C temperature, they found that when the soft segment of the TPU was exposed to high temperature, the chains of molecules were stretched and oriented by external forces, allowing them to be morphed into any shape.181,182 After the removal of the external forces and below the temperature of the soft segment, the molecules are in a glassy state, leaving their deformed shape in a temporary state. When heated above the soft segment temperature, the soft segment melts, and the deformed structure will automatically return to its original printed shape.

4.4.3 X-ray diffraction. X-ray diffraction (XRD) is a rapid analysis technique to obtain detailed information on the structure and physical properties of crystalline in materials.183 During the measurement, an X-ray beam is directed to the sample. Lattice satisfying the Bragg's law in the crystals in the sample will diffract the X-rays as eqn (5).184
 
= 2d[thin space (1/6-em)]sin[thin space (1/6-em)]θ (5)
where n is the reflection series; λ is the wavelength of the incident beam; d is the lattice spacing; θ is the angle between the incident beam and the reflecting crystal plane.

In addition, XRD can be used to measure the average spacing between atoms, to determine the grain direction, degree of crystallinity and the crystal structure of unknown materials.183

In biomass-derived 3D printing materials, it is usually used to determine the concentration of the crystals in the biomass composites through XRD spectra.87,129,136,139,170

Tanase-Opedal et al. reported that neat PLA exhibits a broad peak in XRD spectra which indicated a low crystallinity of PLA.87 As the addition of lignin, two new peaks appear in the XRD pattern which indicates PLA was further crystallized because of the lignin as a nucleating agent. Long et al. determined the diffraction peaks in the XRD pattern by observing the characteristic peaks of pure polypropylene (PP) and neat PLA.136 Since there were no differences in the diffraction peaks in the XRD spectra, they concluded that the crystal type did not change during the preparation of bamboo fiber/PP/PLA. In addition, as the PLA increases, the diffraction intensity in the spectrum decreases which indicates that the increase in PLA leads to a decrease in the crystallinity of PP.

4.4.4 Dynamic mechanical analysis. Dynamic mechanical analysis (DMA) is widely used to measure the viscoelastic behavior of polymers. At a specific frequency or temperature, sinusoidal stress or strain is applied to the material, and then the resulting stress or strain inside the material is measured. However, the responding stress owns a phase displacement (θ, and the tan[thin space (1/6-em)]θ is called the damping factor) compared with the given stimulus due to the property of the polymer. The deformation resistance (i.e., the complex modulus) is then deconvoluted into an elastic portion (in phase with the given stimulus, in which the ratio of stress/strain is called storage modulus) and a viscous portion (π/2 out of phase with the given stimulus, in which the ratio of stress/strain is called loss modulus) this technique is used to determine the storage modulus, loss modulus and damping as a function of temperature, time or frequency. The storage modulus (E′) is a measure of the elastic storage energy of a material.185,186 Loss modulus (E′′) is the energy dissipation (usually a loss in the form of heat) when a viscoelastic material deforms.185,186 Damping (tan[thin space (1/6-em)]delta) can characterize the energy dissipation efficiency of a material and can be expressed as E′′/E′.186 The DMA glass transition temperature can be obtained from the intersection of two tangent lines from the storage modulus and the peak temperature of tan[thin space (1/6-em)]delta.186 These properties can be used to determine the viscoelasticity of biomass-derived 3D printed products related to printing behaviors.64,74,93,137,171

Kaynak et al. used DMA to measure the storage modulus of PP/microcrystalline cellulose (MCC, unmodified and surface modified) composites.171 The stiffness of all composites obtained by measuring the Young's modulus increased slightly below the glass transition temperature (Tg).187–189 However, when the temperature was above Tg, the storage modulus of the material decreased slowly with the increase of temperature. The authors discussed that the heating above Tg provides enough thermal energy or activation energy to make the bond in the polymer segment rotate.190

5. Conclusions and perspectives

Recent advancements of 3D printing technologies offer a promising manufacturing strategy by reducing waste generation and energy consumption as well as fabricating wide and complex structures. In particular, utilization of lignocellulosic biomass in 3D printing has been investigated to overcome the current challenges posed by petroleum-based materials such as the shortage of resources and negative impacts on the environment. Diverse researches in printing technologies, printing feedstock, characteristics and applications have been conducted to enhance the properties of biomass-derived 3D printed materials and broaden their applications. Cellulose, lignin, starch, and whole biomass are some of the biomass-derived substances that have been utilized in fused deposition modeling, stereolithography, binder jetting, or direct ink writing 3D printing technologies. These technologies allow the use of low-cost biomass-derived substances to fabricate composites that can serve construction, biomedical, pharmaceutical and food industries.

In this emerging field, the structure–processing–property relationship is still followed, which means the final product was formed hierarchically. Therefore, understanding the structure of biomass component(s) remains important, along with the utilization of the supramolecular structures, like crystallinity, material anisotropy and interfacial interactions need to be well-studied to help reach the target property of the 3D printed biomass-derived materials. For the processing, various parameters such as the printing resolution and part production rate are the areas that have to undergo further engineering for making 3D printing competitive with conventional material fabrication technologies. In the future, the use of compatibilizers and modification of interfacial chemistry may enhance bonding and distribution of biomass-derived fillers with plastics, which could significantly improve the concentration of the fillers with moderate strength to alleviate depletion, at least partially, of the petroleum-based materials. Moreover, characterization techniques that are tailored to the final commercial application can be valuable to better assess the strengths and weaknesses of printed materials. An in-site characterization that has been applied to metal 3D printing is also a possible approach in this field to promote the printing quality.

