You are on page 1of 20

Carbohydrate Polymers 207 (2019) 297–316

Contents lists available at ScienceDirect

Carbohydrate Polymers
journal homepage: www.elsevier.com/locate/carbpol

Current advances and future perspectives of 3D printing natural-derived T


biopolymers

Jun Liua, Lushan Sunb, Wenyang Xuc, Qianqian Wanga, Sujie Yua, Jianzhong Suna,
a
Biofuels Institute, School of the Environment and Safety Engineering, Jiangsu University, 212013, Zhenjiang, China
b
Institute of Textiles and Clothing, The Hong Kong Polytechnic University, Hong Kong, China
c
Johan Gadolin Process Chemistry Centre, c/o Laboratory of Wood and Paper Chemistry, Åbo Akademi University, Turku, FI-20500, Finland

A R T I C LE I N FO A B S T R A C T

Keywords: 3D printing enables the complex or customized structures production in high speed and resolution. However, the
3D printing lack of bio-based materials with user-defined biochemical and mechanical property is a significant barrier that
Additive manufacturing limits the widespread adoption of 3D printing for products fabrication. Development of eco-friendly natural-
Biopolymers derived biopolymers for 3D printing technologies and their promising application in different areas are of huge
Biomass
academic, and environmental interests. This paper reviews the state-of-the-art in terms of 3D printing technology
Biomedical application
using natural-derived feedstocks, including lignocellulose, starch, algae, and chitosan-based biopolymers.
Special consideration is given to the development of lignocellulosic materials, i.e. cellulose, hemicellulose,
lignin, and their derivatives as 3D printing feedstocks. A strategical development roadmap with identified
material property requirements, key challenges, as well as possible solutions was proposed. It serves as guideline
aiming to explore natural-derived biopolymers as novel feedstocks for different 3D printing technologies that
will be potentially applied in various areas.

1. Introduction Wang, Kim, and Kim (2016), and Wong and Hernandez (2012).
3D printing is a growing technology that makes a revolutionary
3D printing, also known as additive manufacture (AM), is a process impact on products fabrication for applications in areas like healthcare
for constructing 3D physical objects from digital models through the and medicine, aeronautics and space, automotive, food industry, art,
successive layer-by-layer deposition of materials such as plastic, metal, textile and fashion, architecture and construction, which has drawn an
ceramics, or even living cells (Bhatia & Ramadurai, 2017). The 3D increasing attention globally (Melchels et al., 2012; Ozbolat, Peng, &
printing concept was first proposed by Charles W. Hull, using a stereo- Ozbolat, 2016; Shishkovsky, 2016; Vanderploeg, Lee, & Mamp, 2016;
lithography to make polymer objects in the 1980s. Nowadays, there Ventola, 2014). However, there are still a lot of challenges need to
many different 3D printing or AM technologies which have been di- overcome before the 3D printing technology can be adopted as a
vided into several process categories according to the standard of ISO/ common fabrication technology and can achieve its full potential.
ASTM 52921:2013 as in shown in Fig. 1, which mainly include the Limited variety of available, environmentally friendly, and printer-
material extrusion, material jetting, binder jetting, powder bed fusion, friendly materials is a key barrier to the wide-scale adoption of 3D
direct energy deposition, vat photo-polymerization, and sheet lamina- printing technologies.
tion. Depend on the printing method and the inks formulation, objects In line with the current focus on the sustainable economy, the ex-
are fabricated in a successive layer-by-layer fashion by dispensing the ploring of natural-derived and renewable biopolymers, instead of those
material with an extruder, by using chemical agents (e.g. binder) or a fossil oil-based plastics, for various products fabrication has received
laser (sintering/melting) (Kalaskar, 2017). Fig. 1 shows the 3D printing tremendous attention. Biomass from marine, woody and agricultural
principle, classification of 3D printing systems and the specific printing residuals, the most abundant renewable feedstocks on earth, has
technologies, as well as the available materials for each printing ap- showed a promising potential as alternatives to fossil resources (Bhatia
proach. For a detailed review of different 3D printing technologies, it is & Ramadurai, 2017; Fernandes, Pires, Mano, & Reis, 2013; Liu,
available for readers to read further relevant reviews by Gao et al. Korpinen, Mikkonen, Willför, & Xu, 2014; Zhang, Liu, Cui, & Chen,
(2015), Huang, Leu, Mazumder, and Donmez (2014), Mandrycky, 2015). Development of biomass-based materials instead of fossil oil-


Corresponding author.
E-mail address: jzsun1002@ujs.edu.cn (J. Sun).

https://doi.org/10.1016/j.carbpol.2018.11.077
Received 15 June 2018; Received in revised form 21 November 2018; Accepted 23 November 2018
Available online 24 November 2018
0144-8617/ © 2018 Elsevier Ltd. All rights reserved.
J. Liu et al. Carbohydrate Polymers 207 (2019) 297–316

Fig. 1. Categories of 3D printing technologies and commonly used materials. Modified from Wijk and Wijk (2015), Copyright (2015), IOS Press.

based plastics for different 3D printing technologies provides an op- This paper reviewed the state-of-art in the development of natural-
portunity to realize a truly sustainable and recycling economy (Wijk & derived biopolymers, with a main focus on the polysaccharides, such as
Wijk, 2015). By integration with the 3rd generation of biorefinery cellulose, hemicellulose, and their derivatives or composites, for dif-
concept, each component of the biomass can be extracted with their ferent 3D printing technologies and their potential applications in dif-
sophisticate native structure and tailored into versatile 3D printing ferent fields. The aims are to identify some opportunities for replace-
materials in different forms for addictive manufacturing of variety of ment of fossil oil-based plastics with natural-derived biopolymers for
high-value-added products or a further application in various areas, 3D printing, and further to promote a transformation of product de-
such as healthcare and medicine, aeronautics and space, automotive, velopment and their manufacturing in a sustainable way. We first
food industry, art, textile and fashion, architecture and construction. outline the material properties that are key for different targeting ap-
Using natural-derived biopolymers as 3D printing feedstocks, e.g. cel- plication areas, mainly including medicine and pharmacy, customized
lulose, hemicellulose, lignin, starch, alginate, chitosan, and derivatives food fabrication, textile and apparel products, followed by re-
of them, not only meet the demand for sustainability, but also truly presentative research and development of natural-derived biopolymers
reduce a possibility in the side/negative effects associated with some for 3D printing. The current technological challenges and future per-
synthetic polymers in biomedical applications, such as degradability, spectives of 3D printing with natural-derived biopolymers for applica-
recyclability, harmful breakdown products, released additives, and tion in different areas are also proposed.
decreased cell attachment (Chia & Wu, 2015; Guvendiren, Molde,
Soares, & Kohn, 2016; Li et al., 2014; Park, Lih, Park, Joung, & Han,
2017).

298
J. Liu et al. Carbohydrate Polymers 207 (2019) 297–316

2. Key properties required for 3D printing with natural-derived medicine and pharmacy include, but not limited to, printability, bio-
biopolymers in different areas compatibility, degradability (safe degradation byproducts and good
degradation kinetics), exhibiting tissue biomimicry, and appropriate
2.1. Medical and pharmaceutical products mechanical properties (Derby, 2012; Guvendiren et al., 2016; Murphy
& Atala, 2014).
3D printing offers significant advantages for medical and pharma- Materials should have a sound processability, and are printer-
ceutical applications due to the ability to fabricate patient specific and friendly, e.g. shear thinning behavior and high zero shear viscosity,
custom-made medical products, equipment, and drugs (e.g. dosage allowing accurate and easy 3D plotting of constructs with fine resolu-
forms, release profiles, and dispensing). Current and future research tion, high shape fidelity, and structural stability (Kyle, Jessop, Al-
interests in the 3D printing of biomedical and pharmaceutical products Sabah, & Whitaker, 2017; Sultan, Siqueira, Zimmermann, & Mathew,
mainly include: 1) tissue and organ fabrication; customization of 2017).
prostheses, implants, and anatomical models; 2) personalized drug Biocompatibility regarding the cellular and tissue compatibility of
dosage forms, drug delivery, drug screening and discovery in pharma- the ink components, printed structure, any leachables or degradation
ceutical research (Derby, 2012; Melchels et al., 2012; Murphy & Atala, by-products from the printed structure or the material should be eval-
2014; Ventola, 2014; Zhang, Fisher, & Leong, 2015). 3D printing uated (Liu, Willför, & Mihranyan, 2017). Normal biocompatibility tests
technologies for medical and pharmaceutical applications would include in vitro tests of DNA damage, cytotoxicity, cell proliferation,
mainly involve with extrusion-based printing, inkjet-based printing, quantification of specific proteins, as well as necessary in vivo tests
particle fused-based 3D printing, and light-assisted 3D printing according to the medical device categorization by tissue contact and
(Guvendiren et al., 2016; Kalaskar, 2017). Different printing ap- contact duration (Ratner, Hoffman, Schoen, & Lemons, 2013).
proaches and their targeted applications will certainly pose some dif- For the scaffolds that are not aimed for permanent implants, the
ferent requirements for material properties. However, all biopolymers printed structures are supposed to degrade into biocompatible by-pro-
for 3D printing application in medicine and pharmacy need to meet a ducts in a controlled manner, permitting the cells to produce their own
series of strict requirements, considering both physiological conditions extra cellular matrix or allowing the incorporated components to reach
and their associated interactions with the local body environment the desired release profiles (Axpe & Oyen, 2016; Ursan, Chiu, & Pierce,
(Kalaskar, 2017). Fig. 2A shows the key material property requirements 2013).
for 3D printing applications in medicine and pharmacy. In general, the Tissue biomimicry poses requirements for the printed constructs not
material property requirements for 3D printing applications in only mimic the natural shape of the organs and tissues, but also the

Fig. 2. Material property requirements and the associated factors to be considered for bioprinting (A) and 3D food printing (B). Adapted with permission from
Roopavath and Kalaskar (Kalaskar, 2017), Copyright (2017), Elsevier; (Godoi et al., 2016), Copyright (2016), Elsevier.

299
J. Liu et al. Carbohydrate Polymers 207 (2019) 297–316

complex, heterogeneous nature of those tissues, including the desired increasing research turn to focus on the application of 3D printing to
biological functionality, sufficient strength and stiffness to maintain functionalize and modify different types of textile substrates.
structural integrity, macro- and micro-porous architecture, which turns Marrying different types of fabrics and 3D printing offers the pos-
into the key challenges for 3D printing biopolymers for biomedical sibility to develop innovative multicomponent smart textiles by de-
application, especially in tissue engineering (Munaz et al., 2016; Zhang, position of functional polymers or intelligent materials onto textiles
Fisher et al., 2015). More detailed requirements of the 3D printing without compromising on quality and flexibility of the fabric
materials or the printed structures should be considered for a specific (Grimmelsmann, Meissner, & Ehrmann, 2016; Korger et al., 2016). 3D
target application and its printing approach. printed smart textiles can find many applications such as wearable
electronics with heart-rate monitoring and high speed communications,
2.2. Customized food fabrication wearables with protection purposes for soldier, firefighters and astro-
nauts, unique sportswear that manages body temperature, smart ban-
3D printing for food fabrication offers opportunities to manufacture dages, and virtual reality gloves (Gao et al., 2017; Hashemi Sanatgar,
food products with customization in shape, color, flavor, texture and Campagne, & Nierstrasz, 2017; Hashemi Sanatgar, Cayla, Campagne, &
even nutrition (Sun et al., 2015). Generally, materials for 3D food Nierstrasz, 2017).
printing should be edible, printable, rigid and stiff enough to support its The FDM deposition is the most common approach to functionalize
own weight and the weight of subsequently deposited layers without textiles with thermoplastic polymers, e.g. ABS, PLA and Nylon. Direct-
deformation or shape change, and resist to post-processing (e.g. baking, write has been applied to print materials that do not possess the ther-
cooking, and frying). Fig. 2B shows the general material requirements moplastic properties required in the FDM method, e.g. cellulosic ma-
for 3D food printing, which include printability, applicability, and post- terials. SLS printing of different types of powdered materials, e.g.
processing feasibility. Controlling the physical-chemical, rheological, thermoplastic polyurethane (TPU), could be adopted to fabricate soft
structural and mechanical properties of the materials is the key to en- fabrics for a fashion application (Beecroft, 2016; Tenhunen et al.,
sure the accurate precise, and smooth printing process (Godoi, Prakash, 2018).
& Bhandari, 2016; Lin, 2015; Sun et al., 2015). A variety of natural- Polymer deposition depends strongly on the combination of textiles
derived biopolymers such as cellulose, hemicellulose, pectin, starch, and polymers. For effective deposition of polymers onto fabrics, areas
alginate, agarose, and chitosan show potential to be utilized in 3D food that need to be considered include: (a) sufficient mechanical properties
printing owing to their nontoxicity, edibility, high abundancy, bioac- of the printed objects and the interaction with fabrics; (b) drapability of
tivity (e.g. serve as dietary fibers), and long history of utilization in the printed fabric for free movement; (c) flexibility of the printed fab-
traditional food production (Alec, 2016; Godoi et al., 2016; Lille, rics when subject to daily use; (d) washability (Hashemi Sanatgar,
Nurmela, Nordlund, Metsä-Kortelainen, & Sozer, 2018). Campagne et al., 2017; Hashemi Sanatgar, Cayla et al., 2017). It is
Currently, four types of 3D printing approaches available for 3D unequivocal that further research of new materials, compatibility be-
food printing, including extrusion based printing, selective sintering tween fabrics and polymer, polymer-textile adhesion and polymer de-
printing (with laser or hot air), binder jetting, and inkjet printing (Liu, position technology would benefit the production of individualized or
Zhang, Bhandari, & Wang, 2017). Each printing method poses different customized functional textiles as well as apparel.
requirements for the materials. Therefore, more requirements upon the
3D food printing of biopolymers depend on the printing method and the 3. 3D printing of natural-derived biopolymers
final products. Natively printable materials like hydrogels, starch,
chocolate, and cake frosting can be printed through an extruder, while 3.1. 3D printing of lignocellulosic biomass materials
some non-printable food by nature such as meat, vegetables, and rice
have to be processed into the form of powder or paste before subjecting The main components of lignocellulosic biomass, including cellu-
to 3D printing. Biopolymers in the form of hydrogel or paste-like slurry lose, hemicellulose, and lignin, represent the most abundant en-
with shear thinning behavior are beneficial for the extrusion based 3D vironmentally friendly biopolymers with the properties of sustain-
food printing, as they possess not only relative low viscosity upon shear ability, non-toxicity, biocompatibility, and biodegradability (Fig. 3)
stress to be easily extruded out from nozzle, but also suitable yield (Fernandes et al., 2013; Liu, Willför, & Xu, 2015). However, the physic-
stress and elastic modulus upon the departure from the nozzle to chemical properties of lignocellulose, e.g. infusibility and insolubility in
maintain a high shape fidelity (Godoi et al., 2016; Liu, Zhang, Bhandari, common solvents, limited their application in the 3D printing process.
Wang, 2017). The main requirements for the selective sintering and Hence, a variety of strategies to modify the lignocellulosic biomass or to
binder jetting based printing of food are that the materials should be update the 3D printing technologies have been adapted to explore the
edible free-flowing powder and have no tendency to agglomerate or to lignocellulosic biomass as a potential feedstock for 3D printing (Fig. 3).
adhere to contact surface. For the selective sintering, the materials A list of 3D printing natural-derived biopolymers, with the focus on the
should have appropriate thermal stability, should be able to fuse to- biomass polysaccharides, is tabulated in Table 1 and further discussed
gether without decomposition during laser scan (Liu, Zhang, Bhandari, in the following sections.
Wang, 2017). Generally, materials for inkjet printing should have low
viscosity to be easily ejected through the print head. 3.1.1. 3D printing of native cellulose-based materials
Cellulose, a homogeneous polysaccharide consists of linear β-(1→
2.3. Textile and apparel products 4)-glucan with intra- and intermolecular hydrogel networks (Fig. 3), is
the most abundant renewable and biodegradable biopolymers world-
3D printing technologies that offer mass customization of products wide. Development of cellulose-based materials as a new type of feed-
on-demand by freedom of shape and design are increasingly being ex- stocks for 3D printing technologies would open a new window to ex-
plored for textile industry to create parts that would be impossible to plore cellulose materials. Currently, the most intensively studied native
produce with conventional techniques (Sun & Zhao, 2017, 2018). Fully cellulose-based 3D printing material is the cellulose nanofibrils (CNF)
printed shoes, dresses, and fashion collection were presented and drawn hydrogel. The CNF hydrogel-based inks are biocompatible and can
increasing interest of designers over the last few years. Generally, ap- mimic the microenvironment of extracellular matrix (ECM), which,
parel and textile products require sufficient flexibility, resilience, and therefore, has been recently considered in biomedical applications (Liu,
tensile strength in accommodating to wearer’s comfort and body Cheng et al., 2016; Liu, Chinga-Carrasco et al., 2016). However, the
movement, which are, unfortunately, hard to meet with most of the high hydrophilicity and the inherently entangled state of CNF limited
currently available materials (Martens & Ehrmann, 2017). Therefore, the increase of ink concentration at a given viscosity and storage

