You are on page 1of 22

Acta Biomaterialia 113 (2020) 1–22

Contents lists available at ScienceDirect

Acta Biomaterialia
journal homepage: www.elsevier.com/locate/actbio

Advancing bioinks for 3D bioprinting using reactive fillers: A review


Susanne Heid, Aldo R. Boccaccini∗
Institute of Biomaterials, University of Erlangen-Nuremberg, Cauerstraße 6, 91058 Erlangen, Germany

a r t i c l e i n f o a b s t r a c t

Article history: The growing demand for personalized implants and tissue scaffolds requires advanced biomaterials and
Received 7 March 2020 processing strategies for the fabrication of three-dimensional (3D) structures mimicking the complexity
Revised 26 June 2020
of the extracellular matrix. During the last years, biofabrication approaches like 3D printing of cell-laden
Accepted 26 June 2020
(soft) hydrogels have been gaining increasing attention to design such 3D functional environments which
Available online 2 July 2020
resemble natural tissues (and organs). However, often these polymeric hydrogels show poor stability and
Keywords: low printing fidelity and hence various approaches in terms of multi-material mixtures are being devel-
Composite hydrogels oped to enhance pre- and post-printing features as well as cytocompatibility and post-printing cellular
Bioprinting development. Additionally, bioactive properties improve the binding to the surrounding (host) tissue at
Cell encapsulation the implantation site. In this review we focus on the state-of-the-art of a particular type of heterogeneous
Bioactive inorganic fillers bioinks, which are composed of polymeric hydrogels incorporating inorganic bioactive fillers. Such sys-
Bioactive glasses
tems include isotropic and anisotropic silicates like bioactive glasses and nanoclays or calcium-phosphates
Laponite
Hydroxyapatite
like hydroxyapatite (HAp), which provide in-situ crosslinking effects and add extra functionality to the
matrix, for example mineralization capability. The present review paper discusses in detail such bioac-
tive composite bioink systems based on the available literature, revealing that a great variety has been
developed with substantially improved bioprinting characteristics, in comparison to the pure hydrogel
counterparts, and enabling high viability of printed cells. The analysis of the results of the published
studies demonstrates that bioactive fillers are a promising addition to hydrogels to print stable 3D con-
structs for regeneration of tissues. Progress and challenges of the development and applications of such
composite bioink approaches are discussed and avenues for future research in the field are presented.

Statement of Significance

Biofabrication, involving the processing of biocompatible hydrogels including cells (bioinks), is being in-
creasingly applied for developing complex tissue and organ mimicking structures. A variety of multi-
material bioinks is being investigated to bioprint 3D constructs showing shape stability and long-term
biological performance. Composite hydrogel bioinks incorporating inorganic bioreactive fillers for 3D bio-
printing are the subject of this review paper. Results reported in the literature highlight the effect of
bioactive fillers on bioink properties, printability and on cell behavior during and after printing and pro-
vide important information for optimizing the design of future bioinks for biofabrication, exploiting the
extra functionalities provided by inorganic fillers. Further functionalization with drugs/growth factors can
target enhanced printability and local drug release for more specialized biomedical therapies.
© 2020 Acta Materialia Inc. Published by Elsevier Ltd.
This is an open access article under the CC BY-NC-ND license.
(http://creativecommons.org/licenses/by-nc-nd/4.0/)

1. Introduction medicine and tissue engineering [1,2]. As cells in the human body
are located within a three-dimensional matrix, this method allows
Biofabrication is gaining increasing attention to create com- to resemble and mimic the original 3D structures and functions
plex 3D structures mimicking tissues and organs in regenerative of living tissues (and organs) [3]. Additionally, with this approach
a platform for drug testing and for the analysis of tissue mor-

Corresponding author: Prof. Aldo R. Boccaccini: Cauerstraße 6, 91058 Erlangen,
phogenesis can be generated as cells are encapsulated in a suit-
Germany. able hydrogel matrix and subsequently processed to 3D structures,
E-mail address: aldo.boccaccini@ww.uni-erlangen.de (A.R. Boccaccini). for example by 3D bioprinting [3–5]. To differentiate between

https://doi.org/10.1016/j.actbio.2020.06.040
1742-7061/© 2020 Acta Materialia Inc. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license.
(http://creativecommons.org/licenses/by-nc-nd/4.0/)
2 S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22

Fig. 1. Distinction between a bioink (left side) and a biomaterial ink (right side). In a bioink, cells are a mandatory component of the printing formulation in the form
of single cells, coated cells or cell aggregates (of one or several cell types), or also in combination with materials (for example seeded onto microcarriers, embedded in
microgels, formulated in a physical hydrogel, or formulated with hydrogel precursors). In case of a biomaterial ink, in principle any biomaterial can be used for printing and
cell-seeding occurs post-fabrication. The images in this scheme are not displayed in scale. (Reproduced from ref. [13] according to the terms of the Creative Commons CC BY
license).

bioprinting and biofabrication the “International Society for Biofab- defined as materials which are capable to include cells and other
rication” suggested a new definition of “Biofabrication” in the con- bioactive components for the use in biofabrication [13].
text of tissue engineering and regenerative medicine within which The need for processing ‘with cells’ has increased the challenges
“Bioprinting” and “Bioassembly” are two complementary technolo- that must be tackled in the development of suitable bioinks. To
gies followed by tissue maturation. Thus biofabrication is defined fulfill as many requirements as possible (regarding bioinks and
as “the automated generation of biologically functional products with bioprinted scaffolds), composite hydrogels consisting of a hydro-
structural organization from living cells, bioactive molecules, bioma- gel matrix and inorganic fillers represent attractive systems, which
terials, cell aggregates such as micro-tissues, or hybrid cell-material have been specially investigated for bone and cartilage tissue re-
constructs, through bioprinting or bioassembly and subsequent tissue construction so far [25–29]. This paper reviews polymeric hydro-
maturation processes” [6]. gels with incorporated bioactive inorganic fillers and encapsulated
Using bioprinting, cells in combination with hydrogels (acting cells for the use in bioprinting, here referred to as “composite
as a temporary extracellular matrix) and active substances can be bioinks”, published in the last six years. In literature, there are
precisely distributed hierarchically and spatially according to the also other definitions for composite bioinks. Some authors have
desired three-dimensional function [7]. Hence, bioprinting provides used the expression for naming mixtures of two or more polymers
free-shaping design in combination with controlled positioning of yielding blends or multi-phase materials [30–32], however in this
materials and cells in one step by avoiding a post-cell-seeding pro- review only hydrogel + inorganic filler composites are considered
cedure, which is commonly done in classical tissue engineering under the expression “composite bioink”. The current literature is
[8,9]. Bioprinting additionally implicates an intimate connection analyzed and discussed regarding printability, printing fidelity and
between cells and materials as well as a higher cell loading effi- cell viability of composite bioinks incorporating bioreactive inor-
ciency with homogenous or tailored cell distribution which cannot ganic fillers, e.g. fillers which are not ‘persistent’ but can change
be achieved by conventional cell culture methods [10]. By choosing their properties as function of time (for example by partially dis-
the appropriate cell type and density, post-printing cell behavior solving in the hydrogel matrix).
can be analyzed including favorable or impaired cell proliferation, Overall, this review intends to present a comprehensive sum-
maintenance of cell integrity and morphology and preservation of mary of the field of composite hydrogels for bioprinting, discussing
phenotype and genotype [11]. Furthermore, the selection of phys- a broad variety of such composite bioinks and their properties, re-
iologically relevant culture systems gives the opportunity to gain cently reported for 3D bioprinting. Only such composite bioinks
insights into the fundamental biological processes of living cells in were taken into account which consist of a polymeric matrix (hy-
3D architectures [12]. drogel) and bioactive inorganic filler(s) and which are bioprinted
As suggested by Groll et al. [13] the term “Bioink“ should be with embedded cells. The cited references found in the databases
clearly defined and differentiated from the term “Biomaterial ink” (as listed below) were grouped by the type of inorganic filler.
considering hydrogels for 3D printing, as displayed in Fig. 1. Bio- Previous comprehensive review papers are available [9,33–36],
material inks can be printed and if needed sterilized and sub- which have described the broad field of bioinks for bioprint-
sequently seeded with cells for applications as scaffold materials ing/biofabrication, however with limited discussion of the type of
or implants or for hybrid scaffolds providing mechanical support composite bioinks on which we focus the present review. Equally,
in 3D printed constructs [14–16]. Bioinks on the other hand are composite hydrogels with inorganic components have been
S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22 3

discussed in previous articles [16,37–40], but not with focus on a combination of both [58]. New multi-material bioinks are cur-
3D bioprinting approaches, e.g. such previous reviews have not rently being developed to not only improve the printability but
covered in detail composite hydrogels incorporating cells for bio- also to enhance features like mechanical resistance and stiffness,
printing. As such, a gap in the current literature is filled with the to tune the degradation rate, or to add bioactive moieties which
present review article. In this literature review, we have focused on show significant impact on cell behavior and tissue formation
the most relevant work published in the last six years regarding [30,59–62]. One particular type of such multi-material bioinks are
the use of nanostructured silica, bioactive glasses, calcium phos- hydrogel-inorganic phase composite systems (composite bioinks),
phate or strontium-based bioactive materials and their applica- which are the subject of this review. By adding different nanopar-
tions in the field of 3D bioprinting. PRISMA guidelines for struc- ticles or anisotropic fillers, e.g. graphene, graphene oxide, carbon
tured reviews were followed and the literature search was carried nanotubes (CNTs), hydroxyapatite (HAp) and other calcium phos-
out on platforms like Web of Science R
, Wiley database, Scopus R
, phates, bioactive glasses, silica nanoparticles, and nanoclays, the
Pubmed, Google Scholar R
, Science Direct, as well as Springer, mechanical and biological characteristics of nanocomposite hydro-
Mendeley R
, IOP Science and science.gov databases. For the liter- gels can be improved [38,63–71]. For example, with the incorpo-
ature search, different combinations of the keywords “composite”, ration of such fillers physicochemical properties affecting bioprint-
“bioink”, “hydrogel”, “biofabrication”, “3D (bio)printing”, “cell encap- ability can be tuned, like viscosity, stiffness, nonlinear viscoelas-
sulation”, “bioactive”, “inorganic”, “filler”, “bioactive glass”, “laponite”, ticity (thixotropy/rheopexy), surface tension or density. Mechani-
“hydroxyapatite”, “tricalcium phosphate”, “clay”, “silica”, “silicate” and cal properties of the bioinks should be adjusted to the used cell
“carbonate” were used. Publications that did not report cell incor- type and the desired cell function. Additionally, due to the different
porated composite hydrogels or where the word “composite” was chemical compositions of the inorganic fillers and the release of
(erroneously) used to indicate blend hydrogels were not consid- specific ions (upon dissolution, in case of bioreactive fillers), most
ered. of these fillers are able to support cell adhesion and proliferation
and some of them also influence differentiation of stem cells, pro-
genitor cells or cell lines, leading to, for example, osteogenic (e.g.
2. Composite bioinks bioceramics, HAp, nanoclays), chondrogenic (e.g. bioactive glasses),
angiogenic (e.g. some bioceramics, bioactive glasses), adipogenic
In the last years, the development and characterization of novel (e.g. silica nanoparticles) or neurogenic responses (e.g. carbon nan-
bioinks have received increasing attention, as such systems are re- otubes) [72,73]. More detailed presentation and discussion of the
quired to fabricate complex and highly customizable cell-laden bio- biological effects of the chemical and morphological properties of
printed scaffolds for various applications in personalized medicine, different fillers are included further below in the respective sub-
tissue engineering, disease models or chips for high-throughput sections.
screening [41]. Hydrogel composite bioinks are one of the most In the reviewed literature, bioprinting of the designed cell-
suitable strategies for incorporating and combining various hydro- laden composite bioinks has mostly taken place by using
gel and filler properties, which are not attainable by one-phase hy- extrusion-based 3D printers. This technique has a large printing
drogels alone. Indeed, the lack of suitable (bioactive) materials for window ranging from roughly 101 –1013 Pa s [46]. Hence, a variety
bioprinting has been one of the significant drawbacks for a fast of bioinks can be successfully printed. Such bioinks need to exhibit
progress in this area [42,43]. Numerous new strategies are being shear-thinning and viscoelastic behavior and they must rapidly gel
continuously developed to enhance the material features for 3D and maintain their shape after printing. Considering shear-thinning
bioprinting demands, which have been reviewed in recent articles properties, bioinks are conventionally characterized by rheologi-
[30,35,42–52]. cal measurements (storage and loss moduli, tangent delta and re-
Facing a complex biological system in the human body, the re- sponse to shear stress, for example). During printing, the bioink
quirements for biomaterials which can be 3D printed incorporat- matrix protects the cells from damage as too high shear forces
ing cells to achieve complex geometrical designs are manifold and might disrupt the cell membrane (depending on the cell type). Not
challenging [20,21]. The demanding list of requirements of (com- only the bioink properties must be considered carefully, but also
posite) bioinks which are used for bioprinting may be divided into technical issues like e.g. the nozzle size and geometry [58,74]. Em-
pre- and post-printing properties which on the one hand examine bedding inorganic fillers into bioinks changes the rheological prop-
the materialś characteristics necessary for bioprinting and on the erties, but, according to the review of the available literature, the
other hand focus on features of the bioprinted 3D constructs that effect of such filler addition on the shear forces acting on cells and
are required for maintaining cell viability and for inducing (or pre- on the respective cell viability after printing have not been yet fully
venting) a given behavior of encapsulated cells after the printing examined and understood.
process [17–19,41,46,53,54]. After printing, cell containing constructs need to guide the
First of all, such biomaterials need to be biocompatible and cellular response to develop a matured functional tissue. So far,
sterilizable, not eliciting an unresolved inflammatory response. this task remains challenging as the time-dependent mechanical
Usually, bioinks consisting of low-viscosity natural or synthetic properties of the native ECM must be matched as close as pos-
biocompatible hydrogels are used for bioprinting. Hydrogels are sible by the (degrading) 3D printed construct [46]. In this con-
hydrophilic, physically or chemically crosslinked polymers. They text, it becomes a paramount requirement to carefully select the
possess a high-water content, which provides a suitable environ- hydrogel and the fillers to be used in composite bioink formula-
ment similar to the natural extracellular matrix (ECM) for cell tions. Biodegradability, bioactivity, internal porosity, interconnected
function and viability [55,56]. Nonetheless, pure hydrogels often porosity (for vascularization, ingrowth of surrounding tissue and
exhibit a lack of cell-binding sites (especially synthetic polymers), nutrients, waste and oxygen transport) are key characteristics of
relatively low mechanical and structural stability and they of- 3D bioprinted constructs which affect cell proliferation while the
ten degrade too fast or in an unpredictable manner, which nar- biodegrading construct leaves space for the expansion of the de-
rows their use as bioinks for long-term cell cultivation or for veloping ECM [22,23,75]. Gradient structures can furthermore help
the transfer to in vivo applications [57]. Hence, endeavors have to guide cell migration along the bioprinted scaffolds [24]. All such
been made to design composite bioinks which combine the pos- time-dependent features of 3D bioprinted scaffolds are affected by
itive features of biocompatible, low-viscosity hydrogels with the the chemical composition and physical properties of the chosen in-
biofunctionality of inorganic fillers or higher-viscous hydrogels or organic fillers.
4 S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22