Conflicts of interest

The authors declare no conflict of interest.

Acknowledgements

This work was supported by State University of New York, College of Environmental Science and Forestry (SUNY ESF). This work was funded, in part, by the Laboratory Directed Research and Development Program of Oak Ridge National Laboratory, managed by UT-Battelle, LLC, for the U.S. Department of Energy under Contract DE-AC05-00OR22725. The United States Government retains and the publisher, by accepting the article for publication, acknowledges that the United States Government retains a nonexclusive, paid-up, irrevocable, worldwide license to publish or reproduce the published form of this manuscript, or allow others to do so, for United States Government purposes. The Department of Energy will provide public access to these results of federally sponsored research in accordance with the DOE Public Access Plan (http://energy.gov/downloads/doe-public-access-plan). The views and opinions of the authors expressed herein do not necessarily state or reflect those of the United States Government or any agency thereof. Neither the United States Government nor any agency thereof, nor any of their employees, makes any warranty, expressed or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately owned rights.

References

  1. O. Diegel, in Comprehensive Materials Processing, ed. S. Hashmi, C. J. V. Tyne, G. F. Batalha and B. Yilbas, Elsevier, 1st edn, 2014, vol. 10, ch. 2, pp. 3–18 Search PubMed.
  2. X. Wang, M. Jiang, Z. Zhou, J. Gou and D. Hui, Composites, Part B, 2017, 110, 442–458 CrossRef CAS.
  3. ISO/ASTM52900-15, Standard Terminology for Additive Manufacturing – General Principles – Terminology, ASTM International, West Conshohocken, PA, 2015, https://www.astm.org/Standards/ISOASTM52900.htm Search PubMed.
  4. S. C. Ligon, O. R. Liska, J. Stampfl, M. Gurr and R. Mülhaupt, Chem. Rev., 2017, 117, 10212–10290 CrossRef CAS PubMed.
  5. H. Kodama, Rev. Sci. Instrum., 1981, 52, 1770–1773 CrossRef.
  6. A. Pîrjan and D.-M. Petroşanu, Journal of Information Systems & Operations Management, 2013, 7, 360–370 Search PubMed.
  7. C. G. Yoo, X. Meng, Y. Pu and A. J. Ragauskas, Bioresour. Technol., 2020, 301, 122784 CrossRef PubMed.
  8. J. Pérez, J. Munoz-Dorado, T. De la Rubia and J. Martinez, Int. Microbiol., 2002, 5, 53–63 CrossRef PubMed.
  9. M. Petre, G. Zarnea, P. Adrian and E. Gheorghiu, Resour., Conserv. Recycl., 1999, 27, 309–332 CrossRef.
  10. S. Camarero, M. J. Martínez and A. T. Martínez, Biofuels, Bioprod. Biorefin., 2014, 8, 615–625 CrossRef CAS.
  11. A. P. Heiner, J. Sugiyama and O. Teleman, Carbohydr. Res., 1995, 273, 207–223 CrossRef CAS.
  12. P. Langan, Y. Nishiyama and H. Chanzy, J. Am. Chem. Soc., 1999, 121, 9940–9946 CrossRef CAS.
  13. P. Langan, Y. Nishiyama and H. Chanzy, Biomacromolecules, 2001, 2, 410–416 CrossRef CAS PubMed.
  14. Y. Nishiyama, J. Sugiyama, H. Chanzy and P. Langan, J. Am. Chem. Soc., 2003, 125, 14300–14306 CrossRef CAS PubMed.
  15. W. Chen, G. C. Lickfield and C. Q. Yang, Polymer, 2004, 45, 1063–1071 CrossRef CAS.
  16. S.-Y. Ding and M. E. Himmel, J. Agric. Food Chem., 2006, 54, 597–606 CrossRef CAS PubMed.
  17. A. v. Wijk and I. v. Wijk, 3D printing with biomaterials – towards a sustainable and circular economy, IOS press, 2015 Search PubMed.
  18. J. Liu, L. Sun, W. Xu, Q. Wang, S. Yu and J. Sun, Carbohydr. Polym., 2019, 207, 297–316 CrossRef CAS PubMed.
  19. B. M. Tymrak, M. Kreiger and J. M. Pearce, Mater. Des., 2014, 58, 242–246 CrossRef CAS.