300
J. Liu et al. Carbohydrate Polymers 207 (2019) 297–316

Fig. 3. Molecular structures of cellulose and representative hemicelluloses, as well as their representative 3D printed structures. 3D printed structures are adapted
with permission from: Müller et al. (2016), Copyright (2016), Springer; (Schütz et al., 2017), Copyright (2015), John Wiley and Sons; (Pattinson & Hart, 2017),
Copyright (2017), John Wiley and Sons; (Markstedt et al., 2014), Copyright (2014), Mary Ann Liebert.

modulus. Efforts have been carried out to directly dissolve or utilize 3.1.2. 3D printing of cellulose derivative-based materials
native cellulose in other forms. The super-molecular structure of cellulose, i.e. high crystallinity
Native cellulose cannot be readily dissolved in common solvent degree and rigid intra/intermolecular hydrogen bond networks, re-
systems due to the strong H-bond network and crystalline structure. stricts its application. However, the abundant hydroxyl groups on the
Ionic liquids (ILs) are a new class of cellulose “green” solvents with low cellulose surface opens opportunities for chemical modification or
melting temperature. Dissolution of cellulose is achieved by disruption functionalization, such as esterification, etherification, selective oxi-
of native cellulose H-bond network with new formed H-bonds between dation, graft copolymerization, and intermolecular crosslinking reac-
cellulose and anions, and the hydrophobic interactions with cations. tion, bringing novel opportunities to exploit them for various applica-
Anti-solvent, such as water, ethanol, and acetone, can be added to re- tions such as in food, cosmetic, medicine and pharmacy (Liu, Willför
generate cellulose and recycle the ILs (Gupta & Jiang, 2015). Markstedt, et al., 2015; Qi, 2017).
Sundberg, and Gatenholm (2014) dissolved different cellulose mate- Interfacial compatibility problem between polar native cellulose
rials, including bacterial cellulose (BC), microcrystalline cellulose and nonpolar polymers matrix is the key limitation for the application
(MCC), and dissolving pulp, in an ILs (EmimAc) with solid concentra- of native cellulose which has excellent mechanical performance as re-
tion up to 4% and applied for direct ink writing (DIW) printing. Com- inforcing agents in composites manufacture. To improve the compat-
plex patterns of 2D structures and multilayered cylinders structures ibility with the PLA, surface-modification of cellulose was conducted
were printed and coagulated with water. However, the spatial resolu- using a titanate coupling agent via coordination exchange (neoalkoxy)
tion of the printed structure and the recycle of the ILs still need to be or solvolysis (monoalkoxy) (Murphy & Collins, 2016). The modified
addressed. cellulose was found to improve dispersion and interfacial adhesion
Cellulose nanocrystals (CNC) prepared from acids hydrolysis may between the cellulose and the PLA, leading to the higher mechanical
offer advantages over the cellulose ILs dissolution and the CNF in terms strength of the composite filament than the neat PLA alone. Porous
of the increase of concentration at a given rheology requirement for 3D scaffold prototypes were successfully printed with FDM using the cel-
printing. Siqueira et al. (2017) dispersed the CNC in water to prepare lulose reinforced PLA filaments, offering the opportunity for rapid
different concentration of suspensions/gels inks (0.5–40%) for DIW and production of fully degradable biocomposite 3D prototypes for appli-
yielded cellulose-based structures with high degree of CNC particle cations in the biomedical, automotive, and construction sectors.
alignment along the printing direction, offering the opportunity to print Similarly, to improve the dispersibility of CNC and interfacial
cellulosic architectures with tailored mechanical properties. bonding after DIW and UV polymerization, Siqueira et al. (2017) ap-
Native cellulose can also be utilized in 3D food printing by serving plied acetylation reaction for CNC surface modification by using me-
as dietary fiber, bulk filling agent, or rheological modifying ingredients, thacrylic anhydride. The modified CNC was dispersed in a mixture of
and reinforcing ingredient owing to its indigestibility, excellent me- photopolymerizable monomer and photoinitiator. After DIW printing
chanical performance, and high viscosity of native cellulose. Lille et al. and curing, the yielded structures were found to have a high degree of
(2018) incorporated the CNF gel in a 3D printing ink formula as re- CNC particle alignment along the printing direction, providing the
inforcing ingredient for the development of healthy customized 3D opportunity to print cellulosic architectures with programmable re-
printed foods. CNF was found to improve the shape stability of the inforcement along prescribed directions.
printed structures and to decrease the hardness of the dried objects, (2,2,6,6-Tetramethylpiperidin-1-yl)oxyl (TEMPO)-mediated oxida-
revealing its potential for 3D printing healthy fiber-rich and structured tion and carboxymethylation are two common approaches to prepare
foods. CNF by introducing negative charges on the cellulosic fiber surface
followed by mechanical processing (Liu et al., 2014). The yield CNF

301
Table 1
Natural-derived biopolymers for 3D printing and their potential applications.
J. Liu et al.

Biobased printing Printing/solidification approach Ink formulation Printed structure Potential applications Achievements/comments References
materials

BC/MCC/Dissolving pulp Bioprinting /nonsolvent (water) Dissolution of cellulose in ionic liquid Complex patterns of 2D Spatially tailored 3D gels Printed structure needs a Markstedt et al. (2014)
coagulation (EmimAc), 1-4%. structures, multilayered or membrane structures nonsolvent for coagulation;
Cylinders. production. Vertical supports are needed for
multilayers structures; printing
spatial resolution need to be
enhanced.
CNC/acetylated CNC DIW/DIW + UV photopolymerization Aqueous CNC suspension or gel, 0.5-40% / Grids and block (8 layers). Cellulose-based Programmable reinforcement Siqueira et al. (2017)
10% or 20% acetylated architectures with tailed owing to the CNC particle
CNC + photopolymerizable mechanical properties. alignment along the printing
monomer + photoinitiator. direction.
CNF/alginate Bioprinting/crosslinking with Human chondrocytes laden CNF-alginate Grids, solid disc, human ear Cartilage tissue Forces during mixing and the cross- Markstedt et al. (2015),
100 mM CaCl2 bioinks; and sheep meniscus shaped engineering linking process could result in the Martínez Ávila et al.
Human bone marrow-derived stem cells and structures; decrease of cell viability. Cells (2016), Möller et al.
human nasal chondrocytes-laden CELLINK®. patient-specific auricular embedded in thick structure may (2017)
constructs with 50% open have oxygen/nutrient/waste
porosity. Lattice-structure; diffusion problems.
Lattice-shaped constructs Ca2+ may have negative effect on
the cell viability. Loss of Ca2+
during long-term culture may
weaken the strength of the
structure.
Clinically demonstrated the
neocartilage formation and
stability of the 3D bioprinted cell-

302
laden scaffolds in vivo.
CNF/alginate/hyaluronic Bioprinting/crosslinking with Living adipocytes-laden CELLINK®/ Gridded structure with three Adipose tissue Incorporation of HA lead to the Henriksson et al. (2017),
acid 100 mM CaCl2, or 0.001% (v/v) H2O2 CELLINK-H (CNF: hyaluronic acid, 80:20, layers; engineering; cells to releases more angiogenic Nguyen et al. (2017)
70:30); Six-layer grids Cartilage lesions growth factors which could be
Human-derived induced pluripotent stem treatment benefit the co-culture of vascular
cells (iPSCs) and irradiated human cells for future printing of blood
chondrocytes- laden CNF-alginate (60:40)/ vessel structures for gas and
hyaluronic acid composite inks. nutrient influx in the printed tissue;
The co-printed irradiated human
chondrocytes stimulates the iPSCs
differentiation.
CNF/Alginate sulfate Bioprinting/ crosslinking with Bovine chondrocytes-laden CNF-Alginate Grid-like constructs (2–4 Cartilage tissue The biological performance of the Müller et al. (2016)
100 mM CaCl2 sulfate bioinks (4.2:1, v/v) layers) engineering cells was highly dependent on
printing conditions, e.g. the nozzle
geometry.
CNF/Alginate/Glycerin VTT’s microdispensing system/ Hydrogels composed of biofunctionalized Grid structures Biomedical devices, The biofunctionalized CNF-alginate Leppiniemi et al. (2017)
crosslinking with 90 mM CaCl2 CNF, alginate, and glycerin. wearable sensors, and hydrogel provide a platform for
drug-releasing materials immobilization of bioactive
components in the 3D printed
structure.
CNF Bioprinting/ crosslinking with 0.05 M TEMPO-mediate oxidized CNF, 0.95% / Grid structure, 3D scaffold Wound dressing; Printed structures have limited Håkansson et al. (2016),
CaCl2; Bioprinting; Extrusion-based Carboxymethylated and periodate oxidized (9 layers); Packaging, textiles, resolution. CNF hydrogels Lille et al. (2018), Rees
3D printing CNF, 3.9%; Hydrogel / aerogel / biomedical devices, and collapsed and no defined tracks et al. (2015)
Carboxymethylated CNF(2%) / CNF-carbon conductive films; furniture with could be observed;
nanotube conductive ink; Squares filled with diamond- conductive parts; Incorporation of functional
CNF, starch, milk powder, rye bran, oat/faba like structures. Healthy, fiber-rich, and ingredients into the CNF-based ink
bean protein. structured foods. for 3D printing offers opportunities
to construct a vast verity of new
applications, e.g. functional
(continued on next page)
Carbohydrate Polymers 207 (2019) 297–316
Table 1 (continued)

Biobased printing Printing/solidification approach Ink formulation Printed structure Potential applications Achievements/comments References
J. Liu et al.

materials

packaging, textiles, health care


products, and furniture;
Tip clogging is challenge,
mechanical properties of the
printed structures still need to be
optimized.
Cellulose acetate Modified FDM with filament extruder Dissolution of cellulose acetate in acetone Eyeglass frames, rose, Customized medical Evaporation of acetone limited the Pattinson and Hart (2017)
replaced by a capillary nozzle. (25-35%). Antimicrobial agents. forceps with customized tips instruments manufacturing rate.
containing antimicrobial
agents.
Methylcellulose-alginate Bioplotting, crosslinked with Ca2+ Methylcellulose-3% alginate Scaffolds Regenerative therapy, Temporary integration of Schütz et al. (2017)
e.g. fat, cartilage, defects methylcellulose offers better
at tissue interfaces. printability while keeping the
advantages of low-concentrated
alginate bioink for cell embedding.
Titanate surface modified FDM Titanate surface modified cellulose (0-5%), Porous scaffold prototype Biomedical, automotive, Dispersion and interfacial adhesion Murphy and Collins
cellulose PLA. and construction. between the cellulose and the PLA (2016)
were improved with low ratio of
modified cellulose.
Thermoplastic cellulose Extrusion based 3D printing, DIW Cellulose hydrogel, thermoplastic cellulose Self-standing objects, Clothing design and – DWoC Ali (2015)
derivatives, derivatives, cellulose-based plastics and pulp surface structures on fabrics. manufacture, Medical
hydrogels, cellulose- fibre composites science.
based plastics and
pulp fibre composites
Hemicelluloses (GGM), FDM Blends of GGM and PLA with varied ratio up 3D scaffold (10 layers) Biomedical devices, – Xu et al. (2018)