In the following sections we review and discuss in detail the 3.2. Bioactive glasses – doped and undoped – as inorganic fillers
current state-of the art in the field of composite bioink devel-
opment, focusing on hydrogels incorporating bioreactive inorganic Hench et al. discovered that specific silicate compositions
fillers, and discussing key involved issues such as printability and demonstrated superior biocompatibility and the ability to bind to
cytocompatibility. The reviewed systems are grouped regarding the bone [108]. In contact with biological fluids, bioactive glasses show
inorganic fillers used in each case, as summarized in Table 1. The bioactivity through the formation of a carbonated hydroxyapatite
chemical and physical properties of the used inorganic fillers are layer on the glass surface because of the release of distinct ions.
described at the beginning of each respective subsection. The basic constituents of bioactive glasses are SiO2 , Na2 O, CaO and
P2 O5 but several modifications in composition and doping with
(biologically active) metallic ions can be made for specific applica-
3. Silicate-based composite hydrogels tions [109]. Most applications have considered bone regeneration
approaches, however applications in soft tissue repair are emerg-
3.1. Silica nanoparticles as inorganic fillers ing [110].
In hydrogels, ions released from bioactive glass particles may
Silica-based bioceramics have received great attention in bone induce efficient osteogenic differentiation of encapsulated cells.
tissue engineering, drug and gene delivery and theranostics, as Human adipose stem cells (hASCs) incorporated in gellan gum or
reviewed and summarized in previous studies [101–105]. Chemi- collagen type I hydrogels, respectively, revealed significantly higher
cal modifications can change the properties of (mesoporous) silica cell numbers and higher osteogenic gene expression when cultured
nanoparticles leading to bioinert, bioactive or biocompatible func- in bioactive glass extract based osteogenic medium (BG-OM) [111].
tions [106]. The ionic degradation products of Si-containing con- The strong mineralization and stiffening behavior under these cul-
structs have been reported to lead to high osteoconductivity, while ture conditions showed great potential for bone-like grafts. In ad-
possible cytotoxic side effects still need to be examined [105,106]. dition, gellan gum hydrogels crosslinked with BG extracts also
Silica nanoparticles have also found interest in the field of 3D bio- showed mineralization in pure osteogenic medium which could
printing. Wang et al. [76] investigated the effect of alginate/gelatin not be found in the neat collagen hydrogels under the same condi-
hydrogels supplemented with silica, biosilica and polyphosphate tions. Nevertheless, collagen-hydrogels loaded but not printed with
(polyP•Ca2+ ) on encapsulated SaOS-2 cells (human osteogenic sar- hASCs under BG-OM conditions showed significantly higher os-
coma cells), respectively. For this purpose, round gel cylinders teogenic marker expression than the gellan gum counterparts with
(13 mm diameter, 1.5 mm height) were printed with the composite an elongated and spread morphology as well as the highest miner-
hydrogels using a sterile 3D-Bioplotter (Envisiontec, Gladbeck; Ger- alization and hydroxyapatite formation capability [111]. Based on
many). The addition of polyP•Ca2+ -complex and biosilica, but also such results, bioactive glasses are considered attractive bioactive
to a smaller extent ortho-silicate, resulted in a significant stimu- fillers for developing composite hydrogels for bioprinting.
lation of cell growth after 3 days of incubation in medium. Espe- Often, bioactive glasses are used as pure scaffolds or in com-
cially the bioinik with polyP and biosilica allowed the cells to pro- bination with other materials to induce a hydroxyapatite coating
liferate. Supplementing the incubation medium/ serum with addi- during immersion in simulated body fluid (SBF). This layer pro-
tional bioactive glass nanoparticles (molar ratio of SiO2 :CaO:P2 O5 : motes cell attachment, whereas cells in direct contact with bioac-
55:40:5) with a size of 55 nm led to a significant enhancement of tive glass particles may behave differently [78]. Reakasame et al.
cell proliferation and biomineral formation during incubation for showed that the viability and growth of osteoblast-like MG-63
3 days in standard cell culture medium and further 5 days in os- osteosarcoma cells encapsulated into ADA-keratin-45S5 bioactive
teogenic differentiation medium. The growth potency of cells en- glass (0.5%) composite hydrogel microcapsules were drastically re-
capsulated in the composite bioink was not significantly affected duced during the first 7 days after encapsulation. However, dur-
by incubation either in normal medium or medium with bioactive ing the cultivation for 21 days the cell viability increased and be-
glass particles. came comparable to that of alginate-keratin, but remained never-
Silica nanoparticles (NPs) of size < 100 nm can also be benefi- theless permanently lower compared to the samples without BG
cial for enhancing the physical and mechanical properties of hydro- [78]. The outcomes were in agreement with the results reported
gels. Silica nanoparticles can effectively crosslink anionic polysac- by Rottensteiner et al. [79]. In the latter study the addition of 45S5
charides leading to improved printability and printing fidelity, as BG nanoparticles into ADA-gelatin revealed slight cytotoxicity to
shown in Fig. 2. Anionic polysaccharides, namely alginate (3% bone-marrow derived mesenchymal stem cells (MSCs) cultured on
(w/v)) and gellan gum (3% (w/v)) hydrogels, were printed with 6 the hydrogel films. Moreover, Reakasame et al. found that after 3
wt% cationic silica (SiNP) nanoparticles or aminopropyl-modified weeks of incubation under cell culture conditions the MG-63 cells
aminated cationic SiNPs (AmNPs) by Lee et al. [77]. Gellan gum started to spread, which did not arise in the samples without BG
is reported to have shear thinning as well as fast recovery behav- indicating a possible positive effect of the CaP formation induced
ior after printing [107]. After calcium crosslinking the pure poly- by the dissolution of the 45S5 BG particles [78].
mer hydrogel underwent shrinkage and hence the printing fidelity The effect of stimulating osteogenesis was illustrated by a
was lowered immensely. The addition of silica particles greatly en- comparison of 3D printed poly(ethylene glycol)dimethacrylate
hanced the shear viscosity (1062% increase). The storage modulus (PEGDMA) incorporating bioactive glass particles (45S5; 20 μm),
was also increased by SiNPs (and even more with AmNPs) due to hydroxyapatite (200 nm) or a co-print of both fillers, and bone
a higher crosslinking density and vice versa a decreased pore size. marrow-derived human mesenchymal stem cells, as reported by
Subsequently, the composite bioink with AmNPs was printable into Gao et al. [80]. The addition of bioactive fillers and cells resulted in
stable and defined grid structures [77]. For both bioinks (with and an overall decrease of mechanical properties which were recovered
without AmNPs) the cell viability of printed bovine chondrocytes over the course of 21 days of incubation in cell culture conditions
was recovered to around 90% after one week of cultivation in dif- due to the possible ECM formation by encapsulated hMSCs under
ferentiation medium containing TGFβ -3. Immunostaining and col- osteogenic differentiation. This effect was not observed in PEG-BG
lagen staining confirmed the growth and extracellular matrix se- (bioactive glass containing cylinders) bioinks, which demonstrated
cretion of the printed cells. Moreover, this approach could find ap- the lowest compressive modulus and highest mass swelling ratio
plications for many other polymeric hydrogels, specially consider- among the groups. Cell viability 24 h after printing was highest
ing anionic polysaccharides such as alginate or hyaluronic acid. in PEG-HAp samples. During the incubation period no significant
S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22 5

Table 1
Summary of hydrogel composites containing bioactive inorganic fillers for bioprinting and their potential applications.

Composite Bioink Cells Cell 3D Bioprinting Application Reference


Density/mL Technique (+ Figure)
Hydrogel Inorganic Bioactive Filler

5% (w/v) Alginate - 5% 50 μmoles/L silica/ 50 SaOS-2 (human Not reported Extrusion Bone [76]
(w/v) gelatin μmoles/L biosilica / 100 osteogenic
μmoles/L polyP∗ Ca2+ -> sarcoma cells)
silica

3% (w/v) Alginate - 3% 6 wt% Silica Bovine chondrocytes 1 × 108 Extrusion Tissue engineering and [77], Fig. 2
(w/v) gellan gum regenerative medicine
in general
2.5% (w/v) ADA (alginate 0.5% (w/v) Micro-bioactive MG-63 osteoblasts 1 × 106 Extrusion Bone-tendon interface [78]
dialdehyde) - 0.5% (w/v) glass (μBG, 45S5) regeneration
keratin
2.5% (w/v) ADA - 2.5% 0.1% (w/v) Nano-bioactive rMSCs (Rat bone 2 × 106 Extrusion Bone [79]
(w/v) gelatin glass (nBG, 45S5) marrow derived
mesenchymal stem
cells)
20% (w/v) PEGDMA 2% Bioactive glass/ 2% hMSCs (human 1 × 106 Modified Inket Cartilage/ osteochondral [80]
(poly(ethylene Hydroxyapatite/ 1% mesenchymal stem Printing interface
glycol)dimeth-acrylate BG + 1% HAp (all (w/v)) cells)
4% (w/w) Alginate - 5% 1% (w/w) Bioactive glass SaOS-2/ hBMSCs 5 × 105 Extrusion Bone [81]
(w/w) gelatin (+0.25% (human bone
(w/w) cellulose marrow stem cells)
nanofibers)
5% (w/v) ADA - 5% (w/v) 0.1/ 0.5 wt% Bioactive MG-63 osteoblasts Not reported Extrusion Bone [82], Fig. 3
GEL (ADA: alginate glass nanoparticles
dialdehyde) (BGNPs) or strontium
doped BGNPs
5–10% (w/v) Gelatin 0.01% (w/v) Silicate HUVECs + hMSCs 2 × 106 Extrusion Bone, Vascularization [83], Fig. 4
methacrylate nanoplatelets
(GelMA)-VEGF
10% (w/v) PEG:PEGDTT 4% (w/v) Nanosilicates, HUVECs Not reported Extrusion Drug delivery [84], Fig. 5
(poly(ethylene Laponite
R
XLG
glycol)-dithiothreitol)
4% (w/v) Alginate - 2% 4% (w/v) Montmorillonite BxPC3 (human 2 × 106 Extrusion Large scale bioprinted [85]
(w/v) clay pancreatic cancer scaffolds
carboxymethylcellulose cells)
3% (w/v) Alginate - 3% 3% (w/v) Laponite hTERTMSC (human 5 × 106 Extrusion Bone, drug delivery [86]
(w/v) methylcellulose mesenchymal stem
cell line expressing
hTERT (human
telomerase reverse
transcriptase))
20% (w/v) PEGDA 7% (w/v) Laponite
R
XLG ROBs (Primary rat 4 × 106 Extrusion Bone [87], Fig. 6
(polyethyleneglycol osteoblasts)
diacrylate)
2.0/ 2.5 wt% κ CA 3 - 6 wt% Nanosilicates MC3T3-E1 (Mouse Not reported Extrusion Tissue engineering (TE) [88]
(Laponite R
XLG) preosteoblasts) in general
10% (w/v) GelMA - 1% 2% (w/v) Nanosilicates MC3T3-E1 (Murine 1 × 106 Extrusion Large human tissues, [89], Fig. 7
(w/v) κ CA (laponite) 3T3 preosteoblasts, Bioactive 3D structures
(Kappa-carrageenan) Subclone 4)
7 wt% GelMA 0/ 0.01/ 0.05/ 0.5 wt% hMSCs 3 × 106 Micropatterning Bone, Cranial defects [90]
Nanosilicates
2% (w/v) Gellan Gum 1% (w/v) Laponite C2C12 (mouse 1 × 106 Extrusion in Bone, critical bone [91]
myoblasts) agarose fluid defect filling, skeletal
bed tissue engineering
5/ 7.5/ 10 wt% GelMA 0.5/ 0.75/ 1/ 2 wt% hBMSCs 5 × 106 Extrusion Hard and soft tissue [92]
Laponite R
XLG reparation, drug
delivery,
vascularization
2% (w/v) Chitosan 2% (w/v) MC3T3-E1 3 × 107 Extrusion Bone [93], Fig. 8, 9
compared to 3% (w/v) Nano-Hydroxyapatite
alginate (nHAp)
GelMA - 1 wt% 5% (w/v) hASCs (human 1 × 106 Microextrusion Bone [94], Fig. 10
methacrylated Nano-Hydroxyapatite adipose-derived and
hyaluronic acid stem cells) encapsulation
"by hand"
2 wt% Sodium alginate - 8 1 wt% hASCs 3 × 106 Dispense plotting Bone [95]
wt% gelatin (A2G8) Nano-Hydroxyapatite
2% (w/v) Alginate - 10% 0/ 4/ 8% (w/v) hMSCs 5 × 106 Extrusion Drug delivery, [18], Fig. 11
(w/v) gelatin Hydroxyapatite Microsphere
deposition, Soft tissue
engineering
(continued on next page)
6 S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22

Table 1 (continued)

Composite Bioink Cells Cell 3D Bioprinting Application Reference


Density/mL Technique (+ Figure)
Hydrogel Inorganic Bioactive Filler

2% (w/v) Alginate −2% 0.5% (w/v) hMSCs 2 × 107 Fused deposition Bone [60]
(w/v) gelatin Nano-Hydroxyapatite modeling,
Injection
Gelatin - alginate (8:2 0/ 9.1/ 30 wt% Mouse chondrocytes 2 × 105 Extrusion Cartilage [96]
(w/w) Hydroxyapatite
2/2.5% (w/v) Alginate - 1% 2.5% (w/v) Hydroxyapatite MC3T3-E1 (Mouse 2.5 × 105 Extrusion Bone [97]
(w/v) polyvinyl alcohol calvaria 3T3-E1)
(PVA)
1% (w/v) RGD∗ -alginate 50% (v/v) nHAp-pDNA∗∗ Bone 1 × 106 Extrusion Bone, muscoskeletal [98]
marrow-derived tissue engineering
MSCs
5 wt% Collagen compared 20 wt% β -TCP (tricalcium MC3T3-E1 compared 1 × 107 Extrusion Bone [99]
to 5 wt% alginate phosphate) to hASCs
5 wt% GelMA 0.05/ 0.15/ 0.3/ 0.5% (w/v) hMSCs 5 × 106 Extrusion Bone [100], Fig. 12
SrCO3

RGD – Arginine-glycine aspartic acid modified alginate.
∗∗
pDNA – plasmid DNA.

Fig. 2. A) Schematic presentation of 3D printing with AmNP-based nanocomposite inks. B) Bioprinted bovine chondrocytes in Alg/gellan or Alg/gellan/AmNP bioink were
cultured over 21d and their viability was assessed according to culture time. ∗ P-value < 0.05; ∗ ∗ P-value < 0.01; ns: not significantly different. C) Side view of 3D printing 10
layers using Alg/gellan inks with/without AmNPs. Red arrows indicate just extruded composite hydrogel that was not properly deposited on the underlying printed layer due
to collapse. Top view of the grids (2 × 2 cm) before and after crosslinking, showing shrinkage as the red boxes have the same size. D) Images and 3D scans of 3D printed
ear models after crosslinking, scale bar = 5 mm. E) Digitalized images of crosslinked ear models. The color indicates dimension difference between the printed, subsequently
crosslinked construct and the original 3D digital model. The gray background is the outline of the used 3D design. (Reproduced with permission from ref. [77]. Copyright
2018, American Chemical Society). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

cell proliferation could be observed in all samples but gene ex- bioactive glasses (47.12 SiO2 - 6.73 B2 O3 - 6.77 CaO-22.66 Na2 O-1.72
pression, measured by quantitative PCR, revealed that HAp in bio- P2 O5 - 5 MgO-10 SrO (mol%)) to tailor the rheological properties of
printed PEG-scaffolds was more effective with respect to stimula- the composite hydrogel as well as the response of two different
tion of hMSCs osteogenic differentiation and ECM production com- bone cells (SaOS-2 and hBMSCs (human bone marrow stem cells))
pared to 3D printed PEG-BG constructs. embedded in these composite hydrogels [81]. The 4-layered rect-
Furthermore, different cell types seem to have different sen- angular grids were stable in cell culture medium over the incuba-
sitivities to tolerate 3D bioprinting including the induced shear- tion of 21 d but showed different cell number, viability and pro-
stress and the materialś viscosity for short- and long-term cul- liferation. It was found that the incorporation of BG resulted in
ture [112]. Ojansivu et al. prepared alginate-gelatin based bioinks death of SaOS-2 cells immediately after printing, which is assumed
incorporating wood-based oxidized cellulose nanofibers (CNF) and to be the consequence of high shear forces induced in the bioink
S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22 7

