You are on page 1of 24

Bioprinting 27 (2022) e00219

Contents lists available at ScienceDirect

Bioprinting
journal homepage: www.sciencedirect.com/journal/bioprinting

Extrusion-based 3D printing of bioactive glass scaffolds-process parameters


and mechanical properties: A review
Bhargav Chandan Palivela, Sai Drupadh Bandari, Ravi Sankar Mamilla *
Department of Mechanical Engineering, Indian Institute of Technology Tirupati, Andhra Pradesh, 517506, India

A R T I C L E I N F O A B S T R A C T

Keywords: Tissue regeneration, proliferation, and cell attachment are the objectives of a scaffold for bone and cartilage-
Slurry related defect treatments. The development of novel biodegradable scaffolds is a case of exceptional research.
Extrusion This paper aims to review solid freeform fabrication additive manufacturing techniques based on slurry extrusion
3D printing
for the fabrication of bioactive glass and allied composite scaffolds and their controlled architecture and porosity.
Bioactive glass
Scaffold
Polymer or bioactive glass pastes with ceramic content ranges are also of interest here. Also reviewed the me­
Mechanical properties chanical characteristics, physical properties, and post-processing aspects such as sintering. However, the polymer
type and contents of ceramics establish the response in mechanical, mostly anisotropically. Extrusion-based
additive manufacturing technologies such as direct ink writing fabrication and robocasting are less significant
due to the technical backlog in making highly stable solid loading colloidal suspension inks with bioactive glass
powder. In this review, a discussion is about overcoming these difficulties with dispersant, binder, gelation agent
as processing additives, and ink rheological properties in satisfying the strict necessities of extrusion-based ad­
ditive manufacturing techniques. Flocculation bridging by partial coverage is essential in getting exceptional
performable inks.

1. Introduction the consolidation of the system printed by ceramic inks is by robocasting


[12]. This study reviews implementing this technique to build
The complex functionality requirements in tissue engineering, three-dimensional inorganic or organic hybrids for custom composition
structural, electronics, energy generation, etc., led to the demanding with fine porosity and attractive mechanical behavior. Bioactive glass,
inspiration in fabricating three-dimensional porous structures. Tradi­ with osteoinduction ability similar to a bone’s inorganic component, is
tionally, techniques like gas foaming, particulate leaching, solvent reviewed here. Bioactive glass is being widely on demand for the pro­
casting, etc. [1–4], to synthesize porous structures exist. Interconnec­ duction of biomaterials and allied composites from the discovery of the
tion, lack of proper control on porosity, poor reproducibility, need for first bioactive glass by Hench for its capacity to form an excellent bond
specialized tooling and molds, and consequently unpredictable me­ with osseous tissues generating hydroxyapatite [13].
chanical properties were some of the drawbacks of these traditional
processes. Minimizing or reducing these drawbacks is the critical driving 2. Bioactive glass and preparation methods
force in developing freeform fabrication techniques. These techniques
build a component layer by layer using colloidal inks. The deposition is Hench et al. [14] discovered a contemporary material in 1969 for
according to a design fed by a computer device with great precision and clinical utilization, which established a solid basis for five decades in
consistency, such as hot melt printing, robotic-assisted deposition or bone or tissue reconstruction field research. The bioactive glass with
robocasting, and direct inkjet printing [5–10]. Over the last few years, 45% SiO2, 24.5% Na2O, 24.5% CaO, and 6% P2O5 composition as wt %
robocasting has been one such technique gaining popularity: an named 45S5 was the first to be generated. This material can bond with
extrusion-based additive manufacturing technique [11]. A layer-wise the bone and stimulate bone growth because hydroxyl carbonate apatite
three-dimensional structure at room temperature is computer (HCA) is a similar phase formation of minerals in structure and chemical
controlled and deposited as a thick slurry. Through fluid gel transition, composition to hard tissue. This bone-bonding procedure to the

* Corresponding author.
E-mail address: evmrs@iittp.ac.in (R.S. Mamilla).

https://doi.org/10.1016/j.bprint.2022.e00219
Received 20 March 2022; Received in revised form 19 May 2022; Accepted 31 May 2022
Available online 20 June 2022
2405-8866/© 2022 Elsevier B.V. All rights reserved.
B.C. Palivela et al. Bioprinting 27 (2022) e00219

bioactive glass forms an adhesive interface capable of holding the me­ emulsion method, are produced [20].
chanical interactions with the tissue, which has been well preserved and
found in various bioactive glass compositions [15,16]. In numerous 2.1.4. Laser-spinning
scenarios, the adhesive interfacial strength is similar to or higher than A little volume of parent material is quickly heated, made to melt
cohesive material or bioactive tissue implant strength. The using a powerful power laser, and then cooled with a supersonic gas jet.
bone-bonding mechanism in bioactive glass has been described in the Because heating is under control by cooling at high speed, an amorphous
following five steps [17]. form of the product is the end product of this process. Bio glasses,
namely 52S4.6 and 45S5 nanofibers, are being made with the aid of this
Step (i): Ca2+ and Na + ions are quickly released and interact with the method [19].
solution’s H3O + ions, leading to a rapid pH increment.
2.1.5. Solvent-casting
Si–O–Na+ + H+ + OH− → Si–OH+ + Na+ (aq) + OH− … … … … (1)
Combining polymer and bioactive ceramics in a given organic sol­
vent and pouring solution in a readymade 3D mold to prepare scaffolds
Step (ii): Si (OH)4 is formed by the attack on the silica networks by of composites is called solvent casting. Once cast, the solvent is vented
hydroxyl groups present in solution. out by heating and evaporation. Non-requirement of any specialized
Si–O–Si + H2O → Si–OH + OH–Si … … … … (2) equipment simplifying the fabrication is the main advantage of this
technique. Poly (D, L-lactide) composite with 45S5 bioglass made by
Step (iii): − Si–OH groups called silanols develop a rich layer of silica solvent casting [21,22].
due to re-polymerization reactions on the surface.
Step (iv): Migration of PO3−4 , Ca
2+
forms a CaO–P2O5 film on the 2.1.6. Gas phase (flame spray) synthesis
previously formed silica surface. Since 1940 flame spray synthesis technique has existed [23],
Step (v): CaO–P2O5 film gets crystallized and includes other ions extensively used to produce megatons of silicon and titanium-based
from the solution such as CO−3 and OH− will form a layer of HCA that nanoparticles per year. Nano-size particles are produced at
is hydroxycarbonate apatite. Cao - P2O5 film will form the HCA layer high-temperature ranges more significantly at or above 1000 ◦ C by
when the film crystallizes using the ions such as OH and CO−3 in the creating nuclei, thereby coalescence and condensation of nuclei core,
solution. which leads to the growth of nuclei-core. The starting materials are
metal-organic precursor compounds (Boccaccini et al., 2010).

3. Slurry extrusion-based additive manufacturing of scaffolds


2.1. Preparation methods
Gel casting [24], solvent casting [25], slip casting [26], polymer
2.1.1. Quenching method foam replication [27], and freeze casting [28] are some of the traditional
Quenching or rapid cooling is the traditional method of cooling the methods used to fabricate 3D scaffolds. The solid freeform fabrication
molten liquid at its liquidus temperature to form glasses. The cooling (SFF) techniques can precisely control the scaffold design by CAD over
rate must be comfortably high enough to produce nucleation crystals traditional techniques. Few of them are stereolithography [29], selective
and develop crystallites. Usually, a platinum crucible calculates con­ laser sintering [30], 3D printing [31], fused deposition modeling [32],
stituent compounds in oxides or carbonates at 600 ◦ C–800 ◦ C in an extrusion [33] fabrication, and robocasting [12,34–37]. However, in
electrically operated furnace. At temperatures ranging from 1400 ◦ C to this review, 3D printing, robocasting, and extrusion additive
1600 ◦ C, the calcined charge is melted, followed by this homogenized manufacturing based on slurries are dealt with and discussed. However,
molten liquid quenching by pouring between two copper plates at room the discussion is restricted to bioactive glass scaffolds extrusion, as
temperature to form glass [18]. shown in Fig. 1, classifying the manufacturing process of ceramic scaf­
folds. Few critical conventional manufacturing processes, their limita­
2.1.2. Sol-gel method tions, and microstructures of scaffolds are as shown in Fig. 3.
The popular alternative for the melt-quenching technique is the sol- Direct write assembly, direct ink writing, robocasting, and free-
gel technique [19]. The sol-gel method’s chemistry-based synthesis forming extrusion are different synonyms of extrusion-based tech­
route is the compositional precursors undergoing hydrolysis and niques found in the literature. A paste is generally extruded onto a
condensation reactions at room temperature to form a gel. The sol-gel building table during fabrication through a nozzle. In detail, slurry in the
produces bioactive silicon dioxide-based glasses at lower temperatures form of a paste using air pressure from the nozzle tip is filled in a car­
using alkoxide of metals like Tetraethyl ortho-silicates (TEOS) as parent tridge and extruded, as shown in Fig. 2. By moving in x- and y-directions
materials. The steps involved in the sol-gel process are hydrolysis in using computer program control, scaffolds production is through the
basic or acidic conditions of an appropriate metal alkoxide. Then inhi­ layer-by-layer deposition. The paste remains stable, though, as it de­
bition of additives such as network modifiers to obtain the desired posits through a small nozzle on the building table. For this to happen,
product composition in solution followed by aging gelation of solution slurry with the required rheological behavior is essential in this method.
thereby obtaining dry powder by freeze-drying of gel. Extra-gaseous Scaffolds production is by adjusting the direction of the extruded rod by
substances are vent out by calcining the powder. specific angle variation from one layer to the other layer; besides, the
pore geometry variation is by the lay-down angle modification like 0/
2.1.3. Micro-Emulsification method 60/120◦ or 0/90◦ . The CAD data determines the size of pores, and the
Two non-miscible liquids stabilized by transparent surfactant mole­ rod’s diameter depends on the diameter of the nozzle and the deposition
cules, dispersion is isotropic and stable thermodynamically, is called a speed. Once fabricated and dried, densification of structures is achieved
micro-emulsion. Aqueous droplets in the water-based micro-emulsions mainly by de-binding and sintering after printing the scaffold [34].
act as a nanoreactor inside which the droplets have the required re­ Fig. 4 shows the parameters influencing bone tissue growth.
actants to contact each other. The parent hydroxide particles are Optimizing the following process parameters involved with
generated primarily in a microemulsion system. The desired oxide sys­ extrusion-based techniques are required to obtain suitable scaffolds per
tem is obtained at the required temperature, and calcination and drying the challenging requirements [12,29–37].
of parent powder at an appropriate temperature are over.
SiO2–CaO–P2O5 bioglass nanoparticles, with the help of the micro- • Rheological characteristics of the slurry

2
B.C. Palivela et al. Bioprinting 27 (2022) e00219

Fig. 1. Classification of manufacturing processes of ceramic scaffolds [12,29–37].

• Piston air pressure


• Diameter of the nozzle
• Paste temperature
• Velocity of the paste under extrusion
• Gap between building table and nozzle tip
• Strut diameter and inter-strut distance.

4. Review on extrusion-based additive manufacturing of


bioactive glass scaffolds

The influence of several variables on mechanical properties and di­


mensions of the printed scaffolds is presented in the following sections,
followed by in vitro and in vivo studies.

4.1. Influence of printing parameters and binders

The size of the ceramic particles played a significant role in deciding


Fig. 2. The slurry extrusion-based additive manufacturing [34]. the minimum achievable diameter of the thread in the deposition of
scaffolds at ambient temperature with inorganic component 45S5
bioactive glass and organic component polycaprolactone (PCL). The
printing speed and piston speed are related by relation (3) as

Fig. 3. Method of fabrication, process, limitations, and microstructures of scaffolds [38–42].

3
B.C. Palivela et al. Bioprinting 27 (2022) e00219

Fig. 4. Parameters influencing bone tissue growth [38–42].

φ
Vp = ( t )2 Vω (3) strength besides adding bio-activity to the implant [43]. SPCL with 70
φs wt % PCL, 30 wt % starch compression molded with BaG in layers
Vp-piston speed, φt-tip diameter, φs-syringe diameter, and Vw produced using single-screw extrusion resulting in at least 50% high
-printing speed. mechanical properties than the non-reinforced specimens with strength
Paste dried up in the nozzle for the low printing speed, clogging it in shear as 56% higher than that without BaG [44]. For bioactive glass
while depositing. On the contrary, at high printing speed, the line scaffolds in the freeze extrusion deposition process, the paste discharge,
diameter is not analogous; besides, the lines are discontinuous. For thin paste viscosity, and extrusion pressure are as per the relation (4) as
capillaries, swelling variation is by speed increment of the printing, and π R4 P
the diameter of the lines could be decreased four times, not affecting the η= (4)
8LQ
size of the pore [34] significantly. Starch poly ε caprolactone (SPCL) and
bioactive glass (BaG) 1–98 fiber composite have high initial mechanical R -nozzle diameter (mm), L-nozzle opening length (mm), Q

Fig. 5. Mesoporous bioactive glass scaffolds pore morphology and microstructure [45].

4
B.C. Palivela et al. Bioprinting 27 (2022) e00219

-discharge rate (mm3/sec), and P -extrusion pressure (Pa). With 13–93 4.2. Influence of shrinkage and porosity after sintering on printed
bioactive glass, grid-like microstructure scaffolds are freeze extrusion dimensions
deposited and sublimated, maintaining elastic modulus and compressive
strength within the human cortical bones range [33]. The rate of degradation and ability to convert into hydroxyapatite
Mesoporous bioactive glass scaffolds using polyvinyl alcohol (PVA) (HA) with no morphological and post-conversion particle size changes
as a binder by varying pore morphology from cubic to hexagonal are are strongly dependent on the size of the strut. The size of the particle is
extrusion deposited, as shown in Fig. 5. A considerable improvement in inversely proportional to the dissolution and reaction rate of the
mechanical strength and toughness with compressive strength was 200 bioactive glass scaffold [62]. Incorporating ten wt% bioactive glass
times [45] compared to the traditional polyurethane templated and sole particles less than 50 μm into PCL increased the scaffolds’ strength and
inorganic scaffolds [46], as depicted in Fig. 6. On the other hand, Yun Young’s modulus [63]. 13–93B3 and 13–93 scaffolds, with microscopic
et al. [47] and Garcia et al. [48] fabricated scaffolds without second structures oriented, have lesser compression strength than the grid-like
sintering following binder burnout with larger pores controllable for microstructure for segmental rat femurs defects [64]. 13–93 bioactive
tissue ingrowth and cell seeding. glass scaffold with polymer impregnated polycaprolactone (PCL), which
Using rheology studies of Herschel Bulkley model, 6P53B bioactive is ductile for denser strut structures, maximizing damage tolerance even
glass scaffold and 20 wt % F-127 solution which is homogenized and under bending stresses, rough manual milling, toughening, and enabling
pre-cooled in an ice bath, thereby lowering viscosity (Fig. 7(a)) which machining. Polylactic acid (PLA) for higher compressive strength and
nicely fits the shear rate and viscosity relation (5), [49]. stiffness, except in the case of bending strength. A comparison between
the raw polymer and bioactive glass impregnated polymer with the aid
τ = τy + Kγn (5) of uniaxial compressive stress-strain curves, as depicted in Fig. 11 [65].
where τ - shear stress, K -consistency index, τy -yield stress, γ -shear rate, Composite scaffolds from FastOs® bioactive glass-reinforced in the PLA
and "n" a representation (<1) for shear-thinning fluid. Here τy is a radial matrix led to a significant direct proportional increment in all material
varying stress parameter for extrusion-based continuous deposition properties such as elastic gradient and yield stress; however, the plateau
approach. Regardless of the composition of the ink, the temperature is stress remained approximately constant.
directly proportional to the load, that is, 5–10 N at 0 ◦ C–40 ◦ C, respec­ In particular, an alkali-free FastOs®BG composition containing
tively. However, with the initial steadiness following the steady-state CaMgSi2O6 (diopside) (Di), Ca5(PO4)3F (fluorapatite) (FA), 3CaO•P2O5
flow as depicted in the shadowed zone in Fig. 7(b). The morphology (tricalcium phosphate) (TCP) that is Di70–FA10–TCP20 (wt.%) desig­
of the anatomic pattern printed is as in Fig. 8 (a). Scaffolds from 6P53B nated as FastOs®TCP20 was availed [67]. However, the summary of
bioactive glass printing in the cooling environment caused nozzle bioactive glass composition and bioactive glass type reference is pre­
clogging and water segregation when extruding the gel-like inks through sented in Fig. 12 and appendix table 2, respectively. Micro-computed
the narrower tips (less than 100 μm). However, lower viscosity inks with tomography data of robocasting bioactive glass and PCL scaffold
micrometer size ceramic particles, capillaries in the range of 30 μm analyzed using finite element analysis. The outcomes revealed low
extruded without clogging [50], and 30 μm filaments made from inks elastic stiffness than spongy bone tissue, naturally triggering stress
containing barium titanate nanoparticles, Li et al. [52]. concentration at the tissue interface and in-homogeneous scaffold
Chitosan (Chit) is a polysaccharide with medical applications, deformation. Novel algorithms of erosion mechanism that is surface
including wound dressing, tissue engineering, and carriers of drugs [53, erosion and bulk erosion model showed a direct correlation between
54]. Chit with ten wt% nano bioactive glass (n-BG), other polymers [55] porosity, strength, stiffness, and increase in porosity of scaffold gradu­
imparted higher bio-activity to the composite with dual pore structure ally, material properties decay exponentially [68]. Bioceramic scaffolds
retention in the frozen atmosphere as low as − 50 ◦ C. The stem of the with high strength calcium silicate (CSi) added with x% BG resulting in
scaffold displayed a higher microporosity structure with a pore size CSi-BGx (x = 1,2,3), are produced using low melting point bioactive
ranging from fewer to 10 μm and shrinkage when an organic solvent is glass (BG). The morphologies of the pores are rectangular, Archimedean
used [56], and the morphology is as shown in Fig. 9. Using a single, chord, parallelogram, and honeycomb structures. In particular sole
multi-functional additive carboxymethylcellulose (CMC) of high mo­ honeycomb pore, CSi-BG1, displayed higher compressive strength than
lecular weight and viscosity range as 0.3–0.4 Pas to 1–3 Pas, respec­ the CSi-BG0 and other structures of the pores. No remarkable variation is
tively, two wt% aqueous solutions enabled increased solid loading up to noticeable in the case of the parallelogram pore morphologies [69].
45 vol%. Overcoming intrinsic aqueous difficulties of processing 45S5® PLA coated robocasting of 45S5 bioactive glass displayed higher
bioactive glass is possible, thereby retaining higher compressive toughness and strength than scaffolds bare in bending and compression
strength [57] (Fig. 10 (a)) far superior to solid free form fabricated [86]. Glass compositions 13–93, SBP-3 and ICIE16 with an extended
scaffolds at 50% porosity [58]. thermal processing window, network connectivities modified. The
average number of Si–O–Si bonds bridging per silicon atoms of 2.84,