  20. P. Tran, T. D. Ngo, A. Ghazlan and D. Hui, Composites, Part B, 2017, 108, 210–223 CrossRef CAS.
  21. Q. Sun, G. M. Rizvi, C. T. Bellehumeur and P. Gu, Rapid Prototyp. J., 2008, 14, 72–88 CrossRef.
  22. R. Melnikova, A. Ehrmann and K. Finsterbusch, IOP Conf. Ser.: Mater. Sci. Eng., 2014, 62, 012018 Search PubMed.
  23. B. Caulfield, P. E. McHugh and S. Lohfeld, J. Mater. Process. Technol., 2007, 182, 477–488 CrossRef CAS.
  24. C. R. Garcia, J. Correa, D. Espalin, J. H. Barton, R. C. Rumpf, R. Wicker and V. Gonzalez, Prog. Electromagn. Res. Lett., 2012, 34, 75–82 CrossRef.
  25. N. A. Nguyen, S. H. Barnes, C. C. Bowland, K. M. Meek, K. C. Littrell, J. K. Keum and A. K. Naskar, Sci. Adv., 2018, 4, eaat4967 CrossRef CAS PubMed.
  26. K. Markstedt, A. Escalante, G. Toriz and P. Gatenholm, ACS Appl. Mater. Interfaces, 2017, 9, 40878–40886 CrossRef CAS PubMed.
  27. G. Siqueira, D. Kokkinis, R. Libanori, M. K. Hausmann, A. S. Gladman, A. Neels, P. Tingaut, T. Zimmermann, J. A. Lewis and A. R. Studart, Adv. Funct. Mater., 2017, 27, 1604619 CrossRef.
  28. Y. Shao, D. Chaussy, P. Grosseau and D. Beneventi, Ind. Eng. Chem. Res., 2015, 54, 10575–10582 CrossRef CAS.
  29. M. K. Hausmann, P. A. Ruhs, G. Siqueira, J. Lauger, R. Libanori, T. Zimmermann and A. R. Studart, ACS Nano, 2018, 12, 6926–6937 CrossRef CAS PubMed.
  30. M. K. Hausmann, G. Siqueira, R. Libanori, D. Kokkinis, A. Neels, T. Zimmermann and A. R. Studart, Adv. Funct. Mater., 2020, 30, 1904127 CrossRef CAS.
  31. V. C. F. Li, C. K. Dunn, Z. Zhang, Y. L. Deng and H. J. Qi, Sci. Rep., 2017, 7, 8018 CrossRef PubMed.
  32. V. C. F. Li, A. Mulyadi, C. K. Dunn, Y. L. Deng and H. J. Qi, ACS Sustainable Chem. Eng., 2018, 6, 2011–2022 CrossRef CAS.
  33. C. C. Qian, L. H. Li, M. Gao, H. Y. Yang, Z. R. Cai, B. D. Chen, Z. Y. Xiang, Z. J. Zhang and Y. L. Song, Nano Energy, 2019, 63, 103885 CrossRef CAS.
  34. J. P. Wang, A. Chiappone, I. Roppolo, F. Shao, E. Fantino, M. Lorusso, D. Rentsch, K. Dietliker, C. F. Pirri and H. Grutzmacher, Angew. Chem., Int. Ed., 2018, 57, 2353–2356 CrossRef CAS PubMed.
  35. D. Kam, M. Chasnitsky, C. Nowogrodski, I. Braslaysky, T. Abitbol, S. Magdassi and O. Shoseyov, Colloids Interfaces, 2019, 3, 46 CrossRef CAS.
  36. D. Kam, M. Layani, S. B. Minerbi, D. Orbaum, S. A. Ben Harush, O. Shoseyov and S. Magdassi, Adv. Mater. Technol., 2019, 4, 1900158 CrossRef CAS.
  37. Y. N. Jiang, X. D. Xv, D. F. Liu, Z. Yang, Q. Zhang, H. C. Shi, G. Q. Zhao and J. P. Zhou, Bioresources, 2018, 13, 5909–5924 CAS.
  38. S. W. Pattinson and A. J. Hart, Adv. Mater. Technol., 2017, 2, 1600084 CrossRef.
  39. N. B. Palaganas, J. D. Mangadlao, A. C. de Leon, J. O. Palaganas, K. D. Pangilinan, Y. J. Lee and R. C. Advincula, ACS Appl. Mater. Interfaces, 2017, 9, 34314–34324 CrossRef CAS PubMed.
  40. V. C. F. Li, X. Kuang, A. Mulyadi, C. M. Hamel, Y. L. Deng and H. J. Qi, Cellulose, 2019, 26, 3973–3985 CrossRef CAS.
  41. J. T. Sutton, K. Rajan, D. P. Harper and S. C. Chmely, ACS Appl. Mater. Interfaces, 2018, 10, 36456–36463 CrossRef CAS PubMed.
  42. S. Infanger, A. Haemmerli, S. Iliev, A. Baier, E. Stoyanov and J. Quodbach, Int. J. Pharm., 2019, 555, 198–206 CrossRef CAS PubMed.
  43. J. Credou and T. Berthelot, J. Mater. Chem. B, 2014, 2, 4767–4788 RSC.
  44. R. J. Moon, A. Martini, J. Nairn, J. Simonsen and J. Youngblood, Chem. Soc. Rev., 2011, 40, 3941–3994 RSC.
  45. N. Kayra and A. Ö. Aytekin, in Cellulose-Based Superabsorbent Hydrogels, ed. M. I. H. Mondal, Springer International Publishing, Cham, 2018, pp. 1–28 Search PubMed.