303
PLA to 25% of GGM. tissue engineering.
Starch Particle binding, post-processing to A three-powder blend consists of 50% corn Porous cylindrical and Tissue engineering; Biocompatibility of the scaffolds Lam et al. (2002), Lille
remove unbound powders and to starch, 30% dextran, 20% gelatin, and water- rectangular scaffolds, solid Food 3D printing needs to be studied in the future; et al. (2018), Liu, Zhang,
enhance mechanical strength; based binder; cylinder scaffolds; Starch working as thickening / Bhandari, Yang (2017),
FDM + enzymatic crosslinking / Sodium caseinate, apple pectin, potato Multi-layer lattice structure gelling agent, rheological modifier, Schutyser et al. (2018),
Extrusion-based 3D printing starch, sucrose / Potato starch, mashed / Apple, heart-shape, bear and carbohydrate source. Yang et al. (2018)
potatoes / Potato starch, lemon juice head, π, and Chinese
character / Cylinder
structure, anchor, lizard,
star, cycle, and pyramid
shape.
PLA FDM + cold atmospheric plasma post- PLA, biomacromolecules (e.g. protein, Scaffolds Bone tissue engineering Post-processing, surface Kao et al. (2015), Park
processing / FDM + surface coating growth factor, RGD), bioactive filler (e.g. functionalization / coating, et al. (1998), Serra,
hydroxyapatite, bioactive glass) incorporation of (in-)organic Planell et al. (2013),
components to tune the Senatov et al. (2016),
biochemical, material-cellular Wang et al. (2016) Zhang
responds, mechanical, and et al. (2016)
morphological properties of PLA-
based scaffolds.
Alginate, oxidized Bioprinter, custom-made, piston- Sodium periodate oxidized alginate, RDG Lattice structure, Tissue engineering, Effect of bioink properties (e.g. Bendtsen et al. (2017),
alginate driven deposition system, on a peptides-conjugated alginate; 5 х 5 dot array; fabrication of complex oxidization degree, viscosity, and Chung et al. (2013) Duan
temperature-controlled plate at 4 °C, Alginate, Human glioma U87-MG cells-laden Branched vascular and clinically sized soft density) on the printability, et al. (2013), Daly et al.
crosslinked with Ca2+; alginate at 4% w/v; structures; tissues; structural stability, and (2016), Diogo et al.
A dual syringe extrusion-based 3D NIH 3T3 mouse fibroblast cells-laden Square grid patterns, Tissue engineering, degradability were investigated. (2014), Gudapati, Yan,
printer. Three crosslinking stages with alginate solution, 1-3% w/v; dumbbell-shaped structure, heart valves Diffusion of crosslinker Ca2+ may Huang, and Chrisey
80 mM CaCl2, 100 mM CaCl2, 60 mM Porcine aortic valve interstitial cell (VIC) and Tri-leaflet heart valves; replacement; have negative effect on Ca2+ (2014),
BaCl2 for the purpose of enhancing human aortic root smooth muscle (SMS) cell- Three layers lattice Myoregenerative sensitive cell types; Jia et al. (2014), Tabriz,
printability, rigidity after printing, laden alginate (5%)/gelatin (6%) hydrogels; structure; application; Three crosslinking stages allow for Hermida, Leslie, and Shu
and long-term stability; Pre-crosslinked alginate and composite ink 7-layer Bone tissue engineering; tuning the bioink printability, (2015), Wu et al. (2016)
(continued on next page)
Carbohydrate Polymers 207 (2019) 297–316
Table 1 (continued)

Biobased printing Printing/solidification approach Ink formulation Printed structure Potential applications Achievements/comments References
J. Liu et al.

materials

Matrix-assisted pulsed-laser of alginate (1, 2, 4% w/v) gelatin (10% w/v), porous, cylindrical scaffolds; mechanical properties, and long-
evaporation direct-writing, 4:1; Vertebrae constructs; term structural stability;
crosslinking with 2% (w/v) CaCl2; Mouse calvaria 3T3-E1 cells-laden alginate- Rectangular construct Cell viability was found to be
3D bioprinter, modified from a Fab@ polyvinyl-alcohol-hydroxyapatite hydrogel affected by the alginate gelation
Home RP platform, crosslinked in bioink; processes, including alginate
300 mM CaCl2; Arg-Gly-Asp adhesion peptides incorporated concentration, gelation time, and
Extrusion-based bioprinting, and gamma-irradiated alginate bioink, laser fluence. Process-induced cell
crosslinked with 2% w/v CaCl2; reinforced with polycaprolactone fibers β- injury arises from mechanical
3D bioprinting; tricalcium phosphate/alginate hybrid stress, thermal effect of laser
Rapid prototyping (Fab@home) materials; fluence, and chemical effect of
Human corneal epithelial cells-laden Ca2+;
collagen/gelatin/alginate hydrogel bioink Mechanically robust living tri-
leaflet heart valves were bioprinted
by encapsulating VIC and SMS
cells;
Pre-crosslinked alginate ink was
not suitable for printing due to low
resolution and fluid-like properties
of the printed scaffolds;
Further in vitro studies of cells
differentiation to regenerate new
bone are needed;
Complex solid organs can be
fabricated by bioprinting
precursors that have the capacity to

304
mature into their adult
counterparts over time in vivo;
Further surface functionalization to
facilitate bioadhesion and to
improve mechanical properties
using different polymers or
ceramics are needed;
The degradation profile of the
bioprinted constructs can be tuned
by altering the ratio of sodium
citrate/sodium alginate,
Agarose/collagen 3D drop-on-demand printing with the Agarose (0.25-1.5%) and type I collagen Columns structure (2-3 cm) Drug discovery Impact of hydrogel structure, Köpf et al. (2016)
supporting effect of perfluorocarbon (0.2%) stiffness, and composition on cell
(cooled down to 10 °C to accelerate morphology and possibility of
the agarose gelation) mechanical reinforcement of the
blend hydrogel should be further
investigated.
Agar/alginate Modified extrusion-based 3D printing, Agar, alginate, acrylamide, N,N’- Dogbone and meniscus Load-bearing tissue – Wei et al. (2015)
crosslinking with 0.5 M CaCl2 and UV methylenebis(acrylamide), photo-initiator substitutes
irradiation
Agarose Custom-made extrusion-based 3D Human mesenchymal stem cells-laden Vascular bifurcation and Tissue engineering and Promising approach allows for Duarte Campos et al.
printing in a hydrophobic high- agarose hydrogel (1.5%); cylinder structure; regenerative medicine; fabrication of compositionally and (2013), Norotte et al.
density fluid (C12F27N2); Smooth muscle cells and fibroblasts-laden Vascular tubular grafts; Tissue architecturally intricate tissue; (2009), Navarro and
Extrusion-based bioprinting; multicellular spheroids or cylinders Cylinder, box, and ear fabrication Higher resolution is needed to print Garcia (2016), Wu et al.
concomitantly were printed with agarose structure. complex structure. (2017)
rods template followed by post fusion to
form vascular tubes;
Hela cells-laden agarose hydrogel (1, 1.25,
1.4%)
(continued on next page)
Carbohydrate Polymers 207 (2019) 297–316
J. Liu et al. Carbohydrate Polymers 207 (2019) 297–316

hydrogels with shear thinning behavior and high zero shear viscosity
have been tested as inks for 3D bioprinting. Rees et al. (2015) prepared

Demirtaş et al. (2017)


Almeida et al. (2014),

(2017), Morris et al.


Elviri, Foresti et al.
the CNF hydrogels with TEMPO-mediated oxidation and a combination
of carboxymethylation and periodate oxidation and used to print 3D
porous structures for wound dressing application. Both CNF-based inks
References

were found to inhibit bacterial growth, suggesting the potential appli-


(2016),
cation as wound dressing materials. The CNF prepared from TEMPO-
chemistry, topography and scaffold mediated oxidation failed to print constructs with acceptable resolution
and desired tracks due to the limited low consistency (0.95%) but re-
Wrinkle formation and its impact

The macrophage responses were


found to be affected by material
scaffold biodegradation kineties

performance in mechanical and


Chitosan-based hydrogel were latively high viscosity. The one prepared by carboxymethylation and
on cell growth, as well as the

cellular activities supporting.


periodate oxidation was suitable for being used as bioink in terms of
Achievements/comments

still need to investigate;

high consistency (3.9%) and appropriate rheological properties. The


found to have superior

printed 3D structures have fine tracks with open porosity and have the
potential to carry and release antimicrobial components for wound
dressing. Carboxymethylated CNF with high surface charge and higher
architecture;

consistency (2%) which holds the 3D shape after deposition has been
investigated as ink to print 3D structures (Håkansson et al., 2016). By
controlling the solidification process, e.g. air drying, air drying with
surfactants, solvent exchange before drying, and freeze drying, 3D
Potential applications

printed structures with desired mechanical, surface texture, and porous


structure can be obtained (Fig. 4A). Incorporation of functional in-
Soft tissue repair;
Tissue or organ

Cartilage tissue

gredients, e.g. conductive carbon nanotubes, offers additional func-


regeneration;

Engineering.
Bone tissue

tionality to the printed products. As shown in Fig. 4B, C, the conductive


repairing;

ink which composed of carbon nanotubes and CNF can be printed in the
CNF hydrogel matrix in different layers with fine resolution. After
drying, two over crossing conductive lines with decent electrical con-
orthogonal-diagonal double
Multi-layers scaffolds, thick

ductivity were separated by the insulation CNF layers, suggesting a


Orthogonal displaced and

potential for fabrication of sustainable commodities such as packaging,


Disk shaped hydrogel,
Multi-layers scaffolds;

textiles, biomedical devices, and furniture with conductive parts.


Printed structure

Pattinson and Hart (2017) demonstrated the manufacture of fully


layer scaffold;

1 mm x 6 mm

dense cellulose-based materials using cellulose acetate by 3D printing.


Ear scaffold;

The cellulose acetate has 80% of the surface hydroxyl groups sub-
stituted by acetate groups which effectively disrupted the hydrogen
film;

bonding network of native cellulose and made it possible to be dis-


solved in acetone with high consistency (25–35%). Solid cellulose
MC3T3-E1 pre-osteoblast cell laden chitosan
(2% w/v, in 0.1 M acetic acid) and chitosan-
and 1:15, and photoinitiator (Irgacure 819);
8% chitosan hydrogel (in an acidic mixture:

polyethylene glycol diacrylate at 1:5, 1:10,


40 vol. % acetic acid, 10 vol. % lactic acid,

Chitosan (6% w/v) and chitosan-raffinose

acetate structures with isotropic strength and high toughness were built
Photocurable resin consist of chitosan :

upon the evaporation of acetone. By incorporation of antimicrobial


agents, direct 3D printing of cellulose acetate-based objects with tai-
3% chitosan (in 2% acetic acid);

lored biochemical functionality (e.g. antimicrobial properties) can be


solution (in 2% acetic acid);

achieved, enabling customization of medical instruments in short time


hydroxyapatite hydrogels
and 3 wt. % citric acid);

(e.g. forceps with customized tips, Fig. 4D).


Methylcellulose has been utilized by Schötz et al. (2017) to enhance
Ink formulation

the temporary printability of a 3% alginate which has limited viscosity.


The addition of methylcellulose significantly enhanced the bioink
viscosity, enabling accurate and easy 3D bioplotting of constructs with
tailored architecture and of high shape fidelity, after which the me-
thylcellulose was released from the scaffolds during the following cul-
tivation. The embedded mesenchymal stem cells in the 3D printed
dried at 50 °C for 24 h, neutralized in
Extrusion-based 3D printing, vacuum

(Customized from a FDM), frozen on


the cooling substrate and gelation in

Extrusion-based bioprinting, thermal

methylcellulose-alginate scaffolds were found to maintain their differ-


Direct ink printing, in situ cross-
linking with NaOH (8% w/v) in
Printing/solidification approach

Stereolithography, cured using

entiation potential with high viability. The temporary integration of


crosslinking with the glycerol
a 1 M NaOH solution for 2 h;

Extrusion-based 3D printing

methylcellulose into the alginate-based bioink with low concentration


allowed the generation of scaffolds with high shape fidelity and stabi-
lity while keeping the advantages of low-concentrated alginate bioink
phosphate at 37 °C
KOH (8% w/v);

for cell embedding.


ethanol (70%);
405 nm laser;

A research team from Aalto University working within the project of


“Design Driven Value Chains in the World of Cellulose” studied 3D FDM
printing of thermoplastic cellulose derivatives (Ali, 2015). Printability
of the pure cellulose derivative was improved through plasticization
which reduced melt viscosity, improved the layer adhesion, and low-
ered glass transition temperature, allowing manufacture of a variety of
Table 1 (continued)

cellulose-based 3D structures. Similarly, within the same project, oxi-


Biobased printing

dized cellulose hydrogel, cellulose-based plastics, and pulp fiber com-


posites were studied for textile printing (Fig. 4E, F) (Kataja, 2015;
materials

Chitosan

Tenhunen et al., 2018). Structure with desired look or feel, e.g. hard or
soft, strong or brittle, stiff of flexible, porous or dense texture, can be
printed directly on substrates or fabrics or form self-standing structures

305
J. Liu et al. Carbohydrate Polymers 207 (2019) 297–316

Fig. 4. 3D printed structures using cellulose


derivatives. 3D printed and freeze-dried car-
boxymethylated CNF chair (A); Schematic
drawing and the printed carboxymethylated
CNF structure, where the black conductive
lines are clearly separated by the translucent
CNF (B); Images of the flexible electrical con-
ductive film after drying in air (C); Miniature
forceps with customized tips printed using
cellulose acetate containing antimicrobial dyes
(D); 3D printed cellulosic structure on fabrics
and multiple layers textiles (E, F). 3D printed
structures are adapted with permission from:
Håkansson et al. (2016)), Copyright (2016),
John Wiley and Sons; Pattinson and Hart
(2017), Copyright (2017), John Wiley and
Sons; 3D printed fabrics photos from VTT,
Finland.

