You are on page 1of 16

REVIEW

Biofabrication www.advhealthmat.de

‘Printability’ of Candidate Biomaterials for Extrusion Based


3D Printing: State-of-the-Art
Stuart Kyle, Zita M. Jessop, Ayesha Al-Sabah, and Iain S. Whitaker*

biological and physical sciences, including


Regenerative medicine has been highlighted as one of the UK’s 8 ‘Great Tech- cell biology, molecular biology, mate-
nologies’ with the potential to revolutionize patient care in the 21st Century. rials science, engineering and clinical
Over the last decade, the concept of ‘3D bioprinting’ has emerged, which medicine.[2]
In recent years, the ability to print
allows the precise deposition of cell laden bioinks with the aim of engineering
biological ‘inks’, rather than plastic and
complex, functional tissues. For 3D printing to be used clinically, there is metal inks of traditional 3D printing,
the need to produce advanced functional biomaterials, a new generation of has resulted in the birth of the new bio-
bioinks with suitable cell culture and high shape/print fidelity, to match or printing research field.[3] The global 3D
exceed the physical, chemical and biological properties of human tissue. With bioprinting market was estimated to be
$487 million in 2014 and is predicted
the rapid increase in knowledge associated with biomaterials, cell-scaffold
to reach $1.82 billion by 2022 (Grand
interactions and the ability to biofunctionalize/decorate bioinks with cell rec- View, 2015). 3D bioprinting allows the
ognition sequences, it is important to keep in mind the ‘printability’ of these potential to replicate complex native-like
novel materials. In this illustrated review, we define and refine the concept tissue architecture in the laboratory with
of ‘printability’ and review seminal and contemporary studies to highlight the potential to biomanufacture physio­
the current ‘state of play’ in the field with a focus on bioink composition and logically relevant multicellular tissue con-
structs on demand. The ultimate success
concentration, manipulation of nozzle parameters and rheological properties.
of this platform technology depends on
the process itself, plus answers the funda-
mental scientific questions regarding the
1. Introduction most appropriate blend of tissue-specific cell source, suitable
scaffold and ideal microenvironment.[4]
Congenital or acquired tissue loss and dysfunction is an enor- The ultimate goal of 3D bioprinting is to produce functional
mous socioeconomic burden for healthcare systems globally. biomimetic composite tissues. A key tenet of this paradigm
In the Western World, hundreds of thousands of people are is that cell biology in 2D and 3D culture are significantly dif-
on transplant waiting lists, and although hundreds of millions ferent and in order to fabricate a successful tissue engineering
are registered as organ donors there is simply not an adequate substrate, 3D constructs must best match native tissue in vitro.
supply (www.unos.org; www.organdonation.nhs.uk). Globally, it Contemporary high resolution 3D printers are able to do this by
is impossible to estimate the number of individuals who would controlling the nano-, micro- and macrostructure.[5,6] Current
benefit from functional tissue engineering to replace and bioprinting technology enables cells and scaffolds to be printed
restore function. simultaneously, allowing accurate control of cell distribution,
Regenerative medicine has the potential to revolutionize high resolution, scalability and cost-effectiveness.[7]
patient care in the 21st Century.[1] One of its major branches, Several bioprinting strategies have been explored with
tissue engineering, holds promise to repair or replace human the ultimate goal of fabricating functional tissue constructs
tissues and organs in order to restore normal function. Tissue (Table 1). As extrusion based bioprinting is the most widely
engineering is a multidisciplinary field that involves aspects of used in current research and the commercial sphere, this article
focuses on the applicability of contemporary biomaterials with
this strategy in mind.
Dr. S. Kyle, Ms. Z. M. Jessop, Dr. A. Al-Sabah, Prof. I. S. Whitaker One of the current key research questions is what the
Reconstructive Surgery & Regenerative Medicine Group (ReconRegen)
ideal properties of a ‘bioink’ actually are. The characteristics
Institute of Life Sciences
Swansea University Medical School of an ideal bioink can be seen in Table 2. It is essential that
Swansea SA2 8PP, UK suitable bioinks have the desired functional and mechanical
E-mail: iainwhitaker@fastmail.fm properties that closely match the tissue it is to replace. A suc-
Dr. S. Kyle, Ms. Z. M. Jessop, Prof. I. S. Whitaker cessful, printable bioink must have suitable properties that
The Welsh Centre for Burns and Plastic Surgery maintain the bioink and cell integrity post-printing. It is well
Morriston Hospital
Swansea SA6 6NL, UK documented that various cell phenotypes are exquisitely sensi-
tive to subtle variations in mechanical properties of the micro-
DOI: 10.1002/adhm.201700264 environment. Hence many bioinks have been developed and

Adv. Healthcare Mater. 2017, 1700264 1700264  (1 of 16) © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advhealthmat.de

engineered for specific cell types. Moreover, printing parame-


ters must be clearly defined for different bioinks and cell types. Stuart Kyle is currently a
The term ‘bioink’ refers to (1) the ‘biopaper’ which generally Clinical Lecturer in Plastic
is hydrogel-based and provides structural and mechanical sup- Surgery at Swansea University
port, and (2) specialized cells, such as stem cells. In order to Medical School, UK.
create a functional 3D construct, the ‘bioink’ must be carefully Following his Wellcome Trust-
selected in addition to an appropriate bioprinter (Figure 1). funded PhD (Biomaterials)
The biofabrication window paradigm (Figure 2) defines the in 2010 from the University
compromise between suitability for fabrication (printability), of Leeds, he completed his
and the ability to encapsulate cells and maintain cell via- medical degree (MBChB
bility.[10,15] Traditionally there has been a compromise between with Honors) in 2014. He
the ideal bioink for 3D printing and cell culture. Many lacked was awarded the prestig-
shape fidelity and printability, had inadequate mechanical ious Foulkes Foundation
strength post-printing and had poor bioactivity and biodeg- Fellowship (one of only 275 in the past 40 years). He has
radability profiles. Higher concentrations of some polymeric published widely in the field of biomaterials and nano-
bioinks introduces more cross-linking and stiffness. However technology and his current research interests are in the
these networks can often hinder cell migration, proliferation development of novel biomaterials for 3D bioprinting for
and matrix deposition. Conversely, less dense networks can lead use in reconstructive surgery.
to poor shape fidelity as it can be difficult to maintain shape
post-printing. Therefore more traditional bioinks have com- Ayesha Al-Sabah is a Post-
promised with a moderate degree of crosslinking, but with far Doctoral Research Fellow in
from ideal cell culture environments. More advanced bioinks the Reconstructive Surgery
are now being designed to significantly improve printability and Regenerative Medicine
and biocompatibility. These can be achieved through careful Research Group, Swansea
control of various physical, chemical and biological proper- University Medical School,
ties, such as rheology (viscosity, shear-thinning, viscoelasticity), UK where she supports
biofunctionalization, biodegradation, gelation kinetics and cell a large team of Clinical
function (cytocompatibility, cell adhesion, migration, prolifera- Researchers. Following the
tion, differentiation). award of her PhD from Cardiff
There is a wealth of current literature reviewing bioprinting University (Cartilage Tissue
techniques, candidate biomaterials for use as bioinks, and the Engineering), she was a Post-
applications of bioprinting for tissue engineering and transla- Doctoral Researcher in Cardiff before joining the team in
tional purposes. There is a paucity of articles concentrating on Swansea. Her research interests are currently in the field of
the ‘printability’ of bioinks, which is especially important when mesenchymal tissue engineering with a focus on stem cell
complex 3-dimensional structures are required, to our knowl- biology, bioenergetics and 3D bioprinting.
edge there are no comprehensive reviews in the literature.
Iain Whitaker is Professor of
Plastic Surgery and Director
of the Reconstructive
2. Concept of ‘Printability’ Surgery and Regenerative
Medicine Research Group
The core processes of 3D bioprinting can be seen in
(ReconRegen) based at
Figure 3. The product (printed biological constructs) must be
Swansea University Medical
precise (with high resolution, shape fidelity) and reproducible.
School. After reading medi-
Yet how are printability, fidelity and reproducibility measured,
cine at Cambridge University
and is it adequate? Key studies that have identified parameters
he completed Plastic Surgery
that potentially help control and measure ‘printability’ of var-
training in the UK, Sweden,
ious biomaterials are highlighted in Table 3. The desired struc-
USA, Australia & France and
ture of the finished construct is highly precise and, in order to
was awarded a PhD from the University of Utrecht. He
mimic these structures, it is necessary for the printing process
was the first plastic surgeon to receive the Rowan Nick’s
to be equally precise and for the construct to maintain this
Award to work In Melbourne, followed by the European
shape over a period of cell culture.
Association of Plastic Surgeons Young Plastic Surgeon
Printability is an important concept as the product of the bio-
Scholarship to work in Paris. His current research interests
printing process must mimic both the cellular architecture and
involve 3D biomanufacture and tissue engineering.
shape. This is particularly critical when tissues of an intricate
nature are required, such as a heart valve or ear. Geometry is
as important in tissue structure as it is function. Heart valves
for example have to be durable whilst allowing appropriate novo valve scaffolds have been engineered over the years which
blood flow at varying valvular pressures. For this reason, spa- involved manually shaping leaflet geometries using various
tial and mechanical heterogeneity is needed. A number of de polymers, mostly based on polylactide,[45] polyglycolide,[46]

Adv. Healthcare Mater. 2017, 1700264 1700264  (2 of 16) © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advhealthmat.de

Table 1.  Comparison of four types of bioprinting modalities.