Fig. 6. Compressive strength vs. scaffold deformation for mesoporous bioactive glass scaffolds (a) 3D printed and (b) conventional polyurethane templating [45].

5
B.C. Palivela et al. Bioprinting 27 (2022) e00219

Fig. 7. Viscosity vs. shear rate (a) and load vs. displacement curve of the ink (b) [50].

Fig. 8. Anatomic pattern (inset gross view of the sintered scaffold) (a) [50], Morphology of direct ink written WB-15 (defective silica glass) with silicone binder (b
and c) [51].

Fig. 9. Morphology of the Chitosan/nano bioactive glass robotic dispensed scaffolds [56].

3.01, and 2.13, respectively, resulted in viscosity increase directly pro­ below natural bone under dry and wet conditions [88]. Inks made from
portional to wt% of Pluronic, F-127. Ink composition variations resulted CMC as binder or dispersant and ε -PCL polymer infiltrated into 45 vol%
in an increase in the force of extrusion in SBP-3, ICIE 16, decrease in the milled powders of 45S5 Bio glass®. The scaffolds sintered via melt
case of 13–93. Besides, tip shape also played a significant role [82]. impregnation obtained denser semi-continuous composites besides via
However, the summary of printing parameters and bioactive glass dip-coating up kept pre-designed macroporosity. Strength in compres­
printed type reference is presented in Fig. 14 and appendix table 2, sion of the composite scaffold platform and the toughening effect under
respectively. Nanoparticles (NP) functionalization with the aid of the flexural stresses varied directly with the PCL concentration compa­
3-Aminopropyltriethoxysilane (APTES) reagent on chains of carboxylic rable to cancellous bone properties [89]. Titanium fibers having mean
acid (COOH) groups was carried out [87]. tensile strength within the 246 MPa–370 MPa range would reinforce the
The scaffolds were frozen at − 20 ◦ C and freeze-dried for three days brittle glass matrix by transferring the bending stresses to the fibers from
after printing. The age of the APTES reagent, lesser than two months, the matrix [90]. 13–93 bioactive glass scaffold that is silicate-based, on
was critical in successful nanoparticles functionalization. The addition reinforcement with 0.4 vol% Ti fiber showed 70% increase in fracture
of calcium and zinc did not significantly affect the particle’s zeta po­ toughness and 40% increase in flexural strength, modulus, and tough­
tential, nor did the addition of cations jeopardize the ability of func­ ness similar to human trabecular bone in comparison to the without
tionalized nanoparticles. The scaffolds with 20 wt % NP maintained fibers [91]. Sol-gel prepared Strontium hardystonite Gahnite (Sr-HT
structure better than without NP with higher compressive strength well Gahnite) [92,93] with a unique triphasic microstructure, with mean

6
B.C. Palivela et al. Bioprinting 27 (2022) e00219

Fig. 10. Robocasted different scaffolds geometries from a 45S5 Bio glass® (a) [57], WB-15 (defective silica glass) with silicone binder (b) [51] and composite
three-dimensional scaffolds deposited with a internal hierarchical pore structure (c and d) [59], Ethanol milled 13–93 bioactive glass scaffold pre (white) and post
sintering (black) (e) [60], 0, 0.4, 0.8 and 2.0 wt% CuO doped 13–93 bioactive glass (f) [61].

uniform grids, re-distribution of longer filaments starting from layers in


the middle and up to the bottom and top layers resulted in a decrease in
tensile stress within base lying layers and vice-versa in the case of the
shorter filament case, as depicted in Fig. 13 [96].
Scaffolds fabricated from 45S5 bioglass with PCL coating concen­
tration (5–30% (w/v)) consisting of a tetragonal cylindrical rod mesh.
Scaffolds showed a four-fold increase in the toughness and compressive
strength and directly proportional to concentration from ~3.0 to ~12
MPa and viscosity of 1–4Pas. Primarily polymer concentration and
secondarily process temperature, solvent selection determined the ulti­
mate mechanical properties of the polymer-coated scaffolds [97]. Mes­
oporous bioactive glass (MBG) polycaprolactone (MBG-PCL) scaffolds
stimulated colonization with textural properties: specific surface area
(SSA) (m2⋅g− 1), specific pore volume (SPV) (cm3⋅g− 1), size of pores (SP)
(nm) as ~0.7, 0.003 and ~10 respectively. A significant increase in the
differentiation and cell proliferation is noticeable, allowing cell move­
ment along with surface-reaching upper stratus faster than in bare PCL
having SSA as ~1.2 and bare MBGs’ SSA, SPV, SP as ~178, ~0.29, and
~5.6, respectively [98]. Finite element study combining 13–93 bioac­
tive glass scaffold geometry and polymer placement layer with ~20 μm
Fig. 11. Uniaxial compressive stress-strain curves [65]. mesh size, quadratic tetrahedral elements, and layers in multiple of the
filaments within the long direction of beam (filament diameter = 330
μm) and in the orthogonal direction as shorter filaments (S). Among
particle size 1 μm displayed failure probability of around 0.01. In the
various multiple layers, especially L4S1 structure besides adherent
case of scaffolds of calcium phosphate with porosity similarly, the
surface layer of PLA from 400 μm to 800 μm thickness provided reli­
likelihood of failure was 0.8 [94], and the scaffolds with ~60% porosity
ability for bone repair with higher load-bearing capacities, strength
withstood 106 cycles for cyclic stresses of 1–10 MPa. As and when
under compression as ~88 MPa and strength under bending as ~34 MPa
porosity increased by ~70%, only hexagonal patterned scaffold with­
[99]. Finite element simulation on bioactive glass 13–93 silicate-based
stood 106 cycles, highlighting hexagonal scaffold conditions equivalent
scaffolds with 5 mm cube shape length, pore width ranging from 300
to everyday physiological stresses exceeding all technical foam mate­
to 900 μm, layer spacing ranging from 100 μm to 200 μm, and grid-like
rials and metals and alloys with a range of strength within human
microstructure revealed good mechanical response in compression.
cortical bone. However, with the increase in the cyclic stress amplitude
Elastic modulus was as high as 13 MPa when the Poisson’s ratio was
from 3 MPa to 30 MPa, lowering the average life of fatigue of the scaf­
0.26. The compressive strength was higher at a higher pore width,
folds from 60 to 70% porosity is evident, which in turn, also varied with
medium layer spacing, and at a constant load applied in the direction
the pattern type such as zigzag, curved, hexagonal, and rectangular
perpendicular to the deposition plane [100]. However, a summary of the
[95]. Finite element modeling revealed a cross-section of long bones
number of layers deposited and scaffolds dimensions and the aspect
(Fig. 13 (a)) of more minor porous limbs around ~5–10% outermost
ratio in the case of cylindrical scaffolds, shrinkage, and the corre­
region to better mimic human long cortical bone. The inner area with
sponding bioactive glass type reference is in Fig. 15(a–d) and appendix
porosity range reported for human cancellous bone showed flexural and
table 2, respectively. Bioactive glass composition with alkali-free Fas­
compressive strength double the value of microstructure containing
tOs®BG of diopside fluorapatite tricalcium phosphate prepared via melt
uniform grids, comparable in vivo capacity. Compared to the model with

7
B.C. Palivela et al. Bioprinting 27 (2022) e00219

Fig. 12. Various studies reported bioactive glass compositions (Refer to Appendix table 2 for bioactive glass type reference).

Fig. 13. Dimensions of the model (a), three-dimensional planar view of the load application at four points (red arrow) (b) in finite element method simulations [96].
(For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)

quenching route enabled higher solid loading as high as 47 vol%. temperature of gelation with increased glass loading within the inks. As
Adding two wt% binders as hydroxypropyl methyl-cellulose (HPMC) the size of the particles used is larger, the inherent porosity of the strut
and one wt% Aristoflex®TAC thickener as gelation agent to the starting was to limit the mechanical strength of the produced scaffolds following
suspension significantly enhanced the apparent viscosity for extrusion the sintering operation. Morphologies of the scaffolds fabricated from
[101]. The as-received HA nanopowders with elongated shape of ~20 the different compositions are as depicted in Fig. 16 [78].
nm, calcined with the spherical shape of ~200 nm, and sol-gel prepared A two-syringe system to print the three-dimensional porous struc­
closer to the spherical shape of ~10 nm. They are used as a reinforce­ tures or as a cell carrier to generate a scaffold cellularized with PCL.
ment phase within a polycaprolactone matrix, deposited on the 45S5 13–93B3 glass composites as scaffold material and hydrogel pluronic as
bioactive glass strut surface. Among all the sol-gel hydroxyapatite par­ supporting material. The glass composite filaments showed an increase
ticles, reinforcement resulted in a 200% increment in both the in the dwell time from ~2 min to 4 min required to fabricate
compressive strength and the toughness compared to bare or no coating. PCL+1393B3 composite and hydrogel, respectively, for each layer.
However, whether filler particle addition influenced positively and the Measurements of weight loss and micrographs confirmed hydrogel
scaffold response biologically needs to be addressed [102]. A compari­ presence after seven days of soaking within the culture media amidst
son made between PSrBG, ICIE16 compositions containing less than 50 pluronic unstable nature in case of the multi syringe printing [107].
mol % silicon dioxides, thereby giving network connectivity of silicates Alkali-free bioactive glass composition consisting of 30 wt% trical­
closer to that of the 45S5 with 13–93 bioactive glass constituents, cium phosphate (TCP) and 70 wt% diopsides (Di) revealed that it is
showed that the Pluronic-F127 can be availed as a binder universally possible to reach higher elastic modulus values even with the lower solid
regardless of the reactivity of the glass. Then again, the printability of loading of 40% provided increment in the amount of CMC to 3 wt%
the ink is affected by the particle size, besides a stronger correlation besides additives addition. On the other hand, decreasing solid loading
between the force required for extrusion of the inks loaded with glass in the presence of 2 wt% of CMC caused accentuated decrease in elastic
and particle distribution. There is a reduction in the Pluronic F-127 modulus. A deeper analysis of the steric effect of additives could aid in

8
B.C. Palivela et al. Bioprinting 27 (2022) e00219

Fig. 14. Printing parameters reported in various studies (#Extrusion force (N)) (Refer appendix table 2 for bioactive glass type reference).

controlling the suspension’s rheological properties [70]. The weight loss extending the span of particle size distribution while reducing mean
outcomes of bioactive borate glass in PCL/13-93B3 composite indicated particle size with an increase in BPR [125]. Enhanced biological activity
faster dissolution in the case of solvent-based scaffolds. The formation of is noticeable by honeycomb extrusion of hydroxyapatite bioactive glass
the apatite-like layers in melt extrusion and solvent-based three-di­ that is BG1 as ten wt% stoichiometric Bio glass®, BG2 as calcium excess
mensional printing is noticeable. There is an increase in stiffness and Bio glass® or CAN as Canasite scaffolds [77]. Doping HSSGG with five
decrease in yield strength after glass addition irrespective of solvent or wt% Cu2+ (Cu5) or five wt% La3+(La5) significantly enhanced the
melt extrusion-based, which is evident. A four-fold and two-fold increase scaffold’s mechanical properties. Rheology properties of solid loaded 40
in elastic modulus is observed in melt extrusion-based and solvent-based vol% inks were affected, affecting neither paste printability nor shape
scaffolds, respectively. retention as printed scaffolds. By doping Cu5 or La5 in parent glass and
Upon bioactive glass addition to the melt PCL, the average elastic particularly in Cu5, increments up to 221% were possible compared to
modulus increased drastically from ~4.5 times, whereas solvent-based those fabricated from HSSGG [73]. Indirect inkjet 3D printing, as
was ~1.98 times. Also, the slower rate of glass dissolution, brittle fail­ depicted in Fig. 17, combined with systems of freeze-drying.
ure after glass addition in case of melt extrusion-based and tensile fail­ A novel porous hierarchical silk fibroin (SF) made micro bioactive
ure, and faster dissolution of glass in the case of solvent-based highlight glass (μBG) exhibited excellent compressive strength and modulus in
potential for three-dimensional bioprinting. Then again, the porosity dried conditions and with no considerable effect on mechanical char­
volume and compressive strength percentage decreased with a decrease acteristics in wet conditions retaining flexibility without any fragility.
in the sintering temperature [107]. Scaffolds fabricated from 45S5 Bio The printed scaffold geometry is depicted in Fig. 10 (c and d). The
glass® (80–100%) and β-TCP (β-tricalcium phosphate) composite using surface roughness and contact angle of silk fibroin, silk fibroin nano
CMC as a single multi-functional polymer with values of the strength bioactive glass (SF-n-BG) besides SF-μBG are ~ 4, ~37, and ~44.5 nm,
within the typical ranges of human cortical bones. However, density ~79, ~63.65, and ~73.9 o, respectively, as shown in Fig. 18 [59].
decrement, and in the case of pure TCP, where other than CMC additives Scaffolds based on fine silica defective powders of glass
hampered sintering of β-TCP and densifying, negatively affected robo­ (morphology-figure 8(b), 8 (c)) that mix with a reactive silicone binder,
casting parts’ mechanical performance [123]. High-silica-content sol-­ as depicted in Fig. 19, exhibited excellent strength in compression based
gel bioactive glass’s (HSSGG) suspensions with three various solid on firing within the air. The firing caused shrinkage homogeneously
volume fractions in vol% as 30, 35, and 40 vol% exhibited throughout without edges rounding, and the colloidal silica as an extra
shear-thinning characteristics and accentuated to more modulus of filler in the formulation of inks led to the refinement of the phase as­
elasticity of ink at 40 vol % suitable for printing [72]. The quaternary sembly [51].
bioactive glass of higher silica content modeled by wet milling and In the fabrication of 45S5 bioactive glass content ranging from 44.8
ethanol as controlling process agents led to setting a limit of commi­ to 67.5 wt% with the aid of CMC as a binder and as a dispersant, 67.5 wt
nution for an average size of the particles in the order of 1.0–1.5 μm with % 45S5 bioactive glass displayed the most defined extrusions. On the
a significant effect of milling time [124]. Scaffolds with a high apoptosis other side of the coin, lower viscosities accumulate more ink at cross-
13–93B3 borate bioactive glass with 20 wt% PCL addition were pro­ sections, and ink filling holes besides 60 wt % bioactive glass exhibi­
duced using near field electro-spinning technique with an electric field ted shear-thinning qualities required for ceramic extrusion-based
of 10 kV/cm, reported higher bioactivity [109]. The temperature of printing [122].
calcination and BPR, that is, the ratio of balls to powder upon wet
milling in the aqueous media of bioactive glass that is sol-gel, negatively
affected porosity fraction and specific surface area. Thereby also

9
B.C. Palivela et al. Bioprinting 27 (2022) e00219

Fig. 15. (a–d). Scaffold dimensions and shrinkage from various studies (Refer to appendix table 2 for bioactive glass type reference).