and can find applications from medical science to personalized sports- for 3D bioprinting and the results suggested that the shear forces during
wear. mixing and the cross-linking process could result in the decrease of cell
viability. Although the cell viability can be recovered after 7 days
3.1.3. 3D printing of cellulose composites-based materials culture, cells embedded in the structures (e.g. thickness exceeds 150-
Incorporation of other ingredients into the cellulose-based ink for- 200 μm) may gradually lose viability due to the limited diffusion
mulations may significantly improve ink properties such as processa- oxygen/gas, nutrients, growth factors, and waste-product (Ventola,
bility, printability, mechanics, and bioactivity (Guvendiren et al., 2014).
2016). CNF hydrogels have ideal structural similarity to ECM and ex- To overcome this issue, Martínez Ávila, Schwarz, Rotter, and
cellent rheological properties for 3D printing. However, the pure CNF Gatenholm (2016) printed a chondrocyte-laden patient-specific auri-
printed structures lack high shape fidelity (fine line resolution), suffi- cular construct with open porosity for oxygen, nutrients, and waste
cient mechanical properties for handling during transplantation, and diffusion (Fig. 5D). The bioprinted constructs showed an excellent
long-term structure stability, limiting its application in the 3D printing shape and size stability after bioprinting and long-term 3D culture and
of complex scaffolds in different biomedical applications (Gatenholm the CNF-alginate bioink was found to support redifferentiation of
et al., 2016). Alginate also possesses shear thinning behavior at high human chondrocytes, reestablishing and maintaining their chondro-
shear forces, but the low viscosity at zero shear rate makes it difficult to genic phenotype, suggesting the usefulness of the bioink for auricular
print 3D constructs. Inks composed of CNF and alginate keep the shear cartilage tissue engineering and many other biomedical applications.
thinning behavior and high zero shear viscosity of CNF while the al- Similarly, Nguyen et al. (2017) designed the human-derived in-
ginate allows cross-link by divalent ions to maintain a long-term shape duced pluripotent stem cells (iPSCs) and irradiated human chon-
fidelity and structural integrity after 3D printing. drocytes laden CNF-alginate/hyaluronic acid composite inks to mimic
Markstedt et al. (2015) optimized and evaluated the printability and the cartilaginous tissue for potential cartilage lesions treatment. In the
biocompatibility of this composite ink. Anatomically shaped cartilage composite inks, the CNF was supposed to mimic the bulk collagen
structures, such as a human ear and sheep meniscus, were 3D printed matrix, alginate stimulates proteoglycans, and hyaluronic acid serves to
with high shape fidelity and stability (Fig. 5A–C). Human nasoseptal substitute the hyaluronic acid in native cartilage. Bioink of CNF-algi-
chondrocytes were incorporated into the CNF-alginate composite ink nate was found to maintain the pluripotency of the stem cell and to

306
J. Liu et al. Carbohydrate Polymers 207 (2019) 297–316

Fig. 5. 3D printed human ear (A and D) and


sheep meniscus (B and C) with CNF-alginate
bioink; Photograph of an hNC-laden scaffold
after 14 days of implantation (E). Adapted with
permission from Markstedt et al. (2015),
Copyright (2015), American Chemical Society;
Martínez Ávila et al. (2016), Copyright (2016),
Elsevier; Möller et al. (2017), Copyright
(2017), Wolters Kluwer Health.

support the new generation of cartilaginous tissue with collagen ex- 3.1.4. 3D printing of hemicelluloses-based materials
pression and high cell density, suggesting the bioprinting of iPSCs with Hemicelluloses, which mainly include xylans, glucomannans, ara-
CNF-alginate bioinks as a promising treatment to repair damaged car- binans, galactans, and glucans, are a type of heterogeneous poly-
tilage. saccharides and are widely distributed in biomass. Hemicelluloses have
Alginate sulfate with the potential to supports chondrocyte pheno- long been studied and widely used in different areas, such as food and
type has also been incorporated into CNF hydrogel to formulate bioinks feed, medicine and pharmaceutics, and papermaking due to their bio-
for cartilage bioprinting (Müller, Öztürk, Arlov, Gatenholm, & Zenobi- compatibility, biodegradability, nontoxicity, and some specific ther-
Wong, 2016). The composite bioink was found to promote bovine apeutic activities (Liu, Willför et al., 2015). However, utilization of
chondrocytes spreading, proliferation, and collagen II synthesis. How- hemicelluloses either in native or their functionalized products as
ever, the bioprinting process significantly compromised cell prolifera- feedstocks for 3D printing has rarely been studied.
tion, especially in the case that uses nozzles and valves with a small Recently, a spruce wood hemicellulose (O-acetyl galactogluco-
diameter and high extrusion pressure and stress. mannan, GGM) was utilized to partially replace the synthetic PLA as
Currently, the CNF-alginate composite ink formulation has been feedstock in 3D FDM printing (Xu et al., 2018). The blends of hemi-
successfully commercialized with the brand name of CELLINK® and has celluloses and PLA with a varied ratio up to 25% of GGM were evenly
been evaluated in vitro and in vivo with human chondrocytes, human mixed with a solvent casting approach and were extruded into filaments
dermal fibroblasts and keratinocytes, neural cells, mesenchymal stem by hot melt extrusion. 3D scaffold prototypes were successfully printed
cells derived from bone marrow, and adipose tissue and iPSC cells de- from the composite filaments by FDM 3D printing (Fig. 3). As a pioneer
rived from chondrocytes (Gatenholm et al., 2016). For example, exploration, this study demonstrated the feasibility of applying bior-
Henriksson, Gatenholm, and Hagg (2017) bioprinted the living adipo- enewable hemicelluloses as a novel material candidate for FDM 3D
cytes-laden CELLINK® bioink and a bioink composed of nanocellulose printing, which has explored a new route to utilize hemicellulose-based
and hyaluronic acid (CELLINK-H) for 3D bioprinting in adipose tissue biopolymer in 3D printing for versatile applications in, but not limited
engineering. The adipocytes laden in 3D printed structures, especially to, biomedical devices.
the CELLINK-H, produced better adipogenic differentiation and more By mimicking the hemicelluloses’ natural affinity to cellulose,
mature cell phenotype with larger lipid droplets than conventional 2D Markstedt, Escalante, Toriz, and Gatenholm (2017) mixed a tyramine
culture system, suggesting a promising method for adipose tissue en- substituted xylan with the CNF to introduce cross-linking property of
gineering. The clinical study has recently been conducted by trans- the all-wood-based inks for 3D printing. The tyramine substitution de-
planting the bioprinted cell-laden CELLINK® bioink scaffold in a sub- gree of the xylan-tyramine and the mixture ratio were found to influ-
cutaneous pocket of mice (Möller et al., 2017) (Fig. 5E). The ence the printability and cross-linking density of the all-wood-based
histological, immunohistochemical, and mechanical analysis suggested inks, and the swelling properties of the printed structure, opening the
that the 3D bioprinted scaffolds have excellent structural integrity, opportunity to tune the mechanical and structural properties of the
shape fidelity, and good mechanical properties after 60 days of im- printed object and might even transfer into 4D printing.
plantation and the scaffolds can result in cartilage synthesis, suggesting
the potential application of the CELLINK® bioinks in 3D bioprinting
cartilage tissue for application in reconstructive surgery.

307
J. Liu et al. Carbohydrate Polymers 207 (2019) 297–316

(caption on next page)

308
J. Liu et al. Carbohydrate Polymers 207 (2019) 297–316

Fig. 6. Molecular structures of starch and PLA and their representative 3D printed structures (A); Molecular structures of alginate and agarose and their re-
presentative 3D printed structures (B); Molecular structures of chitin and chitosan and the representative 3D printed structures (C); 3D printed structures are adapted
with permission from: Lam et al. (2002), Copyright (2002), Elsevier; Schutyser et al. (2018), Copyright (2017), Elsevier; Liu, Zhang, Bhandari, Yang (2017),
Copyright (2017), Elsevier; Yang et al. (2018), Copyright (2018), Elsevier; Wang et al. (2016), Copyright (2016), Elsevier; Senatov et al. (2016), Copyright (2016),
Elsevier; Serra, Mateos-Timoneda et al. (2013), Copyright (2013), Taylor & Francis; Liberski (2016) HBKU Press; Duarte Campos et al. (2013), Copyright (2013), IOP
Publishing; Norotte et al. (2009)), Copyright (2009), Elsevier; Wei et al. (2015), Copyright (2015), Royal Society of Chemistry; Hong et al. (2015), Copyright (2015),
John Wiley and Sons; Chavanne et al. (2013), Copyright (2013), Walter de Gruyter; Li et al. (2009), Copyright (2009), SAGE Publications; Wu (2016), Copyright
(2016), Elsevier; Morris et al. (2016), Copyright (2016), Springer.

3.2. 3D printing of starch-based materials prostheses, drug-based delivery systems, and anatomical models fabri-
cation (Davachi & Kaffashi, 2015; Guvendiren et al., 2016; Modjarrad &
Starch is another type of abundant polysaccharide produced by Ebnesajjad, 2014; Pilla, 2011; Saini, Arora, & Kumar, 2016; Tyler,
higher plants as energy storage and is composed of the linear amylose Gullotti, Mangraviti, Utsuki, & Brem, 2016). Depend on the ratio of
and highly branched amylopectin with structures of α-(1→4) linked PLLA and PDLA, the mechanical performance and crystalline structure
glucan and the α-(1→4) linked glucan with α-(1→6) branch linkages, of the polymers can be tailored for specific requirements of different
respectively (Fraser-Reid, Tatsuta, & Thiem, 2008) (Fig. 6A). Utilization biomedical applications. For example, sutures and orthopedic devices
of starch or starch-based polymers, e.g. thermoplastic starch and PLA, which need high mechanical strength and toughness can be fabricated
as feedstocks in 3D printing technologies for a variety of application with a high ratio of PLLA, while drug delivery systems that require a
draw increasing interest from both academic research and industry porous and monophasic matrix can be produced with the amorphous
application owing to its renewability, biodegradability, biocompat- PDLA (Shah Mohammadi, Bureau, & Nazhat, 2014).
ibility, and high abundance in nature (Davachi & Kaffashi, 2015; Li Although the abovementioned benefits of PLA render the board
et al., 2014; Zeng, Li, & Du, 2015). applications in biomedical and pharmaceutical areas, some dis-
A three-powder blend ink consisted of 50 wt.% corn starch 30 wt.% advantages of PLA, such as the hydrophobic nature, low ultimate
dextran and 20 wt.% gelatin and water-based binders were adapted for elongation strain, lack of cell motif sites, small particle and the acidic
3D printing of porous scaffolds for tissue engineering application (Lam, byproduct of PLA degradation which induce foreign-body inflammatory
Mo, Teoh, & Hutmacher, 2002). After 3D printing, post-processing by reactions, may cause clinical complications and limited its use in bio-
heat treatment (dried at 100 °C for 1 h) was applied to sinter the three- medicine (Suganuma & Alexander, 1993). Therefore, increasing efforts
blend powdered particles together via necking connections, enhancing have been carried out to enhance the hydrophilic properties, to increase
the strength of the scaffolds and also the water resistance. Although the the cell motif sites, to introduce bioactivity, to improve the ultimate
fabricated porous scaffolds which consisted of natural polymers and elongation strain, and to address the formation of acidic biodegradation
water-based binder showed potential in tissue engineering, further products (Manavitehrani et al., 2016; Ulery, Nair, & Laurencin, 2011).
biocompatibility study is needed. For example, Wang et al. (2016) printed the PLA-based scaffolds and
In 3D food printing, starch is commonly utilized as a thickening/ treated with a cold atmospheric plasma (CAP) modification for en-
gelling agent or rheological modifier, and also serves as an important hancing hydrophilicity and cell motif sites. The ends of the PLA chains
carbohydrate source. Incorporation of starch into the ink formulation (−CH3) on the surface were reacted with the reactive oxygen species
improves the ink printability by offering the ink with shear thinning (peroxides, superoxide, hydroxyl radical, and atomic oxygen) and were
behavior and also the shape stability of the printed structure. For ex- converted to hydrophilic groups, e.g. −CH2OH, −CHO, −COOH.
ample, starch has been applied to facilitate the 3D printing of sodium Meanwhile, the CAP treatment raised the surface temperature of the
caseinate (Schutyser, Houlder, de Wit, Buijsse, & Alting, 2018), mashed PLA to ca. 40 °C, which partially molt and reform the polymer, allowing
potatoes (Liu, Zhang, Bhandari, & Yang, 2017), lemon juice gel (Yang, for a conversion from micro-scale to nano-scale surface features. Results
Zhang, Bhandari, & Liu, 2018), and a protein and fiber-rich food ma- showed that the enhanced surface hydrophilicity and nano-scale
terials, (Lille et al., 2018) by working as thickening/gelling agent. roughness by CAP surface modification significantly promoted both
Similar to cellulose, the structure of starch also has abundant hy- osteoblast (bone forming cells) and mesenchymal stem cell attachment
droxyl groups, (Fig. 6A) offering numerical chemical modification/ and proliferation, suggesting promising application in bone tissue en-
functionalization possibilities to produce starch derivatives or starch- gineering (Wang et al., 2016).
based polymers or plastics for 3D printing (BeMiller & Whistler, 2009). Except for the physical treatment, surface functionalization with
For instance, a thermoplastic starch (TPS) / acrylonitrile-butadiene- bioactive molecules is another approach for improving biomaterials
styrene (ABS) copolymers was prepared by using a compatibilizer bioactivity. For instance, Kao et al. (2015) functionalized 3D printed
(styrene maleic anhydride) to generate hydrogen bonds as well as PLA scaffolds via a mussel inspired surface coating with polydopamine
strong van der Waals force between TPS and ABS and were subjected to to accelerate protein adsorption and cell cycle of the human adipose-
filament extrusion and FDM 3D printing. The results revealed that the derived stem cells. Similarly, surface coating of PLA scaffolds by phy-
TPS/ABS filaments and thereof printed 3D structures have superior sical sorption and covalently bonding of collagen and polymers (poly-
mechanical properties and thermal resistance than that of commercial ethyleneoxide-polypropyleneoxide copolymers), or by incorporation of
ABS, suggesting that the potential of using biomass polymeric materials bioactive glass have also been adapted to improve scaffolds bioactivity
and their derivatives with excellent 3D printing processibility and and to tune cell response (Park, Ben, & Griffith, 1998; Serra, Mateos-
physical performance would be highly promising in the future (Kuo Timoneda, Planell, & Navarro, 2013; Serra, Planell, & Navarro, 2013).
et al., 2016). The surface coated or modified PLA scaffolds are demonstrated to
As shown in Fig. 6A, through fermentation with micro-organisms regulate cell organization, adhesion, proliferation, and to induce os-
and polymerization processes, polylactic acid (PLA) can be produced teogenesis and angiogenesis differentiation, providing a very promising
from starch in a sustainable and a green approach. Owing to a variety of tool to regulate stem cell behavior, and may serve as an effective stem
benefits, including biocompability, nontoxicity, biodegradability, ease cell delivery carrier for bone tissue engineering (Kao et al., 2015; Lin &
of processability, low cost, excellent mechanics, and the green feature Fu, 2016).
of its synthesis routes from renewable resources, PLA makes up the Modification or functionalization of PLA to tune the mechanical
majority of the FDM 3D printing feedstock and has drawing increasing performance of the 3D printed PLA-based scaffolds also play a key role
interest from biomedical application, such as scaffolds, implants, in its biomedical application. Senatov et al. (2016) designed and