Fig. 3. Bright-field microscopy images and life/ dead fluorescent staining (green: live cells, magenta: dead cells) of MG-63 cells after incubation for 24 h in various ADA-
GEL-BG composite hydrogels [82]. The cell viability of respective samples normalized to ADA-GEL without BGNPS (control) is shown on the left. Results are presented
as arithmetic mean ± standard deviation and the pairwise comparison of the means was performed with the Bonferroni’s test (post hoc comparison). (Reproduced with
permission from ref. [82]. Copyright 2016, IOP Publishing Ltd.). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version
of this article.)

through the addition of rigid bioactive fillers. Mixing of the hy- where the particles can act as local carrier for growth factors or
drogel with small amounts of CNF enhanced the printability re- therapeutic drugs, thus exploiting the dual effect of biologically ac-
markably whereas BG may enhance osteogenic capacity of encap- tive ions and drug release in the proximity of cells [82].
sulated cells, as shown via alkaline phosphatase (ALP) activity in-
crease. Unlike the behavior of SaOS-2 cells, hBMSCs incorporated 3.3. Anisotropic nanosilicates as inorganic fillers
in hydrogel containing BG and CNF maintained good viability but
also limited proliferation over 14 d [81]. Nanosilicates (NS), including for example laponite (trademark of
Leite et al. [82] focused on the integration of MG-63 cells in the company BYK Additives Ltd.), halloysite, montmorillonite, or
alginate-dialdehyde/ gelatin (ADA-GEL) hydrogel combined with hectorit, are ultrathin anisotropic nanomaterials with a high as-
bioactive glass nanoparticles. In previous studies, ADA-GEL already pect ratio (disk-shaped geometry of 20 - 50 nm in diameter and
had been proven to be cytocompatible in combination with cell 1 - 2 nm thickness) and a high degree of functionality for appli-
seeding, encapsulation or as 3D printed constructs [69,113–117]. cations in biomedicine [70,121–124]. Due to their shape and ori-
As ADA-GEL itself lacks bioactive behavior (mineralization abil- entation under shear, these nanosilicates can act as inorganic ad-
ity), bioactive glass nanoparticles (BGNPs) were integrated to pro- ditives to improve the mechanical and biological properties of 3D
mote the deposition of a calcium phosphate layer, which should (bioprinted) scaffolds, or they may be used as therapeutic drug de-
lead to enhanced osseointegration of the construct [118]. ADA-GEL livery agents [125,126]. Beyond these, laponite, having the struc-
was compounded with ternary formulations of 0.1% and 0.5% (w/v) tural formula Na+ 0.7 [(Mg5.5 Li0.3 )Si8 O20 − (OH)4 ]− 0.7 , is extensively
BGNPs (SiO2 –CaO-P2 O5 ) or strontium doped BGNPs (SiO2 –CaO- researched for the development of composite hydrogels for 3D bio-
P2 O5 -SrO), respectively [82]. Strontium was used due to its favor- printing [125]. Especially Laponite R
XLG is studied for biomedi-
able effects on osteogenic stimulation as well as in vivo bone for- cal applications due to its lower heavy metal content. As the hy-
mation ability [119,120]. The addition of BGNPs led to a decreas- drous silicate contains elements such as magnesium, zinc, lithium
ing crosslinking time and thus a smaller processing window com- or iron that are also found in natural bone, its nontoxic degrada-
pared to ADA-GEL, possibly due to the release of divalent cations tion products [Na+ , Mg2+ , Si(OH)4 , Li+ ] can easily be absorbed by
and their interaction with the guluronic parts of ADA. The in vitro the body [127,128]. During 3D printing, the anisotropic clay ma-
biomineralization study of bioplotted composite grid structures terials align and thus significantly affect the microstructure and
(15 mm edge length, 4 layers) showed bioactive behavior lead- the rheological properties of the bioink [86,129]. Hence, nanosil-
ing to the formation of an apatite layer on the surface during in- icates are readily used due to their shear thinning characteris-
cubation in SBF. Moreover, these composite bioinks revealed the tics which protect embedded cells in the hydrogel from too high
possibility to sustainably release pharmacological drugs and they detrimental stresses during the printing process [123,124,130]. This
displayed similar cell viability of encapsulated MG-63 osteoblasts anisotropic interaction between the polymer and nanosilicates is
compared to neat ADA-GEL after 24 h of incubation (Fig. 3). The based on the presence of both positive and negative charges on the
special characteristic of BG fillers, being dissolvable or bioreactive, NS surfaces, which allow the formation of a physically crosslinked
remains to be exploited in future bioink strategies, not only for network with anionic, cationic or natural polymers [131–133]. Fur-
bone regeneration (bioprinted) scaffolds. An interesting approach thermore, the optical transparency of nanosilicates in aqueous me-
is the incorporation of mesoporous BG particles in the hydrogel, dia can be beneficial for imaging the subsurface cellular behavior
8 S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22

to design complex printed tissues [131]. Beside these advanta- [85] blended a novel hybrid ink by mixing alginate, carboxymethyl-
geous properties, toxicological studies have shown that high con- cellulose and montmorillonite clay. The printability of this com-
centrations of nanosilicates can reduce cell proliferation in vitro position was tested with various complex 3D geometries and dif-
[134]. Consequently, only low concentrations of nanosilicates (0.1 fering patterns. The bioink showed reproducible printability and
– 7%(w/v)) are generally used when applied as inorganic fillers in shape fidelity. Moreover, 84% of printed human pancreatic can-
hydrogel bioinks. cer cells were alive after 7 days of culture. In comparison, Ahlfeld
Xavier et al. [135] examined a gelatin-based hydrogel contain- et al. [86] used laponite instead of montmorillonite clay for
ing bioactive nanosilicate for enhanced matrix mineralization for extrusion-based printing at physiological conditions. Printed scaf-
the regeneration of bone defects. In vitro encapsulation of NIH folds demonstrated excellent shape fidelity even in scaffolds with a
MC3T3 preosteoblasts for 4 days showed cytocompatible proper- height of 2 cm. More complex structures were plotted by changing
ties which supported cell adhesion, growth and osteogenic dif- the layer orientation between 20° and 90°. Scanning electron mi-
ferentiation without the usage of growth factors. The composite croscopy images showed dense, smooth surfaces which remained
hydrogel could be bioprinted to precisely designed scaffolds with predominantly unaltered after 21 days of culture. Mechanical test-
a stable porous structure [135]. This result is in good agreement ing revealed that non-porous, solid scaffolds had a higher compres-
with previous studies which had found that silicate nanoplatelets sive strain and Young’s modulus than porous ones. However, after
significantly led to osteogenic differentiation of encapsulated hM- 21 days the mechanical properties of both scaffold types drasti-
SCs in GelMA hydrogels in growth-factor free medium [90,121]. cally decreased. Cell-laden constructs preserved their morphologi-
A different approach to mimic the complex architecture of tis- cal properties and showed a cell viability of 70% to 75% at the end
sues with multiscale vascular networks was carried out by Byam- of the culture period. Furthermore, laponite could enhance the BSA
baa et al. [83]. They developed a functional vascular structure (bovine serum albumin) and VEGF release kinetics of the hydrogel
within 3D printed bone tissue-like fibers consisting of GelMA hy- blend and by designing an appropriate inner porosity of the bio-
drogels with different amounts of vascular endothelial growth fac- printed scaffolds the release was sustained.
tor (VEGF) and 100 μg/mL silicate nanoplatelets. A pyramidal ar- Hong et al. [140] printed a PEG-alginate-nanoclay (Laponite R

rangement of printed cylinders was designed to represent differ- XLG) hydrogel and infiltrated the porous structure with HEK (hu-
ent cellular and chemical compositions to fabricate a gradient ma- man embryonic kidney) cells encapsulated in a collagen matrix. Af-
terial for vascularization. The central core cylinder having a fast ter 7 days the cell viability still remained very high (≈ 95%) within
degradation rate and lower stiffness created the internal lumen this environment. Furthermore, the printed mesh was highly de-
and consisted of 5% (w/v) GelMALOW . A co-culture of endothelial formable. After stretching to 300% of its original length, relaxation
cells (GFP-HUVECs) and human mesenchymal stem cells was car- led to almost complete recovery of the initial shape. The unique
ried out. The other cylindrical fibers were printed from a stiffer nature and the ability of nanoclays to improve the structure,
hydrogel made of 10% (w/v) GelMAHIGH , osteoinductive silicate properties and performance of polymeric matrices have encour-
nanoplatelets and embedded hMSCs. During 7 days of static cul- aged research groups to further develop this material for 3D bio-
ture the inner core degraded and HUVECs and hMSCs formed a printing with encapsulated cells. A polyethylene glycol diacrylate
densely packed network along the lumen-like hollow structure dis- (PEGDA)/Laponite R
XLG nanoclay bioink was recently co-printed
played in Fig. 4. Further maturation and osteogenic differentiation by Zhai et al. [87] in combination with hyaluronic acid sodium salt
were achieved by medium perfusion for 5 more days and another by extrusion printing. The rectangular scaffolds showed suitable
9 days of static culture. Real-time polymerase chain reaction dis- print fidelity as a result of the increased nanocompositeś strength
played an enhanced expression of angiogenesis-related gene CD31, and flow characteristics [87]. Within bioprinted structures, nan-
so that an organized, mature vascular niche was formed within oclays have additionally been shown to directly influence cell be-
the printed model. The approach of Byambaa et al. [83] resulted havior, which is depicted in Fig. 6. Primary rat osteoblasts (ROBs)
in a material that supported vasculogenesis and coincidently os- encapsulated in a PEG-clay composite hydrogel revealed a high cell
teogenesis. As can be seen in Fig. 5, Peak et al. [84] combined viability of more than 95% up to 7 days of culture [87]. Using a
2D silicate nanoplatelets with a poly(ethylene glycol)-dithiothreitol two-channel printing setup, efficient delivery of nutrients and oxy-
(PEGDTT) bioink with the goal to receive a shear-thinning bioink gen to the growing cells, uniform cell distribution and deposition
that has the capability to load and sustainably release therapeutic efficiency were achieved. Both in vitro and in vivo, the printed rat
proteins from a 3D printed structure. The inclusion of PEGDA into osteoblasts differentiated, which was considered to be enhanced
the hydrogels slowed down the degradation of the printed scaf- by the release of bioactive ions from the laponite clay, especially
folds. In agreement with previous studies, the interactions between Mg2+ and Si4+ ions [87].
laponite and the polymer network prevented swelling of the con- Wilson et al. [88] designed new self-supporting composite
struct [136]. Within this 3D composite material, the nanosilicate hydrogels from natural κ CA (Kappa-carrageenan; 2.5 wt%) stiff-
particles acted as a therapeutic delivery platform. Due to its nega- ened with 3 - 6 wt% nanosilicate addition (Laponite R
XLS) and
tive and positive charges laponite can interact with a huge range of crosslinked with potassium ions. The resulting mechanical prop-
biomolecules [137], like e.g. vascular endothelial growth factor, fi- erties of the stable constructs were in the same order of mag-
broblast growth factor (FGF), bone morphogenic protein 2 (BMP-2) nitude as cartilage tissue or blood vessels which undergo con-
or platelet-derived growth factor [126,138,139]. To analyze the bi- stant mechanical load. Complex structures were printed, like 30-
ological activity of the released VEGF and FGF from biocompatible layered cylinders, demonstrating a high shape stability of the
3D printed composites, the migration of HUVECs across a transwell shear-thinning bioink at physiological temperatures, fine lattice
membrane was measured. The PEGDTT/nSi scaffold with encapsu- networks with high resolution or anatomical-scale models of noses
lated growth factors showed similar behavior compared to the pos- and ears. Bioprinting using mouse preosteoblasts MC3T3-E1 led to
itive control of exogenously delivered growth factor. an even distribution of cells in the printed strands as shown by
As mentioned above, biofabrication approaches involving 3D CellTracker red or green, respectively, using fluorescent microscopy.
bioprinting are considered to develop tissue-like functional con- As κ CA does not have cell binding moieties, printed cells showed
structs by layer-by-layer deposition of hydrogels with aligned limited spreading while maintaining their metabolic activity dur-
and homogenously distributed living cells. The incorporation of ing 7 days incubation in cell culture medium in humidified atmo-
anisotropic fillers is being put forward for supporting not only the sphere. Thus, the κ CA-nanosilicate bioink could either be used for
printing process but also to promote cell alignment. Habib et al. 3D printing non-adherent cells like chondrocytes or it could be
S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22 9

Fig. 4. Bioprinting of bone mimetic 3D architecture containing osteogenic and vasculogenic niches as well as formation of HUVECs/hMSCs-lined perfusable hollow lu-
men structure. The construct, containing both GFP-HUVECs and hMSCs, was co-cultured for 7 days, followed by medium perfusion for 5 days. A) Complex bone tissue
architecture consists of a perfusable vascular lumen lined with HUVECs which can be fabricated within a pyramidal bioprinted construct by arranging individual rods of
VEGF-functionalized GelMA bioinks with different mechanical strengths. Silicate nanoparticles were embedded in the hMSCs-laden three outer layers of cylinders to induce
osteogenic differentiation of hMSCs into bone tissue. The VEGF was covalently conjugated into the three outer layers of the cylindrical hydrogels with concentrations of
17.1, 34.2, and 68.5 ng mL−1 as determined with ELISA. B) Schematic 3D printing of independent cell-laden cylinders using the NovoGen MMX Bioprinter (Organovo). C)
Cross-sectional view of the whole bioprinted construct with Live/Dead staining shown from cross- and top-view of the encapsulated cells. D) DAPI and α -SMA stainings of
a HUVEC-lined vessel-like lumen within the bioprinted construct with cross- and top-sectioned confocal micrographs of central vessel at day 12 post culture. Encapsulated
endothelial cells lined the vascular walls (green fluorescence) and hMSCs were differentiated into pericytes (red fluorescence). E) Formation and lining of endothelial cells
inside the central channel. F) Immunostaining of endothelial cells and α -SMA-expressing hMSCs in the inner part of the lumen. G) Vascular hollow lumen network before
and after perfusion with fluorescent microbeads at day 7 post culture. (Reproduced with permission from ref. [83]. Copyright 2017, WILEY-VCH). (For interpretation of the
references to colour in this figure legend, the reader is referred to the web version of this article.)

modified with RGD-containing sequences to enhance cell migra- improved mechanical properties, flexibility, elasticity, toughness,
tion, spreading and proliferation [88]. bioactivity and flow properties, which are the result of the syner-
Taking this issue into account, a follow up work was done by gistic combination of nanocomposite and ionic covalent entangle-
Chimene et al. [89], who developed a novel and unique nanoengi- ments. Moreover, a Herschel-Bulkley fluid behavior model showed
neered ionic-covalent entanglement (NICE) bioink with outstand- the effect of the nanoclay altering the flow behavior of the NICE
ing printability and mechanical properties by taking advantage of bioink. A relatively low influence of stresses on cells during ex-
a novel reinforcement technique, which is schematically shown trusion bioprinting was demonstrated. Subsequently, encapsulated
in Fig. 7. The NICE bioink is composed of 10% w/v GelMA, 1% 3D printed mouse preosteoblasts proliferated well and maintained
w/v κ CA, 2% w/v nanosilicates and 0.25% w/v Irgacure2959 pho- high cell viability (≈90%) over a long-term incubation of 120 days.
toinitiator and was demonstrated to be printable to freestanding, NICE bioinks are thus an attractive composite system for replica-
high aspect ratio scaffolds of 3 cm in diameter and 150 layers in tion of human tissues and for developing bioactive 3D structures,
height with an excellent shape fidelity without the need of addi- where the presence of the inorganic filler (nanosilicate) confers the
tional support materials. After UV-crosslinking, the constructs had bioink outstanding properties.
10 S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22