4.3. Influence of sintering temperature secondary porosity than the CPC and 8% CPC-MBG scaffolds [103].
Phosphate-based bioactive glasses at two sizes of particles less than 38
Primarily two scales of local and global densifying were observed μm and 125–250 μm with Pluronic® F-127 as a binder showed crystal­
using synchrotron X-ray tomography. With the aid of four-dimensions lization peaks post sintering. The mechanical strength of 10 MgO +5 SrO
imaging, quantification of sintering viscous flow is of interest in the and 15 MgO +5 SrO is in the range reported for cancellous bone and
porous scaffold of bioactive glass 13–93. Sintering was with the assis­ modulus in trabecular bone. Ion release was inversely proportional to
tance of a two-stage process. A continuation of stage 2 for a prolonged the particle size. After one day, ion release became constant for less than
time is stage 3, observing that the local sintering dominates in stage 1 38 μm particle size, confirming ion dissolution dependency on the par­
and stage 2, whereas global sintering in stage 3 [76]. Composites syn­ ticle size [71]. Functionally graded bioactive glass scaffold (BG30) had a
thesized from non-aqueous carrier liquid designated as cl, based on compressive strength near the lower limit of human trabecular bone. It
particles of MBG and calcium phosphate cement (CPC) paste. CPC with increased after reinforcement with nanofiber with six layers printed,
the addition of 10 wt% mesoporous bioactive glass (MBG) by changing maintaining center to center distance between the strut as 1.25 mm, 1.5
the amounts of cl and scaffolds containing 4%, 6%, 8% calcium phos­ mm, and 1.875 mm for the first, second, and sixth layers, respectively
phate cements-mesoporous bioactive glass composites are of interest. [85]. Scaffolds were manufactured with the aid of 47.5 B glass by
MBG composition-controlled release of Sr2+ and Ca2+ ions improving superimposing 20 identical layers 8.91 mm long with 16 straight lines,
degradation due to increased surface area. Not dependent upon the each tilted by 90o in reference with the one underlying and outer six
needle diameter, there were no changes in the profile of extrusion of from each side spacing 0.636 mm apart besides inner tubes 0.51 mm
MBG with a fraction of less than 45 μm in comparison to the incorpo­ apart setting Weibull scale parameter as σ0 = 0.14 [79]. Phosphate glass
rated mesoporous bioactive glass with all fractions lesser than 180 μm. foams attained fracture energy of 20 kJ/m3 [126] and increased to 150
Therefore, concluding that for mesoporous bioactive glass particles, less kJ/m3 by polyethylene burn-off [127]. Baino obtained a significantly
than 180 μm show a more secondary role for lower concentrations of higher value of 544 kJ/m3 than Vitale Brovarone et al., using sponge
calcium phosphate cements-mesoporous bioactive glass composite ink replication to produce SiO2–CaO–Na2O–Al2O3 based glass-ceramic
extrusion. At high concentrations, particle size influenced the extrusion scaffolds [128]. Solvent-dependent extrusion printing, as depicted in
behavior. On the other side of the coin, the tendency of MBG particle size Fig. 20, for the fabrication of cellular besides acellular scaffolds
lesser than 45 μm to agglomerate by cluster formation hindered homo­ mimicking bone tissue with the accuracy of ±10 μm using PLA and PLA
geneous MBG-CPC paste preparation. CPC-cl scaffold showed more + B3 showed good potential for tissue engineering applications.

10
B.C. Palivela et al. Bioprinting 27 (2022) e00219

Fig. 16. (a–c) Green bodies before sintering, (d–e) after sintering and row (a) z-direction cross-sections of struts, row (b) x-y struts, row (c) packing of particles, row
(d) after sintering, row (e) surface at higher magnification after sintering [78].

The yield strength of bare PLA was more than that of the PLA+ 33 wt appendix table 2. Limitations of melt electro writing (MEW), when
% B3 glass. The PLA+ 50 wt % B3 glass displayed a higher value than the combined with the prediction model required for extruding PCL incor­
bare PLA and PLA +33 wt % B3 glass [106]. Scaffolds with a solid porated into a composite with SrBG (strontium bioactive glass), are
loading of 40% volume at a calcining temperature of 800 ◦ C and mes­ overcome. With the aid of a predictive framework and using the
opores lesser than 30 nm followed by 800 ◦ C sintering for 2 h showed non-Newtonian rate of the shear model, the average velocity for
compressive strength in the range reported for human cancellous bone extruding polymer from the nozzle is 2.39 mm min− 1.
at 36% porosity. The mean particle size reduction that is most accen­ Viscosity lower than raw composite, shear-thinning, and higher
tuated was at balls to powder ratio of 15. Most good rheological char­ quality was noticeable from shear viscosity profiles for 1, 0.25 g PCL or
acteristics could be obtained at as low as 25 vol% solid loading and balls 33 wt % SrBG dissolved in the 1 mL of CHCl3 that is chloroform (CF)
to powder ratio of 5–10. The calcination temperature of 600 ◦ C resulted [81]. Comparison between three-dimensional printing and porogen
in an elastic modulus of ~1 MPa and specific surface area, pore volume burn-off technique using ammonium hydrogen carbonate (NH4HCO3) as
as ~100–130 m2 g− 1, 0.34–0.48 cm3 g− 1, respectively [129]. Like this the porogen is of consideration here. Two groups, namely silicate glass
study, the summary of sintering and binder burnout studies is as in B12.5 (12.5 mol% of SiO2 substituted by B2O3) and B12.5-Mg5-Sr10
Fig. 21(a–c), with corresponding bioactive glass type reference in (replacing the CaO with 5 mol % of magnesium oxide (MgO) and 10

11
B.C. Palivela et al. Bioprinting 27 (2022) e00219

Fig. 17. Schematic diagram of fabricating 3D scaffolds [59].

Fig. 18. Atomic force microscopy images of a) silk fibroin, b) silk fibroin nano bioactive glass and c) silk fibroin micro bioactive glass [59].

Fig. 19. Schemes of processing ceramics of glass that is devitrification conventionally (left) in comparison to the heat treatment of glass mixtures that are silica
defective or silicone (right) [51].

mol % of strontium oxide (SrO)), are of interest. Composition with 30 ions. However, Mg concentration remained constant regardless of the
vol% (glass powder): 70 vol% (porogen), which is non-toxic, is intro­ scaffold composition and type under the dynamic conditions. The pore
duced into the scaffolds before sintering, revealing three-dimensionally area (μm2), width (μm), length (μm), porosity (%), and volume (mm3)
printed scaffolds with Mg and Sr exhibited the best cell viability per­ are lower for mixed 3D in comparison to B12.5 3D [108]. Fracture
formance dependent upon the size of the scaffolds limiting the release of toughness value using the extended finite element method with

12
B.C. Palivela et al. Bioprinting 27 (2022) e00219

Fig. 20. (a) Bioprinter schematic, (b) Solvent-based extrusion bioprinting process for cellularized scaffolds [106].

computational fluid dynamics simulations of the ceramic Sr-HT-Gahnite tomography-based finite element modeling, with scaffolds, printed fi­
scaffolds is estimated to be within the range near the human cortical bers along with the two directions in perpendicular on the parallel layers
bone. The porosity was 55%, 0.275 mm strut radius, and DAM (design with 90o tilting in between the two layers. The selected design param­
for additive manufacturing) parameters h/r (strut radius (r), over­ eters majorly affecting strength, stiffness including relative density,
lapping depth(h)) equal to 0.5 and 0.9 as depicted in Fig. 22 [132]. porosity, fiber misalignment, and fiber spacing between the layers, are
among the most prominent ones. Strength; elastic modulus is in agree­
4.4. Influence of process parameters and temperature on mechanical ment with that found on the ideal geometry at similar porosity macro­
properties scopically and partially validating the model showed that the
computational procedure proposed was able to estimate how the me­
Borosilicate glass with higher alkaline oxide and alkali content, as chanical characteristics of the scaffolds vary with varying geometrical
milled glass powder with 15 wt %, 25 wt %, and 35 wt % mixed with 30 parameters [112]. Three-dimensional scaffolds with pores that are
wt% pluronic stock solution revealed direct proportionality between the highly ordered and with diameters of ca. 4 nm, 30 μm–80 μm range with
concentration of glass inks and viscosity of the pluronic solution. Elastic the interconnections of 2 μm–4 μm and 8 μm–9 μm, respectively, besides
and viscous modulus in pluronic solutions changed from 27 to 17%, the ultra larger pores of ca.400 μm provided scaffolds with the capability
leading the way for the converse trends that is elastic modulus can re- of biologically hosting the active molecules [47]. Mesoporous bioactive
build up to 73.9%, 71.1%, and 97.2%, followed by shearing under glass substituted scaffolds (MBG_Scs) with Ce2O3, ZnO around 0.4% or
500% and the left-out structure undeformed for 81 s for increasing the 2.0%, and Ga2O3 around 0.2% or 1.0% by varying molar percentage
solid loading of the ink [75]. Porous scaffolds fabricated from concentration for the proliferation of cells and vascular ingrowth syn­
yttria-stabilized zirconia (YSZ) powder, optimized ink with the 48 vol% thesized from un-substituted scaffolds (B_Sc) with combined rapid pro­
solid loading enabled for producing the defect-free scaffolds with totyping technique and evaporation induced self-assembly. Textural
considerable shape retention and good ability for sintering. The properties evaluation showed a slight decrease in the pore and surface
as-sintered scaffolds exhibited macro size of the pore and macroporosity area volume with no considerable effect on pore diameter. The porosity
suitable for the growth of the bone. Upon simulated body fluid (SBF) hierarchy has resulted with the aid of non-ionic P-123 surfactant as
immersion of the scaffolds sintered for shorter periods, the extensive mesostructured agent directing. PCL as a microporous template and for
ability of the mineralization is evident in coating the sintered scaffolds’ ultra larger macropores’ rapid prototyping [113]. Novel composite
surface with the bioactive glass. Therefore, confirming their ability to bio-inks with the aid of β-TCP (TCP-tricalcium phosphate) and 45S5
bond to the tissues and thus suitable for load-bearing applications [110]. bioactive glass powder with carboxymethyl cellulose as a unique
Slurry with a solid loading of 10 wt% for deposition as a coating on the multi-functional additive for processing, sintered at temperatures
surface onto the YSZ scaffold filaments, upon adding coagulating agent 1000 ◦ C–1100 ◦ C. In particular, sintering at 1100 ◦ C improved 50–100%
displayed elastic modulus as high as more significant than or equal to 10 strength. However, the sample with 80 vol% 45S5 bioactive glass dis­
MPa [133]. played no strength improvement associated with an increase in tem­
These scaffolds performed exceptionally well in physiological senses perature beyond 1000 ◦ C. Models with lower or null 45S5 bioactive glass
such as standing, walking, and even in more extreme cases such as contents revealed material porously with barely any effect of sintering
running, jumping, etc. The ultimate maximum generated stress is ~70 between two particles. The bending strengths and densities of the rods
MPa undergone by femur bone was lower than the mean stress sintered are noticeable to increase beyond 60 vol % of 45S5 bioactive
[111–113]. As per the Weibull statistics, the scaffolds are 99.8% reliable glass content, thereby diminishing porosity, which was barely notice­
when the difference between bioactive glass coating and YSZ substrate able at a composition of 80 vol%. Theoretically, the relative density of
coefficient of thermal expansion (CTE) is 0.3 × 10− 6 K− 1. As a result, the pure TCP samples with CMC as a sole additive was sixty-eight percent of
coating would be undergoing slighter stresses in compression and upon the density. The strength values of glass containing 80%–100% 45S5
reduction in temperature, beneficial for enhancing the glassy materials’ glass composition are even above or within the range reported for the
mechanical characteristics compared to that of the other different bio­ performance of human cortical bone. The addition of 45S5 bioglass to
glasses with a coefficient of thermal expansion as ~14 × 10− 6 K− 1 to 15 tricalcium phosphate was not defined to be benefitted exclusively for
× 10− 6 K− 1. A coating on top of the scaffolds sintered with more enhancing the density and strength characteristics. However, 20 vol %
bioactive glasses induces mineralization of the surface extensively in TCP and 80 vol % 45S5 bioactive glass contents within the sintered ink
vitro within two to three weeks [137]. Microcomputed at 1000 ◦ C was optimal in fabricating the better composite material both

13
B.C. Palivela et al. Bioprinting 27 (2022) e00219

human compact and trabecular bones.

4.5. In vitro studies

Simulated body fluid (SBF) has initially made by Kokubaand co-


workers [143]. 70 wt % 6P53B glass with PLA tested for 20 days in
SBF revealed a pH increase to 7.8, precipitation of crystals of apatite,
and reduction of stiffness in the perpendicular direction of the plane of
printing by 40% [34]. BaG-SPCL and SPCL scaffolds showed equal ten­
sile strength around 10–14 MPa after two weeks of testing; at five weeks,
the pH of the BaG-SPCL specimen was 0.05 U higher than SPCL scaffolds
[44]. Enhancement of bone bioactivity showed variation with the
amount of BG powder in PCL biopolymer. Seven days of immersion in
SBF led to an almost complete apatite mineral phase with no
de-lamination, and the duration is hence considered optimal. Perfusion
dynamic culturing had the edge over static culturing in populating stem
cells derived from human adipose called hASCs in regular notations on
the BG-PCL biopolymer scaffold [104]. The suitable substrate condition
for the growth of cells and blood circulation is optimally evident to be
macro-channeled pores with fiber thickness of ~301 μm and with the
dimension of channel ~352 μm for uniformly embedded BG particles in
PCL scaffold [130]. Post 2 weeks of immersion in SBF, Chit (chitosan)
with ten wt% n-BG (nano bioactive glass) are made use of for making
composite scaffold in nano-level with microporous structure. The pore
sizes up to 10 μm demonstrated dual-pore micro and macro-structure
with phosphate mineral formation and mechanical properties are yet
to be laid stress upon for bone tissue engineering [56]. The bioactive
glass (BG) in micro-particulate form into chitosan enhances cellular
functioning and in-vitro mineralization [144].
2εςf (κ.a)
UE = (6)

where UE is the particles’ electrophoretic mobility, η is the solution


viscosity, and ε is the dielectric constant. In the equation approximation
method, the Smoluchowski method is of interest, and f (κ.a) was
considered as1.5.
Zeta potential for 7,10, and 30 days in-vitro on 13–93B3 indicated
the possibility of controlling bioactivity and degradation rate by modi­
fying the S/S ratio (scaffold to SBF ratio). The weight loss of ~67%,
~15% besides ~19% for 10 mg/ml, 2 mg/ml and 1 mg/ml, ratios,
respectively after 7 days. Considerable weight loss (~67%) with HA
crystal formation was noticeable for 2 mg/ml at 30 days, the mechanism
as depicted in Fig. 24 [145]. The zeta potential (ζ), with the aid of the
Henry equation, gives the value of zeta potential (ζ) from equation (6) as
stated above [146]. When immersed in phosphate-buffered saline (PBS)
aqueous medium or exposed to the humid atmosphere, an impressive
ability to return to its deposition for the dried scaffolds and in-vitro on
gelatine bioactive glass led to morphological changes on the filament
surface as a function of time [84]. Gelatine’s swelling property was the
attributed reason for this regaining size ability of the scaffold [147], and
it may have future application in minute invasive surgery. Using
Murine-derived preosteoblast cell line (MC3T3) cells under osteogenic
conditions on PCL/SrBG and PCL/45S5 with the inclusion of BG pro­
motes precipitation of calcium phosphates (CaP). Scaffolds were
non-cytotoxic with supporting cell attachment and proliferation. Using
5 M NaOH for 1 day degradation showed ~48.6%, ~12.1% and ~1.6%
loss of the original mass by PCL/SrBG, PCL/45S5 and PCL respectively
Fig. 21. (a–c). Sintering and binder burnout studies (Refer to appendix table 2 [140]. On the contrary, the degradation kinetics are similar for 13–93
for bioactive glass type reference). scaffolds that are ethanol milled and water milled for immersion in SBF.
The water milled scaffolds’ weight loss was comparatively higher,
in reliability and strength [123]. The summary of mechanical properties resulting in a more significant reduction in strength, and HA excres­
of human cortical, cancellous bone, and bioactive glass type reference cences are more numerous and more effective. Both of the scaffolds,
from various studies are in Fig. 23(a–d) and appendix, table 2, respec­ after eight weeks, exhibited higher compressive strength values than
tively. Table 1 shows the summary of mechanical properties ranges of human trabecular bone. Fig. 10 (e) shows the printed ethanol and water
milled samples pre and post sintering [60]. PCL composites with coated