Bioprinting modality Advantages Disadvantages Applications Reference


■ Fast printing speeds (1–10,000 ■ Require low viscosity (3–12 mPa.s) ■ Fabrication of 3D cell-laden [8–10]
droplets s−1) environments to prevent clogging constructs in a drop by drop
■ Low cost ■ Low thermal conductivity manner
■ Small volume droplets (picoliters) ■ Small volumes droplets may restrict ■ Tissue engineering of blood

■ Intricate structures with high scaling up to fabricate larger organ vessels, cartilage, bones and
resolution (10–50 µm) structures neurons
■ Readily adapted with multiple ■ Difficult printing vertical 3D structures

nozzles for multimaterial printing


■ Fast gelation to allow for rapid
fabrication

■ High resolutions (≈1 µm) and ■ Limited to photosensitive materials ■ Uses lasers/projectors that [11]
cell viabilities (>85%) ■ Possibility of DNA damage by UV polymerize light sensitive polymer
■ Increased fabrication speed light, but as this method typically materials which enable 3D
■ Reduced fabrication time uses UV-visible light (anywhere from structures to be created in
■ No limitation of viscosities 365–530 nm), many studies have bottom-up assembly
■ Low cost shown no/little damage to cells ■ Tissue engineering of blood

or DNA for the brief duration vessels, liver, bone and cartilage
of exposure ■ Organ on a chip

■ High resolutions ■ Fast gelation kinetics and fast ■ Uses a laser as the energy [9,12]
(pico- to microscale) moving stage for fabrication source to propel and deposit
■ Small printing volumes can be troublesome bioinks onto a collecting
(pico-to-nanoliter) ■ High cost substrate in a precise manner
■ No dispensing nozzle so no ■ Complex design ■ Tissue engineering of bone,

clogging issues ■ Heat from laser may damage cells adipose tissue, skin and blood
■ Reduces issues of shear stress vessels
which can affect cell viability

■ Multiple cell/material delivery ■ Slow print speed (µm s−1) ■ Produces a continuous bioink [13,14]
■ Higher viscosity biomaterials ■ Critical timing of gelation time filament onto a stationary sub-
can be printed (30–6 × 107 mPa.s) ■ Specific matching of material strate from a microscale syringe
■ Room temperature processing densities to preserve intricate shapes tip driven by either pneumatic
■ Direct incorporation of cells and ■ Small diameter nozzles (150 µm) pressures or mechanical force
homogenous distribution of cells can lower cell viability (68.6%) ■ Tissue engineering of blood

■ Large and intricate structures vessels, muscle, bone, cartilage,


with moderate resolutions heart valves, liver and neurons
(200–1000 µm) ■ Organ on a chip

■ Shape fidelity is high post-printing

Adv. Healthcare Mater. 2017, 1700264 1700264  (3 of 16) © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advhealthmat.de

Table 2.  Ideal properties of bioinks.

Property Examples Reference


Form scaffolds with adequate mechanical strength, maintaining shape ■ Ionic crosslinking [14,16,17]
fidelity at a high resolution ■ Photochemical crosslinking
■ Sacrificial scaffolds
■ Quantifying shear stress during bioprinting
■ Concomitantly printed scaffolds
Exhibit tunable gelation to facilitate extrusion ■ Co-axial nozzle [17–20]
■ Inks loaded with
○ nanoparticles

○ spheroids

○ micro-carriers

■ Blending

Form scaffolds that resemble the native cellular microenvironment ■ Natural [17,21,22]
○ Alginate

○ Collagen

○ Gelatin

○ Agarose

○ Chitosan

○ Cellulose

○ Hyaluronic acid
■ Synthetic

○ Poly(lactic acid)
○ Poly(lactic acid-co-glycolic acid)
○ Polyethylene glycol

○ Polycaprolactone

○ Polyurethane

○ de novo designed peptides

Exhibit biocompatibility ■ Improved cell adhesion and viability [5,17,19,23–25]


■ Improved cell binding
■ Support differentiation and matrix deposition

■ Matrix remodeling

Be amenable to chemical modifications (functionalization/decoration) ■ RGD modification [17,25,26]


■ Cyclic RGDS
■ PEGylation

■ MMP-sensitive sequences

Capable of large scale synthesis with minimal batch-to-batch variation ■ Fabrication of large-volume constructs using 3D printing [17,27]
system capable of providing enhanced printability

polyhydroxybutyrates[47] and polyhydroxyalkanoates.[48] Hence, after importing Standard Tessellated Language file (STL) geom-
a more automated manufacturing technology that could fabri- etries acquired using micro-CT. As print paths were too wide
cate tissue engineered scaffolds more rapidly with higher for filling in smaller regions, careful control and manipulation
anato­mical precisions was required. One group implemented of print paths (height and/or width) is critical for improving
an on-board photocrosslinking system that enabled simulta- print resolution and shape fidelity. Hence in order to optimize
neous 3D extrusion printing and curing of hydrogels in com- future studies, smaller nozzle diameters to increase print reso-
plex aortic geometries.[38] Native anatomic and axisymmetric lution, and a lower extrusion rate may be required. In analysing
aortic valve geometries with 12–22 mm inner diameters were CT slices to assess internal geometric fidelity, they found that
3D printed with poly-ethyleneglycol-diacrylate (PEG-DA) hydro- overlapping area percentage decreased as the size of the image-
gels supplemented with alginate (Figure 4A–F). 3D printing derived valve scaffolds decreased. They also reported on the
geometric accuracy/shape fidelity was quantified and com- importance of swelling in printed constructs. The scaffolds in
pared using micro computerized tomography (micro-CT) this study were found to expand outward but not inward, pos-
(Figure 4G–H). It was found that the 3D printing strategy sibly suggesting surface tension on the inner walls prevented
generated scaffolds with high geometric precision, but accu- this. In fact, surface tension is an important factor when con-
racy decreased somewhat with reduced size. Higher shape sidering printability, where contact angles between two media
fidelity was seen with larger valves (22 mm), compared to can be measured and manipulated.[7] Surface tension has been
smaller valves but more overprint error (regions printed out- shown to be important in droplet formation in inkjet-printing
side the target print area) was seen for 12 mm valves, which and this could be extended to extrusion-based systems, since
suggested that print paths were too wide. Print paths are soft- the affinity of the nanofluid for the substrate can be formulated
ware generated dimensions of path height and width defined from the solid surface tensions.[49] Surface tension of the bioink

Adv. Healthcare Mater. 2017, 1700264 1700264  (4 of 16) © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advhealthmat.de

pore diameter and filament width at higher


percentage compositions. However, quanti-
tatively there were large variations in dimen-
sions compared to the intended dimensions.
Visually on the other hand, better resolution
of pores and widths was evident (Figure 5).