Fig. 5. Printing and release of therapeutics in 3D. a) The high surface area and charged characteristics of nanosilicates are able to sequester protein therapeutics within the
3D printed structure. Therapeutics will be released during degradation of printed networks. b) The release of fluorescently labeled protein from 3D printed structure was
monitored over 28 d in PEGDTT (10% wt vol−1 )/nSi (4% wt vol−1 ) hydrogels (no PEGDA in formulation). Printed structure dimensions are ≈ 10 mm in diameter and 15 mm in
height (significantly larger mass than used in degradation studies). c) Sustained release of therapeutics was observed during initial time period, while complete protein was
released within 28 d in PEGDTT (10% wt vol−1 )/nSi (4% wt vol−1 ) hydrogels (n = 3, ∗ p < 0.05, one-way ANOVA with Tukey post-hoc testing). (Reproduced with permission
from ref. [84]. Copyright 2019, WILEY-VCH). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Another method to create (ECM)-mimicking GelMA hydrogels laponite-containing hydrogels. Moreover, they analyzed similar is-
reinforced with nanosilicates is the micropatterning technique sues with laponite incorporated GelMA bioinks and printed them
[141–144]. For preparing microchannel units (150 μm height) en- with hBMSCs [92]. The incorporation of laponite nanoclay influ-
capsulating hMSCs, the composite material was exposed to UV enced the visible-light mediated crosslinking of the 3D fabricated
light while using a photomask with arrays of linear microchannels. constructs as demonstrated by a faster degradation compared to
The resultant pattern showed high interconnected porosity for all pure GelMA. After parameter optimization, a shear-thinning, bio-
compositions, a reduced degradation rate and an increasing com- compatible bioink was developed with increased shape fidelity
pressive modulus with increasing nanosilicate concentrations (from post-crosslinking and exhibiting cell viability over a 21 d period
0% to 0.5%). In vitro cytocompatibility studies with the GelMA-nSi of incubation under cell culture conditions. Adsorption and release
composite material showed uniform spreading except for the group of BSA and lysozyme in GelMA-laponite hydrogels revealed charge
with 0.5% nSi. At concentrations of 0.01% and 0.05% nSi the sam- dependent kinetics of the growth factors in digestive environment
ples did not show a difference of in vitro cytocompatibility com- due to interaction with the dipolar charging of laponite. Finally,
pared to the control. The same concentrations could demonstrate a these bioinks showed promising results for stimulating angiogene-
positive effect on osteogenic differentiation of hMSCs in a 3D ma- sis, vascular network penetration and osteogenic differentiation in
trix. Subcutaneous implantation of the bioprinted scaffold in im- the CAM assay [92].
munocompetent rats revealed the compositeś biocompatibility and The available studies in the literature clearly indicate the high
low localized immune responses indicating application potential in potential of bioinks incorporating nanosilicate fillers for bioprint-
the clinic, for e.g. to support bone repair [90]. ing and applications are expected to expand in the future. Explor-
Novel techniques to print complex functionalized 3D scaffolds ing different types of nanosilicates, also loaded with biomolecules,
in a one-step procedure seem to be attractive for applications in incorporated in printable hydrogels appears to be a very promis-
skeletal tissue engineering. Cidonio et al. [91] combined Laponite
R
ing avenue for future research. Nevertheless, long-term studies like
XLG with gellan gum (GG) to receive highly stable constructs with degradation or swelling kinetics as well as shape fidelity charac-
tunable swelling capacity printed in an agarose fluid bed as tem- terizations or ion/ drug release were often done with casted disks
porary support before crosslinking. In vitro, printed C2C12 promy- made from composite hydrogels so far. As 3D printed scaffolds
oblasts with sustained viability of around 80% yielded functional mostly have a more complex structure, their architecture, topol-
capacity, as demonstrated by ALP expression, and proliferation over ogy and interconnected porosity may play an important role when
an incubation period of 21 d. For the first time, the fluid bed determining these properties and when designing more complex,
was used as functional loading platform for the localization of patient-specific tissue implants.
drugs during 3D printing. Compared to pure GG, printed laponite-
GG showed promising localization and sustained release of lyso- 4. Calcium phosphate-based nanocomposite hydrogels
some and bovine serum albumin, which were used as analogues
for BMP-2 and VEGF, respectively. With a chick chorioallantoic Calcium phosphates, especially HAp, have been intensively
membrane (CAM) assay the potential of angiogenesis and integra- investigated and used as inorganic fillers in bioprinting for
tion of VEGF-containing laponite-GG printed scaffolds was evalu- bone regeneration, since around 60 wt% of bone is made of
ated. Significant vascular infiltration was stimulated and guided by HAp (Ca10 (PO4 )6 (OH)2 ) [145]. Such composite hydrogels have
S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22 11

Fig. 6. A) 3D-bioprinted scaffold prepared by the two-channel method with gentian violet stained PEG-Clay and Rhodamine stained HAp. B) Live/dead assay of primary rat
osteoblasts (ROBs) 1 d after printing within PEG4K-Clay scaffolds, I) viable ROBs (green, calcein AM), II) dead ROBs (red, EthD-1), and III) merged. C) Fluorescently stained
ROBs after 3D-biofabrication (PEG4K-Clay-P) and traditional seeding (PEG4K-Clay-S) on PEG4K-Clay scaffolds after culturing for 0, 1, 3, 5, and 7 d. D) SEM image of 3D-
bioprinted ROBs in PEG4K–Clay scaffolds after 7 d of culture. The red rectangle shows ROBs that were marked with yellow color. E, I) Live/dead assay and II) cytoskeleton
staining of 3D-biofabricated ROBs within PEG4K-Clay scaffolds after 7 d of culture. F) Cell counting Kit-8 (CCK-8) analysis of ROBs after 3D-biofabrication (PEG-Clay-P) or
traditionally seeding on PEG-Clay scaffolds (PEG-Clay-S) on days 1, 3, 5, and 7 d of culture. G) ALP activity of biofabricated ROB containing PEG-Clay scaffolds after culturing
for 4, 7, 14, and 21 d. The insert images show respective ALP stainings of ROBs on I) PEG4K-Clay-P and II) PEG10K-Clay-P. Asterisks (∗ ) denote significant differences (∗ p <
0.05, ∗ ∗ p < 0.01). (Reproduced from ref. [87] according to the terms of the Creative Commons Attribution CC BY license). (For interpretation of the references to colour in
this figure legend, the reader is referred to the web version of this article.)

demonstrated very good integration with cells and with the sur- bone morphogenetic protein. HAp can also enhance ALP activity in
rounding tissue, exhibiting bioactivity, osteoconductive properties mesenchymal stem cells, which is important as ALP activity plays
and osteogenic ability [146,147]. The strong chemical and struc- a central role in the early mineralization process associated with
tural similarity to the composition of the inorganic phase of bone bone formation [57,149]. Nevertheless, the relatively slow degra-
allows HAp to form a strong bond with surrounding bone [148]. dation and low mechanical strength limit the use of HAp as sin-
Moreover, HAp can be used as drug delivery system or to stimu- gle component material in bone tissue engineering, especially for
late the endogenous expression of osteogenic growth factors like load-bearing applications. Consequently, HAp particles are used as
12 S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22

Fig. 7. NICE bioinks combine nanocomposite reinforcement and ionic-covalent entanglement reinforcement mechanisms to create a tough, elastic, and highly printable bioink.
(a) NICE bioinks made from GelMA, κ CA and nanosilicates create a dually reinforced hydrogel network that behaves as a solid at low shear stresses and improves shear
thinning characteristics during bioprinting. Post-crosslinking, ICE and nanosilicate reinforcement synergistically improve the bioink’s mechanical strength. In TEM images
two-dimensional nanosilicate particles show a uniform morphology. (b) Encapsulated cells align parallel to the 3D printed NICE scaffolds after 30 days in culture. (c) 3D
printing of NICE bioink enables freestanding structures which are mechanically (film) and physiological (bifurcated vessel) stable and have high structural fidelity (3D printed
ear). (d) Crosslinked stiff and elastomeric structures can support more than 50-times their own weight (scale bar = 1 mm). (Modified from [89]. Copyright 2018, American
Chemical Society).

reinforcing agent in tough and flexible polymer matrices [150]. Me- bone tissue engineering. Wenz et al. [94] added 5 wt% HAp with a
chanical properties are of interest since it has been shown that size of 50 nm - 1 μm to a GelMA/ methacrylated hyaluronic acid
cells differentiate according to matrix stiffness [23,151]. and used their composite hydrogel to encapsulate primary human
Nanoparticle-reinforced bioactive hydrogels often show adipose-derived stem cells (hADSCs). Simple grid structures with
beneficial mechanical and biological properties compared to an edge length of 10 mm were printed with hydrogels contain-
microparticle-reinforced ones. To study this effect, Demirtaş ing HAp particles (referred to as HAp-inks) and without particles
et al. [93] prepared four different bioinks – alginate, alginate- (GelMA-inks) and they were incubated under cell culture condi-
hydroxyapatite (20 mg/ml), chitosan and chitosan-hydroxyapatite, tions for 28 days. HAp-inks revealed a significant upregulation of
as shown in Fig. 8. 3% (w/v) alginate-based hydrogels were mixed osteogenic markers and an extensive matrix production which ad-
with 1% (w/v) CaSO4 to start the internal gelation for enhanced ditionally led to an increase of elastic and viscous composite hy-
printability. With the addition of nano-HAp the viscosity of both drogel properties. These effects were further enhanced by cultur-
bioink types increased. Mechanical properties were shown to be ing the different inks in osteogenic differentiation medium, as dis-
higher for chitosan-HAp compared to alginate-HAp. The addition played in Fig. 10. Furthermore, it was found that the way of mixing
of HAp resulted in a 3- to 6-fold increased Young’s modulus for (stirring, sonication and iterated stirring) varied the size distribu-
both hydrogels, respectively. HAp particles were also found to be tion of the HAp particles strongly, ranging from a broad bimodal
homogeneously distributed within the polymer matrix and the distribution to distinct peaks which correlated to particle sizes of
bioprinted scaffolds showed overall pore structures (Fig. 9). Cell 40 nm and 1 μm.
studies with the four groups resulted in a general high viability Wang et al. [95] reported similar outcomes for sodium algi-
(average 89–93% for all samples on day 3, 90–95% after 9 days nate/gelatin/ nHAp composite hydrogels. In their study on the
incubation) and MC3T3-E1 pre-osteoblasts proliferated during osteogenic effects of nano-hydroxyapatite in 3D constructs, they
the 21 days of cultivation. It could be clearly seen that bone-like found that nano-hydroxyapatite could promote human adipose de-
nanostructured HAp increased the cell proliferation over time and rived stem cells (hASCs) to osteogenic differentiation in the 3D-
chitosan-bioinks were advantageous over alginate inks. Printed printed environment, which is consistent with current and pre-
chitosan-HAp hydrogels had peak expression levels for early and vious studies [146,152,153]. With a 3D Bioplotter reticular scaf-
late stage osteogenic markers and pre-osteoblasts were shown to folds were fabricated, intended for the repair of bone tissue de-
mineralize and differentiate osteogenically after 21 days of culture. fects. hASCs were encapsulated prior to printing as they still have
Adding HAp particles to polymer-based hydrogels enlarges the proliferation and osteogenic differentiation potential after bioprint-
range of applications, leading specifically to new approaches for ing and subsequently the 3D constructs were incubated in two
S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22 13

Fig. 8. Schematic diagrams of alginate and chitosan hydrogels as bioinks for 3D bioprinting for bone tissue engineering applications. (a) Mixing of alginate-nHAp solution
with CaSO4 solution results in a printable form of alginate solution. (b) MC3T3-E1 preosteoblast cells and nHAp are incorporated in a printable alginate-nHAp solution and
placed in 3D printer syringe. Disk-shaped scaffolds (6 mm diameter × 1 mm thickness) are printed and subsequently placed in a CaCl2 solution for fast ionic gelation. (c)
Chitosan-nHAp solution (pH: 4.0) is mixed with GP salt to generate a printable form of chitosan solution (pH: 6.95–7.0). (d) MC3T3-E1 pre-osteoblast cells and nHAp are
mixed with a printable form of chitosan-nHAp solution and placed in a syringe. A printed disk-shaped hydrogel is placed to 37 °C and 5% CO2 for thermal ionic gelation.
(Reproduced with permission from ref. [93]. Copyright 2017, IOP Publishing Ltd).

Fig. 9. SEM micrographs of hydrogels without cells and with MC3T3-E1 cells at different times of culture; alginate hydrogels (a) without cells, (e) day 7, (i) day 14 × 10 0 0
and × 2500, respectively; alginate-HAp hydrogels (b) without cells, (f) day 7, (j) day 14 × 10 0 0 and × 2500, respectively; chitosan hydrogels (c) without cells, (g) day 7, (k)
day 14 × 10 0 0 and × 2500, respectively; chitosan-HAp hydrogels (d) without cells, (h) day 7, (l) day 14 × 10 0 0 and × 2500, respectively. (Scale bars indicate 20 μm, inset
figures scale bars indicate 10 μm). (Reproduced with permission from ref. [93]. Copyright 2017, IOP Publishing Ltd).

different media. In comparison to proliferation medium, HAp par- ies in the back sub-cutaneous area of nude mice following Hall
ticles could promote the osteogenic differentiation of hASC cells et al. [154] demonstrated ectopic bone formation after 8 weeks
incorporated in 3D bioprinted scaffolds when cultivated in os- of implantation for scaffolds printed with cells. The printed
teogenic medium for 14 days. Cell viabilities after 1 and 7 days 3D constructs revealed regular pores for vascularization as well
of culture were around 88–90% and were not significantly dif- as interconnected pores for cell growth and proliferation. Nev-
ferent between hydrogels with and without HAp. In vivo stud- ertheless, limitations for alginate/gelatin/hydroxyapatite used in
14 S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22

Fig. 10. Evaluation of bone matrix formation in printed grid constructs according to ref. [94]. Hydrogel sections on days 1 and 14 or 28 of culture under osteogenic (osteo)
or control conditions (ctrl) were stained for the expression of the matrix proteins fibronectin and collagen I (day 28), as well as the bone-specific markers ALP and OPN
(day 14). (a) Scale: 200 μm. The antibody-labeled proteins are shown in red, the DAPI-stained nuclei are shown in blue. (b) Tile scans made of 4 × 5 single pictures. Scale:
20 0 0 μm. (Reproduced with permission from ref. [94]. Copyright 2017, IOP Publishing Ltd). (For interpretation of the references to colour in this figure legend, the reader is
referred to the web version of this article.)