14
B.C. Palivela et al. Bioprinting 27 (2022) e00219

Fig. 22. a) Computational fluid dynamics simulations with multiple representative volume elements and pressure distribution, b) Experimental setup [132].

calcium phosphate (CaP) and 50% 45S5 or SrBG when immersed in 5 M viability by preventing excessive fast ion release [121]. HASCs with a
NaOH (sodium hydroxide) solution, the PCL/50-SrBG and PCL/50-45S5 PCL-bioactive borate 13–93B3 glass composite scaffold for up to 2 weeks
scaffold struts appeared intact after 6 h with ~39.84% and ~15.5% of in a culture medium demonstrated the controlled bioactive glass release
loss in original mass respectively post disintegrating for 24 h. PCL/CaP during degradation with ~23% weight loss. After 24 h, ~70% was cells’
and PCL scaffolds lost ~24.2% and ~24.6% of their initial scaffold mass, viability, followed by ~58% after one week [105]. Fine MG63 viability
respectively, when immersed for over seven days [148]. 13–93 silicate rate with no cytotoxic nature was evident for HSSGG sol-gel-based glass
glass porous scaffolds doped with various concentrations of copper from produced using robocasting. The surface layer in spider web shape,
0 wt % to 2 wt % copper oxides (Fig. 10 (f)) to estimate the copper ions’ whose thickness was morphologically increasing with the progression of
releasing effects from the glass on the response of osteoblast cells in vitro time and with the gradual change in pH in consecutive four stages at
besides on regeneration of the bone and also apoptosis in vivo. A com­ lower rates, was noticeable after 72 h [72]. Bioactive silicate glass
parison is by availing a type of protein that is bone morphogenetic (47.5B) scaffolds with F-127 acting as a binder in simulated body fluid
protein-2 (BMP-2). (SBF) established a progressive decrease in the bio-activity of forming
On protein addition to the scaffold that is one μh per defect or apatite with compressive stress in the range reported for the cancellous
scaffold with morphology as depicted in Fig. 25, concluding that there bone after four weeks [114]. Melt-derived bioactive glass (47.5 B)
was less considerable or little effect of copper ions in lower concentra­ scaffolds’ ability to form apatite post soaking in SBF and favorable
tions on the regeneration of the bone. The attention of ions of copper mechanical strength within and beyond the range reported for human
above specific threshold values are toxic to the osteoblasts, and the bone spongy bone under the compression. The properties of architecture like
regeneration, delivery of BMP-2 from ions of copper was less effective 50 vol % total porosity, 174 μm larger width of the pore, 139 μm as
than from 13 to 93 glass fabricated scaffolds in bone regeneration smaller width of the pore to support their compatibility as bone
stimulation [61]. Sintering cycles on 45S5 bio-glass and CMC composite substituting material [80]. The in vitro behavior of Sr0 and Sr50
during immersion in SBF showed faster degradation kinetics. It bioactive 3D printed glass scaffolds were examined via dissolution study
improved the overall conversion factor to calcium carbonates and and compared with ICIE16 printed glass. Phosphate bioactive glasses
phosphates with conversion reactions happening at the strut surfaces. It containing strontium ((20-x) SrO) and magnesium oxides (x MgO) in
also results in scaffolds’ compressive strength degradation and tough­ mol% showed enhanced thermal properties. FTIR result revealed that
ness because of faster volumetric diffusion. Extrusion of 45S5 bioglass the pH of the SBF remained unchanged for two weeks in SBF of
sintered at 1000 ◦ C when cells cultured with MC3T3 cells showed low-intensity x = 20 glass with CaP layer formation. Compared to
densifying and crystallization with enhanced cell proliferation and ICIE16, the Sr0 and Sr50 glass scaffolds had a relatively lower pH

15
B.C. Palivela et al. Bioprinting 27 (2022) e00219

Fig. 23. (a–d). Mechanical properties reported in various studies (Refer to appendix: table 2 for bioactive glass type reference).

Table 1
Summary of the mechanical properties of the human bone.
Bone type Compression Flexure Fracture Tensile Porosity Young’s Strain to
toughness (MPa strength (%) Modulus (GPa) failure (%)
Strength Elastic Strength Flexural
m1/2) (MPa)
(MPa) Modulus (MPa) Modulus
(GPa) (GPa)

Cortical or compact 100-150 10-20 [117] 135-193 9-16 [117] 2-12 [117] 50-151 [141] 5-10 7-30 [142] 1-3 [142]
bone [117] [117] [141]
Cancellous or 2-12 [117] 0.1–5 [117] 10-20 – 0.1–0.8 [117] 1-5 [141] 50-90 0.5–0.05 5-7 [142]
trabecular or [117] [141] [142]
spongy bone

increase and slower dissolution rate for four weeks. On the other hand, a of duration in SBF. The micrographs of the surface of glass particles also
slower HCA formation on the scaffold’s surface was noticeable with show Sr0 leads to an early HCA layer formation at week one relative to
samples immersed up to seven days which might be due to the higher pH Sr50, where HCA was only visible at week 2. The same may be due to the
environment of ICIE16 or the combination of pH and higher magnesium low pH environment caused by the lower dissolution rate and the
content of the Sr0 glass. High purity silica (SiO2), phosphate(P2O5), strontium content in Sr50 glass [71]. The viability investigation of
magnesium (MgO), and modifying oxides that are equivalent to car­ mesenchymal stem cells (MSCs) derived from human-adipose stem cells
bonate were required/used for the glass synthesis. The silica gel layer (hASCs) is carried out. The investigation is in dynamic conditions
attracted more Ca2+ and the PO3- 4 to form the calcium-phosphate-rich combining a bio-ink hydrogel of alginate-gelatine (1:1) installed be­
layer. The printed group showed a relatively small amount of bone tween polylactic acid (PLA), borate bioactive glass 13–93B3, also called
ingrowth compared with foam and control groups after four weeks of B3 glass with ratios of weight as 50% and 33%. Three different config­
implantation. Both Sr0 and Sr50 glass resulted in rich urations, namely PLA + B3 glass + bio-ink, bio-ink, PLA + bio-ink. A
calcium-phosphate layer precipitation on the glass surface at four weeks non-uniform ASC viability with top layers greater than 80% and bottom

16
B.C. Palivela et al. Bioprinting 27 (2022) e00219

Fig. 24. Conversion and degradation in 13–93B3 (a) Charge on the surface just immersed (b) Layer of amorphous calcium phosphate at scaffold to simulated body
fluid ratio 10 mg/ml (c) Amorphous calcium phosphate layer at scaffold to simulated body fluid ratio 1 mg/ml (d) Hydroxyapatite at 1 mg/ml [145].

Fig. 25. Scanning electron micrograph with copper oxide (0.8 wt %) doping (a) and glass filament surface, post immersion in the simulated body fluid of the scaffold
for seven days at 37 ◦ C (b) of 13–93 scaffold [61].

layers as ~60% for all scaffold groups were noticeable in live/dead ingrowth of the bone compared to the scaffolds produced from SBP-3
assessment. Because of hypoxic akin ambient generated to a level with [82]. The summary of mechanical properties in vitro and in vivo with
the existence of PLA and polylactic acid + filaments of B3 as the corresponding immersion time in vitro and in vivo from various
non-uniformities are mere observable in scaffolds of bio-ink [106]. studies was plot along with the ranges of human cortical and cancellous
Copper (Cu) doped 13–93 released Cu ions dose-dependently in SBF, cell bones from the literature in Fig. 26 (a and b).
proliferation with alkaline phosphatase (ALP) activities concerning
MC3T3-E1 cells is related to wt% of copper oxide (CuO). For 0.4 to 0.8
4.6. In vivo studies
wt %, the proliferation and % new bone infiltrated in rat calvarial de­
fects with no noticeable changes with high significance at six weeks. For
Doping bioactive glass with elements in trace amounts like copper,
2% CuO the inhabitation of % new bone growth and newer blood vessels zinc, and strontium stimulated apoptosis and osteogenesis [134,135,
area in fibrous tissues was significantly higher. Morphogenetic bone
136]. As degradation of the glass takes place, converting into HA in vivo,
protein-2 usage (1 μg/defect) for loading was considerably higher, but elements at acceptable levels therapeutically are released. 13–93 B3
two wt % copper oxide was harmful to the cells besides damaging the
borate bioactive glass by replacing all silicon dioxide in 13–93 glass with
regeneration of the bone. The higher deviations were evident in B12.5 the di-boron trioxide, converting faster than silicate-based 13–93 glass
than in mix groups of porogen at a mass loss in SBF at 336 h with
to hydroxyapatite [151]. Porous 13–93 bioactive glass silicate-based,
~4.28% and ~2.45%, respectively. SBF immersion of B12.5 and mix studied in vivo (rat subcutaneous) testing environments showed the
with magnesium (Mg) and strontium (Sr) concentrations revealed Sr
modulus and strength rapidly decreased in the first two weeks. Still,
concentration in B12.5 glass remained null, and Mg concentration more slowly, the decrease in vitro was smaller than in vivo [117]. A
decreased for B12.5 and increased for mix glass scaffold in static con­ comparison of regeneration within the implanted rat calvarial defect
dition. In contrast, in dynamic conditions, Mg concentration remained with the modified surfaces and bone morphogenetic protein-2 loaded
constant regardless of scaffold types. 13–93 scaffolds of the glass that are pre-treated and as fabricated scaf­
The magnesium and strontium 3D printed scaffolds showed excellent folds revealed that conversion was faster in case of as-fabricated in vivo
performance when 72 h pre-incubation was put to hold. The ion burst as in comparison to the pre-treated scaffold. Before the implantation,
the size of the scaffold limits the ion release in the cell viability test amorphous calcium phosphate material was significantly and effectively
[108]. ALP activity, proliferation, human bone marrow-derived enhancing the capacity of the scaffold for supporting the new bone
mesenchymal stem cells (hBMSCs) osteogenesis expression, and adhe­ formation with no interrelation between pre-treatment time and BMP
sion with the more acceptable ability of phosphates formation stimu­ loading. The % new bone formation for pre-treated 1–6 days varied from
lated by inhibition of MBG into calcium sulfate hydrate (CSH) in porous ~45 to 57% (3 days), whereas in pre-treated and loaded for 1–6 days,
3D CSH/MBG scaffolds [115]. In vitro analysis depicted that the the percentage of newer bone formation varied from ~61 to 64% (6
composition chemistry of ICIE16 was more akin to the original days). The scaffolds’ grid-like microscopic structure can support bone
composition of 45S5 and shown to stimulate and undergo better regeneration better than orient, trabecular mesh-like, and fibrous

17
B.C. Palivela et al. Bioprinting 27 (2022) e00219

other hand, the values of t (mm), CoVs are found to be higher in the case
of RC structures and lower in the case of FR scaffolds and SGF scaffold
values falling in the median ranges of both the RC and FR scaffolds.
Average amounts of new bone within the implanted defects with grid
13–93B3 was 26%, oriented 13–93B3 was 28%, and 13–93 grid scaffolds
were 25%. On the other hand, the mean amount of newer bone within
the defects implanted with autografts was 38%. The total Von Kossa
(VK) positive area inside the fault of grid-like 13–93B3 was 32%, in the
13–93B3 oriented was 37%, for 13–93 grid-like was 38% and for auto­
graft 40% with conversion % 5,9 and 13 respectively. The formation of
newer blood vessels was found in all four groups all over the area with
the defect. The sections stained with periodic acid Schiff (PAS) are
shown in Fig. 27. The difference in the area of the new blood vessel was
the least significant. The oriented 13–93B3 scaffold showed the newer
blood vessels area highest at 8%, while the newer area of blood vessels
was 5% in the grid 13–93B3 besides in case of grid-like 13–93 scaffolds
was 4%, autograft was 5%. Out of all high concentration of newer
cartilage at the area of the defect was seen in 13–93 grid-like scaffolds
(18%), which is higher than 13–93B3 grid-like structure (8%) but nearly
with oriented 13–93B3 scaffolds (12%). It was noticeable that 13–93
grid-like scaffolds showed high amounts (18%) of cartilage within the
defects in comparison to that of the 13–93B3 grid-like (8%) and auto­
graft (8%), suggesting these bioactive glasses conversion is defect site
dependant, and 13–93B3 is the opt able choice of all within the given
time frame of formation [64]. No significant difference was evident for
13–93B3 and 13–93 glass in the Von Kossa area that is positive as the
hydroxyapatite conversion to the 30% and 25%, respectively, when
implanted for twelve weeks in a rat calvarial defect [141].When sub­
jected to a force five times the weight of an average rat animal, the
compressive stress generated on the cross-sections of rat-femur was
1Mpa.
Three scaffolds fabricated from bioactive glass (25–28%) were
interesting to find the best fit, but a little dependence was noticeable on
composition and microstructure. In one study, the scaffolds of 13–93 B3
with fibrous microstructure displayed an improved capacity for regen­
eration of the bone [158] than that of the 13–93 made scaffolds con­
taining the same microscopic structure. In the other study, grid-like
13–93 scaffolds were more capable of regenerating the bone than the
Fig. 26. Various studies reported in vitro and in vivo mechanical properties (a) 13–93 oriented scaffolds [118]. The same was with the Von Kossa area,
Compressive strength (MPa) and (b) Flexural strength (MPa), Elastic modulus which is positive in defects from 32% to 38%, showed no significant
(MPa) with corresponding immersion time of scaffolds (Refer to appendix: table difference. However, other studies revealed that when implanted inside
2 for bioactive glass type reference). the sites of rat subcutaneous, the cancellous scaffold’s conversion to
hydroxyapatite was faster considerably than the scaffolds of 13–93 with
microscopic structure. The amount of forming new bone in vivo was the a similar microscopic structure [159]. Twenty-four weeks of testing on
same, approximately or greater than that in the orient, trabecular, and rat calvarial scaffolds with three groups of implants, namely
fibrous microstructures at 12 weeks [118]. Normalized oxygen diffu­ as-fabricated, pre-treated in a phosphate solution, and loading with one
sivity was compared in the scaffolds obtained from extrusion, foam μg/morphogenetic defect protein-2. Observations are the degradation
replication (FR), and sol-gel foaming by performing the Lattice Monte rate dropped from 6 to 24 weeks with no considerable difference in
Carlo simulations. FR method-made scaffolds exhibit the highest oxygen positive Von Kossa area generation. The scaffold loaded with BMP-2 has
diffusivity due to the interconnected pores and higher porosity. These enhanced bone regeneration and the number of blood vessels and areas
structures ensured a better supply of oxygen for the newly generated within the new bone at 6–12 weeks post-implantation. The
tissues among the types of the scaffolds considered as per the numerical as-fabricated, pre-treated, and BMP-2 loaded scaffolds for 12 weeks
forecasting [152]. The diffusivities effectively are obtained and con­ showed a new bone layer varied from Ca/P atomic ratio of ~1.71–1.73.
trasted for the robocasting (RC) structures made from 45S5 bio The least for BMP-2 loaded, the variation in the layer of glass converted,
glass-PCL [120], sol-gel foamed (SGF) bioactive glass 70S30C (70 mol% rich silica layer and un converted glass layer are ~1.50–1.61 with
SiO2, 30 mol% CaO) [149,150,153]. It is noticeable that the oxygen pre-treated as least, ~2.3–2.6 with as-fabricated as highest and
transport, nutrients, biomolecules, and waste in the scaffold’s tissues are ~5.6–5.8 with same for pre-treated and bone morphogenetic protein-2
necessary to grow the new tissue and for efficient or proper functioning loaded, respectively. 13–93 bioactive glass fabricated strong porous
[154,155]. In these lines, the ImageJ [156] and plugin BoneJ [157] were scaffolds that are loaded with acceptable levels clinically of bone
made use of for calculating the connection density "dc," the mean size of morphogenetic protein-2 could generate promising implants for the
the pore "s" and also the mean thickness of the strut "t," coefficients of structural healing of loaded defects of the bone within a relevant time
variation for pore size "CoVs" and coefficients of variation of strut clinically [119]. When implanted in vivo in rat calvarial defect, scaffolds
thickness "CoVt" respectively, besides also volume fraction of the tissue displayed lower regeneration capacity with volume fraction increment
"Ф" and diffusivity (D). The values of D, Ф, dc (mm− 3), s(mm), and CoVt of 13–93B3 glass inside the scaffolds. The percentage of newer bone
are lower in RC structures. Higher in the case of FR scaffolds and on the inside the implanted defects with the scaffolds fabricated from 13 to 93
was considerably higher than that of the defects with 13–93B3