2.2. Computer Simulation and Modeling

There is no question that quantitative anal-


ysis of post-printing structures is difficult,
and it has often been trial and error. In order
to achieve a successful 3D bioprinting proce-
dure however, parameters must be carefully
tuned. Many reports have focussed on illus-
trating how bioink properties and printing
parameters can either affect cell viability or
bioink printability following printing. Yet it is
not until the last couple of years that studies
have tried to highlight the importance of
the two combined. Good progress has been
made in trying to semi-quantify printability
using mathematical models or more quanti-
tative methods using computational analysis.
Extrusion-based bioprinting systems use
mini-tissues as bioinks.[51,52] Mini-tissues
combine biomimicry with complex, self-
assembly processes to produce large-scale
Figure 1.  Functional 3D constructs can be fabricated through careful control of the (1) ‘bioink’
functional tissue. Biological structure forma-
(specialized cells + biopaper) and (2) the bioprinter.
tion can be characterized by the coalescence
of discrete mini-tissues in the form of multi-
can also influence printing quality, resolution and width of cellular spheroids or cylinders.[52] Shape evolution during tissue
printed lines.[50] Hockaday et al (2012) therefore proposed that fusion[53] has been shown to be analogous to merging of liquid
hydrogel scaffolds should therefore be printed thinner than the drops.[51,54] This liquid analogy has been exploited to develop
target size so that the expanded final state matches the native an experimental-theoretical-computational formalism.[55,56]
model.[38] This study demonstrated that 3D printing of ana- One group[37] have made great progress in this area using a
tomical scaffolds with high shape fidelity was possible but por- computer simulation method known as the cellular particle
trayed the complexities associated with these techniques. dynamics (CPDs). This method can predict the shape evolution
of multicellular systems that undergo shape-changing biome-
chanical relaxation (Mathematic details of the CPD formalism
2.1. Pore and Filament Dimensions can be found in[55]). Furthermore, it has been developed with
the intention to make extrusion bioprinting more predictive,
Whilst is it clear that the properties of bioinks critically influ- thus reproducible and more time efficient.[37] They found that
ence their delivery and the integrity of the structure formed, complex, mathematical modeling via CPD simulations can be
the concept of ‘printability’ has only been investigated in a used to predict post-printing structure formation even in the
handful of papers. One early study investigated the printability case of volume changing bioink units. They went on to highlight
of alginate and gelatin with calcium chloride as a cross linking that tubes printed with cylindrical bioink formed nearly twice
agent in an extrusion-based printing system.[41] They found as fast as one printed with spheroidal bioink. Figure 6 shows
that the properties were changed favourably when gelatin was snapshots from the start and end of CPD simulations through
included at 2 and 4% Alg-Gel. In addition to other methods of the fusion of stacked rings formed by spherical and cylindrical
testing printability such as rheology and ink consistency (dis- cellular aggregates. CPD simulation was then used to predict
cussed later), they compared sample dimensions inputted into the tube formation for tubes built with spherical (21.7 hours)
the software and those obtained after fabrication. Although and cylindrical (11.8 hours) bioink units. Albeit, this would
they had obviously tried to make some headway in measuring appear to be more appropriate for tubular organ structures such
pore diameters and filament widths, it simply was not robust as fabricating vascular grafts. It certainly appears to be a sig-
enough as a means to correlating intended dimension parame- nificant step in the right direction by being the conduit between
ters to those that were actually measured. The study highlighted needing a 3D tissue engineering construct and potentially pro-
that alginate-gelatin showed better resolution with respect to ducing an accurate, reproducible structure with high fidelity.

Adv. Healthcare Mater. 2017, 1700264 1700264  (5 of 16) © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advhealthmat.de

Figure 2. The biofabrication window. Advanced bioinks are being rationally designed to have excellent biocompatibility with high shape fidelity con-
structs. There has been a paradigm shift from the more traditional bioinks which have seen a stand-off between shape fidelity and biocompatibility
(top left and bottom right). Stiffer, more dense polymeric networks result in better printable bioinks but lead to a poor cell culture microenvironment.
Adapted with permission.[10,15] Copyright 2016, Springer.

More recently, Ouyang et al. (2016) have systematically chloride) over a range of temperatures. In ideal gelation con-
studied the influence of bioink properties and printing para­ ditions, the extruded filament would demonstrate clear mor-
meters on bioink printability and embryonic stem cells viability phology with smooth surface and constant width in three
using extrusion-based cell printing.[24] The study used three dimensions resulting in regular grids and square holes in the
extrusion states: under gelation, proper gelation and over gela- fabricated construct. In under gelation conditions, a more
tion (Figure 7A). In this study, gelation is defined as hydrogel liquid-like state was created where upper and lower layers
fixation using post-processing chemical crosslinking (calcium would fuse and create holes that are more circular. They then

Figure 3.  Core fabrication process for 3D bioprinting technology.

Adv. Healthcare Mater. 2017, 1700264 1700264  (6 of 16) © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advhealthmat.de

Table 3.  Key parameters for measuring printability.

Parameter(s) Conclusions Reference


Rheology • Printing accuracy determined by objectively quantifying printability over a 1 cm2 area and evaluated in terms of adhering [28]
to the programmed 1 cm2 dimensions
• Printing accuracy was calculated by dividing 1 cm2 by the actual printed area and multiplying by 100%
• More accurate (>95%) printing was found in higher viscosity bioinks (chitosan, chitosan-collagen,
methylcellulose-hyaluronan) which prevented them from spreading out over the printing area
• More viscous bioinks could only be printed via syringe deposition, compared to less viscous bioinks which were
printed using pressure-driven printers
• Oxidized alginate biofunctionalized with cell recognition motif, RGD resulted in a practical method of controlling viscosity [29]
and biodegradation
• Optimal range of viscosity suitable for high fidelity printing was found to be between 400 mm2 s−1 and 3000 mm2 s−1
• Hydroxyapatite made the bioink more viscous and difficult to print [30]
• Lower viscosity hydrogels led to watery, soft structures that could not hold their shape post-printing [31]
• Highly viscous materials were difficult to print and deposit smoothly
• Better rheological properties achieved through structural support of nanocellulose and cross-linking ability of alginate [32]
• Viscoelasticity was found to be the decisive factor for printability in microextrusion-based 3D printing technology [33]
• Incorporating collagen into the bioink enhances extrudability, and alters rheological properties to allow better control over [34]
shape fidelity
• Gelation viscosity remained on the same level when bioink concentration was constant, suggesting that gelation viscosity [24]
could act as an intrinsic property parameter representing bioink rheology
• Longer gelation times resulted in poor printability
• High resolution bioprinting of viscous materials was possible at lower shear stress [35]
• All samples of alginate with calcium chloride showed shear thinning properties [36]
• Incorporating graphene oxide increased viscosity and led to different thixotropic properties with a better quality of the
printed construct
Bioink composition • Stable constructs created when gelatin added to alginate [30]
and concentration
• Incorporating nanocellulose into alginate improved shape fidelity [32]
• Ideal printing conditions that gave best resolution were with 7.5% gelatin with 1% alginate and a printing temperature [24]
of 27.5 and 30 °C
Computer simulation/ • Predicts shape evolution of multicellular systems that undergo shape-changing relaxation [37]
modeling • Printing using cylindrical bioinks form twice as fast as using spheroidal bioinks
• Fluid-dynamic modeling used to control shear stress at the nozzle site, which could be adjusted by varying printing [35]
pressure, hydrogel viscosity and nozzle diameter
Nozzle variables • Smaller nozzle diameters can lead to better print resolution [38]
• Print scaffolds thinner than target size as post-printing expansion occurs
• Smaller diameter nozzles do not necessarily enhance printing resolution [39]
• Higher pressures may be necessary if printing with fine nozzles and highly viscous materials to overcome shear resistance
at the cost of resolution loss
• Addition of nozzle tip heaters allowed homogenous deposition of hydrogel and eliminated clogging at tip [30]
• Control of nozzle parameters and diffusion rate at lattice intersections is important for printing intricate structures with [40]
high shape fidelity
• Air pressure was found to be the most important parameter as it determined extrusion output
Pore and/or filament • Correlated intended and measured dimensions [41]
dimensions • Quantitatively, large variations between intended and measured dimensions
• Better resolution of filament widths and pore size with alginate-gelatin bioinks by controlling temperature of formulations
Grid geometry • Solidified filament morphology demonstrated with sufficient mechanical strength to support the deposition of upper layers [42]
• Semi-quantifiable method developed to assess printability based on fibre/grid morphology [24]
Effect of printing angles • Shorter distances between nozzle tips and cross-linking agents such as calcium chloride led to better mechanical [43]
on shape fidelity properties, support and printing quality when printed at varying angles for printing complex, 3D structures
• Printing quality was worse with acute angles compared to obtuse and right angles [40]
Resolution determined • Incorporation of hydroxyapatite led to enhanced visibility of hydrogel and enhanced resolution of intricate microstructures [30]
by micro CT

Adv. Healthcare Mater. 2017, 1700264 1700264  (7 of 16) © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advhealthmat.de

Table 3. Continued.