3D bio-printing technology have been noted, which mainly were terial including encapsulated hMSCs for the reconstruction of large
related to the relatively low mechanical properties of the hydro- bone defects. One scaffold could be printed in less than 10 min
gels [155,156]. and possessed a highly porous structure with pore sizes of around
Inorganic fillers may also help to enhance the visibility and ac- 400 μm and strut diameters in the range of 320 μm. Incubation in
cessibility of printed and implanted hydrogel scaffolds using med- SBF showed nucleation and growth of apatite crystals on the sur-
ical detection methods. Wüst et al. [18] for example prepared 2% face after 12 days which confirmed the scaffoldś bioactivity. The
(w/v) alginate/ 10% (w/v) gelatin hydrogels with final HAp concen- apatite formation is considered essential to support the osteogenic
trations of 8, 4% (w/v) for applications in bone tissue engineer- differentiation of hMSCs [157,158] and hence an expedited osseoin-
ing, as shown in Fig. 11. Gelatin was added due to its thermo- tegration for in vivo applications may be expected from these scaf-
sensitive physical crosslinking behavior. When printing the com- folds [159].
posite ink with a temperature-controlled printhead at 40 °C onto a Related research using gelatin-alginate hydrogel with nHAp and
cooled surface, gelatin provided instantaneous stability, initial in- mouse chondrocytes for bioprinting was performed by Fan et al.
creased viscosity and mechanical stability which led to superior [96]. Relatively high amounts of nHAp (up to 30 wt%) resulted in
structural stability and a more accurate structure of the printed a cell viability of around 70% post-printing. Interestingly, it rose up
constructs. In this approach, the long-term stability of the printed to 99% after 10 days of incubation in standard cell culture condi-
scaffolds is provided by the slower ionic crosslinking of the al- tions. The mechanism behind such favorable cell behavior and the
ginate with CaCl2 . Against the expectations, the incorporation of interaction between filler and cells were not further explained. Me-
varying HAp concentrations did not result in large differences of chanical properties were found to be higher correlating with the
the Youngś modulus. An additional benefit was that hydrogels con- filler content and degradation could be adjusted with the addition
taining HAp showed superior radiopacity and thus visibility of the of bioactive fillers.
printed scaffolds in μCT images. With respect to bioprinting and Another study which combines hydrogels and synthetic poly-
the two-step post-processing treatment, the authors also investi- mers was conducted by Bendtsen et al. [97]. An alginate-polyvinyl
gated the cell compatibility of their new composite hydrogel with alcohol (PVA)-hydroxyapatite hydrogel for 3D bioprinting bone tis-
encapsulated human mesenchymal stem cells (hMSCs). They fabri- sue engineering constructs was designed with improved rheologi-
cated simple cubic structures using a modified open-source two- cal features. CaCl2 crosslinking led to shrinkage of around 24% due
syringe 3D printer. The comparison of the three compositions did to further crosslinking of the alginate chains. Degradation studies
not reveal any differences or reverse cytocompatibility effects with in cell culture medium resulted in a swelling of around 32% com-
an average cell viability of 84–85% over 3 days of in vitro cell cul- pared to nearly no swelling observed in the control (2.5% (w/v) al-
ture. ginate with gelation controlling agents 0.15% (w/v) Na2 HPO4 and
Using the same alginate/ gelatin/ nHAp composite hydrogel, 0.20% (w/v) CaSO4 but without PVA-HAp). Swelling was caused by
Hernandez et al. [60] designed a hybrid system of printed poly- an ion exchange between calcium ions crosslinked with the algi-
caprolactone (PCL) gyroid scaffolds with infiltrated composite ma- nate G-blocks in the egg-box structure and the sodium ions in the
S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22 15

Fig. 11. Schematic diagram of the two-step hydrogel gelation mechanism based on reversible thermal gelation of gelatin and irreversible chemical gelation of alginate.
(A) Biofabrication of heated hydrogel precursor including living cells onto a cold substrate. (B) Instantaneous gelation and primary strand stability take place due to the
temperature drop, which leads to solidification of the gelatin. (C) Post-printing immersion of the whole construct in a CaCl2 bath to crosslink the alginate present in the
hydrogel precursor. Chemical crosslinking is conducted in a cold environment to maintain construct stability until the procedure is completed. (D) Long-term stability is
ensured by the crosslinked alginate and the cooling plate can be removed. (Reproduced with permission from ref. [18]. Copyright 2014, Elsevier Ltd). (For interpretation of
the references to colour in this figure legend, the reader is referred to the web version of this article.)

surrounding medium [160–162]. All bioprinted scaffolds remained to their good printability and high cell viabilities in vitro (close to
intact for 14 days of incubation. The cell viability of MC3T3-E1 100%) with cell spreading and osteogenic differentiation of MSCs,
pre-osteoblasts was assessed directly after printing (~95%) and af- these novel bioinks based on gene-activated compounds might be
ter crosslinking inside the calcium bath (~77%) for the composite superior to classical therapeutic drug or protein delivery matri-
bioink and for the control (~60% and 22.5%, respectively). Reasons ces and they could induce the expression of biologically functional
for the low viability in the control could be attributed to lacking proteins. Hence, toxic doses or the dispersion of the drugs due to
porosity, fast degradation or to physical forces exerted on cells. fast degradation of the proteins could be prevented in future mus-
A different approach to the ones described above is presented culoskeletal applications [98].
in the research of Cunniffe et al. [98]. Nano-hydroxyapatite com- If the viscosity of the prepared bioink is lower compared to the
plexes were used as a vector for plasmid DNA (nHAp-pDNA) de- one in extrusion-based printing, Guillemot et al. [163] suggested
livery within an alginate hydrogel that was co-printed with poly- biological laser printing as a technique with high precision and a
caprolactone into a grid like structure. Bone marrow-derived MSCs good combination of different biological materials. In his studies
were encapsulated in the gene-activated bioink and the osteogenic solutions of biopolymers, nHAp or cells were sprayed on a target
capability was analyzed over the course of 28 days in vitro as with the advantage of enabling higher fractions of cells per droplet
well as during subcutaneous implantation in nude mice for 4 and compared to ink-jet printing.
12 weeks post-printing, respectively. The polycaprolactone mesh Besides HAp, tricalcium phosphates (TCP) were recently re-
around the bioink enhanced the mechanical stability of the bio- searched in collagen and alginate bioinks as they can trigger the
printed construct. In vitro results demonstrated sustained pro- differentiation of adipose stem cells without using an osteogenic
tein expression for up to 14 days post-printing. While nHAp has medium [99,164]. For bioprinting of mechanically stable 3D porous
osteoinducing properties, it effectively delivered the therapeutic structures with encapsulated MC3T3-E1 or hASCs the temperature
genes BMP-2 and transforming growth factor (TGF-β 3) addition- during bioprinting was set below the gelation temperature, which
ally to pDNA, enhancing local osteogenesis of MSCs in vitro and in- reduced damage to the encapsulated cells resulting from high wall
ducing homogeneous matrix mineralization. Vascularized and min- shear stress in the microsized nozzle. Addition of up to 20 wt%
eralized tissues were also achieved after in vivo implantation (12 β -TCP revealed initial cell viabilities of more than 90% before and
weeks) with an increased effect observed on cellularized gene- after crosslinking. Pore size shrinkage correlated with an increase
activated bioprinted constructs compared to cell-free controls. Due of β -TCP. Compared to a control which was cultured in osteogenic
16 S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22

Fig. 12. Printability of a) 5 wt% GelMA and b) Sr-GelMA bioinks (scale bars: 1 mm). Latter one can be printed into stable 3D grid structures with interconnected porosity
as optically analyzed by c) LM and d) SEM, respectively. e) Imaging of scaffold geometry throughout 28 days of in vitro culture displayed fiber shape retention of Sr-GelMA.
f) Cell viability was shown to be similar comparing hMSC-laden Sr-GelMA cast disks and bioprinted Sr-GelMA constructs which is also supported by g) live/dead staining
of bioprinted hMSCs. Osteogenic differentiation was evaluated by: h, i) Alizarin Red (red) and j) OCN (green), Col I (red) and cell nuclei (DAPI; blue) staining. Dashed white
lines indicate the edge of the Sr-GelMA filament. (Mean ± SD, n = 3 ∗ (p<0.05); Scale bars: a, b, c, e, g): 1 mm, d, h): 400 μm, I, j): 200 μm). (Modified from [100]. Copyright
2019, Elsevier Ltd). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

medium but did not contain the bioceramic inclusions, compos- a low extent it is also present in the human plasma [165]. As ex-
ite constructs demonstrated significant osteogenic gene expression. amined in previous studies, low doses of strontium ranelate hinder
It was therefore concluded that dissolution products from β -TCP, osteoclast activity and favor the proliferation of osteoblast cells,
which were released into the cell culture medium during the 21 leading to a dose-dependent increase in bone-volume and mechan-
day culture period, could promote osteogenic differentiation of the ical properties [166,167]. Thus, Sr is already used as a treatment
cells present in the bioprinted composite structure [99]. of osteoporosis as it can integrate with bone hydroxyapatite and
The available studies on composite hydrogels incorporating CaP reduce bone resorption by increasing the number of sites of os-
particles for bioprinting anticipate that these bioinks will continue seous formation and decreasing the number of active osteoclasts
to attract the attention of researchers, particularly for bone tissue [148,168,169]. Strontium-doped bioactive glass, hydroxyapatite, and
regeneration. Nevertheless, issues like specific ion release from the calcium phosphate (scaffolds) therefore show great promise as
CaP fillers and their influence on cell viability, proliferation and mi- bone graft substitute materials [119,120,170–175].
gration have to be examined in future studies to investigate the ef- Given these properties of Sr, strontium carbonate has been con-
fect of the direct contact between reactive fillers and encapsulated sidered as inorganic filler in bioinks intended for bone regener-
cells on the overall (time-dependent) cell compatibility of the con- ating constructs. Alcala-Orozco et al. recently designed a novel
structs, which has not yet been considered in detail in the pub- bioink consisting of 5 wt% GelMA and 1.5 mg/mL crystalline stron-
lished literature. tium carbonate (Sr-CO3 ) having the potential for guiding vascu-
lar ingrowth and serving as an adequate 3D matrix for cell func-
5. Strontium carbonate containing composite hydrogels tionality (Fig. 12) [100]. Strontium carbonate was synthesized in
a self-assembly process resulting in multi-rod star-like structures
The alkaline earth metal strontium can normally be found in that were incorporated into GelMA precursor solution. It could be
the human skeleton, because it is a “bone-seeking element” and to shown that Sr did not affect the crosslinking reaction of GelMA
S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22 17

yielding a soft bioink with tailorable rheological properties. More- lease of alkaline ions, needs to be prevented to assure sufficient
over, holding times of 30–60 min prior to bioprinting did not cell viability post-printing. The mechanical properties usually ex-
significantly affect cell viability which is important when print- hibit a proportional increase with filler concentration. Moreover, a
ing large-scale implants for clinical purpose. 3D bioprinted 10- too high concentration of inorganic particles strongly increases the
layered grid structures preserved their shape, geometry and pore viscosity and hence leads to enhanced printing pressures as well
integrity during 28 days static culture in vitro with cell viabilities as high shear forces on the cells. So far, the exact mechanisms of
above 90%. Culturing 3D bioprinted scaffolds in osteogenic medium particle interactions with the surrounding hydrogel matrix are still
demonstrated that encapsulated hMSCs were able to differenti- unknown. Probably, filler surface charges may interact with oppo-
ate osteogenically with an increase of deposition of bone minerals sitely charged functional groups on polymer chains. Thereby, the
in the hydrogel matrix. For future research, the translation of Sr- network and crosslinking density might be affected, leading to im-
GelMA hydrogels for load bearing applications is of special inter- proved printing characteristics.
est in bone tissue engineering. In this case the designed composite The differentiation of various cell types in composite hydro-
inks could be combined with thermoplastic polymers or bioceram- gels is complex and depends on the filler concentration, as higher
ics [100]. amounts of inorganic particles lead to an increase of the stiffness
of the printed hydrogel making likely the formation of hard tissue
6. Discussion while low viscosity hydrogels usually promote the differentiation
of encapsulated cells to form soft tissues. Furthermore, the release
By using 3D bioprinting it is in principle possible to develop of a multitude of specific ions from the bioactive fillers like for ex-
cell-laden tissue scaffolds of suitable mechanical stability, biolog- ample silicon or calcium will have an (usually desired) influence on
ical compatibility and degradability, exhibiting 3D structural and cell behavior. However they could accumulate in the environment
biochemical complexity to closely mimic natural tissues. In this re- causing local deviations from the physiological pH and potentially
view, bioinks based on hydrogel - rigid inorganic filler composites significant variations in the ECM chemical composition. Dissolution
with bioactive properties were discussed based on the available of ionic products from the surface of the bioactive filler usually
recent literature. Owing to their similar mineral phase to that of leads to mineralization of the hydrogel matrix [176]. The formation
the human body, certain inorganic materials such as hydroxyap- of hydroxyapatite, which is based on the creation of silanol bonds
atite, β -TCP, and bioactive glasses are known to have osteogenic followed by the attachment of calcium and phosphate ions, pro-
and mineralization effects and hence most reviewed studies have motes protein adsorption, leading to anchorage, proliferation, and
focused on the development of bone tissue-like constructs. These subsequent cell differentiation [176,177]. Silicon has been shown to
materials react with physiological fluids and form tenacious bonds enhance the formation of vacuoles in osteoblast cells [178]. Addi-
to hard (and in some cases soft) tissues through cellular activity tionally, calcium phosphate-based materials can either release or
and they are therefore known as „bioactive“ [99]. In the reviewed take up calcium ions from the environment. At the cellular level,
studies, encapsulated cells were in general not harmed neither by when cells attach to bioactive particles their surface characteristics
the printing process itself nor through encapsulation into the 3D change. This approach, identified also as “cells grab on particles”,
composite hydrogel matrices. In such bioinks, cells showed high has been recently discussed by Abalymov et al. [179]. As an exam-
viabilities after cultivation under normal cell culture conditions, as ple, it is possible that cells directly attached onto inorganic fillers
reported in most of the reviewed studies. Nevertheless, other as- will be able to get more calcium locally, compared to the calcium
pects specifically related to this type of hydrogel-inorganic filler concentration measured in the surrounding cell culture medium.
systems must be considered in more detail in future studies. For This process competes with the crosslinking of certain hydrogels,
example, the release of specific ions from the inorganic fillers must like alginate, which also take up calcium. Following such complex
be quantitatively studied in detail as some of them may have toxic, interactions, it is not trivial to measure the concentration of cal-
antibacterial or pH enhancing effects in relation to their concentra- cium in the bioink, especially the free ion concentration which is
tion and thus, direct contact with inorganic components in a hy- not bonded by the alginate chains and is available for interaction
drogel matrix may be harmful to cells. Such issues should be ad- with cells. Future research needs to look deeper into the inter-
dressed in further studies. Indeed, bioreactive fillers behave differ- action of released ions from bioreactive fillers and cell differenti-
ently to “persistent” fillers (e.g. carbon nanotubes) as they dissolve ation in the bioink, as the process is not completely understood
during the time of contact with the hydrogel matrix and release so far.
ionic dissolution products. Moreover, bioreactive inorganic particles Also the shape of the nanofiller is highly relevant as it has a
can be doped with therapeutic drugs or certain biologically active profound effect on mechanical properties and printability of such
ions like e.g. B3+ , Zn2+ , Sr2+ , Ag+ , Cu2+ or Mg2+ to adjust the com- composite bioinks, as clearly observed when considering nanosili-
posite hydrogels to specific needs for tissue regeneration, for ex- cates/ clays. Current research on composite bioinks lacks a distinct
ample to induce angiogenesis, trigger specific cellular responses or answer regarding the mechanisms by which filler shape and mor-
simply to act as crosslinking ions. phology affect mechanical and rheological properties. Most studies
Additionally, the inorganic components are suitable to enhance have assumed the formation of a house-of-cards structure, how-
various properties of bioinks in terms of post-printing stability, ever this effect does not generally arise at the low concentrations
mechanical properties, printing fidelity, osteogenic differentiation usually considered in the reviewed studies. It is possible to state
(when considered for bone tissue engineering) and vascularization. that by varying filler surface charges, orientation of particles and
Regarding the mechanical stability, in general, the addition of in- release of (crosslinking) ions, the mechanical properties of the hy-
organic fillers resulted in higher strength, shape recovery and en- drogel network can be tuned to match the requirement of particu-
ergy dissipation under compressive stresses. Here, structural char- lar applications. Clearly, dedicated research is necessary to investi-
acteristics like the particle size and the respective filler distribu- gate these complex interactions in composite bioinks.
tion in the hydrogel matrix have been recognized as important pa- Moreover, the design of composite materials offers an ex-
rameters affecting the mechanical behavior of the composite ink. ceptional opportunity to combine biodegradability and bioactiv-
Moreover, the release of e.g. bivalent ions from some of the fillers ity in optimized bioprinted tissue-engineering scaffolds. Composite
may lead to a (partial) crosslinking of respective composite bioinks bioinks incorporating bioreactive fillers allow the creation of biore-
(e.g. alginate, ADA or gellan gum) over time. Nevertheless, a lo- sorbable and bioactive structures with tailored time-dependent
cal increase in pH, which might arise for example from the re- physical and mechanical properties, which can be engineered in
18 S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22