18
B.C. Palivela et al. Bioprinting 27 (2022) e00219

Fig. 27. Light transmission images of sections stained with periodic acid Schiff in a critically sized segmental rat femoral defects that are implanted within the
scaffolds of the bioglass (A) grid 13–93B3, (B) oriented 13–93 B3, (C) grid-like 13–93, and (D) autologous bone post-implantation after 12 weeks. Asterisk indicates
bioactive glass, and arrow indicates blood vessel at scale bar = 100 μm (b) and toluidine blue-stained. Asterisk indicates bioglass, and arrow signs cartilage. Scale bar
= 500 μm (b) [64]. (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)

implanted scaffolds, which is higher in both the compositions for 12 interconnections of pores with porosity, allowing necessary cell growth.
weeks than six weeks. On the other hand, the 13–93 glass was partially This review article discussed bioactive glasses (BGs), including their
converted to hydroxyapatite post twelve weeks of implantation in vivo, composition, processing, properties, and medical applications in tissue
whereas 13–93B3 converted almost fully within six weeks. The lowest regeneration. For decades, BGs have been used in the clinical field to
capacity of the 13–93B3 scaffolds for supporting bone regeneration repair injured trabecular and cancellous bones surgically. They are close
anticipates the higher concentration of boron released from the glass, to natural bones with biocompatibility in physical and mechanical as­
converting rapidly [131]. MBG-PCL macro-porous scaffold in an osteo­ pects. However, the applications are limited because their acute fracture
porotic sheep model with one wt% of Zoledronic acid incorporated into toughness is less than 2 MPa m1/2 and is brittle. But, glass ceramics and
scaffolds. For evaluation of components of inflammation response, a bioceramics with polymer composites have displayed a great potential
graded scale was made use of having five units depending upon the concerning parent glasses in enhancing the required mechanical prop­
distribution and density (absent/moderate/mild/marked/very marked, erties like fracture toughness and strengths.
scored from 0 to 4). A four-graded scale depended on the density of
blood vessels (absent/moderate/mild/marked/scored 0–3) of the 6. Future directions
vascular nature. The H and E (Hematoxylin-Eosin) slide scan at lower
magnification power (x 4), identifying hot spots around three in number AM’s significant advantage is the freedom and ability to fabricate
with the higher vascular density and counting the number of vessels in geometries specific to a patient in micro-macro levels to enhance bone
20× field magnification. The MBG-PCL-zol for 30 days in phosphate regeneration by bio-mimicking the bone tissue. Using 3D printed bio­
buffer saline (PBS) displayed a significant change of around ~ ten times ceramics with hydrogels gives the ability to utilize a variety of materials
in the specific surface area, ~2 times in the pore size, and ~0.1 times in to construct clinically close bio-functional bone substitutes with the
connection with the pore volume. The Zoledronic acid-loaded scaffolds ability to generate vascular networks. The features of size lesser than
in Saos 2 (osteoblast-like cells) cells inculcated apoptosis, impeding 100 μm and scaffolds with intricate internal details are still challenging
differentiation of osteoclast cells with dependence on time besides to fabricate because of the duration of fabrication or cracking tendency
inhibiting bone healing, inflammatory response intensely in osteopo­ in post-processing stages. There are still significant gaps in duration
rotic sheep [116]. In vivo analysis depicted that the composition optimization, dimensional accuracy enhancements, and sintering mod­
chemistry of ICIE16 was more akin to the original composition of 45S5 ifications, which can control fabricating at the nano-scale. There are
and shown to stimulate and undergo better ingrowth of the bone numerous implementable ideas and works to extend scaffolds for clinical
compared to the scaffolds produced from SBP-3 [82]. For silk fibroin and and many other applications. In these lines, the following statements are
silk fibroin bioactive glass scaffolds, the cell culture experiments put by this paper:
revealed that the viability of Human bone marrow mesenchymal stem
cells on the incorporated scaffolds with nano bioactive glass was higher (i) Research strengthening on material prospects: The scaffold’s per­
obviously than the silk fibroin at 3 and 7 days. There was no influential formance is the primary function that is highly dependent on
significance of BG particle size. Hence, the attention is laid on combining material and its properties, which establishes the limits in
the indirect printing in three scaffold dimensions and freeze-drying to working and fabrication. Fiber compounding, growth factors, and
fabricate hierarchically bioactive glass porous scaffolds with silk fibroin composites can effectively enhance material and biological
as an essential water-soluble polymer for bone regenerating applications properties.
[59]. (ii) Manifold forming technology integration: Optimization of process
parameters and multiple kinds of sintering variants are excellent
5. Conclusions prospects for enhancing mechanical properties and fabricating a
high-quality BG ceramic scaffold.
Additive manufacturing (AM) systems with extrusion-based material (iii) Interdisciplinary research: By producing BG scaffolds artificially is
fabrication techniques favoring the printing of scaffolds are of interest in a multi-disciplinary stream of research dealing with engineering
this paper. This review gives a glimpse of these techniques in processing and medicine. Extrusion AM of BG and its allied composites are
the extensive range of biomaterials maintaining superior control on

19
B.C. Palivela et al. Bioprinting 27 (2022) e00219

the new provokes to enhance comprehensive theory, and the work reported in this paper.
methods could be created and promoted.
Acknowledgments
The attention on the research in extrusion AM on BG scaffolds is
progressively increasing. So, it can be of urge that the future of BG The authors are thankful to ‘‘ELSEVIER” for granting copy right
scaffold fabrication will be highly dependent upon the development of permission to the figures that has been used in this paper (License
porous tissue-specific 3D scaffold that enhances in vivo tissue regener­ numbers: 5295240416289 [38], 5295240583295 [39], 52952
ation. There is a need to understand correlative structure-property re­ 40767957 [40], 5295240890460 [41], 5295241005181 [42],
lations to develop and ensure the mechanical reliability of AM fabricated 5295241126792 [45], 5295241293000 [50], 5295241425872 [51],
scaffolds. 5295250117213 [56], 5295250215547 [57], 5295250322178 [59],
5295250442923 [60], 5295250541327 [61], 5295250640142 [64],
CRediT authorship contribution statement 5295250763486 [65], 5295250857377 [78], 5295250962497 [96],
5295251082555 [106], 5295251176741 [132], 5295251289446
Bhargav Chandan Palivela: Investigation, data curation, visualiza­ [145]).
tion, conceptualization, writing-original draft. Authors are grateful to the Department of Science and Technology
Sai Drupadh Bandari: Visualization and writing-original draft. (DST)-Innovation of Science Pursuit for Inspire Research (INSPIRE) for
Ravi Sankar Mamilla: Supervision, writing - review & editing. fellowship support (IF190716) and BIRAC (BT/GUAR-GUM0031/01/
19).
Declaration of competing interest
Appendix: Table 2
The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influence

Bioactive glass type reference

Reference Bioactive glass type Reference Bioactive glass type Reference Bioactive glass type

[33] 13-93 (Freeze extrusion fabricated) [95]-3 Sr-HT Gahnite (Hexagonal) (50% [59]-2 SF-BG (Micro particles)
porosity)
[45] MBG + PVA [96]-1 FEM-Generated (variable porosity, [51] WB-15/MK
Poisson’s ratio = 0.25)
[36,50]-1 6P53B + 127-pluronic (measured parallel [96]-2 FEM-Generated (Grid like) [103] 8 %-MBG-CPC
to scaffold surface)
[36,50]-2 6P53B + 127-pluronic (measured [100]-1 FEM-Simulated 13–93 pore width [79,80] 47.5B + Pluronic F-127
perpendicular to scaffold surface) 300 μm
[74]-1 13-93+ (ethyl cellulose, polyethylene [100]-2 FEM-Simulated 13–93 pore width [85]-1 BG30 (Functionally graded bioactive
glycol) 600 μm glass)
[74]-2 13–93B3+ (ethyl cellulose, polyethylene [100]-3 FEM-Simulated 13–93 pore width [85]-2 BG30 (After reinforcing scaffold with
glycol) 900 μm the nanofiber)
[64]-1 13-93 grid like + Pluronic® F-127 [70]-1 Di-TCP Pore size (μm) (300) [129]-1 HSSGG (300 μm)
[64]-2 13–93B3 Grid like + Pluronic® F-127 [70]-2 Di-TCP Pore size (μm) (400) [129]-2 HSSGG (400 μm)
[64]-3 13–93B3 Oriented/+Pluronic® F-127 [70]-3 Di-TCP Pore size (μm) (500) [129]-3 HSSGG (500 μm)
[60]-1 13-93 + CMC-250 (water milled) [83] ICIE 16 + 25 wt% Pluronic F-127 [110] Yttria-stabilized zirconia (YSZ)
[60]-2 13-93 + CMC-250 (ethanol milled) [84] Gelatine-bioactive glass [138]-1 MBG
[68] 45S5 Bio glass®+ PCL+1 wt % CMC [57] 45S5+ CMC-35 +CMC-250 [138]-2 Sr-MBG
[69]-1 CSi-BG0 (Archimedean Chords) [117] 13–93 [32,139] PCL
[69]-2 CSi-BG1 (Archimedean Chords) [120] 45S5 Bio glass® + 1 wt% CMC-250 [106]-1 PLA
[88]-1 Np alginate nano composite (dry) [140]-1 PCL/45S5 [106]-2 PLA + B3 (33 wt%)
[88]-2 Np alginate nano composite (wet) [140] PCL/SrBG [106]-3 PLA + B3 (50 wt %)
[89]-1 45S5+ CMC-250 PCL 20% [102] 45S5 BG + PCL + Nano HA (Sol gel) [34] 70 wt% 6P5B3+ PLA (Mechanical
properties)
[89]-2 45S5+CMC-250 PCL 25% [101]-3 FastOs® BG Filament (500 μm) [102]-1 45S5 BG + PCL + Nano HA (Calcined)
[89]-3 45S5+CMC-250 PCL 30% [77]-1 BG1(10 wt% stoichiometric Bio [132] Sr-HT Gahnite
glass®)
[91]-1 (Mechanical 13-93+Ti-Fibre 0 vol% [77]-2 BG2(calcium excess Bio glass®) [44] 6xSPCL+5xBaG
properties)
[91]-2 (Mechanical 13-93+Ti-Fibre 0.4 vol% [77]-3 CAN (Canasite) [101]-1 FastOs® BG Filament (200 μm)
properties)
[95]-1 Sr-HT Gahnite (Hexagonal) (70% [114] 47.5B + Pluronic F-127 [101]-2 FastOs® BG Filament (300 μm)
porosity)
[95]-2 Sr-HT Gahnite (Hexagonal) (60% [59] SF-BG (Nano particles) [71]-4 Phosphate BG (15 MgO, 5 SrO)
porosity)
[62] 13–93B3 [70] Di70-TCP30 [71]-5 Phosphate BG (20 MgO, 0 SrO)
[74] 45S5 [72,73] HSSGG [75] Borosilicate Glass
[62,76] 13–93 [51]-1 WB [82] SBP-3
[78] PSrBG [51]-2 WB-15 [103] 8%-MBG-CPC
[81] SrBG [71]-1 Phosphate BG (0 MgO, 20 SrO) [79] 47.5 B + 27.5 wt% F-127 Pluronic
[82,83] ICIE16 [71]-2 Phosphate BG (5 MgO, 15 SrO) [106]-1 Acellular PLA + B3 (printing
parameters)
[85] BG30+(20 wt%) Pluronic-127 [71]-3 Phosphate BG (10 MgO, 10 SrO) [106]-2 Cellularized PLA + bio-ink (printing
parameters)
[34] BG 45S5(20%) +PCL [91] # 13-93 + Ti fibre [106]-3 PLA + B3+bio-ink (printing
(Printing parameters, Scaffolds parameters)
dimensions and shrinkage)
(continued on next page)

20
B.C. Palivela et al. Bioprinting 27 (2022) e00219

(continued )
Reference Bioactive glass type Reference Bioactive glass type Reference Bioactive glass type

[104] BG-PCL [95] Sr-HT Gahnite [108] B12.5 & B12.5-Mg5-Sr10


[56] 10 wt% nBG + Chit [105] 13–93B3 + PCL:CF [112] Silicate Glass
[64] 13-93 and 13–93B3 (printing parameters) [101] FastOs® BG [113] MBG_Scs
[60] 13-93 + CMC-250 [107] 13–93B3 + PCL [115] CSH/MBG
[69] CSi-BG1 [71] Phosphate BG + Pluronic® F-127 [116] MBG-PCL
[88] NP-alginate nano-composite [109] 13–93B3+ 20 wt%PCL [73] HSSGG
[89] 45S5 Bio glass®+ε-PCL + CMC-250 [111] 13-93+Pluronic® F-127 [36,50] 6P53B + 20 wt % Pluronic
[34,68,88] 45S5(20%) +PCL [95] Sr-HT Gahnite [108]-2 B12.5-Mg5-Sr10
[62,64,73] 13–93B3 + (ethyl cellulose, poly ethylene [123] 45S5 Bio glass® +β-TCP + CMC [113] MBG_Scs
glycol)
[57,96,120–122] 45S5+ CMC [130] ICIE16 + 25 wt% Pluronic F-127 [91]-1 13-93+Ti-fibre (Sintering and binder
burn out)
[116–118] 13–93 +(20 wt%) Pluronic-127 [79,114] 47.5B + Pluronic F-127, 27.5 wt% [91]-2 13-93+Ti-fibre (Sintering and binder
burn out)
[86] 45S5+PLA [76]-1 13-93 (multi-stage) [91]-3 13-93+Ti-fibre (Sintering and binder
burn out)
[61,106,110,128] 13-93(40%vol) +(20 wt%) Pluronic-127 [76]-2 13-93 (multi-stage) [85]-1 BG30+(20 wt%) Pluronic-127
(Sintering and binder burnout)
[117]-1 13-93 (SBF) (RAT FEMOR) (2 weeks) [108]-1 B12.5 [85]-2 SiO2–P2O5+ Pluronic® F127
(Sintering and binder burnout)
[117]-2 13-93 (SBF) (RAT FEMOR) (12 weeks) [115]-1 CSH/MBG0 [50] 6P53B + 127-pluronic
[80]-1 47.5B (SBF) (2 weeks) [115]-2 CSH/MBG20 [114]-1 47.5B + Pluronic F-127 (2 weeks)
[80]-2 47.5B (SBF) (4 weeks) [115]-3 CSH/MBG40 [114]-2 47.5B + Pluronic F-127 (4 weeks)
[45]-1 MBG (pore size) [115]-4 CSH/MBG60 [82]-3 ICIE 16
[45]-2 MBG (pore size) [62]-1 13–93B3 + (ethyl cellulose, [91] 13-93+Ti-fibre Diameter = 16 μm
polyethylene glycol)
[45]-3 MBG (pore size) [62]-2 13–93B3 + (ethyl cellulose, [98] MBG-PCL
polyethylene glycol)
[36] 6P53B + 20 wt% F-127 Pluronic [65] 13-93 + PCL or PLA [109] 13–93B3 + 20 wt% PCL
[74] 13-93 + (ethyl cellulose, polyethylene [66] PCL–FastOs®BG [114] 47.5B + 127-Pluronic
glycol)
[85] BG30+(20 wt%) Pluronic-127 [82]-1 13–93 [47] SiO2–P2O5+ Pluronic® F127
[82]-2 SBP-3