Parameter(s) Conclusions Reference


• Used to quantify scaffold structural properties, such as fiber diameter, which was used as a metric to assess printing quality [44]
• Linear regression models used to determine the impact of material and printing parameters on printing accuracy
• Using a direct melt extrusion system and comparing needle size on printing quality,
○ material properties were found to play a more important role in printing quality with smaller diameter needles

○ printing parameters, such as pressure and speed became more significant in determing fiber quality when using large

diameter needles
• 3D printing geometry accuracy was quantified and compared using micro-CT (slice-by-slice comparison) [38]
• Micro-CT generated slices were compared to the corresponding image-derived layers of the original STL files in the XY-plane
• Slices were overlapped and compared using Boolean operations
Oxygen concentration in • Oxygen inhibition had a detrimental effect on shape fidelity which was overcome by increasing UV or visible light [34]
printing environment photoinitiator concentration, or light irradiation intensity

Figure 4.  Printing heterogeneous valve and scaled valve constructs. A) Porcine aortic valve model was B) printed with PEG-DA hydrogels. C) Scaffolds
were printed with inner diameters of 22, 17 and 12 mm. D) Axisymmetric valve model was E) printed with two blends of hydrogels F) and at 22, 17
and 12 mm inner diameter. Fidelity analysis of fully hydrated printed valve scaffold geometry compared to model geometry using microCT-derived STL
files. G) Heat maps and average % frequency histograms representing surface deviations between printed porcine scaffold versus model geometries
and H) printed simplified scaffolds versus model geometries. Warmer colours indicate positive deviation while cooler colours indicate negative devia-
tion in millimetres. % frequency histograms show the deviations that lie within ± 10% tolerance (green), and outside the tolerance (blue and pink).
Reproduced with permission.[38] Copyright 2012, IOP Publishing.

Adv. Healthcare Mater. 2017, 1700264 1700264  (8 of 16) © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advhealthmat.de

Figure 5.  Macrostructure of alginate-gelatin scaffolds with concentrations of (A) 1%, (B) 2% and (C) 4%. Extrusion printing of alginate-gelatin scaffold
is seen in (D). A mechanically robust 2% alginate-gelatin scaffold after crosslinking with calcium chloride is seen in (E). Dimensions of alginate-gelatin
scaffolds +/− calcium chloride is highlighted in (F). Reproduced with permission.[41] Copyright 2013, The Royal Society of Chemistry.

applied mathematical formulae as seen in Equations (1) and (2) π 1 L2


to define bioink printability (Pr). Pr = . = (2)
4 C 16 A
4π A (1) They use the principles that circles have the highest circu-
C=
L2 larity (C) so C is equal to 1, and the closer values are to 1, the
more circular they are (Equation (1)). They then define printa-
bility (Pr) based on a square using Equation (2). They state that
for an ideal gelation condition or perfect printability status, the
interconnected channels of the constructs would demonstrate
a square shape, with a Pr value of 1. Hence, larger Pr values
would correlate with a greater degree of gelation of the bioink
and vice versa. The study used varying compositions of gelatin
and alginate at different nozzle temperatures. It highlighted
very nicely how important controlling design parameters are
when trying to printing materials with defined, intricate struc-
ture and high fidelity. Under proper gelation conditions, smooth
and uniform filaments were continuously extruded, which
resulted in grid constructs with distinguished layers (Figure 7B).
Figure 6.  Tube formation from CPD simulations at the start (t = 0) and When they analysed printability, Pr over time, they found
end (t = 620). Time (t) units are CPD related and when mathemati-
that the 7.5% gelatin with 1% alginate at 27.5 °C and 30 °C,
cally processed, correlate with tube formation times of 21.7 hours and
11.8 hours for spherical and cylindrical bioink units, respectively. Different were the proper printing temperatures and 25 °C resulted in
colours are used to emphasize the absence of mixing during the fusion over gelation after 10 minutes (Figure 7C). They then went on
process. Reproduced with permission.[37] Copyright 2015, IOP Publishing. to demonstrate that cell-laden constructs showed high shape

Adv. Healthcare Mater. 2017, 1700264 1700264  (9 of 16) © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advhealthmat.de

Figure 7.  Bioink printability assessment under different printing parameter combinations. A) Evaluation of printability (Pr) under different gelation
states – under-, proper- and over-gelation. B) Optical microscope images at 30 mins. C) Semi-quantified Pr value construct for gelatin/alginate
bioink with different concentrations under different printing temperatures. Scale bars are 1 mm. Reproduced with permission.[24] Copyright 2016, IOP
Publishing.

fidelity and clear contours, and that embryonic stem cells were printing 3D structures. Hydrogels were created using alginate
uniformly distributed in the hydrogel filaments. In terms of and gelatin, with calcium chloride as the crosslinking
assessing printability based on the degree of gelation, this study agent. They demonstrated that line printing was affected by
has made excellent progress in trying to help define printability, (1) air pressure (the most important parameter as it determines
or at least provide some quantification so that a trial and error extrusion output), (2) nozzle feed rate and (3) distance between
approach is minimized. It further demonstrates how careful nozzle and substrate. Printing quality was also found to be
control of design and process parameters can lead to constructs worse with acute angles compared with obtuse and right angles,
with precise and accurate uniformity. but could be improved when the extrusion rate is controlled by
the motor speed (Figure 8A–D; Table 3). They found that in
sharp angle printing, there was an overlap problem in which
2.3. Printer/Nozzle Parameters and Printing Angles extrusion of the hydrogel was doubled. To overcome this, the
extrusion rate was reduced. The double extrusion at the overlap
A more recent study has tried to address the apparent lack of (Figure 8C) could therefore be released by doubling the nozzle
systematically investigating and controlling printing para­meters feedrate and creating a more uniform extrusion (Figure 8D).
as a means of printing structures with high fidelity, reso­lution In order to produce cell-laden scaffolds, lattice structures
and accuracy.[40] They highlighted important factors that must have been frequently used and so precise control of printing
be considered in order to produce accurate structures, from quality is critical. They found that line distance and the inter-
printing lines in 1D to printing lattices in 2D to ultimately section area of the lines affected lattice quality, in relation to

Adv. Healthcare Mater. 2017, 1700264 1700264  (10 of 16) © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advhealthmat.de

Figure 8.  High fidelity 3D printing of alginate/gelatin hydrogels. Overlap in sharp corner printing. Line shape of (A) obtuse and (B) right angle printing.
C) Double extrusion at overlap area. D) Uniform extrusion by doubling the nozzle federate in the overlap area. Hydrogel diffusion at the intersections
of lines. The model and shape of lattice intersections when line distance was (E) 5 mm and (F) 2 mm. Hydrogel diffusion and fusion when lattice
scaffolds had line distances of (G) 4 mm, (H) 3 mm, (I) 2 mm and (J) 1 mm. Printability of 3D hydrogel scaffolds. K) Digital model of 3D scaffold.
L) An outline of the first layer of the scaffold with line width 2 mm. M) First layer of the scaffold shape and N) three views of hydrogel scaffolds formed
at different angles (30 layers printed with a height of 15 mm, fidelity of the top surface is approximately 80%). Reproduced with permission.[40] Copy-
right 2016, Nature Publishing Group.