such a way that their resorption rate in the body matches the for- [4] I.T. Ozbolat, Y. Yu, Bioprinting toward organ fabrication: challenges and future
mation rate of new tissue (exploiting the interaction of the hydro- trends, IEEE Trans. Biomed. Eng. 60 (3) (2013) 691–699, doi:10.1109/TBME.
2013.2243912.
gel matrix and the intrinsic reactivity of the inorganic filler). As the [5] V. Mironov, V. Kasyanov, R.R. Markwald, Organ printing: from bioprinter
filler changes in shape and composition as function of time post- to organ biofabrication line, Curr. Opin. Biotechnol. 22 (5) (2011) 667–673,
printing, the overall properties of the construct vary as well and doi:10.1016/j.copbio.2011.02.006.
[6] J. Groll, T. Boland, T. Blunk, J.A. Burdick, D.-.W. Cho, P.D. Dalton, B. Derby,
a time-dependent behavior of the constructs is expected, which G. Forgacs, Q. Li, V.A. Mironov, L. Moroni, M. Nakamura, W. Shu, S. Takeuchi,
can be tuned, for example, by considering the requirement of the G. Vozzi, T.B.F. Woodfield, T. Xu, J.J. Yoo, J. Malda, Biofabrication: reappraising
growth of new tissue. the definition of an evolving field, Biofabrication 8 (1) (2016) 013001, doi:10.
1088/1758-5090/8/1/013001.
In general, bioinks for 3D bioprinting should be adjustable to
[7] F. Guillemot, V. Mironov, M. Nakamura, Bioprinting is coming of age: report
the specific application, the operation site as well as the cho- from the International Conference on Bioprinting and Biofabrication in Bor-
sen cell types to guarantee optimal tissue formation and regen- deaux (3B’09), Biofabrication 2 (1) (2010) 010201, doi:10.1088/1758-5082/2/
1/010201.
eration. Using composite bioinks consisting of bioactive inorganic
[8] B.R. Ringeisen, R.K. Pirlo, P.K. Wu, T. Boland, Y. Huang, W. Sun, Q. Hamid,
fillers incorporated in hydrogels provides an extra degree of free- D.B. Chrisey, Cell and organ printing turns 15: diverse research to commer-
dom to approach the complexity of natural tissues and thus leads cial transitions, MRS Bull. 38 (10) (2013) 834–843, doi:10.1557/mrs.2013.209.
to promising steps towards the development of more advanced [9] V. Mironov, R.P. Visconti, V. Kasyanov, G. Forgacsc, C.J. Drake, R.R. Mark-
wald, Organ printing: tissue spheroids as building blocks, Biomaterials 30 (12)
patient-specific 3D matrices. (2009) 2164–2174, doi:10.1016/j.biomaterials.2008.12.084.
[10] T. Billiet, M. Vandenhaute, J. Schelfhout, S. Van Vlierberghe, P. Dubruel, A re-
view of trends and limitations in hydrogel-rapid prototyping for tissue en-
7. Final remarks and conclusions gineering, Biomaterials 33 (26) (2012) 6020–6041, doi:10.1016/j.biomaterials.
2012.04.050.
[11] G. Cidonio, M. Glinka, J.I. Dawson, R.O.C. Oreffo, The cell in the ink:
We have reviewed the recent literature in the field of 3D bio-
improving biofabrication by printing stem cells for skeletal regenerative
printing of composite bioinks containing bioreactive fillers. The medicine, Biomaterials 209 (2019) 10–24, doi:10.1016/j.biomaterials.2019.04.
main features which are required for such bioinks, consisting of 009.
[12] J.W. Haycock, 3D Cell Culture, Humana Press, 2011.
polymeric hydrogels and bioactive inorganic fillers of different
[13] J. Groll, J.A. Burdick, D.-.W. Cho, B. Derby, M. Gelinsky, S.C. Heilshorn,
chemical composition, were presented and the impact of the filler T. Jüngst, J. Malda, V.A. Mironov, K. Nakayama, A. Ovsianikov, W. Sun,
characteristics on various bioink properties were discussed. Exam- S. Takeuchi, J.J. Yoo, F. Woodfield, A definition of bioinks and their dis-
ples from recent literature (the past six years) were displayed, con- tinction from biomaterial inks, Biofabrication 11 (2019) 13001, doi:10.1088/
1758-5090/aaec52.
firming the great potential of such composite bioinks for bone and [14] W. Schuurman, V. Khristov, M.W. Pot, P.R. Van Weeren, W.J.A. Dhert, J. Malda,
cartilage repair/ regeneration, skeletal tissue engineering or drug Bioprinting of hybrid tissue constructs with tailorable mechanical properties,
delivery, while more research is needed for soft tissue engineer- Biofabrication 3 (2) (2011) 021001, doi:10.1088/1758-5082/3/2/021001.
[15] J.H. Shim, J.Y. Kim, M. Park, J. Park, D.W. Cho, Development of a hybrid scaf-
ing applications. It was shown that the incorporation of bioac- fold with synthetic biomaterials and hydrogel using solid freeform fabrica-
tive fillers often resulted in superior mechanical, rheological or tion technology, Biofabrication 3 (3) (2011) 034102, doi:10.1088/1758-5082/
functional properties. Overall, bioprinted composite hydrogels have 3/3/034102.
[16] M.S. Saveleva, K. Eftekhari, A. Abalymov, T.E.L. Douglas, Volodkin Dmitry,
been shown to lead to high cell viability exhibiting advantageous B.V. Parakhonskiy, A.G. Skirtach, Hierarchy of Hybrid Materials — The Place
impact on cell differentiation post-printing. Given the complexity of Inorganics- in -Organics in it, Their Composition and Applications, Front.
of composite bioinks, it is very important to understand the corre- Chem. 7 (April) (2019) 1–21, doi:10.3389/fchem.2019.00179.
[17] M. Müller, J. Becher, M. Schnabelrauch, M. Zenobi-Wong, Nanostructured
lation between the bioactive inorganic filler characteristics (e.g. ion
Pluronic hydrogels as bioinks for 3D bioprinting, Biofabrication 7 (3) (2015)
release, morphology) and the bioink properties as well as cell be- 035006, doi:10.1088/1758-5090/7/3/035006.
havior, where deeper insights and broader knowledge need to be [18] S. Wüst, M.E. Godla, R. Müller, S. Hofmann, Tunable hydrogel composite with
two-step processing in combination with innovative hardware upgrade for
gained. Such in-depth knowledge of composite bioinks and their
cell-based three-dimensional bioprinting, Acta Biomater 10 (2) (2014) 630–
biological performance will allow a better comparison and tuning 640, doi:10.1016/j.actbio.2013.10.016.
of different bioink systems, with the aim to highlight the most [19] F.P.W. Melchels, W.J.A. Dhert, D.W. Hutmacher, J. Malda, Development and
promising compositions and applications for 3D bioprinted con- characterisation of a new bioink for additive tissue manufacturing, J. Mater.
Chem. B 2 (16) (2014) 2282–2289, doi:10.1039/c3tb21280g.
structs in the future. [20] E.L. Chaikof, H. Matthew, J. Kohn, A.G. Mikos, G.D. Prestwich, C.M. Yip, Bioma-
terials and scaffolds in reparative medicine, Ann. N. Y. Acad. Sci. 961 (2002)
96–105, doi:10.1111/j.1749-6632.2002.tb03057.x.
Declaration of Competing Interest [21] L.G. Griffith, Emerging Design Principles in Biomaterials and Scaffolds for Tis-
sue Engineering, Ann. N. Y. Acad. Sci. 961 (1) (2002) 83–95, doi:10.1111/j.
1749-6632.2002.tb03056.x.
The authors declare no conflict of interest.
[22] V. Karageorgiou, D. Kaplan, Porosity of 3D biomaterial scaffolds and osteoge-
nesis, Biomaterials 26 (27) (2005) 5474–5491, doi:10.1016/j.biomaterials.2005.
02.002.
Acknowledgements [23] A.G. Mikos, J.S. Temenoff, Formation of highly porous biodegradable
scaffolds for tissue engineering, Electron. J. Biotechnol. 3 (2) (20 0 0)
The authors thank the German Research Foundation (DFG); Col- 140–144 [Online]. Available: https://scielo.conicyt.cl/scielo.php?pid=
S0717-345820 0 0 0 0 020 0 0 03&script=sci_arttext&tlng=en .
laborative Research Center SFB/TRR225 ’From the fundamentals [24] A. Motealleh, B. Çelebi-saltik, N. Ermis, S. Nowak, A. Khademhosseini,
of Biofabrication to functional tissue models’) – project number N.S. Kehr, 3D printing of step-gradient nanocomposite hydrogels for con-
326998133 - TRR 225 (subproject B03) for financial support. trolled cell migration, Biofabrication 11 (4) (2019) 45015, doi:10.1088/
1758-5090/ab3582.
[25] J.F. Mano, R.A. Sousa, L.F. Boesel, N.M. Neves, R.L. Reis, Bioinert, biodegrad-
References able and injectable polymeric matrix composites for hard tissue replacement:
state of the art and recent developments, Compos. Sci. Technol. 64 (6) (2004)
[1] R.E. Horch, A. Weigand, H. Wajant, A.R. Boccaccini, A. Arkudas, Biofabrikation 789–817, doi:10.1016/j.compscitech.20 03.09.0 01.
– neue Ansätze für den artifiziellen Gewebeersatz, Biofabrication : new ap- [26] A.R. Boccaccini, J.J. Blaker, V. Maquet, R.M. Day, R. Jéróme, Preparation
proaches for tissue regeneration, Handchirurgie, Microchirurgie . Plast. Chir. and characterisation of poly(lactide-co-grycolide) (PLGA) and PLGA/Bioglass R

50 (2018) 93–100 https://doi.org/10.1055/s- 0043- 124674. composite tubular foam scaffolds for tissue engineering applications,
[2] K. Dzobo, N.E. Thomford, D.A. Senthebane, H. Shipanga, A. Rowe, C. Dandara, Mater. Sci. Eng. C 25 (1) (2005) 23–31, doi:10.1016/j.msec.2004.03.
M. Pillay, K. Shirley, and C.M. Motaung, Advances in Regenerative Medicine 002.
and Tissue Engineering : innovation and Transformation of Medicine, Stem [27] H. Niiranen, T. Pyhältö, P. Rokkanen, M. Kellomäki, P. Törmälä, In vitro and
Cells Int. 2018 (2018) Article ID 2495848. doi: 10.1155/2018/2495848. in vivo behavior of self-reinforced bioabsorbable polymer and self-reinforced
[3] B. Derby, Printing and Prototyping of Tissues and Scaffolds, Science 338 bioabsorbable polymer/bioactive glass composites, J. Biomed. Mater. Res. Part
(6109) (2012) 921–926, doi:10.1126/science.1226340. A 69A (4) (2004) 699–708, doi:10.1002/jbm.a.30043.
S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22 19