309
J. Liu et al. Carbohydrate Polymers 207 (2019) 297–316

fabricated a PLA-based porous scaffold with a shape memory effect by and by assisting deposition of a sacrificial layers (e.g. PEG) that can be
FDM 3D printing for potential application as a self-fitting implant in removed after printing (Armstrong, Burke, Carter, Davis, & Perriman,
bone defect replacement. PLA which serves as bioresorbable matrix and 2016; Lee et al., 2014). The slow degradation kinetics of native alginate
the bioactive filler hydroxyapatite (HA) powder (15 wt.%) were mixed can be addressed by oxidation with sodium peroxide or periodate
in a screw extruder and the PLA/HA composite filaments were extruded (Boontheekul, Kong, & Mooney, 2005; Grigore et al., 2014; Jia et al.,
for FDM 3D printing. The 3D printed PLA-based porous scaffolds were 2014; Kong et al., 2004), by controlling the alginate molar mass and
shown to have the shape memory effect after the heat activation which distribution (Boontheekul et al., 2005; Kong et al., 2004), by enzyme
changed the polymer chain mobility. The scaffolds were found to be assisted degradation (e.g. alginate lyase) (Wong, Preston, & Schiller,
able to withstand up to three compression-heating-compression cycles 2000), or by incubation with sodium citrate that accelerates the dis-
without delamination. However, the higher activation temperature solve of alginate hydrogel by chelating calcium ions from the CaCl2
(70 °C) for the shape memory effect than body temperature is an ob- cross-linked alginate hydrogel (Wu et al., 2016). The bioinert property
stacle for the application and therefore further study to lower the ac- of native alginate can be addressed by conjugation of cell attachment
tivation temperature is needed (Senatov et al., 2016). Similarly, Zhang ligands (e.g. peptide sequence Arg-Gly-Asp, i.e. RGD), or by in-
et al. (2016) prepared a PLA/HA (15 wt.%) composite scaffolds with a corporation of other biomacromolecules, such as fibrin, collagens, ge-
pore size of 500 μm and 60% porosity by a mini-deposition system latin, chitosan, hyaluronic acid, avidin protein, and various growth
(FDM) (Jiang, Shi, Li, & Sun, 2012). Incorporation of HA can not only factors to offer desired bioactivities (Abhay et al., 2013; Chung et al.,
enhance the mechanical performance of the scaffolds (Jiang et al., 2013; Daly et al., 2016; Grigore et al., 2014; Jia et al., 2014; Kundu,
2012), but also find to offer the scaffolds with better biocompatibility, Shim, Jang, Kim, & Cho, 2015; Leppiniemi et al., 2017).
biodegradability, and osteoinductive activity than the pure PLA scaffold Agar and its purified agarose are galactose-based polysaccharides
to enhance bone formation, showing potential application in bone derived from red algae and seaweed such as Gracilaria and Gelidium.
tissue engineering as promising candidate for bone defect repair (Zhang Agarose is a linear polysaccharide composed of alternating (1→4)-
et al., 2016). linked (3,6)-anhydro-α-L-galactose and (1→3)-linked β-D-galactose
(Fig. 6B), while agar is a mixture of the predominant agarose and the
3.3. 3D printing of algae-based materials agaro-pectin that composed of partially sulfated (1→3)-linked D-ga-
lactose (Serwer, 1983; Stephen, Phillips, & Williams, 2006). These ga-
Algae-based materials, e.g. alginate, agarose, and carrageen which lactose-based polysaccharides, especially the agarose, have proven
extracted from brown and red algae, represent another group of natu- suitable for bioprinting owning to the cytocompatibility, stable me-
rally-derived biopolymers from marine biomass for 3D printing inks chanical properties, and mild temperature dependent gelling condition
formulation, especially for 3D bioprinting (Axpe & Oyen, 2016). with the starting gelling temperature of 37 °C and fully crosslinking at
Alginate is one of the most widely explored seaweed anionic poly- 32.7 °C (Duarte Campos et al., 2013; Ferris, Gilmore, Wallace, & in het
saccharides composed of β-(1→4)-linked D-mannuronic acid (M block) Panhuis, 2013; Malda et al., 2013).
and α-L-guluronic acid (G block) units arranged with varying propor- Different cell lines have been encapsulated into the agar, agarose,
tion of GG, MG, and GM blocks (Fig. 6B) (Liu, Willför et al., 2015). It and their composite hydrogel with other polymers such as alginate,
has been recognized as the most commonly employed material as collagen, matrigel, and chitosan as bioinks in bioprinting for potential
bioink formulation in 3D bioprinting for a variety of biomedical and application in tissue engineering, regenerative medicine, and drug
pharmaceutical applications, such as wound healing, cartilage repair, discovery (Duarte Campos et al., 2013; Köpf, Campos, Blaeser, Sen, &
bone regeneration and drug delivery, owning to the benefits of non- Fischer, 2016; Landers, Hubner, Schmelzeisen, & Mulhaupt, 2002;
toxicity, biocompatibility, biodegradability, non-antigenicity, and ease Landers, Pfister et al., 2002; Maher, Keatch, Donnelly, Mackay, &
and mild of gelation (Axpe & Oyen, 2016; Chimene, Lennox, Kaunas, & Paxton, 2009; Navarro & Garcia, 2016; Norotte, Marga, Niklason, &
Gaharwar, 2016; Duan, Hockaday, Kang, & Butcher, 2013; Lee & Forgacs, 2009; Tan et al., 2014; Wei et al., 2015) (Fig. 6B).
Mooney, 2012; Liberski, 2016; Sun & Tan, 2013). By mixing with
multivalent cations, e.g. Ca2+, alginate is ionically crosslinked to form 3.4. 3D printing of chitosan-based materials
a hydrogel, offering an ECM mimicking environment for cells in-
corporation and survival, good printability, structural integrity and Chitosan, a cationic polysaccharide, comprises of β-(1→4)-linked N-
mechanical stiffness. By adjusting or utilize alginate with different G/M acetyl-D-glucosamine and deacetylated D-glucosamine units (Fig. 6C). It
ratio, molar mass, solid content, as well as cell density, the rheological, is the deacetylated form of chitin that naturally exists in exoskeletons of
mechanical, the macro- and micro- structural properties of the alginate- crustaceans (crab, shrimp shells, and insects), as well as the cell wall of
based bioink can be tuned to fit the requirements of different bio- fungi and yeast, which is the second richest natural polymer in nature
printing methodologies and different biomedical applications (Fig. 6B) after cellulose (Dutta, 2016; Kumar, Muzzarelli, Muzzarelli, Sashiwa, &
(Axpe & Oyen, 2016; Grigore, Sarker, Fabry, Boccaccini, & Detsch, Domb, 2004; Rinaudo, 2006). Chitosan-based biomaterials with unique
2014; Liberski, 2016; Luo et al., 2015). advantages have drawn increasing attention for application in medical
However, some undesirable features, including poor long-term and pharmaceutical areas, such as tissue engineering, drug delivery,
structural integrity and mechanical properties, slow and uncontrolled and wound healing, owing to their biocompatibility, bioresorbability,
degradation kinetics, and the bioinert property (non-cell-adhesion by biodegradability, antimicrobial activity, mucoadhesivity, and low
itself), should be carefully considered in the design and development of toxicity, as well as natural abundance (Dutta, 2016; Elviri, Bianchera,
alginate-based bioink for bioprinting (Chung et al., 2013). For instance, Bergonzi, & Bettini, 2017).
the poor mechanical performance could be addressed by combining Generally, chitosan is soluble in acidic condition, e.g. 2% acetic
ionic and covalent crosslinking (Bakarich, Panhuis, Beirne, Wallace, & acid, and the amine groups in chitosan are protonated to confer the
Spinks, 2013; Duan, Kapetanovic, Hockaday, & Butcher, 2014; Grigore poly-cationic behavior. The chitosan chains expand into a semi-rigid
et al., 2014; Kesti et al., 2015; Kong, Kaigler, Kim, & Mooney, 2004; rod conformation due to ionic repulsion between the charged groups
Rutz, Hyland, Jakus, Burghardt, & Shah, 2015), by incorporation of (NH3+) and thus the chitosan ink exhibits shear thinning behavior
other component such as hydroxyapatite (Bendtsen, Quinnell, & Wei, which is beneficial for the extrusion-based 3D printing (Wu, Maire,
2017; Wöst, Godla, Möller, & Hofmann, 2014), polycaprolactone (Daly Lerouge, Therriault, & Heuzey, 2017). Chitosan or chitosan-based
et al., 2016; Shim, Lee, Kim, & Cho, 2012), β-tricalcium phosphate composites in the form of hydrogel or paste, filaments or strip, have
(Diogo, Gaspar, Serra, Fradique, & Correia, 2014), and nanocellulose been applied in a variety of 3D printing technologies for potential bone
(Markstedt et al., 2015; Martínez Ávila et al., 2016; Möller et al., 2017), tissue engineering, cartilage tissue regeneration, and drug screening

310
J. Liu et al. Carbohydrate Polymers 207 (2019) 297–316

(Fig. 6C). These printing approaches including: FDM (Wu, 2016), ste- different biopolymers may bring a challenge with respect to its struc-
reolithography (Morris, Nimbalkar, Younesi, McClellan, & Akkus, tural replacement by cell-secreted ECM. Therefore, the balance between
2016), and different nozzle extrusion-based 3D printing technologies the composite ink properties, and the resulted structural and functional
such as double-nozzle assembling (Li et al., 2009), rapid prototyping changes of the printed objects should be carefully considered
robotic dispensing (Ang et al., 2002), direct-printing (Almeida et al., (Guvendiren et al., 2016; Köpf et al., 2016; Ning & Chen, 2017). In
2014), and low temperature manufacturing (Elviri, Foresti et al., 2017), addition, one key challenge for 3D printing biopolymers in biomedical
However, the poor mechanical resistance and degradability in application, especially in tissue engineering, is the tissue biomimicry.
physiological conditions, as well as the acidic dissolution made it dif- Manipulation of multiple biopolymers and cell types to print vascu-
ficult to incorporate living cells into the chitosan-based inks in bio- larized and metabolically active thick tissues, such as, kidney, heart,
printing for biomedical applications. Neutralization and gelation in lung, and liver tissues, that can reproduce a complex architecture and
alkaline condition (Almeida et al., 2014; Demirtaş, Irmak, & composition of natural tissue, is still an essential challenge in tissue
Gömöşderelioğlu, 2017; Elviri, Foresti et al., 2017; Wu et al., 2017), engineering (Ning & Chen, 2017; Novosel, Kleinhans, & Kluger, 2011;
derivatization, crosslinking, or a combination with other polymers, are Ventola, 2014). To date, proof-of-concept evaluations towards the 3D
required prior to a cell seeding or implantation (Chavanne et al., 2013; printing of natural-derived biopolymers for a particular biomedical
Liu, Chang, & Lin, 2015; Pandey, Singh, Momin, & Bhavsar, 2017; Yang application have been successfully demonstrated in a variety of mate-
et al., 2016). For instance, hydroxyapatite, pectin, and genipin have rial forms, including beads, filaments, fibers, channels, sheets, rolls,
been incorporated into chitosan for mechanical reinforcement of grids, and porous 3D constructs that mimic the tissue components.
chiton-based inks or 3D printed scaffolds in bone tissue engineering However, these printed structures are often avascular (lacks blood
(Ang et al., 2002; Chavanne et al., 2013; Demirtaş et al., 2017; Liu, vessels), aneural (no neurons and nerves), and alymphatic (no lym-
Chang et al., 2015). Combination with other natural or synthetic phatic system). Direct or indirect extrusion-based bio-printing, micro-
polymers, such as alginate, gelatin, fibrinogen, laminin, and thermo- fluidic and bioreactor techniques, have been developed to fabricate
plastics, would potentially offers many opportunities to overcome these micro or macro-channels to form vasculatures. However, there is still a
challenges, and may allow to introduce some desired properties (Li great deal of work to cope with regarding this critical limitation before
et al., 2009; Yan et al., 2005; Yang et al., 2016; Zhu, Li, Guan, Schreyer, printing a functional tissue (Ning & Chen, 2017; Novosel et al., 2011;
& Chen, 2010). Ventola, 2014; Zhang, Fisher et al., 2015). Another major challenge is
the limited capabilities of current 3D printers to process natural-derived
4. Current challenges and Strategical Roadmap biopolymers for various biomedical applications. Processing speed and
resolution still lag behind the designed optimal levels in many cases.
Despite the remarkable achievements having been reported for a For instance, angiogenesis guiding in an expected path requires precise
variety of 3D printing technologies in the past decade, which have also deposition of channels in designated regions for further vascularization
expanded their advances or progresses in many types of material ex- in bioprinting for tissue engineering application. Thus, 3D printers
plorations and improvements, the feedstock advances, no doubt, remain capable of precise printing of multiple biopolymers and cell types with
as a key and unique foundation for a large-scale application with 3D high resolution, processing speed, and more precise robotic systems are
printing technologies in many promising areas. As a matter of fact, truly needed (Guvendiren et al., 2016).
some challenges with respect to the material printability, functionality, 3D food printing shows potential for customized food designs,
and safety (biosafety and environmental safety) still need to be fully personalized nutrition, simplifying food supply chain, and broadening
addressed both including therapeutic and nontherapeutic applications. of the available food materials but also poses challenges for the pro-
Particularly, it would be noteworthy that our attention during devel- cessing of natural-derived biopolymers as 3D printing feedstock (Lin,
oping biopolymers as 3D printing feedstocks should be firstly put on 2015). Natural-derived biopolymers, such as starch, cellulose, hemi-
those challenges regarding their specific requirements for a 3D printing cellulose, alginate, agarose, and chitosan, can be potentially employed
method as well as a targeted application. Therefore, in future, more as an ink formulation for 3D food printing. However, there is still a long
efforts are truly required to make by following a biopolymer-based inks way for 3D food printing using biopolymers from the lab to kitchen.
development strategical roadmap, as shown in Fig. 7, in order to meet Limited available printable biopolymers that are compatible with tra-
numerous interdependent and technical requirements of different target ditional food ingredients and conventional manufacturing processes
application, including those that lead to an optimal printing, superior results in the high cost and infancy of the 3D food printing. Therefore,
structural and functional outcomes by tuning the chemical, physical, 3D food printing is currently mainly applied for customized food,
and biological properties of those potential and promising natural-de- military and space food, elderly food, confectionery or sweets, instead
rived biopolymers (Jakus, Rutz, & Shah, 2016). of for our daily food due to a high cost. For instance, such technology is
Challenges in the 3D printing of biopolymers applied in medical currently being considered by NASA to be used for printing of instant
area remain a large requirement to improve their material defects and food in spacecraft for longer duration space missions. However, with
properties, i.e. printability, mechanical properties, biocompatibility, key problems solved, 3D printed food will have the potential impact of
tissue biomimicry, degradation kinetics, and safe degradation by- reducing global resource and providing alternative food source. The
products (Fig. 2A). The inherent physic-chemical properties of some main challenges of 3D food printing using biopolymers including: 1)
biopolymers, e.g. infusibility and insolubility in common solvents, strict food safety safeguards of the biopolymers from difference sources;
certainly pose the intractable challenges in the 3D printing process. To 2) appropriate rheological properties of biopolymers for processing and
address these issues, adapting or modifying commercial printers, printing and self-supporting; 3) acceptable printing precision and ac-
modifying natural-based biopolymers, developing novel solvents sys- curacy; 4) manipulation of multiply biopolymers for personalized nu-
tems, incorporating other functional or bioactive polymers, as well as trition or flavors (Godoi et al., 2016; Liu, Zhang, Bhandari, Wang,
advancing different post-processing technologies (such as surface 2017).
coating, and plasma radiation), should be put on our research agenda 3D printing of apparel and textile products using bio-based mate-
with a strategical roadmap identified for their relative importance and rials holds advantages of quick product development, and personalized
interdependence (Fig. 7) (Chia & Wu, 2015; Melchels et al., 2012). design in a sustainable and renewable way, but also pose challenges to
Besides, blending different types of biopolymers aiming for a compli- designing, large scale production, post-processing, and most im-
mentary advantage is another alternative approach or strategy to for- portantly the safety concern. As for the apparel and textile products, 3D
mulate composite inks with favorable processability, printability, me- printing shows high potential in an efficient and sustainable supply
chanics, and bioactivities. However, biodegradation discrepant of chain through areas including reduced cost, lead time in production,