Adv. Healthcare Mater. 2017, 1700264 1700264  (11 of 16) © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advhealthmat.de

diffusion rates. Hence larger line distances led to a drop in dif- light irradiation intensity. They also incorporated a small con-
fusion rate. Diffusion rates appeared to have a direct affect on centration of collagen I (0.6 wt %) which was found to enhance
shape fidelity as when reticular shapes were printed one on top extrudability of the bioink, and alter rheological properties to
of another, the pore would radially narrow along the second allow for better control over shape fidelity. In the Vis+Ru+SPS
hydrogel layer line and did not change axially. Overlapped system, cell cytocompatibility was found to be significantly
hydrogels diffused due to gravity which resulted in diffusion improved and the resulting porous biofabricated constructs
between lattice intersections (Figure 8E–F). It was also apparent were found to be less susceptible to the effects of oxygen inhibi-
that for larger structures with larger line distances within lat- tion and incomplete crosslinking. This was evident by the lower
tices, diffusion and fusion would be mitigated (Figure 8G–J), percentage change in fiber diameter compared to the UV+l2959
yet for smaller, more intricate structures this may prove more system post-swelling (Figure 9).
problematic. Thus translating this in to 3D printing to produce
high shape fidelity objects was possible (Figure 8K–N).
2.5. Viscosity and Rheology

2.4. Oxygen Concentration in the Printing Environment Viscosity is another key parameter to consider for successful
3D bioprinting. Many groups report on viscosity and rheo­
A novel method in trying to improve print fidelity has recently logy studies in conjunction with printability assessment for
been reported.[34] This study was one of the first to highlight bioprinting materials. Hydrogels have received much more
the importance of oxygen inhibition on print fidelity in photo- attention than other biomaterials as viscosity can be carefully
polymerized gelatin-methacryloyl hydrogel constructs. Oxygen controlled.[22,58,59] Hydrogels must be sufficiently viscous to
can act as a radical scavenger. In certain applications this has retain their shape during the printing process and they should
many benefits, however in 3D bioprinting it can often lead to have crosslinking capabilities so that the 3D structure is not
incomplete or insufficient crosslinking which may affect the altered after printing. Low viscosity materials have been shown
resulting shape fidelity.[18,57] It has been shown that free radi- to be more attractive for bioprinting as cells can grow well in
cals formed from photolysis of photoinitiators are converted the low pressure environment.[60] However, it is well known
to peroxyl radicals that do not react with ester bonds.[57] These that 3D printed structures tend to collapse due to low viscosity.
peroxyl radicals can form hydroxyperoxides or alcohols which Although many research groups have focussed on adjusting
prevent covalent crosslinking. Inert atmospheres involving the concentration and molecular weight of various hydrogels
nitrogen purging has been used widely by many to prevent in order to change viscosity, shape fidelity has often not been
entry of oxygen, yet this would be unsuitable in 3D bioprinting achieved after printing. In this respect, hydrogels have often
involving cell-laden constructs. Lim et al. (2016) utilized light- been printed simultaneously with other materials in order to
activated polymerization systems consisting of water-soluble increase structural fidelity. Fugitive materials have also been
photoinitiators ruthenium (Ru) and sodium persulfate (SPS) employed in order to maintain or improve shape fidelity with
compounds which can absorb photons in the visible light low viscosity bioinks.[58] Recently, nanocellulose has been
range.[34] They compared this system to the conventional UV/ combined with alginate to prepare sponges used for adipose
Irgacure 2959 system and found that oxygen inhibition had tissue[61] and for cell encapsulation.[62] Markstedt et al. (2015)
a detrimental effect on shape fidelity which was overcome by found that in order to improve shape fidelity of alginate, nano-
increasing UV or visible light photoinitiator concentration, or cellulose can be incorporated into the bioink formulation.[32]

Figure 9.  3D plotting of gelatin+methacryloyl+collagen constructs that have been photopolymerized in UV+l2959 and Vis+Ru/SPS systems. Different
shape fidelity can be seen post-irradiation and post-swelling with the different systems. Reproduced with permission.[34] Copyright 2016, American
Chemical Society.

Adv. Healthcare Mater. 2017, 1700264 1700264  (12 of 16) © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advhealthmat.de

They found that the low zero-shear viscosity of alginates gave print-head and additional tip heaters allowed homogenous
poor shape fidelity when printing. Nanocellulose alone had a deposition of hydrogels. Thus eliminating clogging issues at
higher viscosity at low shear rates, however when printed the dispense tips, and also maintaining liquidity of the bioink so
shape was destroyed when mechanical force applied. The com- that the printing substrate could be cooled down sufficiently to
bination of better rheological properties using nanocellulose guarantee instantaneous gelation of gelatin. This then created
with the cross-linking ability of alginate yielded the best results. stable 3D constructs that maintained the shape of the prede-
Photocrosslinking has emerged as a promising strategy fined geometry.
to address several challenges associated with lower viscosity An understanding of shear stress is another important factor
bioinks, cell viability and printability. Since light is minimally to consider in bioprinting applications. Studies have demon-
invasive, spatiotemporally controlling the application of light in strated that shear stress can be influenced by varying nozzle
order to crosslink photosensitive cell-laden bioinks with high diameters, printing pressures and viscosity of the material to be
shape fidelity and without significantly reducing cell viability printed.[64] It has also been well documented that shear stress
has been reported.[63] In one study, photocrosslinkable hydro- plays a pivotal role in cell biology, particularly in relation to cell
gels were generated using methyacrylated hyaluronic acid and signaling[65] and differentiation.[66] Recent research has shown
methyacrylated gelatin, and human aortic interstitial cells were that hydrogel viscosity and nozzle size directly affect shear
encapsulated within the hydrogels for 3D bioprinting.[31] They stress.[35] They presented a microvalve-based bioprinting system
found that systematically varying hydrogel concentration had for the manufacturing of high resolution, multi-material 3D
a significant impact on viscosity and material stiffness. When structures (Figure 10A). They then used fluid-dynamic mode­
viscosity values were lower than 100 Pa⋅s, the material was too ling to precisely control shear stress at the nozzle site, which
fluidic to hold the printed shape, and when values were higher could be adjusted by varying the printing pressure, hydrogel
than 104 Pa⋅s, materials were too viscous and difficult to deposit viscosity and nozzle diameter. Subsequently, it was highlighted
smoothly. how cell viability and proliferation potential were affected by
The combination of alginate and gelatin with varying con- different levels of shear stress (Figure 10B). At lower rates of
centrations of hydroxyapatite have been 3D bioprinted to form shear stress (<5 kPa), L929 mouse fibroblasts remained viable
hydrogels for use in bone tissue engineering. One group sug- at 96%, however this viability decreased to 91% and drasti-
gested that combining gelatin and alginate improves instanta- cally to 76% at shear stress ranges from 5–10 kPa and >10 kPa
neous stability by maintaining initial bonding between single respectively. They further demonstrated that high-resolution
layers due to the gelation of gelatin as well as long-term stability bioprinting of viscous materials is possible at a low shear stress
of the whole construct by chemically crosslinking alginate.[30] level, whereby computer aided design software was trans-
The addition of hydroxyapatite led to enhanced visibility of the formed into multi-layered drop models using custom designed
hydrogel when seen with micro CT and led to higher resolution algorithms. Although this method was specific to microvalve-
of intricate microstructures. Introducing hydroxyapatite into based bioprinting, and was not extrusion-based, it certainly is
the alginate-gelatin composite rendered solutions more viscous pioneering in using fluid-dynamics and mathematical analysis
and more difficult to handle and print. They emphasized that to understand the importance of shear stress on high resolu-
viscosity affects the printing process primarily at the dispense tion, drop-on-demand bioprinted 3D constructs.
tip which is the most critical location during extrusion. They It is evident that very few reports in the literature have dis-
also highlighted that control of the tip temperature affected vis- cussed the relationship between rheological properties of
cosity of the bioink, insomuch that the dual action of a heatable bioinks and 3D printability. We have seen that alginates have

Figure 10.  A) 3D bioprinting with bi-phasic support liquid for the manufacturing of thin-walled, multi-layered hydrogel structures. B) Schematic illus-
tration of the velocity and shear stress distribution as well as the stress impinged on cells inside a bioprinter nozzle. Reproduced with permission.[35]