  46. G. Li, A. G. Nandgaonkar, Y. Habibi, W. E. Krause, Q. Wei and L. A. Lucia, RSC Adv., 2017, 7, 13678–13688 RSC.
  47. V. Kuzmenko, E. Karabulut, E. Pernevik, P. Enoksson and P. Gatenholm, Carbohydr. Polym., 2018, 189, 22–30 CrossRef CAS PubMed.
  48. N. D. Sanandiya, Y. Vijay, M. Dimopoulou, S. Dritsas and J. G. Fernandez, Sci. Rep., 2018, 8, 8642 CrossRef PubMed.
  49. Y. Y. Li, H. L. Zhu, Y. B. Wang, U. Ray, S. Z. Zhu, J. Q. Dai, C. J. Chen, K. Fu, S. H. Jang, D. Henderson, T. Li and L. B. Hu, Small Methods, 2017, 1, 1700222 CrossRef.
  50. W.-T. Cao, C. Ma, D.-S. Mao, J. Zhang, M.-G. Ma and F. Chen, Adv. Funct. Mater., 2019, 29, 1905898 CrossRef CAS.
  51. E. Gutierrez, P. A. Burdiles, F. Quero, P. Palma, F. Olate-Moya and H. Palza, ACS Biomater. Sci. Eng., 2019, 5, 6290–6299 CrossRef CAS.
  52. Y. N. Jiang, J. P. Zhou, H. C. Shi, G. Q. Zhao, Q. Zhang, C. Feng and X. D. Xv, J. Mater. Sci., 2020, 55, 2618–2635 CrossRef CAS.
  53. Y. N. Jiang, J. P. Zhou, Z. Yang, D. F. Liu, X. D. Xv, G. Q. Zhao, H. C. Shi and Q. Zhang, J. Mater. Sci., 2018, 53, 11883–11900 CrossRef CAS.
  54. S. Sultan and A. P. Mathew, Nanoscale, 2018, 10, 4421–4431 RSC.
  55. R. D. Chen, C. F. Huang and S. H. Hsu, Carbohydr. Polym., 2019, 212, 75–88 CrossRef CAS PubMed.
  56. L. Liang, S. Bhagia, M. Li, C. Huang and A. J. Ragauskas, ChemSusChem, 2020, 13, 78–87 CrossRef CAS PubMed.
  57. L. Huang, X. Y. Du, S. N. Fan, G. S. Yang, H. L. Shao, D. J. Li, C. B. Cao, Y. F. Zhu, M. F. Zhu and Y. P. Zhang, Carbohydr. Polym., 2019, 221, 146–156 CrossRef CAS PubMed.
  58. W. Y. Xu, X. Zhang, P. R. Yang, O. Langvik, X. J. Wang, Y. C. Zhang, F. Cheng, M. Osterberg, S. Willfor and C. L. Xu, ACS Appl. Mater. Interfaces, 2019, 11, 12389–12400 CrossRef CAS PubMed.
  59. T. Kim, C. Boo, M. Hausmann, G. Siqueira, T. Zimmermann and W. S. Kim, Adv. Electron. Mater., 2019, 5, 1800778 CrossRef.
  60. K. M. O. Hakansson, I. C. Henriksson, C. D. Vazquez, V. Kuzmenko, K. Markstedt, P. Enoksson and P. Gatenholm, Adv. Mater. Technol., 2016, 1, 1600096 CrossRef.
  61. K. Markstedt, K. Hakansson, G. Toriz and P. Gatenholm, Appl. Mater. Today, 2019, 15, 280–285 CrossRef.
  62. Y. Jiang, L. M. Korpas and J. R. Raney, Nat. Commun., 2019, 10, 128 CrossRef PubMed.
  63. N. E. Zander, J. H. Park, Z. R. Boelter and M. A. Gillan, ACS Omega, 2019, 4, 13879–13888 CrossRef CAS PubMed.
  64. H. L. Tekinalp, X. Meng, Y. Lu, V. Kunc, L. J. Love, W. H. Peter and S. Ozcan, Composites, Part B, 2019, 173, 106817 CrossRef.
  65. C. A. Murphy and M. N. Collins, Polym. Compos., 2018, 39, 1311–1320 CrossRef CAS.
  66. S. Shariatnia, A. Veldanda, S. Obeidat, D. Jarrahbashi and A. Asadi, Composites, Part B, 2019, 177, 107291 CrossRef CAS.
  67. X. H. Feng, Z. Z. Yang, S. S. H. Rostom, M. Dadmun, S. Q. Wang, Q. W. Wang and Y. J. Xie, Mater. Des., 2018, 138, 62–70 CrossRef CAS.
  68. B. Huang, H. He, S. N. Meng and Y. C. Jia, Polym. Int., 2019, 68, 1351–1360 CrossRef CAS.
  69. J. Dong, C. T. Mei, J. Q. Han, S. Lee and Q. L. Wu, Addit. Manuf., 2019, 28, 621–628 CAS.
  70. J. Dong, M. C. Li, L. Zhou, S. Lee, C. T. Mei, X. W. Xu and Q. L. Wu, J. Polym. Sci., Part B: Polym. Phys., 2017, 55, 847–855 CrossRef CAS.
  71. L. Y. Li, Y. Chen, T. X. Yu, N. Wang, C. S. Wang and H. P. Wang, Compos. Commun., 2019, 16, 162–167 CrossRef.
  72. W. J. Long, J. L. Tao, C. Lin, Y. C. Gu, L. Mei, H. B. Duan and F. Xing, J. Cleaner Prod., 2019, 239, 118054 CrossRef CAS.
  73. Z. Yang, G. Wu, S. Wang, M. Xu and X. Feng, J. Polym. Sci., Part B: Polym. Phys., 2018, 56, 935–946 CrossRef CAS.