References [13] H. Larry L, P. Julia M, Third-generation biomedical materials, Science (80- 295
(2002) 1014–1017. February.
[14] L.L. Hench, The story of Bioglass, J. Mater. Sci. Mater. Med. 17 (11) (2006)
[1] A.M.A. Ambrosio, J.S. Sahota, Y. Khan, C.T. Laurencin, A novel amorphous
967–978, https://doi.org/10.1007/s10856-006-0432-z.
calcium phosphate polymer ceramic for bone repair: I. Synthesis and
[15] L.C. Gerhardt, A.R. Boccaccini, Bioactive glass and glass-ceramic scaffolds for
characterization, J. Biomed. Mater. Res. 58 (3) (2001) 295–301, https://doi.org/
bone tissue engineering, Materials (Basel) 3 (7) (2010) 3867–3910, https://doi.
10.1002/1097-4636(2001)58:3<295::AID-JBM1020>3.0.CO;2-8.
org/10.3390/ma3073867.
[2] L.M. Mathieu, T.L. Mueller, P.E. Bourban, D.P. Pioletti, R. Müller, J.A.E. Månson,
[16] A.L.B. Maçon, et al., A unified in vitro evaluation for apatite-forming ability of
Architecture and properties of anisotropic polymer composite scaffolds for bone
bioactive glasses and their variants, J. Mater. Sci. Mater. Med. 26 (2) (2015)
tissue engineering, Biomaterials 27 (6) (2006) 905–916, https://doi.org/
1–10, https://doi.org/10.1007/s10856-015-5403-9.
10.1016/j.biomaterials.2005.07.015.
[17] J.R. Jones, Review of bioactive glass: from Hench to hybrids, Acta Biomater. 9 (1)
[3] C.R. Kothapalli, M.T. Shaw, M. Wei, Biodegradable HA-PLA 3-D porous scaffolds:
(2013) 4457–4486, https://doi.org/10.1016/j.actbio.2012.08.023.
effect of nano-sized filler content on scaffold properties, Acta Biomater. 1 (6)
[18] M. Zhu, J. Shi, Q. He, L. Zhang, F. Chen, Y. Chen, An emulsification-solvent
(2005) 653–662, https://doi.org/10.1016/j.actbio.2005.06.005.
evaporation route to mesoporous bioactive glass microspheres for
[4] W. Cheng, H. Li, J. Chang, Fabrication and characterization of β-dicalcium
bisphosphonate drug delivery, J. Mater. Sci. 47 (5) (2012) 2256–2263, https://
silicate/poly(d,l- lactic acid) composite scaffolds, Mater. Lett. 59 (17) (2005)
doi.org/10.1007/s10853-011-6037-z.
2214–2218, https://doi.org/10.1016/j.matlet.2005.02.069.
[19] F. Quintero, et al., Laser spinning of bioactive glass nanofibers, Adv. Funct. Mater.
[5] S.J. Hollister, Porous scaffold design for tissue engineering (vol 4, pg 518, 2005),
19 (19) (2009) 3084–3090, https://doi.org/10.1002/adfm.200801922.
Nat. Mater. 5 (7) (2006) 590.
[20] S. Abdollahi, A.C.C. Ma, M. Cerruti, Surface transformations of bioglass 45S5
[6] T. Dutta Roy, J.L. Simon, J.L. Ricci, E.D. Rekow, V.P. Thompson, J.R. Parsons,
during scaffold synthesis for bone tissue engineering, Langmuir 29 (5) (2013)
Performance of degradable composite bone repair products made via three-
1466–1474, https://doi.org/10.1021/la304647r.
dimensional fabrication techniques, J. Biomed. Mater. Res., Part A 66 (2) (2003)
[21] M. Mozafari, et al., Development of macroporous nanocomposite scaffolds of
283–291, https://doi.org/10.1002/jbm.a.10582.
gelatin/bioactive glass prepared through layer solvent casting combined with
[7] Z. Xiong, Y. Yan, S. Wang, R. Zhang, C. Zhang, Fabrication of porous scaffolds for
lamination technique for bone tissue engineering, Ceram. Int. 36 (8) (2010)
bone tissue engineering via low-temperature deposition, Scripta Mater. 46 (11)
2431–2439, https://doi.org/10.1016/j.ceramint.2010.07.010.
(2002) 771–776, https://doi.org/10.1016/S1359-6462(02)00071-4.
[22] S. Ban, Y. Iwaya, H. Kono, H. Sato, Surface modification of titanium by etching in
[8] S.J. Kalita, S. Bose, H.L. Hosick, A. Bandyopadhyay, Development of controlled
concentrated sulfuric acid, Dent. Mater. 22 (12) (2006) 1115–1120, https://doi.
porosity polymer-ceramic composite scaffolds via fused deposition modeling,
org/10.1016/j.dental.2005.09.007.
Mater. Sci. Eng. C 23 (5) (2003) 611–620, https://doi.org/10.1016/S0928-4931
[23] D.K. Pattanayak, S. Yamaguchi, T. Matsushita, T. Nakamura, T. Kokubo, Apatite-
(03)00052-3.
forming ability of titanium in terms of pH of the exposed solution, J. R. Soc.
[9] T.M.G. Chu, D.G. Orton, S.J. Hollister, S.E. Feinberg, J.W. Halloran, Mechanical
Interface 9 (74) (2012) 2145–2155, https://doi.org/10.1098/rsif.2012.0107.
and in vivo performance of hydroxyapatite implants with controlled
[24] B. Chen, Z. Zhang, J. Zhang, M. Dong, D. Jiang, Aqueous gel-casting of
architectures, Biomaterials 23 (5) (2002) 1283–1293, https://doi.org/10.1016/
hydroxyapatite, Mater. Sci. Eng. A 435 (436) (2006) 198–203, https://doi.org/
S0142-9612(01)00243-5.
10.1016/j.msea.2006.07.028.
[10] I. Grida, J.R.G. Evans, Extrusion free forming of ceramics through fine nozzles,
[25] Z.G. Tang, R.A. Black, J.M. Curran, J.A. Hunt, N.P. Rhodes, D.F. Williams, Surface
J. Eur. Ceram. Soc. 23 (5) (2003) 629–635, https://doi.org/10.1016/S0955-2219
properties and biocompatibility of solvent-cast poly[ε-caprolactone] films,
(02)00163-2.
Biomaterials 25 (19) (2004) 4741–4748, https://doi.org/10.1016/j.
[11] J.E. Smay, J. Cesarano, J.A. Lewis, Colloidal inks for directed assembly of 3-D
biomaterials.2003.12.003.
periodic structures, Langmuir 18 (14) (2002) 5429–5437, https://doi.org/
[26] R.R. Rao, T.S. Kannan, Dispersion and slip casting of hydroxyapatite, J. Am.
10.1021/la0257135.
Ceram. Soc. 84 (8) (2001) 1710–1716, https://doi.org/10.1111/j.1151-
[12] S. Michna, W. Wu, J.A. Lewis, Concentrated hydroxyapatite inks for direct-write
2916.2001.tb00903.x.
assembly of 3-D periodic scaffolds, Biomaterials 26 (28) (2005) 5632–5639,
https://doi.org/10.1016/j.biomaterials.2005.02.040.

21
B.C. Palivela et al. Bioprinting 27 (2022) e00219

[27] H. Fu, et al., In vitro evaluation of borate-based bioactive glass scaffolds prepared manufacturing of bioactive glass-ceramic scaffolds, Ceram. Int. 45 (11) (2019)
by a polymer foam replication method, Mater. Sci. Eng. C 29 (7) (2009) 13740–13746, https://doi.org/10.1016/j.ceramint.2019.04.070.
2275–2281, https://doi.org/10.1016/j.msec.2009.05.013. [52] Q. Li, J.A. Lewis, Nanoparticle inks for directed assembly of three-dimensional
[28] X. Liu, M.N. Rahaman, Q. Fu, Oriented bioactive glass (13-93) scaffolds with periodic structures, Adv. Mater. 15 (19) (2003) 1639–1643, https://doi.org/
controllable pore size by unidirectional freezing of camphene-based suspensions: 10.1002/adma.200305413.
microstructure and mechanical response, Acta Biomater. 7 (1) (2011) 406–416, [53] M. Dash, F. Chiellini, R.M. Ottenbrite, E. Chiellini, Chitosan - a versatile semi-
https://doi.org/10.1016/j.actbio.2010.08.025. synthetic polymer in biomedical applications, Prog. Polym. Sci. 36 (8) (2011)
[29] W.Y. Yeong, C.K. Chua, K.F. Leong, M. Chandrasekaran, Rapid prototyping in 981–1014, https://doi.org/10.1016/j.progpolymsci.2011.02.001.
tissue engineering: challenges and potential, Trends Biotechnol. 22 (12) (2004) [54] R. Jayakumar, D. Menon, K. Manzoor, S.V. Nair, H. Tamura, Biomedical
643–652, https://doi.org/10.1016/j.tibtech.2004.10.004. applications of chitin and chitosan based nanomaterials - a short review,
[30] B. Stevens, Y. Yang, A. Mohandas, B. Stucker, K.T. Nguyen, A review of materials, Carbohydr. Polym. 82 (2) (2010) 227–232, https://doi.org/10.1016/j.
fabrication methods, and strategies used to enhance bone regeneration in carbpol.2010.04.074.
engineered bone tissues, J. Biomed. Mater. Res. Part B Appl. Biomater. 85 (2) [55] H.H. Lee, H.S. Yu, J.H. Jang, H.W. Kim, Bioactivity improvement of poly
(2008) 573–582, https://doi.org/10.1002/jbm.b.30962. (ε-caprolactone) membrane with the addition of nanofibrous bioactive glass, Acta
[31] W. Zhu, X. Ma, M. Gou, D. Mei, K. Zhang, S. Chen, ScienceDirect 3D printing of Biomater. 4 (3) (2008) 622–629, https://doi.org/10.1016/j.actbio.2007.10.013.
functional biomaterials for tissue engineering, Curr. Opin. Biotechnol. 40 (2016) [56] B. Dorj, J.H. Park, H.W. Kim, Robocasting chitosan/nanobioactive glass dual-pore
103–112, https://doi.org/10.1016/j.copbio.2016.03.014. structured scaffolds for bone engineering, Mater. Lett. 73 (2012) 119–122,
[32] I. Zein, D.W. Hutmacher, K.C. Tan, S.H. Teoh, Fused deposition modeling of novel https://doi.org/10.1016/j.matlet.2011.12.107.
scaffold architectures for tissue engineering applications, Biomaterials 23 (4) [57] S. Eqtesadi, A. Motealleh, P. Miranda, A. Lemos, A. Rebelo, J.M.F. Ferreira,
(2002) 1169–1185, https://doi.org/10.1016/S0142-9612(01)00232-0. A simple recipe for direct writing complex 45S5 Bioglass® 3D scaffolds, Mater.
[33] N.D. Doiphode, T. Huang, M.C. Leu, M.N. Rahaman, D.E. Day, Freeze extrusion Lett. 93 (2013) 68–71, https://doi.org/10.1016/j.matlet.2012.11.043.
fabrication of 13-93 bioactive glass scaffolds for bone repair, J. Mater. Sci. Mater. [58] P. Tesavibul, et al., Processing of 45S5 Bioglass® by lithography-based additive
Med. 22 (3) (2011) 515–523, https://doi.org/10.1007/s10856-011-4236-4. manufacturing, Mater. Lett. 74 (2012) 81–84, https://doi.org/10.1016/j.
[34] J. Russias, E. Saiz, S. Deville, A.P. Tomsia, Fabrication and in-vitro matlet.2012.01.019.
characterization of three-dimensional composite scaffolds by robocasting for [59] M.R. Bidgoli, I. Alemzadeh, E. Tamjid, M. Khafaji, M. Vossoughi, Fabrication of
biomedical applications, Adv. Sci. Technol. 49 (2006) 153–158. https://doi.org/ hierarchically porous silk fibroin-bioactive glass composite scaffold via indirect
10.4028/www.scientific.net/ast.49.153. 3D printing: effect of particle size on physico-mechanical properties and in vitro
[35] J.A. Lewis, Direct-write assembly of ceramics from colloidal inks, Curr. Opin. cellular behavior, Mater. Sci. Eng. C 103 (2019), 109688, https://doi.org/
Solid State Mater. Sci. 6 (3) (2002) 245–250, https://doi.org/10.1016/S1359- 10.1016/j.msec.2019.04.067. November 2018.
0286(02)00031-1. [60] S. Eqtesadi, A. Motealleh, A. Pajares, P. Miranda, Effect of milling media on
[36] Q. Fu, E. Saiz, A.P. Tomsia, Bioinspired strong and highly porous glass scaffolds, processing and performance of 13-93 bioactive glass scaffolds fabricated by
Adv. Funct. Mater. 21 (6) (2011) 1058–1063, https://doi.org/10.1002/ robocasting, Ceram. Int. 41 (1) (2015) 1379–1389, https://doi.org/10.1016/j.
adfm.201002030. ceramint.2014.09.071.
[37] J. Franco, P. Hunger, M.E. Launey, A.P. Tomsia, E. Saiz, Direct write assembly of [61] Y. Lin, W. Xiao, B.S. Bal, M.N. Rahaman, Effect of copper-doped silicate 13-93
calcium phosphate scaffolds using a water-based hydrogel, Acta Biomater. 6 (1) bioactive glass scaffolds on the response of MC3T3-E1 cells in vitro and on bone
(2010) 218–228, https://doi.org/10.1016/j.actbio.2009.06.031. regeneration and angiogenesis in rat calvarial defects in vivo, Mater. Sci. Eng. C
[38] A. Sadeghi, M. Pezeshki-Modaress, M. Zandi, Electrospun polyvinyl alcohol/ 67 (2016) 440–452, https://doi.org/10.1016/j.msec.2016.05.073.
gelatin/chondroitin sulfate nanofibrous scaffold: fabrication and in vitro [62] A.M. Deliormanli, Size-dependent degradation and bioactivity of borate bioactive
evaluation, Int. J. Biol. Macromol. 114 (2018) 1248–1256, https://doi.org/ glass, Ceram. Int. 39 (7) (2013) 8087–8095, https://doi.org/10.1016/j.
10.1016/j.ijbiomac.2018.04.002. ceramint.2013.03.081.
[39] G. Ramanathan, S. Singaravelu, T. Muthukumar, S. Thyagarajan, P.T. Perumal, U. [63] J. Korpela, A. Kokkari, H. Korhonen, M. Malin, T. Narhi, J. Seppalea,
T. Sivagnanam, Design and characterization of 3D hybrid collagen matrixes as a Biodegradable and bioactive porous scaffold structures prepared using fused
dermal substitute in skin tissue engineering, Mater. Sci. Eng. C 72 (2017) deposition modeling, J. Biomed. Mater. Res. Part B Appl. Biomater. 101 (4)
359–370, https://doi.org/10.1016/j.msec.2016.11.095. (2013) 610–619, https://doi.org/10.1002/jbm.b.32863.
[40] M. Costantini, A. Barbetta, Gas foaming technologies for 3D scaffold engineering, [64] L. Bi, B. Zobell, X. Liu, M.N. Rahaman, L.F. Bonewald, Healing of critical-size
in: Functional 3D Tissue Engineering Scaffolds: Materials, Technologies, and segmental defects in rat femora using strong porous bioactive glass scaffolds,
Applications, Woodhead Publishing., 2018, pp. 127–149. Mater. Sci. Eng. C 42 (2014) 816–824, https://doi.org/10.1016/j.
[41] S.M. Mirhadi, N. Hassanzadeh Nemati, F. Tavangarian, M. Daliri Joupari, msec.2014.06.022.
Fabrication of hierarchical meso/macroporous TiO2 scaffolds by evaporation- [65] S. Eqtesadi, A. Motealleh, A. Pajares, F. Guiberteau, P. Miranda, Improving
induced self-assembly technique for bone tissue engineering applications, Mater. mechanical properties of 13-93 bioactive glass robocast scaffold by poly (lactic
Char. 144 (2018) 35–41, https://doi.org/10.1016/j.matchar.2018.06.035. June. acid) and poly (ε-caprolactone) melt infiltration, J. Non-Cryst. Solids 432 (2016)
[42] D.C. Sin, et al., Polyurethane (PU) scaffolds prepared by solvent casting/ 111–119, https://doi.org/10.1016/j.jnoncrysol.2015.02.025.
particulate leaching (SCPL) combined with centrifugation, Mater. Sci. Eng. C 30 [66] T. Fiedler, A.C. Videira, P. Bártolo, M. Strauch, G.E. Murch, J.M.F. Ferreira, On
(1) (2010) 78–85, https://doi.org/10.1016/j.msec.2009.09.002. the mechanical properties of PLC-bioactive glass scaffolds fabricated via
[43] H. Arstila, M. Tukiainen, L. Hupa, H.O. Ylänen, M. Kellomäki, M. Hupa, In vitro BioExtrusion, Mater. Sci. Eng. C 57 (2015) 288–293, https://doi.org/10.1016/j.
reactivity of bioactive glass fibers, Adv. Sci. Technol. 49 (2006) 246–251. https:// msec.2015.07.063.
doi.org/10.4028/www.scientific.net/ast.49.246. [67] A. Goel, S. Kapoor, R.R. Rajagopal, M.J. Pascual, H.W. Kim, J.M.F. Ferreira,
[44] H. Jukola, et al., Development of a bioactive glass fiber reinforced starch- Alkali-free bioactive glasses for bone tissue engineering: a preliminary
polycaprolactone composite, J. Biomed. Mater. Res. Part B Appl. Biomater. 87 (1) investigation, Acta Biomater. 8 (1) (2012) 361–372, https://doi.org/10.1016/j.
(2008) 197–203, https://doi.org/10.1002/jbm.b.31093. actbio.2011.08.026.
[45] C. Wu, Y. Luo, G. Cuniberti, Y. Xiao, M. Gelinsky, Three-dimensional printing of [68] J. Waygood, G.E. Murch, T. Fiedler, Directional and temporal variation of the
hierarchical and tough mesoporous bioactive glass scaffolds with a controllable mechanical properties of robocast scaffold during resorption, J. Mater. Sci. Mater.
pore architecture, excellent mechanical strength and mineralization ability, Acta Med. 26 (9) (2015) 1–8, https://doi.org/10.1007/s10856-015-5560-x.
Biomater. 7 (6) (2011) 2644–2650, https://doi.org/10.1016/j. [69] H. Shao, et al., Bioactive glass-reinforced bioceramic ink writing scaffolds:
actbio.2011.03.009. sintering, microstructure and mechanical behavior, Biofabrication 7 (3) (2015),
[46] A. Bernhardt, F. Despang, A. Lode, A. Demmler, T. Hanke, M. Gelinsky, 35010, https://doi.org/10.1088/1758-5090/7/3/035010.
Proliferation and osteogenic differentiation of human bone marrow stromal cells [70] A. De Marzi, Additive Manufacturing of 3D Porous Alkali-free Bioactive Glass
on alginate - gelatine - hydroxyapatite scaffolds with anisotropic pore structure, Scaffolds for Healthcare Applications,” Master’s Thesis, Materials Engineering,
J. Tissue Eng. Regen. Med. 3 (1) (2009) 54–62, https://doi.org/10.1002/ University of Aveiro, Aveiro, 2017 [Online]. Available: http://hdl.handle.net/10
term.134. 773/25518.
[47] A. García, I. Izquierdo-Barba, M. Colilla, C.L. De Laorden, M. Vallet-Regí, [71] F.A. Bhatti, New Phosphate Bioactive Glasses with Enhanced Thermal Properties
Preparation of 3-D scaffolds in the SiO 2-P 2O 5 system with tailored hierarchical for 3D Scaffold Processing Using Robocasting,” Master of Science Thesis,
meso-macroporosity, Acta Biomater. 7 (3) (2011) 1265–1273, https://doi.org/ Electrical Engineering, Tampere University, Tampere, 2019 [Online]. Available:
10.1016/j.actbio.2010.10.006. http://urn.fi/URN:NBN:fi:tuni-201911045721.
[48] T. Kokubo, H. Takadama, How useful is SBF in predicting in vivo bone [72] B.A.E. Ben-Arfa, A.S. Neto, I.E. Palamá, I.M. Miranda Salvado, R.C. Pullar, J.M.
bioactivity? Biomaterials 27 (15) (2006) 2907–2915, https://doi.org/10.1016/j. F. Ferreira, Robocasting of ceramic glass scaffolds: sol–gel glass, new horizons,
biomaterials.2006.01.017. J. Eur. Ceram. Soc. 39 (4) (2019) 1625–1634, https://doi.org/10.1016/j.
[49] N. Özkan, C. Oysu, B.J. Briscoe, I. Aydin, Rheological analysis of ceramic pastes, jeurceramsoc.2018.11.019.
J. Eur. Ceram. Soc. 19 (16) (1999) 2883–2891, https://doi.org/10.1016/S0955- [73] B.A.E. Ben-Arfa, S. Neto, I.M. Miranda Salvado, R.C. Pullar, J.M.F. Ferreira,
2219(99)00054-0. Robocasting of Cu 2+ & La 3+ doped sol–gel glass scaffolds with greatly
[50] Q. Fu, E. Saiz, A.P. Tomsia, Direct ink writing of highly porous and strong glass enhanced mechanical properties: compressive strength up to 14 MPa, Acta
scaffolds for load-bearing bone defects repair and regeneration, Acta Biomater. 7 Biomater. 87 (2019) 265–272, https://doi.org/10.1016/j.actbio.2019.01.048.
(10) (2011) 3547–3554, https://doi.org/10.1016/j.actbio.2011.06.030. [74] A.M. Deliormanli, M.N. Rahaman, Direct-write assembly of silicate and borate
[51] H. Elsayed, M. Picicco, A. Dasan, J. Kraxner, D. Galusek, E. Bernardo, Glass bioactive glass scaffolds for bone repair, J. Eur. Ceram. Soc. 32 (14) (2012)
powders and reactive silicone binder: interactions and application to additive 3637–3646, https://doi.org/10.1016/j.jeurceramsoc.2012.05.005.