Adv. Healthcare Mater. 2017, 1700264 1700264  (13 of 16) © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advhealthmat.de

been a mainstay as scaffolds for tissue engineering. Yet native entirely on the field (The International Journal of Bioprinting,
alginate is bioinert with a poor degradation profile. This has led Bioprinting, Journal of 3D Printing in Medicine and 3D
many groups to functionalize alginate to make it more bioac- Printing in Medicine) attest to the emerging volume of litera-
tive, and chemically modify it in order to improve biodegra- ture. Early studies focussed on the assessment of ‘printability’ of
dability. Oxidized alginates have been shown to have a better bioinks whilst investigating different biofabrication platforms.
degradation process with promising results for tissue engi- Basic correlations of printability was compared to mechanical
neering applications.[67] Further studies have shown that these properties and biocompatibiity of the bioink. More recently,
biodegradable alginates have potential in 3D bioprinting.[29] more precise control of process parameters during the printing
Oxidized alginate conjugated with the cell recognition motif process have been interrogated and correlated with fidelity and
RGD, and human adipose-derived stem cells were printed in reproducibility of 3D constructs. It is clear that bioink compo-
a point-to-point method. They found that altering oxidation sition and concentration, rheology and manipulation of nozzle
and concentration resulted in a practical method of controlling parameters are key factors in order to evaluate and achieve
the viscosity of degradable alginates. In order to asses printing good printability with high shape fidelity and reproducibility.
resolution, dot arrays were printed using different biodegrad- The potential to biofunctionalize and decorate bioinks with cell
able alginates to examine the effect of viscosity on dot fidelity recognition sequences may enhance physical, chemical and bio-
and size. Hence an optimal range of viscosity suitable for high logical properties of next generation, advanced bioinks.
fidelity printing was found to be between 400 mm2 s−1 and Advanced research methodology and technical advantages
3000 mm2 s−1. Hence rheological properties are key players in pave the way for emerging 4D printing methods, adding in
controlling printability and fidelity of 3D constructs. temporal dimensions. The success of such strategies will allow
In order for a bioink to be printable, it should ideally be biofabricated constructs to adapt and transform to the new
thixotropic i.e. exhibit non-Newtonian shear thinning behav- micro-environment and a range of physico-chemical cues over
iour whereby the higher the stress they experience, the lower time, resulting in accurate biomimicry and smart biomaterials.
their viscosity, and have a reasonable recovery time. Thixotropy This advanced and innovative technology has great potential in
provides information on how quickly and how much viscosity the field of biomedicine, particularly in areas such as regenera-
bioinks recover post-printing, whilst the recovery time is the tive medicine and biomedical devices.
time given to the bioink in order for viscosity to be recovered Despite increased research infrastructures, man power,
during bioprinting. Since not all bioinks are thixotropic, it is funding and technological advances, significant challenges
essential that bioinks must be rheologically evaluated to deter- remain. The design and manufacture of specialized bioprinters
mine the effects of 3D printing on print quality. Studies have is still niche, and high cost, meaning a small number of research
shown that materials that exhibit a linear relationship between groups do not have ready access to them. However, costs have
viscosity and shear rate improves print quality.[5] Hence, finding reduced considerably in the past few years with commercial
relationships between piston speeds of bioprinters, viscosity, bioprinters being sold for ≈$5000, and systems being built with
shear rates and thixotropy are fundamentally important. This open-source designs for <$1000. Development of mechanically
was highlighted by one group who performed rheological appropriate, biomimetic, durable bioinks that encourage cell
studies on 3D printability of alginate-based hydrogels and inves- proliferation and survival whilst maintaining shape fidelity still
tigated the effects of introducing graphene oxide into these remains the holy grail. 4D bioprinting, improvements in print
hydrogels.[36] Graphene oxide is known to have excellent bio- resolution and modification at the molecular scale are key areas
compatibility, mechanical stiffness and is ultra-strong so may of research to watch closely in the future.
be able to modify scaffold properties. This study demonstrated It is in the interests of patients, scientists, clinicians and
that measurable parameters can be defined for quantifying the industry to see translation of 3D bioprinting to routine clinical
quality of 3D printing. For instance, they found that all sam- application, and this will depend on the aformentioned devel-
ples of alginate with calcium chloride showed shear thinning opments in research, reducing the cost and improving repro-
properties. Recovery time decreased as alginate concentration ducibility of the bioprinting process and overcoming significant
increased, but most samples could not recover after a few sec- regulatory/legal hurdles. To date, the FDA has not imposed
onds, and they needed more than 30 seconds to recover their specific restrictions on bioprinting technology which will facili-
viscosities to 83% of the initial values. They found that incor- tate future developments. Bioprinting potentially has to face a
porating graphene oxide increased viscosity and led to dif- dilemma for mass market application. Whilst the technology
ferent thixotropic properties with a better quality of the printed and products have to be scalable in order to be economically
construct. This is another study that has defined measureable viable, the advance of personalized medicine means that a
parameters in order to fabricate high fidelity and accurate 3D number of products will be made to order rather than available
printed constructs. ‘off the shelf’.
3D bioprinting as a field may be considered analogous to the
introduction of tissue engineering over 30 years ago—where
3. Conclusions interdisciplinary researchers including materials scientists,
cell biologists, engineers, computational modelers, biophysi-
Research in the field of 3D bioprinting is very topical, and cists and clinicians came together to try and help the problems
research efforts have accelerated worldwide over the last few with tissue/organ dysfunction and failure. Although the field
years. 3D printing was highlighted in The Economist as having has progressed slower than most hoped, modern technological
enormous potential.[68] The inception of new journals focusing and scientific advances coupled with advances in the internet

Adv. Healthcare Mater. 2017, 1700264 1700264  (14 of 16) © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advhealthmat.de