  74. X. Feng, Z. Yang, S. Chmely, Q. Wang, S. Wang and Y. Xie, Carbohydr. Polym., 2017, 169, 272–281 CrossRef CAS PubMed.
  75. X. H. Feng, Z. H. Wu, Y. J. Xie and S. Q. Wang, Bioresources, 2019, 14, 3701–3716 CAS.
  76. P. Biswas, S. Mamatha, S. Naskar, Y. S. Rao, R. Johnson and G. Padmanabham, J. Alloys Compd., 2019, 770, 419–423 CrossRef CAS.
  77. J. J. Koh, G. J. H. Lim, X. Zhou, X. W. Zhang, J. Ding and C. B. He, ACS Appl. Mater. Interfaces, 2019, 11, 13787–13795 CrossRef CAS PubMed.
  78. T. Nguyen and R. Latkany, Clin. Ophthalmol., 2011, 5, 587 Search PubMed.
  79. B. Arafat, M. Wojsz, A. Isreb, R. T. Forbes, M. Isreb, W. Ahmed, T. Arafat and M. A. Alhnan, Eur. J. Pharm. Sci., 2018, 118, 191–199 CrossRef CAS PubMed.
  80. Y. Yang, H. H. Wang, H. C. Li, Z. M. Ou and G. S. Yang, Eur. J. Pharm. Sci., 2018, 115, 11–18 CrossRef CAS PubMed.
  81. P. A. G. S. Giachini, S. S. Gupta, W. Wang, D. Wood, M. Yunusa, E. Baharlou, M. Sitti and A. Menges, Sci. Adv., 2020, 6, eaay0929 CrossRef CAS PubMed.
  82. J. S. Park, T. Kim and W. S. Kim, Sci. Rep., 2017, 7, 3246 CrossRef PubMed.
  83. S. Shin, H. Kwak and J. Hyun, Carbohydr. Polym., 2019, 225, 115235 CrossRef CAS PubMed.
  84. C. Thibaut, A. Denneulin, S. R. du Roscoat, D. Beneventi, L. Orgeas and D. Chaussy, Carbohydr. Polym., 2019, 212, 119–128 CrossRef CAS PubMed.
  85. A. J. Ragauskas, G. T. Beckham, M. J. Biddy, R. Chandra, F. Chen, M. F. Davis, B. H. Davison, R. A. Dixon, P. Gilna and M. Keller, Science, 2014, 344, 1246843 CrossRef PubMed.
  86. S. Y. Zhang, M. Li, N. J. Hao and A. J. Ragauskas, ACS Omega, 2019, 4, 20197–20204 CrossRef CAS PubMed.
  87. M. Tanase-Opedal, E. Espinosa, A. Rodriguez and G. Chinga-Carrasco, Materials, 2019, 12, 3006 CrossRef CAS PubMed.
  88. J. Dominguez-Robles, N. K. Martin, M. L. Fong, S. A. Stewart, N. J. Irwin, M. I. Rial-Hermida, R. F. Donnelly and E. Larraneta, Pharmaceutics, 2019, 11, 165 CrossRef CAS PubMed.
  89. A. A. Vaidya, C. Collet, M. Gaugler and G. Lloyd-Jones, Mater. Today Commun., 2019, 19, 286–296 CrossRef CAS.
  90. L. X. Liu, M. H. Lin, Z. Xu and M. Q. Lin, Bioresources, 2019, 14, 8484–8498 CAS.
  91. V. Mimini, E. Sykacek, S. N. A. Syed Hashim, J. Holzweber, H. Hettegger, K. Fackler, A. Potthast, N. Mundigler and T. Rosenau, J. Wood Chem. Technol., 2019, 39, 14–30 CrossRef CAS.
  92. N. A. Nguyen, C. C. Bowland and A. K. Naskar, Appl. Mater. Today, 2018, 12, 138–152 CrossRef.
  93. N. A. Nguyen, S. H. Barnes, C. C. Bowland, K. M. Meek, K. C. Littrell, J. K. Keum and A. K. Naskar, Sci. Adv., 2018, 4, eaat4967 CrossRef CAS PubMed.
  94. S. I. Falkehag, J. Marton and E. Adler, in Lignin Structure and Reactions, American Chemical Society, 1966, vol. 59, ch. 7, pp. 75–89 Search PubMed.
  95. F. Ibrahim, D. Mohan, M. S. Sajab, S. B. Bakarudin and H. Kaco, Polymers, 2019, 11, 1544 CrossRef CAS PubMed.
  96. K. Henke and S. Treml, Eur. J. Wood Wood Prod., 2013, 71, 139–141 CrossRef CAS.
  97. M. Rosenthal, C. Henneberger, A. Gutkes and C. T. Bues, Eur. J. Wood Wood Prod., 2018, 76, 797–799 CrossRef CAS.
  98. M. Kariz, M. Sernek and M. K. Kuzman, Eur. J. Wood Wood Prod., 2016, 74, 123–126 CrossRef CAS.
  99. M. Kariz, M. Sernek, M. Obucina and M. K. Kuzman, Mater. Today Commun., 2018, 14, 135–140 CrossRef CAS.
  100. N. Ayrilmis, M. Kariz and M. K. Kuzman, Int. J. Polym. Anal. Charact., 2019, 24, 659–666 CrossRef CAS.
  101. J. V. Ecker, A. Haider, I. Burzic, A. Huber, G. Eder and S. Hild, Rapid Prototyp. J., 2019, 25, 672–678 CrossRef.
  102. M. Kariz, M. Sernek and M. K. Kuzman, Wood Res., 2018, 63, 917–922 CAS.