22
B.C. Palivela et al. Bioprinting 27 (2022) e00219

[75] J.M.F. Nan, Bo, Przemysław Gołębiewski, Ryszard Buczyński, Galindo-Rosales, [99] W. Xiao, M.A. Zaeem, G. Li, B. Sonny Bal, M.N. Rahaman, Tough and strong
J. Francisco, Ferreira, Direct ink writing glass, Preliminary Step Optical Appl. 13 porous bioactive glass-PLA composites for structural bone repair, J. Mater. Sci. 52
(7) (2020) 1636, https://doi.org/10.3390/ma13071636. (15) (2017) 9039–9054, https://doi.org/10.1007/s10853-017-0777-3.
[76] A. Nommeots-Nomm, et al., Four-dimensional imaging and quantification of [100] A.M. Deliormanll, A.H. Deliormanll, Finite element method simulation for the
viscous flow sintering within a 3D printed bioactive glass scaffold using prediction of mechanical properties of three-dimensional periodic bioactive glass
synchrotron X-ray tomography, Mater. Today Adv. 2 (2019), 100011, https://doi. scaffolds, J. Australas. Ceram. Soc. 53 (2) (2017) 299–307, https://doi.org/
org/10.1016/j.mtadv.2019.100011. 10.1007/s41779-017-0037-7.
[77] M. Elbadawi, Z.J. Wally, I.M. Reaney, Porous hydroxyapatite-bioactive glass [101] S.M. Olhero, H.R. Fernandes, C.F. Marques, B.C.G. Silva, J.M.F. Ferreira, Additive
hybrid scaffolds fabricated via ceramic honeycomb extrusion, J. Am. Ceram. Soc. manufacturing of 3D porous alkali-free bioactive glass scaffolds for healthcare
101 (8) (2018) 3541–3556, https://doi.org/10.1111/jace.15514. applications, J. Mater. Sci. 52 (20) (2017) 12079–12088, https://doi.org/
[78] A. Nommeots-Nomm, P.D. Lee, J.R. Jones, Direct ink writing of highly bioactive 10.1007/s10853-017-1347-4.
glasses, J. Eur. Ceram. Soc. 38 (3) (2018) 837–844, https://doi.org/10.1016/j. [102] A. Motealleh, S. Eqtesadi, A. Pajares, P. Miranda, D. Salamon, K. Castkova, Case
jeurceramsoc.2017.08.006. study: reinforcement of 45S5 bioglass robocast scaffolds by HA/PCL
[79] J. Barberi, A. Nommeots-Nomm, E. Fiume, E. Verné, J. Massera, F. Baino, nanocomposite coatings, J. Mech. Behav. Biomed. Mater. 75 (2017) 114–118,
Mechanical characterization of pore-graded bioactive glass scaffolds produced by https://doi.org/10.1016/j.jmbbm.2017.07.012.
robocasting, Biomed. Glas. 5 (1) (2019) 140–147, https://doi.org/10.1515/ [103] A. Richter, Richard Frank, Tilman Ahlfeld, Michael Gelinsky, Lode, Development
bglass-2019-0012. and characterization of composites consisting of calcium phosphate cements and
[80] J. Barberi, et al., Robocasting of SiO2-based bioactive glass scaffolds with mesoporous bioactive glass for extrusion-based fabrication, Materials (Basel) 12
porosity gradient for bone regeneration and potential load-bearing applications, (12) (2019) 2022, https://doi.org/10.3390/ma12122022.
Materials (Basel) 12 (7) (2019), https://doi.org/10.3390/ma12172691. [104] C.H. Oh, S.J. Hong, I. Jeong, H.S. Yu, S.H. Jegal, H.W. Kim, Development of
[81] N.C. Paxton, et al., Rheological characterization of biomaterials directs additive robotic dispensed bioactive scaffolds and human adipose-derived stem cell
manufacturing of strontium-substituted bioactive glass/polycaprolactone culturing for bone tissue engineering, Tissue Eng. C Methods 16 (4) (2010)
microfibers, Macromol. Rapid Commun. 40 (11) (2019) 1–6, https://doi.org/ 561–571, https://doi.org/10.1089/ten.tec.2009.0274.
10.1002/marc.201900019. [105] Caroline Murphy, K. Kolan, W. Li, J. Semon, D. Day, M. Leu, 3D bioprinting of
[82] A.E. Nommeots-Nomm, 3D Printing versus Foaming of Melt-Derived Bioactive stem cells and polymer/bioactive glass composite scaffolds for bone tissue
Glasses for Bone Regeneration, Ph.D. dissertation, Department of Materials, engineering, Int. J. Bioprinting 3 (1) (2017) 53–63, https://doi.org/10.18063/
Imperial College London, London, 2015 [Online]. Available: http://hdl.handle. IJB.2017.01.005.
net/10044/1/50295. [106] K.C.R. Kolan, J.A. Semon, A.T. Bindbeutel, D.E. Day, M.C. Leu, Bioprinting with
[83] X. Shi, Ph.D. dissertation, in: Vitro and in Vivo Behaviour of Bioactive Glass bioactive glass loaded polylactic acid composite and human adipose stem cells,
Scaffold, Department of Materials, Imperial College London, London, 2018 Bioprinting 18 (2020), e00075, https://doi.org/10.1016/j.bprint.2020.e00075.
[Online]. Available: http://hdl.handle.net/10044/1/68257. December 2019.
[84] C. Gao, M.N. Rahaman, Q. Gao, A. Teramoto, K. Abe, Robotic deposition and in [107] K. Kolan, et al., Solvent based 3D printing of biopolymer/bioactive glass
vitro characterization of 3D gelatin-bioactive glass hybrid scaffolds for composite and hydrogel for tissue engineering applications, Procedia CIRP 65
biomedical applications, J. Biomed. Mater. Res., Part A 101 A (7) (2013) (2017) 38–43, https://doi.org/10.1016/j.procir.2017.04.022.
2027–2037, https://doi.org/10.1002/jbm.a.34496. [108] H. Liu, Static vs. Dynamic Dissolution and Cellular Compatibility Tests of
[85] K. Dixit, N. Sinha, Additively manufactured nanofiber reinforced bioactive glass Bioactive Borosilicate Glass Scaffolds,” Master’s Thesis, Bioengineering, Tampere
based functionally graded scaffolds for bone tissue engineering, IEEE Int. Conf. University, Tampere, 2020 [Online]. Available: http://urn.fi/URN:NBN:fi:tuni
Nano/Molecular Med. Eng. NANOMED (2019) 47–51, https://doi.org/10.1109/ -202003022447.
NANOMED49242.2019.9130605, 2020-Novem. [109] K.C.R. Kolan, et al., Near-field electrospinning of a polymer/bioactive glass
[86] S. Eqtesadi, A. Motealleh, F.H. Perera, A. Pajares, P. Miranda, Poly-(lactic acid) composite to fabricate 3D biomimetic structures, Int. J. Bioprinting 5 (1) (2019)
infiltration of 45S5 Bioglass® robocast scaffolds: chemical interaction and its 1–6, https://doi.org/10.18063/ijb.v5i1.163.
deleterious effect in mechanical enhancement, Mater. Lett. 163 (2016) 196–200, [110] A. Gaddam, D.S. Brazete, A.S. Neto, B. Nan, H.R. Fernandes, J.M.F. Ferreira,
https://doi.org/10.1016/j.matlet.2015.10.073. Robocasting and surface functionalization with highly bioactive glass of ZrO2
[87] A. El-Fiqi, J.H. Lee, E.J. Lee, H.W. Kim, Collagen hydrogels incorporated with scaffolds for load bearing applications, J. Am. Ceram. Soc., no. April (2021) 1–12,
surface-aminated mesoporous nanobioactive glass: improvement of https://doi.org/10.1111/jace.17869.
physicochemical stability and mechanical properties is effective for hard tissue [111] A.M. Deliormanlı, M. Türk, H. Atmaca, Response of mouse bone marrow
engineering, Acta Biomater. 9 (12) (2013) 9508–9521, https://doi.org/10.1016/ mesenchymal stem cells to graphene-containing grid-like bioactive glass scaffolds
j.actbio.2013.07.036. produced by robocasting, J. Biomater. Appl. 33 (4) (2018) 488–500, https://doi.
[88] S.L. Greasley, Bioactive Glass Nanoparticles for Therapeutic Applications, Ph.D. org/10.1177/0885328218799610.
dissertation, Department of Materials, Imperial College London, London, 2015 [112] E. Farina, et al., Micro computed tomography based finite element models for
[Online]. Available: http://hdl.handle.net/10044/1/58227. elastic and strength properties of 3D printed glass scaffolds, Acta Mech. Sin.
[89] S. Eqtesadi, A. Motealleh, A. Pajares, F. Guiberteau, P. Miranda, Influence of Xuebao 37 (2) (2021) 292–306, https://doi.org/10.1007/s10409-021-01065-3.
sintering temperature on the mechanical properties of ε-PCL-impregnated 45S5 [113] S. Shruti, A.J. Salinas, G. Lusvardi, G. Malavasi, L. Menabue, M. Vallet-Regi,
bioglass-derived scaffolds fabricated by robocasting, J. Eur. Ceram. Soc. 35 (14) Mesoporous bioactive scaffolds prepared with cerium-, gallium- and zinc-
(2015) 3985–3993, https://doi.org/10.1016/j.jeurceramsoc.2015.06.021. containing glasses, Acta Biomater. 9 (1) (2013) 4836–4844, https://doi.org/
[90] M.J. Donachie, Titanium - A Techincal Guide 99 (5) (2000). 10.1016/j.actbio.2012.09.024.
[91] A. Thomas, K.C.R. Kolan, M.C. Leu, G.E. Hilmas, Freeform extrusion fabrication of [114] F. Baino, J. Barberi, E. Fiume, G. Orlygsson, J. Massera, E. Verné, Robocasting of
titanium fiber reinforced 13–93 bioactive glass scaffolds, J. Mech. Behav. Biomed. bioactive SiO2-P2O5-CaO-MgO-Na2O-K2O glass scaffolds, J. Healthc. Eng.
Mater. 69 (2017) 153–162, https://doi.org/10.1016/j.jmbbm.2016.12.024. (2019), https://doi.org/10.1155/2019/5153136, 2019.
September 2016. [115] X. Qi, et al., Three dimensional printing of calcium sulfate and mesoporous
[92] S.I. Roohani-Esfahani, Y. Chen, J. Shi, H. Zreiqat, Fabrication and bioactive glass scaffolds for improving bone regeneration in vitro and in vivo, Sci.
characterization of a new, strong and bioactive ceramic scaffold for bone Rep. 7 (2017) 2–13, https://doi.org/10.1038/srep42556. January.
regeneration, Mater. Lett. 107 (2013) 378–381, https://doi.org/10.1016/j. [116] N. Gómez-Cerezo, et al., Mesoporous bioactive glass/ϵ-polycaprolactone scaffolds
matlet.2013.06.046. promote bone regeneration in osteoporotic sheep, Acta Biomater. 90 (2019)
[93] S.I. Roohani-Esfahani, et al., Unique microstructural design of ceramic scaffolds 393–402, https://doi.org/10.1016/j.actbio.2019.04.019.
for bone regeneration under load, Acta Biomater. 9 (6) (2013) 7014–7024, [117] X. Liu, M.N. Rahaman, G.E. Hilmas, B.S. Bal, Mechanical properties of bioactive
https://doi.org/10.1016/j.actbio.2013.02.039. glass (13-93) scaffolds fabricated by robotic deposition for structural bone repair,
[94] P. Miranda, A. Pajares, E. Saiz, A.P. Tomsia, F. Guiberteau, Mechanical properties Acta Biomater. 9 (6) (2013) 7025–7034, https://doi.org/10.1016/j.
of calcium phosphate scaffolds fabricated by robocasting, J. Biomed. Mater. Res., actbio.2013.02.026.
Part A 85 (1) (2008) 218–227, https://doi.org/10.1002/jbm.a.31587. [118] X. Liu, M.N. Rahaman, Y. Liu, B.S. Bal, L.F. Bonewald, Enhanced bone
[95] S.I. Roohani-Esfahani, P. Newman, H. Zreiqat, Design and fabrication of 3D regeneration in rat calvarial defects implanted with surface-modified and BMP-
printed scaffolds with a mechanical strength comparable to cortical bone to repair loaded bioactive glass (13-93) scaffolds, Acta Biomater. 9 (7) (2013) 7506–7517,
large bone defects, Sci. Rep. 6 (2016) 1–8, https://doi.org/10.1038/srep19468. https://doi.org/10.1016/j.actbio.2013.03.039.
January. [119] Y. Lin, W. Xiao, X. Liu, B.S. Bal, L.F. Bonewald, M.N. Rahaman, Long-term bone
[96] W. Xiao, M.A. Zaeem, B.S. Bal, M.N. Rahaman, Creation of bioactive glass regeneration, mineralization and angiogenesis in rat calvarial defects implanted
(13–93) scaffolds for structural bone repair using a combined finite element with strong porous bioactive glass (13-93) scaffolds, J. Non-Cryst. Solids 432
modeling and rapid prototyping approach, Mater. Sci. Eng. C 68 (2016) 651–662, (2016) 120–129, https://doi.org/10.1016/j.jnoncrysol.2015.04.008.
https://doi.org/10.1016/j.msec.2016.06.011. [120] S. Eqtesadi, A. Motealleh, P. Miranda, A. Pajares, A. Lemos, J.M.F. Ferreira,
[97] A. Motealleh, S. Eqtesadi, F.H. Perera, A. Pajares, F. Guiberteau, P. Miranda, Robocasting of 45S5 bioactive glass scaffolds for bone tissue engineering, J. Eur.
Understanding the role of dip-coating process parameters in the mechanical Ceram. Soc. 34 (1) (2014) 107–118, https://doi.org/10.1016/j.
performance of polymer-coated bioglass robocast scaffolds, J. Mech. Behav. jeurceramsoc.2013.08.003.
Biomed. Mater. 64 (2016) 253–261, https://doi.org/10.1016/j. [121] A. Motealleh, S. Eqtesadi, A. Civantos, A. Pajares, P. Miranda, Robocast 45S5
jmbbm.2016.08.004. bioglass scaffolds: in vitro behavior, J. Mater. Sci. 52 (15) (2017) 9179–9191,
[98] N. Gómez-Cerezo, S. Sánchez-Salcedo, I. Izquierdo-Barba, D. Arcos, M. Vallet- https://doi.org/10.1007/s10853-017-0775-5.
Regí, In vitro colonization of stratified bioactive scaffolds by pre-osteoblast cells, [122] A. Lenau, Robocasting Bone Scaffolds for Bone Tissue Regeneration, Alfred
Acta Biomater. 44 (2016) 73–84, https://doi.org/10.1016/j.actbio.2016.08.014. University Honors Program Thesis, Department of Material Science and