and advances in international collaboration have brought us Healthcare Mater. 2014, 3, 2004; f) A. X. Sun, H. Lin, A. M. Beck,
closer to creating artificial tissues/organs that closely mimic E. J. Kilroy, R. S. Tuan, Front. Bioeng. Biotechnol. 2015, 3, 115.
native structures. Bioprinting offers the opportunities to fine [12] a) S. Catros, J. C. Fricain, B. Guillotin, B. Pippenger, R. Bareille,
tune such processes to produce a paradigm shift in medical M. Remy, E. Lebraud, B. Desbat, J. Amedee, F. Guillemot,
Biofabrication 2011, 3, 025001; b) M. Gruene, M. Pflaum, C. Hess,
practice.
S. Diamantouros, S. Schlie, A. Deiwick, L. Koch, M. Wilhelmi,
S. Jockenhoevel, A. Haverich, B. Chichkov, Tissue Eng. Part C 2011,
17, 973; c) S. Michael, H. Sorg, C. T. Peck, L. Koch, A. Deiwick,
Acknowledgements B. Chichkov, P. M. Vogt, K. Reimers, PLoS One 2013, 8, e57741;
d) R. Gaebel, N. Ma, J. Liu, J. Guan, L. Koch, C. Klopsch,
Mr Steve Atherton RMIP MIMI, Medical Illustrator, ABMU Health Board M. Gruene, A. Toelk, W. Wang, P. Mark, F. Wang, B. Chichkov, W. Li,
for Figures in Table 1, and Figures 1– 3. G. Steinhoff, Biomaterials 2011, 32, 9218.
[13] a) N. E. Fedorovich, J. R. De Wijn, A. J. Verbout, J. Alblas,
W. J. Dhert, Tissue Eng. Part A 2008, 14, 127; b) A. Skardal, J. Zhang,
L. McCoard, X. Xu, S. Oottamasathien, G. D. Prestwich, Tissue
Conflict of Interest Eng. Part A 2010, 16, 2675; c) C. H. Lee, J. L. Cook, A. Mendelson,
The authors declare no conflict of interest. E. K. Moioli, H. Yao, J. J. Mao, Lancet 2010, 376, 440; d) F. Marga,
K. Jakab, C. Khatiwala, B. Shepherd, S. Dorfman, B. Hubbard,
S. Colbert, F. Gabor, Biofabrication 2012, 4, 022001; e) B. Duan,
L. A. Hockaday, K. H. Kang, J. T. Butcher, J. Biomed. Mater. Res. A
Keywords 2013, 101, 1255; f) L. E. Bertassoni, J. C. Cardoso, V. Manoharan,
A. L. Cristino, N. S. Bhise, W. A. Araujo, P. Zorlutuna, N. E. Vrana,
biofabrication, bioprinting, extrusion-based 3D printing, printability,
A. M. Ghaemmaghami, M. R. Dokmeci, A. Khademhosseini,
printing parameters
Biofabrication 2014, 6, 024105; g) J. Li, M. Chen, X. Fan, H. Zhou,
J. Transl. Med. 2016, 14, 271; h) W. Jia, P. S. Gungor-Ozkerim,
Received: February 28, 2017 Y. S. Zhang, K. Yue, K. Zhu, W. Liu, Q. Pi, B. Byambaa,
Revised: May 2, 2017 M. R. Dokmeci, S. R. Shin, A. Khademhosseini, Biomaterials
Published online: 2016, 106, 58; i) M. Muller, E. Ozturk, O. Arlov, P. Gatenholm,
M. Zenobi-Wong, Ann. Biomed. Eng. 2017, 45, 210; j) B. N. Johnson,
K. Z. Lancaster, I. B. Hogue, F. Meng, Y. L. Kong, L. W. Enquist,
M. C. McAlpine, Lab Chip 2016, 16, 1393.
[1] Z. M. Jessop, A. Al-Sabah, W. R. Francis, I. S. Whitaker, BMC Med. [14] R. Chang, J. Nam, W. Sun, Tissue Eng. Part A 2008, 14, 41.
2016, 14, 115. [15] J. Malda, J. Visser, F. P. Melchels, T. Jungst, W. E. Hennink,
[2] a) R. Langer, Mol. Ther. 2000, 1, 12; b) E. Lavik, R. Langer, W. J. Dhert, J. Groll, D. W. Hutmacher, Adv. Mater. 2013, 25, 5011.
Appl. Microbiol. Biotechnol. 2004, 65, 1. [16] a) W. Schuurman, V. Khristov, M. W. Pot, P. R. van Weeren,
[3] a) V. Mironov, Expert Opin. Biol. Ther. 2005, 5, 1111; b) V. Mironov, W. J. Dhert, J. Malda, Biofabrication 2011, 3, 021001; b) H. W. Kang,
ASAIO J. 2006, 52, e27; c) K. Jakab, B. Damon, A. Neagu, S. J. Lee, I. K. Ko, C. Kengla, J. J. Yoo, A. Atala, Nat. Biotechnol. 2016,
A. Kachurin, G. Forgacs, Biorheology 2006, 43, 509; d) B. Derby, 34, 312; c) L. Bae Hoon, S. Hitomi, C. Nam-Joon, T. Lay Poh, RSC
Science 2012, 338, 921. Adv. 2015, 5, 106094; d) Y. Shanjani, C. C. Pan, L. Elomaa, Y. Yang,
[4] a) Y. Ikada, J. R. Soc. Interface 2006, 3, 589; b) Z. M. Jessop, Biofabrication 2015, 7, 045008; e) D. B. Kolesky, K. A. Homan,
S. Al-Himdani, M. Clement, I. S. Whitaker, Front. Surg. 2015, 2, 52. M. A. Skylar-Scott, J. A. Lewis, Proc. Natl. Acad. Sci. USA 2016, 113,
[5] M. Muller, J. Becher, M. Schnabelrauch, M. Zenobi-Wong, 3179; f) S. Weinandy, S. Laffar, R. E. Unger, T. C. Flanagan, R. Loesel,
Biofabrication 2015, 7, 035006. C. J. Kirkpatrick, M. Van Zandvoort, B. Hermanns-Sachweh,
[6] B. Holmes, K. Bulusu, M. Plesniak, L. G. Zhang, Nanotechnology A. Dreier, D. Klee, S. Jockenhoevel, Tissue Eng. Part A 2014, 20,
2016, 27, 064001. 1858.
[7] C. Mandrycky, Z. Wang, K. Kim, D. H. Kim, Biotechnol. Adv. 2016, [17] Y. Loo, A. Lakshmanan, M. Ni, L. L. Toh, S. Wang, C. A. Hauser,
34, 422. Nano Lett. 2015, 15, 6919.
[8] a) I. T. Ozbolat, Y. Yu, IEEE Trans. Biomed. Eng. 2013, 60, 691; [18] F. P. Melchels, W. J. Dhert, D. W. Hutmacher, J. Malda, J. Mater.
b) K. Christensen, C. Xu, W. Chai, Z. Zhang, J. Fu, Y. Huang, Chem. B 2014, 2, 2282.
Biotechnol. Bioeng. 2015, 112, 1047; c) G. Gao, A. F. Schilling, [19] R. Levato, J. Visser, J. A. Planell, E. Engel, J. Malda,
K. Hubbell, T. Yonezawa, D. Truong, Y. Hong, G. Dai, X. Cui, M. A. Mateos-Timoneda, Biofabrication 2014, 6, 035020.
Biotechnol. Lett. 2015, 37, 2349; d) C. Tse, R. Whiteley, T. Yu, [20] a) C. Colosi, S. R. Shin, V. Manoharan, S. Massa, M. Costantini,
J. Stringer, S. MacNeil, J. W. Haycock, P. J. Smith, Biofabrication A. Barbetta, M. R. Dokmeci, M. Dentini, A. Khademhosseini,
2016, 8, 015017. Adv. Mater. 2016, 28, 677; b) A. Skardal, M. Devarasetty, H. W. Kang,
[9] A. Skardal, A. Atala, Ann. Biomed. Eng. 2015, 43, 730. I. Mead, C. Bishop, T. Shupe, S. J. Lee, J. Jackson, J. Yoo, S. Soker,
[10] D. Chimene, K. K. Lennox, R. R. Kaunas, A. K. Gaharwar, Ann. A. Atala, Acta Biomater. 2015, 25, 24; c) T. Zehnder, B. Sarker,
Biomed. Eng. 2016, 44, 2090. A. R. Boccaccini, R. Detsch, Biofabrication 2015, 7, 025001;
[11] a) H. Lin, D. Zhang, P. G. Alexander, G. Yang, J. Tan, d) X.-F. Wang, P.-J. Lu, Y. Song, Y.-C. Sun, Y.-G. Wang, Y. Wang,
A. W. Cheng, R. S. Tuan, Biomaterials 2013, 34, 331; b) Z. Wang, RSC Adv. 2016, 6, 6832.
R. Abdulla, B. Parker, R. Samanipour, S. Ghosh, K. Kim, Biofab- [21] a) A. Faulkner-Jones, C. Fyfe, D. J. Cornelissen, J. Gardner, J. King,
rication 2015, 7, 045009; c) J. A. Neiman, R. Raman, V. Chan, A. Courtney, W. Shu, Biofabrication 2015, 7, 044102; b) V. Lee,
M. G. Rhoads, M. S. Raredon, J. J. Velazquez, R. L. Dyer, R. Bashir, G. Singh, J. P. Trasatti, C. Bjornsson, X. Xu, T. N. Tran, S. S. Yoo,
P. T. Hammond, L. G. Griffith, Biotechnol. Bioeng. 2015, 112, G. Dai, P. Karande, Tissue Eng. Part C 2014, 20, 473; c) Y. Hu, Y. Wu,
777; d) N. J. Castro, J. O’Brien, L. G. Zhang, Nanoscale 2015, 7, Z. Gou, J. Tao, J. Zhang, Q. Liu, T. Kang, S. Jiang, S. Huang, J. He,
14010; e) S. Schuller-Ravoo, E. Zant, J. Feijen, D. W. Grijpma, Adv. S. Chen, Y. Du, M. Gou, Sci. Rep. 2016, 6, 32184; d) R. Fan, M. Piou,

Adv. Healthcare Mater. 2017, 1700264 1700264  (15 of 16) © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advhealthmat.de