311
J. Liu et al. Carbohydrate Polymers 207 (2019) 297–316

Fig. 7. Strategical roadmap of the development of natural-derived biopolymers for 3D printing.

312
J. Liu et al. Carbohydrate Polymers 207 (2019) 297–316

and increased design possibility and efficiency. However, challenges their functionalized products as feedstocks for different 3D printing
regarding less flexibility, insufficient mechanical properties, poor technologies will be a future trend of 3D printing materials develop-
washability, and poor integration of printed objects and fabrics prohi- ment. Technical advances and the associated feasibility in developing
biting the technology from being a usual process in textile and clothing natural-derived biopolymers as feedstock formulations for a variety of
production (Martens & Ehrmann, 2017). Apparel and textile products 3D printing technologies have been confirmed and demonstrated in
require much more flexibility, resilience, and sufficient tensile strength recent years, at least at a scale of proof-of-concept level. However, there
in accommodating to wearer’s comfort and body movement, which are is still a great deal of work waiting for us to cope with, such as in the
unfortunately hard to meet these performance requirements due to the material processability, degradation, chemical, biological, and me-
limited available printable polymers. Commonly used polymers, e.g. chanical properties before achieving a full potential of biopolymers.
PLA, nylon, or ABS, turns out to be too brittle for the desired fine Once overcoming these challenges, 3D printing technology will no
structures or too hard to be used in typical textile applications such as doubt facilitate a rapid growing for the natural-derived biopolymer
garments. Combination of different types of fabrics and 3D printing applications in a variety of areas. Therefore, it can be predicted that the
could be a good solution to develop innovative smart textiles by de- future materials development in 3D printing will be gradually turned to
position of functional polymers or intelligent materials onto textiles be a utilization of those natural-derived “green” feedstocks.
without compromising on quality and flexibility of the fabric (Gao
et al., 2017; Hashemi Sanatgar, Campagne et al., 2017; Hashemi Acknowledgement
Sanatgar, Cayla et al., 2017).
In summary, owning to the high abundance, low-cost, biocompat- This work was supported by the National Natural Science
ibility and biodegradability, utilization of natural-derived biopolymers, Foundation of China (31772529) and the Start-up Fund for Introduced
such as the lignocellulosic biomass, starch, algae, and chitosan-based Scholar of Jiangsu University (4111370004).
materials, either in native or in a functionalized product as feedstocks
for different 3D printing technologies offers a great deal of their po- References
tentials and will in turn promotes an academic and industrial advance
in 3D printing materials development (Park et al., 2017). 3D printing of Abhay, P., Poldervaart, M. T., Wang, H., van der Stok, J., Weinans, H., Leeuwenburgh, S.
natural-derived biopolymers in biomedical and pharmaceutical appli- C. G., ... Alblas, J. (2013). Sustained release of BMP-2 in bioprinted alginate for os-
teogenicity in mice and rats. PLoS One, 8(8), e72610.
cations is considered as one of the most promising areas for its growth. Alec (2016). Cellulose could act as a building block for 3D printed food. Accessed http://
As a matter of fact, these potential applications mainly include, but not www.3ders.org/articles/20160415-cellulose-could-act-as-a-building-block-for-3d-
limit to: 1) Bioprinting of vascularized and metabolically active tissues printed-food.html.
Ali, H. (2015). Cellulose turning into a supermaterial of the future: Broad-based cooperation
that can reproduce a complex architecture with a unique composition of multiplying the value of Finnish wood. Accessed 12 October 2017 http://www.
natural tissues (such as, kidney, heart, lung, and liver tissues); 2) in situ vttresearch.com/media/news/cellulose-turning-into-a-supermaterial-of-the-future.
bioprinting of implants or living organs in the human body during Almeida, C. R., Serra, T., Oliveira, M. I., Planell, J. A., Barbosa, M. A., & Navarro, M.
(2014). Impact of 3-D printed PLA- and chitosan-based scaffolds on human mono-
operations; 3) Personalized medicines with tailored dose, dose combi-
cyte/macrophage responses: Unraveling the effect of 3-D structures on inflammation.
nation or even the active itself (Ventola, 2014). Acta Biomaterialia, 10(2), 613–622.
Future research interests to expand the limited variety of available Ang, T. H., Sultana, F. S. A., Hutmacher, D. W., Wong, Y. S., Fuh, J. Y. H., Mo, X. M., ...
Teoh, S. H. (2002). Fabrication of 3D chitosan–hydroxyapatite scaffolds using a ro-
natural-derived biopolymers as the novel 3D printing feedstock candi-
botic dispensing system. Materials Science and Engineering: C, 20(1–2), 35–42.
dates may include: 1) Development of novel solvent systems that are Armstrong, J. P. K., Burke, M., Carter, B. M., Davis, S. A., & Perriman, A. W. (2016). 3D
cells-friendly and are capable of dissolving or dispersing natural-de- bioprinting using a templated porous bioink. Advanced Healthcare Materials, 5(14),
rived biopolymers as inks formulations for different 3D printing; 2) 1724–1730.
Axpe, E., & Oyen, M. (2016). Applications of alginate-based bioinks in 3D bioprinting.
Manipulation of multiple biopolymers and incorporation of different International Journal of Molecular Sciences, 17(12), 1976.
biomolecules as composite inks with user-defined and tunable proper- Bakarich, S. E., Panhuis, M. i. h., Beirne, S., Wallace, G. G., & Spinks, G. M. (2013).
ties, including desirable processability, printability, mechanics, bioac- Extrusion printing of ionic–covalent entanglement hydrogels with high toughness.
Journal of Materials Chemistry B, 1(38), 4939.
tivity, and biodegradability; 3) Modification or functionalization of Beecroft, M. (2016). 3D printing of weft knitted textile based structures by selective laser
natural-derived biopolymers or the printed objects via biological, che- sintering of nylon powder. IOP Conference Series: Materials Science and Engineering,
mical, and physical approach to address the printability of the inks, or 137, 012017.
BeMiller, J. N., & Whistler, R. L. (2009). Starch: Chemistry and technology (3rd ed.). Boston:
to offer desired properties; 4) Incorporation of chemical or physical Academic.
stimuli responsive components into natural-derived biopolymers, Bendtsen, S. T., Quinnell, S. P., & Wei, M. (2017). Development of a novel alginate-
making the 3D printed objects programmable to external stimuli such polyvinyl alcohol-hydroxyapatite hydrogel for 3D bioprinting bone tissue engineered
scaffolds. Journal of Biomedical Materials Research Part A, 105(5), 1457–1468.
as pH, temperature, light, electric, and magnetic field, potentially Bhatia, S. K., & Ramadurai, K. W. (2017). 3D printing and bio-based materials in global
bringing forward current 3D printing to future 4D printing (Gao et al., health. Switzerland: Springer International Publishing AG.
2016; Wang, Lee, & Yeong, 2015). Boontheekul, T., Kong, H.-J., & Mooney, D. J. (2005). Controlling alginate gel degrada-
tion utilizing partial oxidation and bimodal molecular weight distribution.
Biomaterials, 26(15), 2455–2465.
5. Conclusions and future perspectives Chavanne, P., Stevanovic, S., Wüthrich, A., Braissant, O., Pieles, U., Gruner, P., ...
Schumacher, R. (2013). 3D printed chitosan/hydroxyapatite scaffolds for potential
3D printing has emerged as a revolutionary technology, promising use in regenerative medicine. Biomedical Engineering/Biomedizinische Technik, 58
(Suppl. 1).
to transform the conventional products fabrication. However, limited Chia, H. N., & Wu, B. M. (2015). Recent advances in 3D printing of biomaterials. Journal
environmentally-friendly printing materials with a high quality of of Biological Engineering, 9(1).
printing performance reamins a major issue need to be overcome before Chimene, D., Lennox, K. K., Kaunas, R. R., & Gaharwar, A. K. (2016). Advanced bioinks
for 3D printing: A materials science perspective. Annals of Biomedical Engineering,
it can be widely adapted for products fabrication in different areas. 44(6), 2090–2102.
Natural-derived biopolymers, including lignocellulosic materials, sea- Chung, J. H. Y., Naficy, S., Yue, Z., Kapsa, R., Quigley, A., Moulton, S. E., ... Wallace, G. G.
weed materials, and materials from crustaceans exoskeletons (e.g. crab, (2013). Bio-ink properties and printability for extrusion printing living cells.
Biomaterials Science, 1(7), 763.
shrimp shells, and insects), represent the most abundant biobased and Daly, A. C., Cunniffe, G. M., Sathy, B. N., Jeon, O., Alsberg, E., & Kelly, D. J. (2016). 3D
renewable feedstocks for different 3D printing technologies. bioprinting of developmentally inspired templates for whole bone organ engineering.
Modification or combination with other ingredients to introduce or Advanced Healthcare Materials, 5(18), 2353–2362.
Davachi, S. M., & Kaffashi, B. (2015). Polylactic acid in medicine. Polymer-Plastics
improve the materials processability and products functionality will Technology and Engineering, 54(9), 944–967.
indeed broaden the application areas of the natural-derived biopoly- Demirtaş, T. T., Irmak, G., & Gümüşderelioğlu, M. (2017). A bioprintable form of chitosan
mers. Utilization of natural-derived biopolymers either in native or hydrogel for bone tissue engineering. Biofabrication, 9(3), 035003.