  103. A. Le Duigou, M. Castro, R. Bevan and N. Martin, Mater. Des., 2016, 96, 106–114 CrossRef.
  104. D. Correa, A. Papadopoulou, C. Guberan, N. Jhaveri, S. Reichert, A. Menges and S. Tibbits, 3D Print. Addit. Manuf., 2015, 2, 106–116 CrossRef.
  105. M. Mirzaee and S. Noghanian, Electron. Lett., 2016, 52, 1656–1658 CrossRef.
  106. P. Li, L. Pan, D. X. Liu, Y. B. Tao and S. Q. Shi, Materials, 2019, 12, 2896 CrossRef CAS PubMed.
  107. S. Guessasma, S. Belhabib and H. Nouri, Polymers, 2019, 11, 1778 CrossRef CAS PubMed.
  108. N. Ayrilmis, Polym. Test., 2018, 71, 163–166 CrossRef CAS.
  109. N. Ayrilmis, M. Kariz, J. H. Kwon and M. K. Kuzman, Int. J. Adv. Manuf. Tech., 2019, 102, 2195–2200 CrossRef.
  110. K. Vigneshwaran and N. Venkateshwaran, Int. J. Polym. Anal. Charact., 2019, 24, 584–596 CrossRef CAS.
  111. Y. Dong, J. Milentis and A. Pramanik, Adv. Manuf., 2018, 6, 71–82 CrossRef CAS.
  112. G. Q. Xie, Y. H. Zhang and W. S. Lin, Bioresources, 2017, 12, 6736–6748 CAS.
  113. W. S. Lin, G. Q. Xie and Z. W. Qiu, Bioresources, 2019, 14, 8689–8700 CAS.
  114. Z. B. Liu, Q. Lei and S. Q. Xing, J. Mater. Res. Technol., 2019, 8, 3741–3751 CrossRef CAS.
  115. G. H. Perry, N. J. Dominy, K. G. Claw, A. S. Lee, H. Fiegler, R. Redon, J. Werner, F. A. Villanea, J. L. Mountain and R. Misra, Nat. Genet., 2007, 39, 1256–1260 CrossRef CAS PubMed.
  116. L. Liu, Y. Meng, X. Dai, K. Chen and Y. Zhu, Food Bioprocess Technol., 2019, 12, 267–279 CrossRef CAS.
  117. Y. J. An, C. F. Guo, M. Zhang and Z. P. Zhong, J. Sci. Food Agric., 2019, 99, 639–646 CrossRef CAS PubMed.
  118. H. Chen, F. Xie, L. Chen and B. Zheng, J. Food Eng., 2019, 244, 150–158 CrossRef CAS.
  119. Z. Liu, M. Zhang, B. Bhandari and C. Yang, J. Food Eng., 2018, 220, 76–82 CrossRef.
  120. L. Zheng, Y. Yu, Z. Tong, Q. Zou, S. Han and H. Jiang, J. Food Process. Preserv., 2019, 43, e13993 Search PubMed.
  121. M. Lille, A. Nurmela, E. Nordlund, S. Metsa-Kortelainen and N. Sozer, J. Food Eng., 2018, 220, 20–27 CrossRef CAS.
  122. P. Tian, F. Yang, Y. Xu, M.-M. Lin, L.-P. Yu, W. Lin, Q.-F. Lin, Z.-F. Lv, S.-Y. Huang and Y.-Z. Chen, Drug Dev. Ind. Pharm., 2018, 44, 1918–1923 CrossRef CAS PubMed.
  123. L. Yang, S. Tang, G. Li, L. Qian, J. Mei, W. Jiang and Z. Fan, Ceram. Int., 2019, 45, 21843–21850 CrossRef CAS.
  124. C. Kuo, L. Liu, W. Teng, H. Chang, F. Chien, S. Liao, W. Kuo and C. Chen, Composites, Part B, 2016, 86, 36–39 CrossRef CAS.
  125. R. A. Paggi, G. V. Salmoria, G. B. Ghizoni, H. D. Back and I. D. Gindri, Int. J. Adv. Manuf. Tech., 2019, 100, 2767–2774 CrossRef.
  126. E. Agnoli, R. Ciapponi, M. Levi and S. Turri, Materials, 2019, 12, 1004 CrossRef CAS PubMed.
  127. F. Chuanxing, W. Qi, L. Hui, Z. Quancheng and M. Wang, Int. J. Food Eng., 2018, 14, 20170297 Search PubMed.
  128. N. A. Nguyen, C. C. Bowland and A. K. Naskar, Data Brief, 2018, 19, 936–950 CrossRef PubMed.
  129. W. Xu, A. Pranovicha, P. Uppstu, X. Wang, D. Kronlund, J. Hemming, H. Öblom, N. Moritz, M. Preis, N. Sandler, S. Willför and C. Xua, Carbohydr. Polym., 2018, 187, 51–58 CrossRef CAS PubMed.
  130. T. D. Ngo, A. Kashani, G. Imbalzano, K. T. Q. Nguyen and D. Hui, Composites, Part B, 2018, 143, 172–196 CrossRef CAS.
  131. M. B. Bazbouz and G. K. Stylios, J. Polym. Sci., Part B: Polym. Phys., 2010, 48, 1719–1731 CrossRef CAS.
  132. K. O. Reddy, B. R. Guduri and A. V. Rajulu, J. Appl. Polym. Sci., 2009, 114, 603–611 CrossRef CAS.
  133. ASTM D638-14, Standard Test Method for Tensile Properties of Plastics, ASTM International, West Conshohocken, PA, 2014, https://www.astm.org/Standards/D638.htm Search PubMed.