[28] J. Yao, S. Radin, P.S. Leboy, P. Ducheyne, The effect of bioactive glass [57] G. Turnbull, J. Clarke, F. Picard, P. Riches, L. Jia, F. Han, B. Li, W. Shu, 3D
content on synthesis and bioactivity of composite poly (lactic-co-glycolic bioactive composite scaffolds for bone tissue engineering, Bioact. Mater. 3 (3)
acid)/bioactive glass substrate for tissue engineering, Biomaterials 26 (14) (2018) 278–314, doi:10.1016/j.bioactmat.2017.10.001.
(2005) 1935–1943, doi:10.1016/j.biomaterials.2004.06.027. [58] M. Müller, E. Öztürk, Ø. Arlov, P. Gatenholm, M. Zenobi-Wong, Alginate
[29] Y.M. Khan, D.S. Katti, and C.T. Laurencin, “Novel polymer-synthesized ceramic Sulfate–Nanocellulose Bioinks for Cartilage Bioprinting Applications, Ann.
composite – based system for bone repair : an in vitro evaluation,” 2004, doi: Biomed. Eng. 45 (1) (2017) 210–223, doi:10.1007/s10439- 016- 1704- 5.
10.10 02/jbm.a.30 051. [59] L.M. Cross, K. Shah, S. Palani, C.W. Peak, K. Akhilesh, Gradient Nanocompos-
[30] D. Chimene, K.K. Lennox, R.R. Kaunas, A.K. Gaharwar, Advanced Bioinks for ite Hydrogels for Interface Tissue Engineering, Nanomedicine Nanotechnology,
3D Printing: a Materials Science Perspective, Ann. Biomed. Eng. 44 (6) (2016) Biol. Med. 14 (2018) 2465–2474, doi:10.1016/j.nano.2017.02.022.
2090–2102, doi:10.1007/s10439- 016- 1638- y. [60] I. Hernandez, A. Kumar, B. Joddar, A Bioactive Hydrogel and 3D Printed
[31] Z. Wang, S.J. Lee, H. Cheng, J.J. Yoo, A. Atala, 3D bioprinted functional and Polycaprolactone System for Bone Tissue Engineering, Gels 3 (3) (2017) 26,
contractile cardiac tissue constructs, Acta Biomater. 70 (2018) 48–56, doi:10. doi:10.3390/gels3030026.
1016/j.actbio.2018.02.007. [61] M. Zhu, J. Zhang, S. Zhao, Y. Zhu, Three-dimensional printing of cerium-
[32] I. Angelopoulos, M.C. Allenby, M. Lim, M. Zamorano, Engineering inkjet bio- incorporated mesoporous calcium-silicate scaffolds for bone repair, J. Mater.
printing processes toward translational therapies, Biotechnol. Bioeng. 117 (1) Sci. 51 (2) (2016) 836–844, doi:10.1007/s10853-015- 9406- 1.
(2019) 272–284, doi:10.1002/bit.27176. [62] A. Gantar, P. Lucilia, J.M. Oliveira, A.P. Marques, V.M. Correlo, R.L. Reis,
[33] M. Hospodiuk, M. Dey, D. Sosnoski, I.T. Ozbolat, The bioink: a comprehen- Nanoparticulate bioactive-glass-reinforced gellan-gum hydrogels for bone-
sive review on bioprintable materials, Biotechnol. Adv. 35 (2) (2016) 217–239, tissue engineering, Mater. Sci. Eng. C 43 (2014) 27–36, doi:10.1016/j.msec.
doi:10.1016/j.biotechadv.2016.12.006. 2014.06.045.
[34] R.R. Jose, M.J. Rodriguez, T.A. Dixon, F. Omenetto, D.L. Kaplan, Evolution of [63] U.G.K. Wegst, H. Bai, E. Saiz, A.P. Tomsia, R.O. Ritchie, Bioinspired structural
Bioinks and Additive Manufacturing Technologies for 3D Bioprinting, ACS Bio- materials, Nat. Mater. 14 (1) (2015) 23–36, doi:10.1038/nmat4089.
mater. Sci. Eng. 2 (2016) 1662–1678, doi:10.1021/acsbiomaterials.6b0 0 088. [64] T. Dvir, B.P. Timko, D.S. Kohane, R. Langer, Nanotechnological strategies for
[35] A. Panwar, L.P. Tan, Current status of bioinks for micro-extrusion-based 3D engineering complex tissues, Nat. Nanotechnol. 6 (1) (2011) 13–22, doi:10.
bioprinting, Molecules 21 (6) (2016), doi:10.3390/molecules21060685. 1038/nnano.2010.246.
[36] H. Li, C. Tan, L. Li, Review of 3D printable hydrogels and constructs, Mater. [65] A.K. Gaharwar, N.A. Peppas, A. Khademhosseini, Nanocomposite hydrogels for
Des. 159 (2018) 20–38, doi:10.1016/j.matdes.2018.08.023. biomedical applications, Biotechnol. Bioeng. 111 (3) (2014) 441–453, doi:10.
[37] S. Utech, A.R. Boccaccini, A review of hydrogel-based composites for 1002/bit.25160.
biomedical applications: enhancement of hydrogel properties by addition [66] J.K. Carrow, A.K. Gaharwar, Bioinspired polymeric nanocomposites for regen-
of rigid inorganic fillers, J. Mater. Sci. 51 (1) (2016) 271–310, doi:10.1007/ erative medicine, Macromol. Chem. Phys. 216 (3) (2015) 248–264, doi:10.
s10853-015-9382-5. 10 02/macp.20140 0427.
[38] A.K. Means, M.A. Grunlan, Modern Strategies To Achieve Tissue-Mimetic, Me- [67] J.I. Dawson, R.O.C. Oreffo, Clay: new opportunities for tissue regeneration and
chanically Robust Hydrogels, ACS Macro Lett. 8 (2019) 705–713, doi:10.1021/ biomaterial design, Adv. Mater. 25 (30) (2013) 4069–4086, doi:10.1002/adma.
acsmacrolett.9b00276. 201301034.
[39] A. Motealleh, N.S. Kehr, Nanocomposite Hydrogels and Their Applications in [68] W. Li, J. Wang, J. Ren, X. Qu, 3D graphene oxide-polymer hydrogel: near-
Tissue Engineering, Adv. Healthc. Mater. 6 (1) (2017) 1600938, doi:10.1002/ infrared light-triggered active scaffold for reversible cell capture and on-
adhm.201600938. demand release, Adv. Mater. 25 (46) (2013) 6737–6743, doi:10.1002/adma.
[40] D. Chimene, R. Kaunas, A.K. Gaharwar, Hydrogel Bioink Reinforcement for Ad- 201302810.
ditive Manufacturing: a Focused Review of Emerging Strategies, Adv. Mater. [69] A.C. Balazs, T. Emrick, T.P. Russell, Nanoparticle Polymer Composites: where
32 (1) (2020) 1902026, doi:10.1002/adma.201902026. Two Small Worlds Meet, Science (80-.). 314 (5802) (Nov. 2006) 1107–1110,
[41] L. Valot, J. Martinez, A. Mehdi, G. Subra, Chemical insights into bioinks for 3D doi:10.1126/science.1130557.
printing, Chem. Soc. Rev. 48 (15) (2019) 4049–4086, doi:10.1039/c7cs00718c. [70] P. Bordes, E. Pollet, L. Avérous, Nano-biocomposites: biodegradable
[42] J. Malda, J. Visser, F.P. Melchels, T. Jüngst, W.E. Hennink, W.J.A. Dhert, J. Groll, polyester/nanoclay systems, Prog. Polym. Sci. 34 (2) (2009) 125–155,
D.W. Hutmacher, 25th anniversary article: engineering hydrogels for biofabri- doi:10.1016/j.progpolymsci.2008.10.002.
cation, Adv. Mater. 25 (36) (2013) 5011–5028, doi:10.1002/adma.201302042. [71] E. Johnson, C. Sheffield, K. Meyers, R.M. Rajachar, Application of Composite
[43] T. Jungst, W. Smolan, K. Schacht, T. Scheibel, J. Groll, Strategies and Molecular Hydrogels to Control Physical Properties in Tissue Engineering and Regenera-
Design Criteria for 3D Printable Hydrogels, Chem. Rev. 116 (3) (2016) 1496– tive Medicine, Gels 4 (2018) 51, doi:10.3390/gels4020051.
1539, doi:10.1021/acs.chemrev.5b00303. [72] A.A. Dayem, H.Y. Choi, G.M. Yang, K. Kim, S.K. Saha, J.H. Kim, S.G. Cho, The
[44] T.S. Jang, H. Do Jung, H.M. Pan, W.T. Han, S. Chen, J. Song, 3D printing of potential of nanoparticles in stem cell differentiation and further therapeu-
hydrogel composite systems: recent advances in technology for tissue engi- tic applications, Biotechnol. J. 11 (12) (2016) 1550–1560, doi:10.1002/biot.
neering, Int. J. Bioprinting 4 (1) (2018) 1–28, doi:10.18063/IJB.v4i1.126. 201600453.
[45] S. Naghieh, M. Sarker, N.K. Sharma, Z. Barhoumi, X. Chen, Printability of 3D [73] I. Ilie, R. Ilie, T. Mocan, D. Bartos, L. Mocan, Influence of nanomaterials on
Printed Hydrogel Scaffolds: influence of Hydrogel Composition and Printing stem cell differentiation: designing an appropriate nanobiointerface, Int. J.
Parameters, Appl. Sci. 10 (1) (2019) 292, doi:10.3390/app10010292. Nanomedicine 7 (2012) 2211–2225, doi:10.2147/IJN.S29975.
[46] F.L.C. Morgan, L. Moroni, M.B. Baker, Dynamic Bioinks to Advance Bioprinting, [74] I.P. Magalhães, P.M. de Oliveira, J. Dernowsek, E.B. Las Casas, M.S. Las Casas,
Adv. Healthc. Mater. (2020) 1901798, doi:10.1002/adhm.201901798. Investigation of the effect of nozzle design on rheological bioprinting prop-
[47] N. Ashammakhi, S. Ahadian, C. Xu, H. Montazerian, H. Ko, R. Nasiri, N. Barros, erties using computational fluid dynamics, Matéria (Rio J.) 24 (3) (2019) e-
A. Khademhosseini, Bioinks and Bioprinting Technologies to Make Heteroge- 12401, doi:10.1590/s1517-707620190 0 03.0714.
neous and Biomimetic Tissue Constructs, Mater. Today Bio 1 (2019) 10 0 0 08, [75] A. Atala, J.J. Yoo, Essentials of 3D Biofabrication and Translation, Elsevier, Ox-
doi:10.1016/j.mtbio.2019.10 0 0 08. ford, 2015.
[48] M. Mehrali, A. Thakur, C.P. Pennisi, S. Talebian, A. Arpanaei, M. Nikkhah, [76] X. Wang, E. Tolba, H.C.S. Der, M. Neufurth, Q. Feng, B.R. Diehl-Seifert,
A. Dolatshahi-Pirouz, Nanoreinforced Hydrogels for Tissue Engineering: bio- W.E.G. Mü Ller, Effect of bioglass on growth and biomineralization of Saos-
materials that are Compatible with Load-Bearing and Electroactive Tissues, 2 cells in hydrogel after 3d cell bioprinting, PLoS ONE 9 (11) (2014) e112497,
Adv. Mater. 29 (8) (2017) 1603612, doi:10.1002/adma.201603612. doi:10.1371/journal.pone.0112497.
[49] D. Kilian, T. Ahlfeld, A.R. Akkineni, A. Lode, M. Gelinsky, Three-dimensional [77] M. Lee, K. Bae, P. Guillon, J. Chang, Ø. Arlov, M. Zenobi-Wong, Exploitation
bioprinting of volumetric tissues and organs, MRS Bull. 42 (08) (2017) 585– of Cationic Silica Nanoparticles for Bioprinting of Large-Scale Constructs with
592, doi:10.1557/mrs.2017.164. High Printing Fidelity, ACS Appl. Mater. Interfaces 10 (44) (2018) 37820–
[50] J. Jang, J.Y. Park, G. Gao, D.W. Cho, Biomaterials-based 3D cell printing for 37828, doi:10.1021/acsami.8b13166.
next-generation therapeutics and diagnostics, Biomaterials 156 (2018) 88– [78] S. Reakasame, D. Trapani, R. Detsch, A.R. Boccaccini, Cell laden alginate-
106, doi:10.1016/j.biomaterials.2017.11.030. keratin based composite microcapsules containing bioactive glass for tissue
[51] P.S. Gungor-Ozkerim, I. Inci, Y.S. Zhang, A. Khademhosseini, M.R. Dokmeci, engineering applications, J. Mater. Sci. Mater. Med. 29 (2018) 185, doi:10.
Bioinks for 3D bioprinting: an overview, Biomater. Sci. 6 (5) (2018) 915–946, 1007/s10856- 018- 6195- 5.
doi:10.1039/c7bm00765e. [79] U. Rottensteiner-Brandl, R. Detsch, B. Sarker, L. Lingens, K. Köhn, U. Kneser,
[52] K. Hölzl, S. Lin, L. Tytgat, S. Van Vlierberghe, L. Gu, A. Ovsianikov, Bioink prop- A.K. Bosserhoff, R.E. Horch, A.R. Boccaccini, A. Arkudas, Encapsulation of
erties before, during and after 3D bioprinting, Biofabrication 8 (3) (Sep. 2016) Rat Bone Marrow Derived Mesenchymal Stem Cells in Alginate Dialdehyde
032002, doi:10.1088/1758-5090/8/3/032002. / Gelatin Microbeads with and without Nanoscaled Bioactive Glass for In
[53] A. Parak, P. Pradeep, L.C. du Toit, P. Kumar, Y.E. Choonara, V. Pillay, Functional- Vivo Bone Tissue Engineering, Materials (Basel) 11 (2018) 1880, doi:10.3390/
izing bioinks for 3D bioprinting applications, Drug Discov. Today 24 (1) (2019) ma11101880.
198–205, doi:10.1016/j.drudis.2018.09.012. [80] G. Gao, A.F. Schilling, T. Yonezawa, J. Wang, G. Dai, X. Cui, Bioactive nanopar-
[54] A. Skardal, Perspective: ‘Universal’ bioink technology for advancing extrusion ticles stimulate bone tissue formation in bioprinted three-dimensional scaf-
bioprinting-based biomanufacturing, Bioprinting 10 (2018) e0 0 026, doi:10. fold and human mesenchymal stem cells, Biotechnol. J. 9 (2014) 1304–1311,
1016/j.bprint.2018.e0 0 026. doi:10.10 02/biot.20140 0305.
[55] N.A. Peppas, J.Z. Hilt, A. Khademhosseini, R. Langer, Hydrogels in biology and [81] M. Ojansivu, A. Rashad, A. Ahlinder, J. Massera, A. Mishra, K. Syverud,
medicine: from molecular principles to bionanotechnology, Adv. Mater. 18 A. Finne-Wistrand, S. Miettinen, K. Mustafa, Wood-based nanocellulose
(11) (2006) 1345–1360, doi:10.1002/adma.200501612. and bioactive glass modified gelatin-alginate bioinks for 3D bioprinting
[56] A.S. Hoffman, Hydrogels for biomedical applications, Adv. Drug Deliv. Rev. 64 of bone cells, Biofabrication 11 (3) (2019) 035010, doi:10.1088/1758-5090/
(2012) 18–23, doi:10.1016/j.addr.2012.09.010. ab0692.
20 S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22