313
J. Liu et al. Carbohydrate Polymers 207 (2019) 297–316

Derby, B. (2012). Printing and prototyping of tissues and scaffolds. Science, 338(6109), 2017 http://cellulosefromfinland.fi/design-driven-value-chains-in-the-world-of-
921–926. cellulose/.
Diogo, G. S., Gaspar, V. M., Serra, I. R., Fradique, R., & Correia, I. J. (2014). Manufacture Kesti, M., Müller, M., Becher, J., Schnabelrauch, M., D’Este, M., Eglin, D., ... Zenobi-
of β-TCP/alginate scaffolds through a Fab@home model for application in bone Wong, M. (2015). A versatile bioink for three-dimensional printing of cellular scaf-
tissue engineering. Biofabrication, 6(2), 025001. folds based on thermally and photo-triggered tandem gelation. Acta Biomaterialia, 11,
Duan, B., Hockaday, L. A., Kang, K. H., & Butcher, J. T. (2013). 3D bioprinting of het- 162–172.
erogeneous aortic valve conduits with alginate/gelatin hydrogels. Journal of Kong, H. J., Kaigler, D., Kim, K., & Mooney, D. J. (2004). Controlling rigidity and de-
Biomedical Materials Research Part A, 101A(5), 1255–1264. gradation of alginate hydrogels via molecular weight distribution. Biomacromolecules,
Duan, B., Kapetanovic, E., Hockaday, L. A., & Butcher, J. T. (2014). Three-dimensional 5(5), 1720–1727.
printed trileaflet valve conduits using biological hydrogels and human valve inter- Köpf, M., Campos, D. F. D., Blaeser, A., Sen, K. S., & Fischer, H. (2016). A tailored three-
stitial cells. Acta Biomaterialia, 10(5), 1836–1846. dimensionally printable agarose–collagen blend allows encapsulation, spreading, and
Duarte Campos, D. F., Blaeser, A., Weber, M., Jäkel, J., Neuss, S., Jahnen-Dechent, W., ... attachment of human umbilical artery smooth muscle cells. Biofabrication, 8(2),
Fischer, H. (2013). Three-dimensional printing of stem cell-laden hydrogels sub- 025011.
merged in a hydrophobic high-density fluid. Biofabrication, 5(1), 015003. Korger, M., Bergschneider, J., Lutz, M., Mahltig, B., Finsterbusch, K., & Rabe, M. (2016).
Dutta, P. K. (2016). Chitin and chitosan for regenerative medicine. India: Springer. Possible applications of 3D printing technology on textile substrates. IOP Conference
Elviri, L., Bianchera, A., Bergonzi, C., & Bettini, R. (2017). Controlled local drug delivery Series: Materials Science and Engineering, 141, 012011.
strategies from chitosan hydrogels for wound healing. Expert Opinion on Drug Delivery, Kumar, M. N. V. R., Muzzarelli, R. A. A., Muzzarelli, C., Sashiwa, H., & Domb, A. J.
14(7), 897–908. (2004). Chitosan chemistry and pharmaceutical perspectives. Chemical Reviews,
Elviri, L., Foresti, R., Bergonzi, C., Zimetti, F., Marchi, C., Bianchera, A., & Bettini, R. 104(12), 6017–6084.
(2017). Highly defined 3D printed chitosan scaffolds featuring improved cell growth. Kundu, J., Shim, J.-H., Jang, J., Kim, S.-W., & Cho, D.-W. (2015). An additive manu-
Biomedical Materials, 12(4), 045009. facturing-based PCL-alginate-chondrocyte bioprinted scaffold for cartilage tissue
Fernandes, E. M., Pires, R. A., Mano, J. F., & Reis, R. L. (2013). Bionanocomposites from engineering. Journal of Tissue Engineering and Regenerative Medicine, 9(11),
lignocellulosic resources: Properties, applications and future trends for their use in 1286–1297.
the biomedical field. Progress in Polymer Science, 38(10–11), 1415–1441. Kuo, C. C., Liu, L. C., Teng, W. F., Chang, H. Y., Chien, F. M., Liao, S. J., ... Chen, C. M.
Ferris, C. J., Gilmore, K. G., Wallace, G. G., & in het Panhuis, M. (2013). Biofabrication: an (2016). Preparation of starch/acrylonitrile-butadiene-styrene copolymers (ABS)
overview of the approaches used for printing of living cells. Applied Microbiology and biomass alloys and their feasible evaluation for 3D printing applications. Composites
Biotechnology, 97(10), 4243–4258. Part B-Engineering, 86, 36–39.
Fraser-Reid, B. O., Tatsuta, K., & Thiem, J. (2008). Glycoscience: Chemistry and chemical Kyle, S., Jessop, Z. M., Al-Sabah, A., & Whitaker, I. S. (2017). ‘Printability’ of candidate
biology (2nd ed.). Berlin: Springer. biomaterials for extrusion based 3D printing: State-of-the-art. Advanced Healthcare
Gao, B., Yang, Q., Zhao, X., Jin, G., Ma, Y., & Xu, F. (2016). 4D bioprinting for biomedical Materials, 6(16), 1700264.
applications. Trends in Biotechnology, 34(9), 746–756. Lam, C. X. F., Mo, X. M., Teoh, S. H., & Hutmacher, D. W. (2002). Scaffold development
Gao, T., Yang, Z., Chen, C., Li, Y., Fu, K., Dai, J., & Hu, L. (2017). Three-dimensional using 3D printing with a starch-based polymer. Materials Science and Engineering: C,
printed thermal regulation textiles. ACS Nano, 11(11), 11513–11520. 20(1–2), 49–56.
Gao, W., Zhang, Y., Ramanujan, D., Ramani, K., Chen, Y., Williams, C. B., & Zavattieri, P. Landers, R., Hubner, U., Schmelzeisen, R., & Mulhaupt, R. (2002). Rapid prototyping of
D. (2015). The status, challenges, and future of additive manufacturing in en- scaffolds derived from thermoreversible hydrogels and tailored for applications in
gineering. Computer-Aided Design, 69, 65–89. tissue engineering. Biomaterials, 23(23), 4437–4447.
Gatenholm, P., Martinez, H., Karabulut, E., Amoroso, M., Kölby, L., Markstedt, K., & Landers, R., Pfister, A., Hübner, U., John, H., Schmelzeisen, R., & Mülhaupt, R. (2002).
Henriksson, I. (2016). Development of nanocellulose-based bioinks for 3D bioprinting of Fabrication of soft tissue engineering scaffolds by means of rapid prototyping tech-
soft tissue. 3D printing and biofabrication1–23. niques. Journal of Materials Science, 37(15), 3107–3116.
Godoi, F. C., Prakash, S., & Bhandari, B. R. (2016). 3d printing technologies applied for Lee, J.-S., Hong, J. M., Jung, J. W., Shim, J.-H., Oh, J.-H., & Cho, D.-W. (2014). 3D
food design: Status and prospects. Journal of Food Engineering, 179, 44–54. printing of composite tissue with complex shape applied to ear regeneration.
Grigore, A., Sarker, B., Fabry, B., Boccaccini, A. R., & Detsch, R. (2014). Behavior of Biofabrication, 6(2), 024103.
encapsulated MG-63 cells in RGD and gelatine-modified alginate hydrogels. Tissue Lee, K. Y., & Mooney, D. J. (2012). Alginate: Properties and biomedical applications.
Engineering Part A, 20(15–16), 2140–2150. Progress in Polymer Science, 37(1), 106–126.
Grimmelsmann, N., Meissner, H., & Ehrmann, A. (2016). 3D printed auxetic forms on Leppiniemi, J., Lahtinen, P., Paajanen, A., Mahlberg, R., Metsä-Kortelainen, S., Pinomaa,
knitted fabrics for adjustable permeability and mechanical properties. IOP Conference T., & Hytönen, V. P. (2017). 3D-printable bioactivated nanocellulose–alginate hy-
Series: Materials Science and Engineering, 137, 012011. drogels. ACS Applied Materials & Interfaces, 9(26), 21959–21970.
Gudapati, H., Yan, J., Huang, Y., & Chrisey, D. B. (2014). Alginate gelation-induced cell Li, S., Xiong, Z., Wang, X., Yan, Y., Liu, H., & Zhang, R. (2009). Direct fabrication of a
death during laser-assisted cell printing. Biofabrication, 6(3), 035022. hybrid cell/hydrogel construct by a double-nozzle assembling technology. Journal of
Gupta, K. M., & Jiang, J. (2015). Cellulose dissolution and regeneration in ionic liquids: A Bioactive and Compatible Polymers, 24(3), 249–265.
computational perspective. Chemical Engineering Science, 121, 180–189. Li, X., Cui, R., Sun, L., Aifantis, K. E., Fan, Y., Feng, Q., ... Watari, F. (2014). 3D-printed
Guvendiren, M., Molde, J., Soares, R. M. D., & Kohn, J. (2016). Designing biomaterials for biopolymers for tissue engineering application. International Journal of Polymer
3D printing. ACS Biomaterials Science & Engineering, 2(10), 1679–1693. Science, 2014, 1–13.
Håkansson, K. M. O., Henriksson, I. C., de la Peña Vázquez, C., Kuzmenko, V., Markstedt, Liberski, A. R. (2016). Three-dimensional printing of alginate: From seaweeds to heart
K., Enoksson, P., ... Gatenholm, P. (2016). Solidification of 3D printed nanofibril valve scaffolds. QScience Connect, 2016(2), 3.
hydrogels into functional 3D cellulose structures. Advanced Materials & Technologies, Lille, M., Nurmela, A., Nordlund, E., Metsä-Kortelainen, S., & Sozer, N. (2018).
1(7), 1600096. Applicability of protein and fiber-rich food materials in extrusion-based 3D printing.
Hashemi Sanatgar, R., Campagne, C., & Nierstrasz, V. (2017). Investigation of the ad- Journal of Food Engineering, 220, 20–27.
hesion properties of direct 3D printing of polymers and nanocomposites on textiles: Lin, C.-C., & Fu, S.-J. (2016). Osteogenesis of human adipose-derived stem cells on poly
Effect of FDM printing process parameters. Applied Surface Science, 403, 551–563. (dopamine)-coated electrospun poly(lactic acid) fiber mats. Materials Science and
Hashemi Sanatgar, R., Cayla, A., Campagne, C., & Nierstrasz, V. (2017). Manufacturing of Engineering C, 58, 254–263.
polylactic acid nanocomposite 3D printer filaments for smart textile applications. IOP Lin, C. (2015). 3D food printing: A taste of the future. Journal of Food Science Education,
Conference Series: Materials Science and Engineering, 254, 072011. 14(3), 86–87.
Henriksson, I., Gatenholm, P., & Hagg, D. A. (2017). Increased lipid accumulation and Liu, I. H., Chang, S.-H., & Lin, H.-Y. (2015). Chitosan-based hydrogel tissue scaffolds
adipogenic gene expression of adipocytes in 3D bioprinted nanocellulose scaffolds. made by 3D plotting promotes osteoblast proliferation and mineralization. Biomedical
Biofabrication, 9(1), 015022. Materials, 10(3), 035004.
Hong, S., Sycks, D., Chan, H. F., Lin, S., Lopez, G. P., Guilak, F., & Zhao, X. (2015). 3D Liu, J., Cheng, F., Grenman, H., Spoljaric, S., Seppälä, J., Eriksson, J. E., & Xu, C. L.
printing of highly stretchable and tough hydrogels into complex, cellularized struc- (2016). Development of nanocellulose scaffolds with tunable structures to support 3D
tures. Advanced Materials, 27(27), 4035–4040. cell culture. Carbohydrate Polymers, 148, 259–271.
Huang, Y., Leu, M. C., Mazumder, J., & Donmez, A. (2014). Additive manufacturing: Liu, J., Chinga-Carrasco, G., Cheng, F., Xu, W., Willför, S., Syverud, K., ... Xu, C. (2016).
Current state, future potential, gaps and needs, and recommendations. Journal of Hemicellulose-reinforced nanocellulose hydrogels for wound healing application.
Manufacturing Science and Engineering, 137(1), 014001. Cellulose, 23(5), 3129–3143.
Jakus, A. E., Rutz, A. L., & Shah, R. N. (2016). Advancing the field of 3D biomaterial Liu, J., Korpinen, R., Mikkonen, K., Willför, S., & Xu, C. (2014). Nanofibrillated cellulose
printing. Biomedical Materials (Bristol, England), 11(1), 014102. originated from birch sawdust after sequential extractions: A promising polymeric
Jia, J., Richards, D. J., Pollard, S., Tan, Y., Rodriguez, J., Visconti, R. P., & Mei, Y. (2014). material from waste to films. Cellulose, 21(4), 2587–2598.
Engineering alginate as bioink for bioprinting. Acta Biomaterialia, 10(10), Liu, J., Willför, S., & Mihranyan, A. (2017). On importance of impurities, potential
4323–4331. leachables and extractables in algal nanocellulose for biomedical use. Carbohydrate
Jiang, W., Shi, J., Li, W., & Sun, K. (2012). Morphology, wettability, and mechanical Polymers, 172, 11–19.
properties of polycaprolactone/hydroxyapatite composite scaffolds with inter- Liu, J., Willför, S., & Xu, C. (2015). A review of bioactive plant polysaccharides: Biological
connected pore structures fabricated by a mini-deposition system. Polymer Engineering activities, functionalization, and biomedical applications. Bioactive Carbohydrates and
and Science, 52(11), 2396–2402. Dietary Fibre, 5(1), 31–61.
Kalaskar, D. M. (2017). 3D printing in medicine. United Kingdom: Woodhead Publishing. Liu, Z., Zhang, M., Bhandari, B., & Wang, Y. (2017). 3D printing: Printing precision and
Kao, C.-T., Lin, C.-C., Chen, Y.-W., Yeh, C.-H., Fang, H.-Y., & Shie, M.-Y. (2015). Poly application in food sector. Trends in Food Science & Technology, 69, 83–94.
(dopamine) coating of 3D printed poly(lactic acid) scaffolds for bone tissue en- Liu, Z., Zhang, M., Bhandari, B., & Yang, C. (2017). Impact of rheological properties of
gineering. Materials Science and Engineering: C, 56, 165–173. mashed potatoes on 3D printing. Journal of Food Engineering, 220, 76–82.
Kataja, K. (2015). Design driven value chains in the world of cellulose (DWoC). Accessed 12. Luo, Y., Zhai, D., Huan, Z., Zhu, H., Xia, L., Chang, J., ... Wu, C. (2015). Three-