  134. E. J. Hearn, in Mechanics of Materials 1, 1997, ch. 1, pp. 1–26 Search PubMed.
  135. H. Ma and J. C. Suhling, J. Mater. Sci., 2009, 44, 1141–1158 CrossRef CAS.
  136. H. Long, Z. Wu, Q. Dong, Y. Shen, W. Zhou, Y. Luo, C. Zhang and X. Dong, Polym. Eng. Sci., 2019, 59, E247–E260 CrossRef CAS.
  137. H. Bi, M. Xu, G. Ye, R. Guo, L. Cai and Z. Ren, Polym, 2018, 10, 1234 Search PubMed.
  138. J. Li, Y. He and Y. Inoue, Polym. Int., 2003, 52, 949–955 CrossRef CAS.
  139. H. Long, Z. Wu, Q. Dong, Y. Shen, W. Zhou, Y. Luo, C. Zhang and X. Dong, J. Appl. Polym. Sci., 2019, 136, 47709 CrossRef.
  140. E. Gkartzou, E. P. Koumoulos and C. A. Charitidis, Manuf. Rev., 2017, 4, 1 Search PubMed.
  141. O. Balkan and H. Demirer, Polym. Compos., 2010, 31, 1285–1308 CAS.
  142. ASTM D256-10(2018), Standard Test Methods for Determining the Izod Pendulum Impact Resistance of Plastics, ASTM International, West Conshohocken, PA, 2018, https://www.astm.org/Standards/D256.htm Search PubMed.
  143. M. T. Demirci, N. Tarakçıoğlu, A. Avcı and Ö. F. Erkendirci, Composites, Part B, 2014, 66, 7–14 CrossRef CAS.
  144. E. Yasa, J. Deckers, J.-P. Kruth, M. Rombouts and J. Luyten, Virtual Phys. Prototyp., 2010, 5, 89–98 CrossRef.
  145. ASTM D790-17, Standard Test Methods for Flexural Properties of Unreinforced and Reinforced Plastics and Electrical Insulating Materials, ASTM International, West Conshohocken, PA, 2017, https://www.astm.org/Standards/D790.htm Search PubMed.
  146. ISO: 527-1, 2019.
  147. ISO: 527-2, 2012.
  148. ASTM D412-16, Standard Test Methods for Vulcanized Rubber and Thermoplastic Elastomers—Tension, ASTM International, West Conshohocken, PA, 2016, https://www.astm.org/Standards/D412 Search PubMed.
  149. ASTM D3039/D3039M-17, Standard Test Method for Tensile Properties of Polymer Matrix Composite Materials, ASTM International, West Conshohocken, PA, 2017, https://www.astm.org/Standards/D3039.htm Search PubMed.
  150. ISO: 37, 2017.
  151. ASTM D882-18, Standard Test Method for Tensile Properties of Thin Plastic Sheeting, ASTM International, West Conshohocken, PA, 2018, https://www.astm.org/Standards/D882.htm Search PubMed.
  152. ISO: 179-1, 2010.
  153. ISO: 179-2, 2011.
  154. ISO: 180, 2019.
  155. ASTM D6110-18, Standard Test Method for Determining the Charpy Impact Resistance of Notched Specimens of Plastics, ASTM International, West Conshohocken, PA, 2018, https://www.astm.org/Standards/D6110.htm Search PubMed.
  156. ISO: 178, 2019.
  157. ASTM E473-18, Standard Terminology Relating to Thermal Analysis and Rheology, ASTM International, West Conshohocken, PA, 2018, https://www.astm.org/Standards/E473.htm Search PubMed.
  158. ASTM E1131-08(2014), Standard Test Method for Compositional Analysis by Thermogravimetry, ASTM International, West Conshohocken, PA, 2014, https://www.astm.org/DATABASE.CART/HISTORICAL/E1131-08R14.htm Search PubMed.
  159. J. S. Revanth, V. S. Madhav, Y. K. Sai, D. V. Krishna, K. Srividya and C. H. M. Sumanth, Mater. Today: Proc., 2020 DOI:10.1016/j.matpr.2019.12.082.
  160. C. A. Wilkie, Polym. Degrad. Stab., 1999, 66, 301–306 CrossRef CAS.
  161. C.-L. Chiang and S.-W. Hsu, J. Polym. Res., 2010, 17, 315–323 CrossRef CAS.
  162. O. Boyron, T. Marre, A. Delauzun, R. Cozic and C. Boisson, Macromol. Chem. Phys., 2019, 220, 1900162 CrossRef.
  163. ISO: 11357-1, 2016.
  164. G. W. H. Höhne, W. F. Hemminger and H.-J. Flammersheim, in Differential Scanning Calorimetry, Springer, 2003, ch. 2, pp. 9–30 Search PubMed.