E. Darling, D. Cormier, J. Sun, J. Wan, J. Biomater. Appl. 2016, 31, [41] J. H. Y. Chung, S. Naficy, Z. Yue, R. Kapsa, A. Quigley, S. E. Moulton,
684; e) S. Reed, G. Lau, B. Delattre, D. D. Lopez, A. P. Tomsia, G. G. Wallace, Biomater. Sci. 2013, 1, 763.
B. M. Wu, Biofabrication 2016, 8, 015003; f) J. H. Shim, K. M. Jang, [42] L. Ouyang, C. B. Highley, C. B. Rodell, W. Sun, J. A. Burdick,
S. K. Hahn, J. Y. Park, H. Jung, K. Oh, K. M. Park, J. Yeom, ACS Biomater. Sci. Eng. 2016, 2, 1743.
S. H. Park, S. W. Kim, J. H. Wang, K. Kim, D. W. Cho, Biofabrica- [43] A. G. Tabriz, M. A. Hermida, N. R. Leslie, W. Shu, Biofabrication
tion 2016, 8, 014102; g) Y. Xu, F. Fan, N. Kang, S. Wang, J. You, 2015, 7, 045012.
H. Wang, B. Zhang, Plast. Reconstr. Surg. 2015, 135, 451; h) H. Lee, [44] T. Guo, T. R. Holzberg, C. G. Lim, F. Gao, A. Gargava,
J. J. Yoo, H. W. Kang, D. W. Cho, Biofabrication 2016, 8, 015011; J. E. Trachtenberg, A. G. Mikos, J. P. Fisher, Biofabrication 2017, 9,
i) B. S. Kim, J. Jang, S. Chae, G. Gao, J. S. Kong, M. Ahn, D. W. Cho, 024101.
Biofabrication 2016, 8, 035013; j) F. Y. Hsieh, H. H. Lin, S. H. Hsu, [45] W. G. Kim, S. K. Cho, M. C. Kang, T. Y. Lee, J. K. Park, Int. J. Artif.
Biomaterials 2015, 71, 48. Organs 2001, 24, 642.
[22] S. Hong, D. Sycks, H. F. Chan, S. Lin, G. P. Lopez, F. Guilak, [46] U. A. Stock, J. E. Mayer Jr., J. Long Term Eff. Med. Implants 2001, 11,
K. W. Leong, X. Zhao, Adv. Mater. 2015, 27, 4035. 249.
[23] a) M. Kesti, M. Muller, J. Becher, M. Schnabelrauch, M. D’Este, [47] C. Stamm, A. Khosravi, N. Grabow, K. Schmohl, N. Treckmann,
D. Eglin, M. Zenobi-Wong, Acta Biomater. 2015, 11, 162; A. Drechsel, M. Nan, K. P. Schmitz, A. Haubold, G. Steinhoff, Ann.
b) K. Dubbin, Y. Hori, K. K. Lewis, S. C. Heilshorn, Adv. Healthcare Thorac. Surg. 2004, 78, 2084.
Mater. 2016, 5, 2488. [48] R. Sodian, J. S. Sperling, D. P. Martin, A. Egozy, U. Stock,
[24] L. Ouyang, R. Yao, Y. Zhao, W. Sun, Biofabrication 2016, 8, 035020. J. E. Mayer Jr., J. P. Vacanti, Tissue Eng. 2000, 6, 183.
[25] A. C. Daly, S. E. Critchley, E. M. Rencsok, D. J. Kelly, Biofabrication [49] S. Vafaei, D. Wen, T. Borca-Tasciuc, Langmuir 2011, 27, 2211.
2016, 8, 045002. [50] S. Vafaei, C. Tuck, I. Ashcroft, R. Wildman, Chem. Eng. Res. Des.
[26] a) A. Skardal, J. Zhang, G. D. Prestwich, Biomaterials 2010, 31, 2016, 109, 414.
6173; b) E. Hoch, T. Hirth, G. E. M. Tovar, K. Borchers, J. Mater. [51] K. Jakab, C. Norotte, B. Damon, F. Marga, A. Neagu,
Chem. B 2013, 1, 5675; c) K. Schacht, T. Jungst, M. Schweinlin, C. L. Besch-Williford, A. Kachurin, K. H. Church, H. Park,
A. Ewald, J. Groll, T. Scheibel, Angew. Chem. Int. Ed. 2015, 54, V. Mironov, R. Markwald, G. Vunjak-Novakovic, G. Forgacs, Tissue
2816; d) R. Lozano, L. Stevens, B. C. Thompson, K. J. Gilmore, Eng. Part A 2008, 14, 413.
R. Gorkin3rd, E. M. Stewart, M. in het Panhuis, M. Romero-Ortega, [52] C. Norotte, F. S. Marga, L. E. Niklason, G. Forgacs, Biomaterials
G. G. Wallace, Biomaterials 2015, 67, 264; e) A. L. Rutz, K. E. Hyland, 2009, 30, 5910.
A. E. Jakus, W. R. Burghardt, R. N. Shah, Adv. Mater. 2015, 27, 1607. [53] J. M. Perez-Pomares, R. A. Foty, Bioessays 2006, 28, 809.
[27] J. S. Lee, B. S. Kim, D. Seo, J. H. Park, D. W. Cho, Tissue Eng. Part C [54] K. Jakab, A. Neagu, V. Mironov, R. R. Markwald, G. Forgacs, Proc.
2017, 23, 136. Natl. Acad. Sci. USA 2004, 101, 2864.
[28] S. V. Murphy, A. Skardal, A. Atala, J. Biomed. Mater. Res. A 2013, [55] I. Kosztin, G. Vunjak-Novakovic, G. Forgacs, Rev. Mod. Phys. 2012,
101, 272. 84, 1791.
[29] J. Jia, D. J. Richards, S. Pollard, Y. Tan, J. Rodriguez, R. P. Visconti, [56] E. Flenner, L. Janosi, B. Barz, A. Neagu, G. Forgacs, I. Kosztin, Phys.
T. C. Trusk, M. J. Yost, H. Yao, R. R. Markwald, Y. Mei, Acta Biomater. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 2012, 85,
2014, 10, 4323. 031907.
[30] S. Wust, M. E. Godla, R. Muller, S. Hofmann, Acta Biomater. 2014, [57] K. Studer, C. Decker, E. Beck, R. Schwalm, Prog. Org. Coat. 2003, 48,
10, 630. 92.
[31] B. Duan, E. Kapetanovic, L. A. Hockaday, J. T. Butcher, Acta [58] D. B. Kolesky, R. L. Truby, A. S. Gladman, T. A. Busbee, K. A. Homan,
Biomater. 2014, 10, 1836. J. A. Lewis, Adv. Mater. 2014, 26, 3124.
[32] K. Markstedt, A. Mantas, I. Tournier, H. Martinez Avila, D. Hagg, [59] S. E. Bakarich, M. i. h. Panhuis, S. Beirne, G. G. Wallace,
P. Gatenholm, Biomacromolecules 2015, 16, 1489. G. M. Spinks, J. Mater. Chem. B 2013, 1, 4939.
[33] Y. Zhao, Y. Li, S. Mao, W. Sun, R. Yao, Biofabrication 2015, 7, 045002. [60] C. Khatiwala, R. Law, B. Shepherd, S. Dorfman, M. Csete, Gene
[34] K. S. Lim, B. S. Schon, N. V. Mekhileri, G. C. J. Brown, C. M. Chia, Ther. 2012, 7, 1.
S. Prabakar, G. J. Hooper, T. B. F. Woodfield, ACS Biomater. Sci. Eng. [61] P. Krontiras, P. Gatenholm, D. A. Hagg, J. Biomed. Mater. Res. B
2016, 2, 1752. 2015, 103, 195.
[35] A. Blaeser, D. F. Duarte Campos, U. Puster, W. Richtering, [62] M. Park, D. Lee, J. Hyun, Carbohydr. Polym. 2015, 116, 223.
M. M. Stevens, H. Fischer, Adv. Healthcare Mater. 2016, 5, 326. [63] S. Knowlton, B. Yenilmez, S. Anand, S. Tasoglu, Bioprinting 2017, 5, 10.
[36] H. Li, S. Liu, L. Li, Int. J. Bioprinting 2016, 2, 54. [64] K. Nair, M. Gandhi, S. Khalil, K. C. Yan, M. Marcolongo, K. Barbee,
[37] A. Shafiee, M. McCune, G. Forgacs, I. Kosztin, Biofabrication 2015, W. Sun, Biotechnol. J. 2009, 4, 1168.
7, 045005. [65] Y. Zhang, B. Liao, M. Li, M. Cheng, Y. Fu, Q. Liu, Q. Chen, H. Liu,
[38] L. A. Hockaday, K. H. Kang, N. W. Colangelo, P. Y. Cheung, B. Duan, Y. Fang, G. Zhang, F. Yu, Cardiovasc. Ther. 2016, 34, 308.
E. Malone, J. Wu, L. N. Girardi, L. J. Bonassar, H. Lipson, C. C. Chu, [66] J. Lu, Y. Fan, X. Gong, X. Zhou, C. Yi, Y. Zhang, J. Pan, J. Cell.
J. T. Butcher, Biofabrication 2012, 4, 035005. Physiol. 2016, 231, 1752.
[39] K. H. Kang, L. A. Hockaday, J. T. Butcher, Biofabrication 2013, 5, [67] K. H. Bouhadir, K. Y. Lee, E. Alsberg, K. L. Damm, K. W. Anderson,
035001. D. J. Mooney, Biotechnol. Prog. 2001, 17, 945.
[40] Y. He, F. Yang, H. Zhao, Q. Gao, B. Xia, J. Fu, Sci. Rep. 2016, 6, [68] Bioprint this, http://gelookahead.economist.com/bioprint/,
29977. (accessed: January, 2017).

Adv. Healthcare Mater. 2017, 1700264 1700264  (16 of 16) © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like