314
J. Liu et al. Carbohydrate Polymers 207 (2019) 297–316

dimensional printing of hollow-struts-packed bioceramic scaffolds for bone re- Saini, P., Arora, M., & Kumar, M. N. V. R. (2016). Poly(lactic acid) blends in biomedical
generation. ACS Applied Materials & Interfaces, 7(43), 24377–24383. applications. Advanced Drug Delivery Reviews, 107, 47–59.
Maher, P. S., Keatch, R. P., Donnelly, K., Mackay, R. E., & Paxton, J. Z. (2009). Schutyser, M. A. I., Houlder, S., de Wit, M., Buijsse, C. A. P., & Alting, A. C. (2018). Fused
Construction of 3D biological matrices using rapid prototyping technology. Rapid deposition modelling of sodium caseinate dispersions. Journal of Food Engineering,
Prototyping Journal, 15(3), 204–210. 220, 49–55.
Malda, J., Visser, J., Melchels, F. P., Jüngst, T., Hennink, W. E., Dhert, W. J. A., & Schütz, K., Placht, A.-M., Paul, B., Brüggemeier, S., Gelinsky, M., & Lode, A. (2017).
Hutmacher, D. W. (2013). 25th anniversary article: Engineering hydrogels for bio- Three-dimensional plotting of a cell-laden alginate/methylcellulose blend: towards
fabrication. Advanced Materials, 25(36), 5011–5028. biofabrication of tissue engineering constructs with clinically relevant dimensions.
Manavitehrani, I., Fathi, A., Badr, H., Daly, S., Negahi Shirazi, A., & Dehghani, F. (2016). Journal of Tissue Engineering and Regenerative Medicine, 11(5), 1574–1587.
Biomedical applications of biodegradable polyesters. Polymers, 8(1), 20. Senatov, F. S., Niaza, K. V., Zadorozhnyy, M. Y., Maksimkin, A. V., Kaloshkin, S. D., &
Mandrycky, C., Wang, Z., Kim, K., & Kim, D.-H. (2016). 3D bioprinting for engineering Estrin, Y. Z. (2016). Mechanical properties and shape memory effect of 3D-printed
complex tissues. Biotechnology Advances, 34(4), 422–434. PLA-based porous scaffolds. Journal of the Mechanical Behavior of Biomedical Materials,
Markstedt, K., Escalante, A., Toriz, G., & Gatenholm, P. (2017). Biomimetic inks based on 57, 139–148.
cellulose nanofibrils and cross-linkable xylans for 3D printing. ACS Applied Materials Serra, T., Mateos-Timoneda, M. A., Planell, J. A., & Navarro, M. (2013). 3D printed PLA-
& Interfaces, 9(46), 40878–40886. based scaffolds: A versatile tool in regenerative medicine. Organogenesis, 9(4),
Markstedt, K., Mantas, A., Tournier, I., Martinez Avila, H., Hagg, D., & Gatenholm, P. 239–244.
(2015). 3D bioprinting human chondrocytes with nanocellulose-alginate bioink for Serra, T., Planell, J. A., & Navarro, M. (2013). High-resolution PLA-based composite
cartilage tissue engineering applications. Biomacromolecules, 16(5), 1489–1496. scaffolds via 3-D printing technology. Acta Biomaterialia, 9(3), 5521–5530.
Markstedt, K., Sundberg, J., & Gatenholm, P. (2014). 3D bioprinting of cellulose struc- Serwer, P. (1983). Agarose gels: Properties and use for electrophoresis. Electrophoresis,
tures from an ionic liquid. 3D Printing and Additive Manufacturing, 1(3), 115–121. 4(6), 375–382.
Martens, Y., & Ehrmann, A. (2017). Composites of 3D-printed polymers and textile fab- Shah Mohammadi, M., Bureau, M. N., & Nazhat, S. N. (2014). Polylactic acid (PLA) bio-
rics. IOP Conference Series: Materials Science and Engineering, 225, 012292. medical foams for tissue engineering. 313–334.
Martínez Ávila, H., Schwarz, S., Rotter, N., & Gatenholm, P. (2016). 3D bioprinting of Shim, J.-H., Lee, J.-S., Kim, J. Y., & Cho, D.-W. (2012). Bioprinting of a mechanically
human chondrocyte-laden nanocellulose hydrogels for patient-specific auricular enhanced three-dimensional dual cell-laden construct for osteochondral tissue en-
cartilage regeneration. Bioprinting, 1–2, 22–35. gineering using a multi-head tissue/organ building system. Journal of Micromechanics
Melchels, F. P. W., Domingos, M. A. N., Klein, T. J., Malda, J., Bartolo, P. J., & Hutmacher, and Microengineering, 22(8), 085014.
D. W. (2012). Additive manufacturing of tissues and organs. Progress in Polymer Shishkovsky, I. V. (2016). New trends in 3D printing. Russian: InTech.
Science, 37(8), 1079–1104. Siqueira, G., Kokkinis, D., Libanori, R., Hausmann, M. K., Gladman, A. S., Neels, A., ...
Modjarrad, K., & Ebnesajjad, S. (2014). Handbook of polymer applications in medicine and Studart, A. R. (2017). Cellulose nanocrystal inks for 3D printing of textured cellular
medical devices. Oxford; San Diego, CA: William Andrew. architectures. Advanced Functional Materials, 27(12), 1604619.
Möller, T., Amoroso, M., Hägg, D., Brantsing, C., Rotter, N., Apelgren, P., & Gatenholm, P. Stephen, A. M., Phillips, G. O., & Williams, P. A. (2006). Food polysaccharides and their
(2017). In vivo chondrogenesis in 3D bioprinted human cell-laden hydrogel con- applications (2nd ed.). Boca Raton, FL: CRC/Taylor & Francis.
structs. Plastic and Reconstructive Surgery – Global Open, 5(2), e1227. Suganuma, J., & Alexander, H. (1993). Biological response of intramedullary bone to
Morris, V. B., Nimbalkar, S., Younesi, M., McClellan, P., & Akkus, O. (2016). Mechanical poly-L-lactic acid. Journal of Applied Biomaterials, 4(1), 13–27.
properties, cytocompatibility and manufacturability of chitosan: PEGDA hybrid-gel Sultan, S., Siqueira, G., Zimmermann, T., & Mathew, A. P. (2017). 3D printing of nano-
scaffolds by stereolithography. Annals of Biomedical Engineering, 45(1), 286–296. cellulosic biomaterials for medical applications. Current Opinion in Biomedical
Müller, M., Öztürk, E., Arlov, Ø., Gatenholm, P., & Zenobi-Wong, M. (2016). Alginate Engineering, 2, 29–34.
sulfate–nanocellulose bioinks for cartilage bioprinting applications. Annals of Sun, J., Peng, Z., Zhou, W., Fuh, J. Y. H., Hong, G. S., & Chiu, A. (2015). A review on 3D
Biomedical Engineering, 45(1), 210–223. printing for customized food fabrication. Procedia Manufacturing, 1, 308–319.
Munaz, A., Vadivelu, R. K., St. John, J., Barton, M., Kamble, H., & Nguyen, N.-T. (2016). Sun, J., & Tan, H. (2013). Alginate-based biomaterials for regenerative medicine appli-
Three-dimensional printing of biological matters. Journal of Science: Advanced cations. Materials, 6(4), 1285–1309.
Materials and Devices, 1(1), 1–17. Sun, L., & Zhao, L. (2017). Envisioning the era of 3D printing: A conceptual model for the
Murphy, C. A., & Collins, M. N. (2016). Microcrystalline cellulose reinforced polylactic fashion industry. Fashion and Textiles, 4(1).
acid biocomposite filaments for 3D printing. Polymer Composites, 39(4), 1311–1320. Sun, L., & Zhao, L. (2018). Technology disruptions: Exploring the changing roles of de-
Murphy, S. V., & Atala, A. (2014). 3D bioprinting of tissues and organs. Nature signers, makers, and users in the fashion industry. International Journal of Fashion
Biotechnology, 32(8), 773–785. Design Technology and Education, 1–13.
Navarro, G., & Garcia, I. (2016). Study of tissue printing parameters for generating Tabriz, A. G., Hermida, M. A., Leslie, N. R., & Shu, W. (2015). Three-dimensional bio-
complex tissue constructs. Journal of Tissue Science & Engineering, 7(2). printing of complex cell laden alginate hydrogel structures. Biofabrication, 7(4),
Nguyen, D., Hägg, D. A., Forsman, A., Ekholm, J., Nimkingratana, P., Brantsing, C., & 045012.
Simonsson, S. (2017). Cartilage tissue engineering by the 3D bioprinting of ips cells in Tan, Y., Richards, D. J., Trusk, T. C., Visconti, R. P., Yost, M. J., Kindy, M. S., ... Mei, Y.
a nanocellulose/alginate bioink. Scientific Reports, 7(1). (2014). 3D printing facilitated scaffold-free tissue unit fabrication. Biofabrication,
Ning, L., & Chen, X. (2017). A brief review of extrusion-based tissue scaffold bio-printing. 6(2), 024111.
Biotechnology Journal, 12(8), 1600671. Tenhunen, T.-M., Moslemian, O., Kammiovirta, K., Harlin, A., Kääriäinen, P., Österberg,
Norotte, C., Marga, F. S., Niklason, L. E., & Forgacs, G. (2009). Scaffold-free vascular M., & Orelma, H. (2018). Surface tailoring and design-driven prototyping of fabrics
tissue engineering using bioprinting. Biomaterials, 30(30), 5910–5917. with 3D-printing: An all-cellulose approach. Materials & Design, 140, 409–419.
Novosel, E. C., Kleinhans, C., & Kluger, P. J. (2011). Vascularization is the key challenge Tyler, B., Gullotti, D., Mangraviti, A., Utsuki, T., & Brem, H. (2016). Polylactic acid (PLA)
in tissue engineering. Advanced Drug Delivery Reviews, 63(4–5), 300–311. controlled delivery carriers for biomedical applications. Advanced Drug Delivery
Ozbolat, I. T., Peng, W., & Ozbolat, V. (2016). Application areas of 3D bioprinting. Drug Reviews, 107, 163–175.
Discovery Today, 21(8), 1257–1271. Ulery, B. D., Nair, L. S., & Laurencin, C. T. (2011). Biomedical applications of biode-
Pandey, A. R., Singh, U. S., Momin, M., & Bhavsar, C. (2017). Chitosan: Application in gradable polymers. Journal of Polymer Science Part B, Polymer Physics, 49(12),
tissue engineering and skin grafting. Journal of Polymer Research, 24(8). 832–864.
Park, A., Ben, W., & Griffith, L. G. (1998). Integration of surface modification and 3D Ursan, I. D., Chiu, L., & Pierce, A. (2013). Three-dimensional drug printing: A structured
fabrication techniques to prepare patterned poly(L-lactide) substrates allowing re- review. Journal of the American Pharmacists Association, 53(2), 136–144.
gionally selective cell adhesion. Journal of Biomaterials Science Polymer Edition, 9(2), Vanderploeg, A., Lee, S.-E., & Mamp, M. (2016). The application of 3D printing tech-
89–110. nology in the fashion industry. International Journal of Fashion Design Technology and
Park, S.-B., Lih, E., Park, K.-S., Joung, Y. K., & Han, D. K. (2017). Biopolymer-based Education, 10(2), 170–179.
functional composites for medical applications. Progress in Polymer Science, 68, Ventola, C. L. (2014). Medical applications for 3D printing: Current and projected uses.
77–105. Pharmacy and Therapeutics, 39(10), 704–711.
Pattinson, S. W., & Hart, A. J. (2017). Additive manufacturing of cellulosic materials with Wang, M., Favi, P., Cheng, X., Golshan, N. H., Ziemer, K. S., Keidar, M., ... Webster, T. J.
robust mechanics and antimicrobial functionality. Advanced Materials & Technologies, (2016). Cold atmospheric plasma (CAP) surface nanomodified 3D printed polylactic
1600084. acid (PLA) scaffolds for bone regeneration. Acta Biomaterialia, 46, 256–265.
Pilla, S. (2011). Handbook of bioplastics & biocomposites engineering applications. USA: Wang, S., Lee, J. M., & Yeong, W. Y. (2015). Smart hydrogels for 3D bioprinting.
Wiley-Scrivener. International Journal of Bioprinting, 1(1), 3–14.
Qi, H. (2017). Novel functional materials based on cellulose. Switzerland: Springer Wei, J., Wang, J., Su, S., Wang, S., Qiu, J., Zhang, Z., & Cong, W. (2015). 3D printing of an
International Publishing AG. extremely tough hydrogel. RSC Advances, 5(99), 81324–81329.
Ratner, B. D., Hoffman, A. S., Schoen, F. J., & Lemons, J. E. (2013). Biomaterials science: Wijk, A., & Wijk, I. (2015). 3D printing with biomaterials-towards a sustainable and circular
An introduction to materials in medicine (3rd ed.). Canada: Academic Press. economy. Amsterdam: IOS Press.
Rees, A., Powell, L. C., Chinga-Carrasco, G., Gethin, D. T., Syverud, K., Hill, K. E., & Wong, K. V., & Hernandez, A. (2012). A review of additive manufacturing. ISRN
Thomas, D. W. (2015). 3D bioprinting of carboxymethylated-periodate oxidized na- Mechanical Engineering, 2012, 1–10.
nocellulose constructs for wound dressing applications. BioMed Research International, Wong, T. Y., Preston, L. A., & Schiller, N. L. (2000). Alginate lyase: Review of major
2015, 925757. sources and enzyme characteristics, structure-function analysis, biological roles, and
Rinaudo, M. (2006). Chitin and chitosan: Properties and applications. Progress in Polymer applications. Annual Review of Microbiology, 54(1), 289–340.
Science, 31(7), 603–632. Wu, C.-S. (2016). Modulation, functionality, and cytocompatibility of three-dimensional
Rutz, A. L., Hyland, K. E., Jakus, A. E., Burghardt, W. R., & Shah, R. N. (2015). A mul- printing materials made from chitosan-based polysaccharide composites. Materials
timaterial bioink method for 3D printing tunable, cell-compatible hydrogels. Science and Engineering: C, 69, 27–36.
Advanced Materials, 27(9), 1607–1614. Wu, Q., Maire, M., Lerouge, S., Therriault, D., & Heuzey, M.-C. (2017). 3D Printing of

315
J. Liu et al. Carbohydrate Polymers 207 (2019) 297–316

microstructured and stretchable chitosan hydrogel for guided cell growth. Advanced efficacy, cytocompatibility and biocompatibility of a 3D-printed osteoconductive
Biosystems, 1(6), 1700058. composite scaffold functionalized with quaternized chitosan. Acta Biomaterialia, 46,
Wu, Z., Su, X., Xu, Y., Kong, B., Sun, W., & Mi, S. (2016). Bioprinting three-dimensional 112–128.
cell-laden tissue constructs with controllable degradation. Scientific Reports, 6, 24474. Zeng, J.-B., Li, K.-A., & Du, A.-K. (2015). Compatibilization strategies in poly(lactic acid)-
Wüst, S., Godla, M. E., Müller, R., & Hofmann, S. (2014). Tunable hydrogel composite based blends. RSC Advances, 5(41), 32546–32565.
with two-step processing in combination with innovative hardware upgrade for cell- Zhang, H., Mao, X., Du, Z., Jiang, W., Han, X., Zhao, D., & Li, Q. (2016). Three dimen-
based three-dimensional bioprinting. Acta Biomaterialia, 10(2), 630–640. sional printed macroporous polylactic acid/hydroxyapatite composite scaffolds for
Xu, W., Pranovich, A., Uppstu, P., Wang, X., Kronlund, D., Hemming, J., & Xu, C. (2018). promoting bone formation in a critical-size rat calvarial defect model. Science and
Novel biorenewable composite of wood polysaccharide and polylactic acid for three Technology of Advanced Materials, 17(1), 136–148.
dimensional printing. Carbohydrate Polymers, 187, 51–58. Zhang, L., Liu, Z., Cui, G., & Chen, L. (2015). Biomass-derived materials for electro-
Yan, Y., Wang, X., Pan, Y., Liu, H., Cheng, J., Xiong, Z., ... Lu, Q. (2005). Fabrication of chemical energy storages. Progress in Polymer Science, 43, 136–164.
viable tissue-engineered constructs with 3D cell-assembly technique. Biomaterials, Zhang, L. G., Fisher, J. P., & Leong, K. (2015). 3D bioprinting and nanotechnology in tissue
26(29), 5864–5871. engineering and regenerative medicine. United States: Academic Press.
Yang, F., Zhang, M., Bhandari, B., & Liu, Y. (2018). Investigation on lemon juice gel as Zhu, N., Li, M. G., Guan, Y. J., Schreyer, D. J., & Chen, X. B. (2010). Effects of laminin
food material for 3D printing and optimization of printing parameters. LWT – Food blended with chitosan on axon guidance on patterned substrates. Biofabrication, 2(4),
Science and Technology, 87, 67–76. 045002.
Yang, Y., Yang, S., Wang, Y., Yu, Z., Ao, H., Zhang, H., & Tang, T. (2016). Anti-infective

316

You might also like