  165. J. A. Forrest and K. Dalnoki-Veress, Adv. Colloid Interface Sci., 2001, 94, 167–195 CrossRef CAS.
  166. ISO: 11357-2, 2013.
  167. ISO: 11357-3, 2018.
  168. R. M. R. Wellen and M. S. Rabello, J. Mater. Sci., 2005, 40, 6099–6104 CrossRef CAS.
  169. A. Adamus-Wlodarczyk, R. A. Wach, P. Ulanski, J. M. Rosiak, M. Socka, Z. Tsinas and M. Al-Sheikhly, Polymers, 2018, 10, 672 CrossRef PubMed.
  170. Y. Tao, H. Wang, Z. Li, P. Li and S. Q. Shi, Mater, 2017, 10, 339 CrossRef PubMed.
  171. B. Kaynak, M. Spoerk, A. Shirole, W. Ziegler and J. Sapkota, Macromol. Mater. Eng., 2018, 303, 1800037 CrossRef.
  172. C. Heitner, D. Dimmel and J. Schmidt, Lignin and Lignans Advances in Chemistry, CRC Press, 2010 Search PubMed.
  173. A. Tojeira, S. S. Biscaia, T. Q. Viana, I. S. Sousa and G. R. Mitchell, in Controlling the Morphology of Polymers, Springer, 2016, ch. 7, pp. 181–207 Search PubMed.
  174. H. Nitz, H. Semke, R. Landers and R. Mülhaupt, J. Appl. Polym. Sci., 2001, 81, 1972–1984 CrossRef CAS.
  175. L. Wang, W. Gramlich, D. Gardner, Y. Han and M. Tajvidi, J. Compos. Sci., 2018, 2, 7 CrossRef.
  176. X. Liu, E. Zong, J. Jiang, S. Fu, J. Wang, B. Xu, W. Li, X. Lin, Y. Xu, C. Wang and F. Chu, Int. J. Biol. Macromol., 2015, 81, 521–529 CrossRef CAS PubMed.
  177. E. Drioli, A. Criscuoli and E. Curcio, in Membrane Contactors: Fundamentals, Applications and Potentialities, Elsevier, 2011, ch. 2, pp. 40–66 Search PubMed.
  178. C. Dwivedi, I. Pandey, H. Pandey, P. W. Ramteke, A. C. Pandey, S. B. Mishra and S. Patil, in Nano- and Microscale Drug Delivery Systems Design and Fabrication, Elsevier, 2017, ch. 9, pp. 147–164 Search PubMed.
  179. V. S. Kulkarni and C. Shaw, in Essential chemistry for formulators of semisolid and liquid dosages., Academic Press, 2015, ch. 2, pp. 5–19 Search PubMed.
  180. W. Wagermaier, K. Kratz, M. Heuchel and A. Lendlein, in Shape-Memory Polymers, ed. A. Lendlein, Springer, Berlin, Heidelberg, 2009, ch. 3, pp. 97–145 Search PubMed.
  181. M. D. Monzón, R. Paz, E. Pei, F. Ortega, L. A. Suárez, Z. Ortega, M. E. Alemán, T. Plucinski and N. Clow, Int. J. Adv. Manuf. Tech., 2016, 89, 1827–1836 CrossRef.
  182. J. Mendez, P. K. Annamalai, S. J. Eichhorn, R. Rusli, S. J. Rowan, E. J. Foster and C. Weder, Macromol, 2011, 44, 6827–6835 CrossRef CAS.
  183. R. Kohli, in Developments in Surface Contamination and Cleaning, William Andrew, 2012, ch. 3, pp. 107–178 Search PubMed.
  184. A. K. Chatterjee, in Handbook of analytical techniques in concrete science and technology, William Andrew, 2001, ch. 8, pp. 275–332 Search PubMed.
  185. J. Runt, A. Pangon, A. Castagna, Y. He and M. Grujicic, in Elastomeric Polymers with High Rate Sensitivity, ed. R. G. Barsoum, William Andrew, 2015, ch. 9, pp. 319–345 Search PubMed.
  186. ASTM D7028-07(2015), Standard Test Method for Glass Transition Temperature (DMA Tg) of Polymer Matrix Composites by Dynamic Mechanical Analysis (DMA), ASTM International, West Conshohocken, PA, 2015, https://www.astm.org/Standards/D7028.htm Search PubMed.
  187. M. He, J. Zhou, H. Zhang, Z. Luo and J. Yao, J. Appl. Polym. Sci., 2015, 132, 42488 Search PubMed.
  188. N. Izzati Zulkifli, N. Samat, H. Anuar and N. Zainuddin, Mater. Des., 2015, 69, 114–123 CrossRef CAS.
  189. S. Spoljaric, A. Genovese and R. A. Shanks, Composites, Part A, 2009, 40, 791–799 CrossRef.
  190. A. Trojanowski, C. Ruiz and J. Harding, J. Phys. IV, 1997, 7, C3-447–C3-452 CrossRef.

Footnote

These authors contributed equally.

This journal is © The Royal Society of Chemistry 2020