23
B.C. Palivela et al. Bioprinting 27 (2022) e00219

Engineering, Alfred University, Alfred, 2019 [Online]. Available: http://hdl.han regeneration (Biofabrication (2013) 5 (045005)), Biofabrication 6 (2) (2014),
dle.net/10829/23405. https://doi.org/10.1088/1758-5082/6/2/029501.
[123] V.P. Galván-Chacón, S. Eqtesadi, A. Pajares, P. Miranda, F. Guiberteau, [141] Q. Fu, E. Saiz, M.N. Rahaman, A.P. Tomsia, Bioactive glass scaffolds for bone
Elucidating the role of 45S5 bioglass content in the density and flexural strength tissue engineering : state of the art and future perspectives, Mater. Sci. Eng. C 31
of robocast β-TCP/45S5 composites, Ceram. Int. 44 (11) (2018) 12717–12722, (7) (2011) 1245–1256, https://doi.org/10.1016/j.msec.2011.04.022.
https://doi.org/10.1016/j.ceramint.2018.04.074. [142] J. Henkel, et al., Bone regeneration based on tissue engineering conceptions-A
[124] B.A.E. Ben-Arfa, I.M. Miranda Salvado, J.R. Frade, R.C. Pullar, Guidelines to 21st century perspective, Bone Res 1 (3) (2013) 216–248, https://doi.org/
adjust particle size distributions by wet comminution of a bioactive glass 10.4248/BR201303002.
determined by Taguchi and multivariate analysis, Ceram. Int. 45 (3) (2019) [143] T. Kokubo, H. Kushitani, S. Sakka, T. Kitsugi, T. Yamamuro, Solutions able to
3857–3863, https://doi.org/10.1016/j.ceramint.2018.11.057. reproduce in vivo surface-structure changes in bioactive glass-ceramic A-W3,
[125] B.A.E. Ben-Arfa, A.S. Neto, I.M. Miranda Salvado, R.C. Pullar, J.M.F. Ferreira, J. Biomed. Mater. Res. 24 (6) (1990) 721–734, https://doi.org/10.1002/
Robocasting: prediction of ink printability in solgel bioactive glass, J. Am. Ceram. jbm.820240607.
Soc. 102 (4) (2019) 1608–1618, https://doi.org/10.1111/jace.16092. [144] D.S. Couto, Z. Hong, J.F. Mano, Development of bioactive and biodegradable
[126] C. Vitale-Brovarone, et al., Resorbable glass-ceramic phosphate-based scaffolds chitosan-based injectable systems containing bioactive glass nanoparticles, Acta
for bone tissue engineering: synthesis, properties, and in vitro effects on human Biomater. 5 (1) (2009) 115–123, https://doi.org/10.1016/j.actbio.2008.08.006.
marrow stromal cells, J. Biomater. Appl. 26 (4) (2011) 465–489, https://doi.org/ [145] A.M. Deliormanl, In vitro assessment of degradation and bioactivity of robocast
10.1177/0885328210372149. bioactive glass scaffolds in simulated body fluid, Ceram. Int. 38 (8) (2012)
[127] O. Bretcanu, F. Baino, E. Verné, C. Vitale-Brovarone, Novel resorbable glass- 6435–6444, https://doi.org/10.1016/j.ceramint.2012.05.019.
ceramic scaffolds for hard tissue engineering: from the parent phosphate glass to [146] M. Kaszuba, J. Corbett, F.M.N. Watson, A. Jones, High-concentration zeta
its bone-like macroporous derivatives, J. Biomater. Appl. 28 (9) (2014) potential measurements using light-scattering techniques, Philos. Trans. R. Soc. A
1287–1303, https://doi.org/10.1177/0885328213506759. Math. Phys. Eng. Sci. 368 (1927) (2010) 4439–4451, https://doi.org/10.1098/
[128] F. Baino, C. Vitale-Brovarone, Mechanical properties and reliability of glass- rsta.2010.0175.
ceramic foam scaffolds for bone repair, Mater. Lett. 118 (2014) 27–30, https:// [147] X.J. Yang, P.J. Zheng, Z.D. Cui, N.Q. Zhao, Y.F. Wang, K. De Yao, Swelling
doi.org/10.1016/j.matlet.2013.12.037. behaviour and elastic properties of gelatin gels, Polym. Int. 44 (4) (1997)
[129] N. Paxton, W. Smolan, T. Böck, F. Melchels, J. Groll, T. Jungst, Proposal to assess 448–452, https://doi.org/10.1002/(SICI)1097-0126(199712)44:4<448::AID-
printability of bioinks for extrusion-based bioprinting and evaluation of PI845>3.0.CO;2-M.
rheological properties governing bioprintability, Biofabrication 9 (4) (2017), [148] P.S.P. Poh, D.W. Hutmacher, B.M. Holzapfel, A.K. Solanki, M.A. Woodruff, Data
https://doi.org/10.1088/1758-5090/aa8dd8. for accelerated degradation of calcium phosphate surface-coated
[130] S.J. Hong, I. Jeong, K.T. Noh, H.S. Yu, G.S. Lee, H.W. Kim, Robotic dispensing of polycaprolactone and polycaprolactone/bioactive glass composite scaffolds, Data
composite scaffolds and in vitro responses of bone marrow stromal cells, J. Mater. Br 7 (2016) 923–926, https://doi.org/10.1016/j.dib.2016.01.023.
Sci. Mater. Med. 20 (9) (2009) 1955–1962, https://doi.org/10.1007/s10856-009- [149] A. Hoppe, N.S. Güldal, A.R. Boccaccini, A review of the biological response to
3745-x. ionic dissolution products from bioactive glasses and glass-ceramics, Biomaterials
[131] Y. Gu, W. Huang, M.N. Rahaman, In vivo evaluation of scaffolds with a grid-like 32 (11) (2011) 2757–2774, https://doi.org/10.1016/j.biomaterials.2011.01.004.
microstructure composed of a mixture of silicate (13-93) and borate (13-93B3) [150] P. Habibovic, J.E. Barralet, Bioinorganics and biomaterials: bone repair, Acta
bioactive glasses, Ceram. Eng. Sci. Proceedings, Am. Ceram. Soc. 35 (5) (2014) Biomater. 7 (8) (2011) 3013–3026, https://doi.org/10.1016/j.
53–64, https://doi.org/10.1002/9781119040392.ch6. actbio.2011.03.027.
[132] A. Entezari, et al., On design for additive manufacturing (DAM) parameter and its [151] M.N. Rahaman, et al., Bioactive glass in tissue engineering, Acta Biomater. 7 (6)
effects on biomechanical properties of 3D printed ceramic scaffolds, Mater. Today (2011) 2355–2373, https://doi.org/10.1016/j.actbio.2011.03.016.
Commun. 23 (March, 2020), https://doi.org/10.1016/j.mtcomm.2020.101065. [152] T. Fiedler, et al., A comparative study of oxygen diffusion in tissue engineering
[133] A. Gaddam, D.S. Brazete, A.S. Neto, B. Nan, J.M.F. Ferreira, Three-dimensional scaffolds, J. Mater. Sci. Mater. Med. 25 (11) (2014) 2573–2578, https://doi.org/
printing of zirconia scaffolds for load bearing applications: study of the optimal 10.1007/s10856-014-5264-7.
fabrication conditions, J. Am. Ceram. Soc. 104 (9) (2021) 4368–4380, https:// [153] S. Midha, T.B. Kim, W. Van Den Bergh, P.D. Lee, J.R. Jones, C.A. Mitchell,
doi.org/10.1111/jace.17874. Preconditioned 70S30C bioactive glass foams promote osteogenesis in vivo, Acta
[134] U.N. Mughal, H.A. Khawaja, M. Moatamedi, Finite element analysis of human Biomater. 9 (11) (2013) 9169–9182, https://doi.org/10.1016/j.
femur bone, Int. J. Multiphys. 9 (2) (2015) 101–108, https://doi.org/10.1260/ actbio.2013.07.014.
1750-9548.9.2.101. [154] D. Hutmacher et al., "State of the Art and Future Directions of Scaffold-Based Bone
[135] K.C.N. Kumar, T. Tandon, P. Silori, A. Shaikh, Biomechanical stress analysis of a Engineering from a Biomaterials Perspective Scaffold-Based Bone Engineering
human femur bone using ANSYS, Mater. Today Proc. 2 (4–5) (2015) 2115–2120, from a Biomaterials Perspective," DOI: 10.1002/term.24.
https://doi.org/10.1016/j.matpr.2015.07.211. [155] T.S. Karande, Effect of Scaffold Architecture on Diffusion of Oxygen in Tissue
[136] A. Dhanopia, M. Bhargava, Finite element analysis of human fractured femur Engineering Constructs," Ph.D. Dissertation, Department of Biomedical
bone implantation with PMMA thermoplastic prosthetic plate, Procedia Eng. 173 Engineering, University of Texas, Austin, 2007 [Online]. Available, http://hdl.
(2017) 1658–1665, https://doi.org/10.1016/j.proeng.2016.12.190. handle.net/2152/3270.
[137] D. Bellucci, V. Cannillo, A. Sola, Coefficient of thermal expansion of bioactive [156] C.A. Schneider, W.S. Rasband, K.W. Eliceiri, NIH Image to ImageJ: 25 years of
glasses: available literature data and analytical equation estimates, Ceram. Int. 37 image analysis, Nat. Methods 9 (7) (2012) 671–675, https://doi.org/10.1038/
(8) (2011) 2963–2972, https://doi.org/10.1016/j.ceramint.2011.05.048. nmeth.2089.
[138] S. Zhao, et al., Three-dimensional printed strontium-containing mesoporous [157] M. Doube, et al., BoneJ: free and extensible bone image analysis in ImageJ, Bone
bioactive glass scaffolds for repairing rat critical-sized calvarial defects, Acta 47 (6) (2010) 1076–1079, https://doi.org/10.1016/j.bone.2010.08.023.
Biomater. 12 (1) (2015) 270–280, https://doi.org/10.1016/j.actbio.2014.10.015. [158] L. Bi, et al., Evaluation of bone regeneration, angiogenesis, and hydroxyapatite
[139] L. Shor, S. Güçeri, X. Wen, M. Gandhi, W. Sun, Fabrication of three-dimensional conversion in critical-sized rat calvarial defects implanted with bioactive glass
polycaprolactone/hydroxyapatite tissue scaffolds and osteoblast-scaffold scaffolds, J. Biomed. Mater. Res. Part A, vol. 100 A, no 12 (2012) 3267–3275,
interactions in vitro, Biomaterials 28 (35) (2007) 5291–5297, https://doi.org/ https://doi.org/10.1002/jbm.a.34272.
10.1016/j.biomaterials.2007.08.018. [159] Q. Fu, M.N. Rahaman, H. Fu, X. Liu, Silicate, borosilicate, and borate bioactive
[140] P.S.P. Poh, D.W. Hutmacher, M.M. Stevens, M.A. Woodruff, Erratum: fabrication glass scaffolds with controllable degradation rate for bone tissue engineering
and in vitro characterization of bioactive glass composite scaffolds for bone applications. I. Preparation and in vitro degradation, J. Biomed. Mater. Res., Part
A 95 (1) (2010) 164–171, https://doi.org/10.1002/jbm.a.32824.

24

You might also like