[82] Á.J. Leite, B. Sarker, T. Zehnder, R. Silva, J.F. Mano, A.R. Boccaccini, Bioplot- cal Applications of Silica, Alumina and Calcium Phosphate-based Nanostruc-
ting of a bioactive alginate dialdehyde-gelatin composite hydrogel containing tured Materials, Curr. Med. Chem. 23 (39) (2016) 4450–4467, doi:10.2174/
bioactive glass nanoparticles, Biofabrication 8 (3) (2016) 035005, doi:10.1088/ 0929867323666161024153459.
1758-5090/8/3/035005. [107] M. Kesti, C. Eberhardt, G. Pagliccia, D. Kenkel, D. Grande, A. Boss, M. Zenobi-
[83] B. Byambaa, N. Annabi, K. Yue, G. Trujillo-de Santiago, M.M. Alvarez, W. Jia, Wong, Bioprinting Complex Cartilaginous Structures with Clinically Compliant
M. Kazemzadeh-Narbat, S.R. Shin, A. Tamayol, A. Khademhosseini, Bioprinted Biomaterials, Adv. Funct. Mater. 25 (48) (2015) 7406–7417, doi:10.1002/adfm.
Osteogenic and Vasculogenic Patterns for Engineering 3D Bone Tissue, Adv. 201503423.
Healthc. Mater. 6 (16) (2017) 170 0 015, doi:10.10 02/adhm.20170 0 015. [108] L.L. Hench, R.J. Splinter, W.C. Allen, T.K. Greenlee, Bonding mechanisms at the
[84] C.W. Peak, K.A. Singh, M. Adlouni, J. Chen, A.K. Gaharwar, Printing Therapeutic interface of ceramic prosthetic materials, J. Biomed. Mater. Res. 5 (6) (1971)
Proteins in 3D using Nanoengineered Bioink to Control and Direct Cell Migra- 117–141, doi:10.10 02/jbm.820 050611.
tion, Adv. Healthc. Mater. 8 (2019) 1801553, doi:10.1002/adhm.201801553. [109] A. Hoppe, N.S. Güldal, A.R. Boccaccini, A review of the biological response to
[85] A. Habib, B. Khoda, Development of clay based novel hybrid bio-ink for 3D ionic dissolution products from bioactive glasses and glass-ceramics, Bioma-
bio-printing process, J. Manuf. Process. 38 (2019) 76–87, doi:10.1016/j.jmapro. terials 32 (11) (2011) 2757–2774, doi:10.1016/j.biomaterials.2011.01.004.
2018.12.034. [110] V. Miguez-Pacheco, L.L. Hench, A.R. Boccaccini, Bioactive glasses beyond bone
[86] T. Ahlfeld, G. Cidonio, D. Kilian, S. Duin, A.R. Akkineni, J.I. Dawson, S. Yang, and teeth: emerging applications in contact with soft tissues, Acta Biomater
A. Lode, R.O.C. Oreffo, M. Gelinsky, Development of a clay based bioink for 13 (2015) 1–15, doi:10.1016/j.actbio.2014.11.004.
3D cell printing for skeletal application, Biofabrication 9 (3) (2017) 034103, [111] K. Vuornos, M. Ojansivu, J.T. Koivisto, H. Häkkänen, B. Belay, T. Montonen,
doi:10.1088/1758-5090/aa7e96. H. Huhtala, M. Kääriäinen, L. Hupa, M. Kellomäki, J. Hyttinen, J.A. Ihalainen,
[87] X. Zhai, C. Ruan, Y. Ma, D. Cheng, M. Wu, W. Liu, X. Zhao, H. Pan, W.W. Lu, S. Miettinen, Bioactive glass ions induce efficient osteogenic differentiation of
3D-Bioprinted Osteoblast-Laden Nanocomposite Hydrogel Constructs with In- human adipose stem cells encapsulated in gellan gum and collagen type I
duced Microenvironments Promote Cell Viability, Differentiation, and Osteo- hydrogels, Mater. Sci. Eng. C 99 (2019) 905–918, doi:10.1016/j.msec.2019.02.
genesis both In Vitro and In Vivo, Adv. Sci. 5 (3) (2018) 1700550, doi:10.1002/ 035.
advs.201700550. [112] L. Ouyang, R. Yao, Y. Zhao, W. Sun, Effect of bioink properties on printability
[88] S.A. Wilson, L.M. Cross, C.W. Peak, A.K. Gaharwar, Shear-Thinning and and cell viability for 3D bioplotting of embryonic stem cells, Biofabrication 8
Thermo-Reversible Nanoengineered Inks for 3D Bioprinting, ACS Appl. Mater. (3) (2016) 035020, doi:10.1088/1758-5090/8/3/035020.
Interfaces 9 (50) (2017) 43449–43458, doi:10.1021/acsami.7b13602. [113] B. Balakrishnan, N. Joshi, A. Jayakrishnan, R. Banerjee, Self-crosslinked oxi-
[89] D. Chimene, C.W. Peak, J.L. Gentry, J.K. Carrow, L.M. Cross, E. Mondragon, dized alginate/gelatin hydrogel as injectable, adhesive biomimetic scaffolds
G.B. Cardoso, R. Kaunas, A.K. Gaharwar, Nanoengineered Ionic-Covalent En- for cartilage regeneration, Acta Biomater. 10 (8) (2014) 3650–3663, doi:10.
tanglement (NICE) Bioinks for 3D Bioprinting, ACS Appl. Mater. Interfaces 10 1016/j.actbio.2014.04.031.
(12) (2018) 9957–9968, doi:10.1021/acsami.7b19808. [114] B. Sarker, R. Singh, R. Silva, J.A. Roether, J. Kaschta, R. Detsch, D.W. Schu-
[90] A. Paul, V. Manoharan, D. Krafft, A. Assmann, J.A. Uquillas, S.R. Shin, A. Hasan, bert, I. Cicha, A.R. Boccaccini, Evaluation of fibroblasts adhesion and prolifera-
M.A. Hussain, A. Memic, A.K. Gaharwar, A. Khademhosseini, Nanoengineered tion on alginate-gelatin crosslinked hydrogel, PLoS ONE 9 (9) (2014) e107952,
biomimetic hydrogels for guiding human stem cell osteogenesis in three di- doi:10.1371/journal.pone.0107952.
mensional microenvironments, J. Mater. Chem. B 4 (20) (2016) 3544–3554, [115] B. Sarker, J. Rompf, R. Silva, N. Lang, R. Detsch, J. Kaschta, B. Fabry, A.R. Boc-
doi:10.1039/c5tb02745d. caccini, Alginate-based hydrogels with improved adhesive properties for cell
[91] G. Cidonio, M. Glinka, J.I. Dawson, R.O.C. Oreffo, Printing bone in a gel: using encapsulation, Int. J. Biol. Macromol. 78 (2015) 72–78, doi:10.1016/j.ijbiomac.
nanocomposite bioink to print functionalised bone scaffolds, Biomaterials 209 2015.03.061.
(2019) 10–24, doi:10.1016/j.biomaterials.2019.04.009. [116] A. Grigore, B. Sarker, B. Fabry, A.R. Boccaccini, R. Detsch, Behavior of Encap-
[92] G. Cidonio, C.R. Alcala-Orozco, K.S. Lim, M. Glinka, I. Mutreja, Y.H. Kim, sulated MG-63 Cells in RGD and Gelatine-Modified Alginate Hydrogels, Tissue
J.I. Dawson, T.B.F. Woodfield, R.O.C. Oreffo, Osteogenic and angiogenic tissue Eng. Part A 20 (15–16) (2014) 2140–2150, doi:10.1089/ten.tea.2013.0416.
formation in high fidelity nanocomposite Laponite-gelatin bioinks, Biofabrica- [117] Y. Hu, L. Liu, Z. Gu, W. Dan, N. Dan, X. Yu, Modification of collagen with a
tion 11 (3) (2019) 035027, doi:10.1088/1758-5090/ab19fd. natural derived cross-linker, alginate dialdehyde, Carbohydr. Polym. 102 (1)
[93] T.T. Demirtaş, I. Gülseren, M. Gümüşderelioğlu, A bioprintable form of chi- (2014) 324–332, doi:10.1016/j.carbpol.2013.11.050.
tosan hydrogel for bone tissue engineering, Biofabrication 9 (2017) 035003, [118] A.R. Boccaccini, M. Erol, W.J. Stark, D. Mohn, Z. Hong, J.F. Mano, Poly-
doi:10.1088/1758-5090/aa7b1d. mer/bioactive glass nanocomposites for biomedical applications: a review,
[94] A. Wenz, K. Borchers, G.E.M. Tovar, P.J. Kluger, Bone matrix production in Compos. Sci. Technol. 70 (13) (2010) 1764–1776, doi:10.1016/j.compscitech.
hydroxyapatite-modified hydrogels suitable for bone bioprinting, Biofabrica- 2010.06.002.
tion 9 (4) (2017) 44103, doi:10.1088/1758-5090/aa91ec. [119] L.A. Strobel, N. Hild, D. Mohn, W.J. Stark, A. Hoppe, U. Gbureck, R.E. Horch,
[95] X.F. Wang, P.J. Lu, Y. Song, Y.C. Sun, Y.G. Wang, Y. Wang, Nano hydroxyapatite U. Kneser, A.R. Boccaccini, Novel strontium-doped bioactive glass nanopar-
particles promote osteogenesis in a three-dimensional bio-printing construct ticles enhance proliferation and osteogenic differentiation of human bone
consisting of alginate/gelatin/hASCs, RSC Adv. 6 (8) (2016) 6832–6842, doi:10. marrow stromal cells, J. Nanoparticle Res. 15 (7) (2013) 1780, doi:10.1007/
1039/c5ra21527g. s11051- 013- 1780- 5.
[96] Y. Fan, T. Shi, X. Yue, F. Sun, D. Yao, 3D Composite Cell Printing [120] A. Hoppe, B. Sarker, R. Detsch, N. Hild, D. Mohn, W.J. Wark, A.R. Boccaccini,
Gelatin/Sodium Alginate/n-HAP Bioscaffold, J. Phys. Conf. Ser. 1213 (4) (2019) In vitro reactivity of Sr-containing bioactive glass (type 1393) nanoparticles, J.
042020, doi:10.1088/1742-6596/1213/4/042020. Non. Cryst. Solids 387 (2014) 41–46, doi:10.1016/j.jnoncrysol.2013.12.010.
[97] S.T. Bendtsen, S.P. Quinnell, M. Wei, Development of a novel alginate- [121] A.K. Gaharwar, S.M. Mihaila, A. Swami, A. Patel, S. Sant, R.L. Reis, A.P. Mar-
polyvinyl alcohol-hydroxyapatite hydrogel for 3D bioprinting bone tissue en- ques, M.E. Gomes, A. Khademhosseini, Bioactive silicate nanoplatelets for os-
gineered scaffolds, J. Biomed. Mater. Res. - Part A 105 (5) (2017) 1457–1468, teogenic differentiation of human mesenchymal stem cells, Adv. Mater. 25
doi:10.1002/jbm.a.36036. (24) (2013) 3329–3336, doi:10.10 02/adma.20130 0584.
[98] G.M. Cunniffe, T. Gonzalez-Fernandez, A. Daly, B.N. Sathy, O. Jeon, E. Alsberg, [122] S.M. Mihaila, A.K. Gaharwar, R.L. Reis, A. Khademhosseini, A.P. Marques,
D.J. Kelly, Three-Dimensional Bioprinting of Polycaprolactone Reinforced Gene M.E. Gomes, The osteogenic differentiation of SSEA-4 sub-population of hu-
Activated Bioinks for Bone Tissue Engineering, Tissue Eng. Part A 23 (17–18) man adipose derived stem cells using silicate nanoplatelets, Biomaterials 35
(2017) 891–900, doi:10.1089/ten.tea.2016.0498. (33) (2014) 9087–9099, doi:10.1016/j.biomaterials.2014.07.052.
[99] W. Kim, G. Kim, Collagen/bioceramic-based composite bioink to fabricate a [123] A.K. Gaharwar, R.K. Avery, A. Assmann, A. Paul, G.H. McKinley, A. Khademhos-
porous 3D hASCs-laden structure for bone tissue regeneration, Biofabrication seini, B.D. Olsen, Shear-thinning nanocomposite hydrogels for the treatment
12 (1) (2020) 015007, doi:10.1088/1758-5090/ab436d. of hemorrhage, ACS Nano 8 (10) (2014) 9833–9842, doi:10.1021/nn503719n.
[100] C.R. Alcala-Orozco, I. Mutreja, X. Cui, D. Kumar, G.J. Hooper, K.S. Lim, [124] A.K. Gaharwar, S. Mukundan, E. Karaca, A. Dolatshahi-Pirouz, A. Patel, K. Ran-
T.B.F. Woodfield, Design and characterisation of multi-functional strontium- garajan, S.M. Mihaila, G. Iviglia, H. Zhang, A. Khademhosseini, Nanoclay-
gelatin nanocomposite bioinks with improved print fidelity and osteogenic Enriched Poly(ɛ-caprolactone) Electrospun Scaffolds for Osteogenic Differen-
capacity, Bioprinting 18 (2020) e0 0 073, doi:10.1016/j.bprint.2019.e0 0 073. tiation of Human Mesenchymal Stem Cells, Tissue Eng. Part A 20 (15–16)
[101] A. Bitar, N.M. Ahmad, H. Fessi, A. Elaissari, Silica-based nanoparticles for (2014) 2088–2101, doi:10.1089/ten.tea.2013.0281.
biomedical applications, Drug Discov. Today 17 (19–20) (2012) 1147–1154, [125] A.K. Gaharwar, L.M. Cross, C.W. Peak, K. Gold, J.K. Carrow, A. Brokesh,
doi:10.1016/j.drudis.2012.06.014. K.A. Singh, 2D Nanoclay for Biomedical Applications: regenerative Medicine,
[102] M. Vallet-Regí, F. Balas, Silica Materials for Medical Applications, Open Therapeutic Delivery, and Additive Manufacturing, Adv. Mater. 31 (2019) 1–28
Biomed. Eng. J. 2 (1) (2008) 1–9, doi:10.2174/1874120700802010001. 1900332, doi:10.1002/adma.201900332.
[103] Z. Li, J.C. Barnes, A. Bosoy, J.F. Stoddart, J.I. Zink, Mesoporous silica nanopar- [126] L.M. Cross, J.K. Carrow, X. Ding, K.A. Singh, A.K. Gaharwar, Sustained and
ticles in biomedical applications, Chem. Soc. Rev. 41 (7) (2012) 2590–2605, Prolonged Delivery of Protein Therapeutics from Two-Dimensional Nanosil-
doi:10.1039/c1cs15246g. icates, ACS Appl. Mater. Interfaces 11 (7) (2019) 6741–6750, doi:10.1021/
[104] M.A. Chowdhury, Silica Materials for Biomedical Applications in Drug Deliv- acsami.8b17733.
ery, Bone Treatment or Regeneration, and MRI Contrast Agent, Rev. J. Chem. [127] D.W. Thompson, J.T. Butterworth, The nature of laponite and its aqueous
8 (2) (2018) 223–241, doi:10.1134/s2079978018020024. dispersions, J. Colloid Interface Sci. 151 (1) (1992) 236–243, doi:10.1016/
[105] J.G. Croissant, Y. Fatieiev, A. Almalik, N.M. Khashab, Mesoporous Silica and 0021-9797(92)90254-J.
Organosilica Nanoparticles: physical Chemistry, Biosafety, Delivery Strategies, [128] J.K. Carrow, L.M. Cross, R.W. Reese, M.K. Jaiswal, C.A. Gregory, R. Kaunas,
and Biomedical Applications, Adv. Healthc. Mater. 7 (4) (2018) 1700831, I. Singh, A.K. Gaharwar, Widespread changes in transcriptome profile of hu-
doi:10.10 02/adhm.20170 0831. man mesenchymal stem cells induced by two-dimensional nanosilicates, Proc.
[106] Y. Ellahioui, S. Prashar, S. Gómez-Ruiz, A Short Overview on the Biomedi- Natl. Acad. Sci. 115 (17) (2018) E3905–E3913, doi:10.1073/pnas.1716164115.
S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22 21

[129] J.L. Dávila, M.A. d’Ávila, Rheological evaluation of Laponite/alginate inks for [153] Q. Li, X. Lei, X. Wang, Z. Cai, P. Lyu, G. Zhang, Hydroxyapatite/Collagen Three-
3D extrusion-based printing, Int. J. Adv. Manuf. Technol. 101 (1–4) (2019) Dimensional Printed Scaffolds and Their Osteogenic Effects on Human Bone
675–686, doi:10.10 07/s0 0170-018-2876-y. Marrow-Derived Mesenchymal Stem Cells, Tissue Eng. Part A 25 (2019) 1261–
[130] Q. Wang, J.L. Mynar, M. Yoshida, E. Lee, M. Lee, K. Okuro, K. Kinbara, 1271, doi:10.1089/ten.tea.2018.0201.
T. Aida, High-water-content mouldable hydrogels by mixing clay and a den- [154] J. Hall, R.G. Soransen, J.M. Wozney, U.M.E. Wikesjö, Bone formation at rhBMP-
dritic molecular binder, Nature 463 (7279) (2010) 339–343, doi:10.1038/ 2-coated titanium implants in the rat ectopic model, J. Clin. Periodontol. 34
nature08693. (5) (2007) 444–451, doi:10.1111/j.1600-051X.2007.01064.x.
[131] B. Ruzicka, Z. Emanuela, L. Zulian, R. Angelini, M. Sztucki, A. Mous- [155] T. Billiet, E. Gevaert, T. De Schryver, M. Cornelissen, P. Dubruel, The 3D
saïd, H. Narayanan, F. Sciortino, Observation of empty liquids and equi- printing of gelatin methacrylamide cell-laden tissue-engineered constructs
librium gels in a colloidal clay, Nat. Mater. 10 (2011) 56–60, doi:10.1038/ with high cell viability, Biomaterials 35 (1) (2014) 49–62, doi:10.1016/j.
NMAT2921 Observation. biomaterials.2013.09.078.
[132] R. Angelini, E. Zaccarelli, F.A. De Melo Marques, M. Sztucki, A. Fluerasu, [156] J. Jia, D.J. Richards, S. Pollard, Y. Tan, J. Rodriguez, R.P. Visconti, T.C. Trusk,
G. Ruocco, B. Ruzicka, Glass-glass transition during aging of a colloidal clay, M.J. Yost, H. Yao, R.R. Markwald, Y. Mei, Engineering alginate as bioink for
Nat. Commun. 5 (2014) 4049, doi:10.1038/ncomms5049. bioprinting, Acta Biomater 10 (10) (2014) 4323–4331, doi:10.1016/j.actbio.
[133] D. Bonn, S. Tanase, B. Abou, H. Tanaka, J. Meunier, Laponite: aging and shear 2014.06.034.
rejuvenation of a colloidal glass, Phys. Rev. Lett. 89 (1) (2002) 157011–157014, [157] P. Müller, U. Bulnheim, A. Diener, F. Lüthen, M. Teller, E.D. Klinkenberg,
doi:10.1103/PhysRevLett.89.015701. H.G. Neumann, B. Nebe, A. Liebold, G. Steinhoff, J. Rychly, Calcium phosphate
[134] S. Maisanaba, S. Pichardo, M. Puerto, D. Gutiérrez-Praena, A.M. Cameán, A. Jos, surfaces promote osteogenic differentiation of mesenchymal stem cells, J. Cell.
Toxicological evaluation of clay minerals and derived nanocomposites: a re- Mol. Med. 12 (1) (2007) 281–291, doi:10.1111/j.1582-4934.20 07.0 0103.x.
view, Environ. Res. 138 (2015) 233–254, doi:10.1016/j.envres.2014.12.024. [158] F. Zhao, W.L. Grayson, T. Ma, B. Bunnell, W.W. Lu, Effects of hydroxyapatite
[135] J.R. Xavier, T. Thakur, P. Desai, M.K. Jaiswal, N. Sears, E. Cosgriff-Hernandez, in 3-D chitosan-gelatin polymer network on human mesenchymal stem cell
R. Kaunas, A.K. Gaharwar, Bioactive nanoengineered hydrogels for bone tissue construct development, Biomaterials 27 (9) (2006) 1859–1867, doi:10.1016/j.
engineering: a growth-factor-free approach, ACS Nano 9 (3) (2015) 3109–3118, biomaterials.2005.09.031.
doi:10.1021/nn507488s. [159] J.L. Moreau, H.H.K. Xu, Mesenchymal stem cell proliferation and differen-
[136] H. Takeno, Y. Kimura, W. Nakamura, Mechanical, Swelling, and Structural tiation on an injectable calcium phosphate - chitosan composite scaffold,
Properties of Mechanically Tough Clay-Sodium Polyacrylate Blend Hydrogels, Biomaterials 30 (14) (2009) 2675–2682 https://doi.org/10.1016/j.biomaterials.
Gels 3 (1) (2017) 10, doi:10.3390/gels3010010. 2009.01.022.
[137] D. Chimene, D.L. Alge, A.K. Gaharwar, Two-Dimensional Nanomaterials for [160] K.Y. Lee, D.J. Mooney, Alginate: properties and biomedical applications, Prog.
Biomedical Applications: emerging Trends and Future Prospects, Adv. Mater. Polym. Sci. 37 (1) (2012) 106–126, doi:10.1016/j.progpolymsci.2011.06.003.
27 (45) (2015) 7261–7284, doi:10.1002/adma.201502422. [161] M.A. LeRoux, F. Guilak, L.A. Setton, Compressive and shear
[138] D.W. Howell, C.W. Peak, K.J. Bayless, A.K. Gaharwar, 2D Nanosilicates Loaded properties of alginate gel: Effects of sodium ions and algi-
with Proangiogenic Factors Stimulate Endothelial Sprouting, Adv. Biosyst. 2 nate concentration, J. Biomed. Mater. Res. 47 (1) (1999) 46–53.
(7) (2018) 180 0 092, doi:10.10 02/adbi.20180 0 092. doi:10.1002/(SICI)1097-4636(199910)47:1<46::AID−JBM6>3.0.CO;2-N.
[139] G. Lokhande, J.K. Carrow, T. Thakur, J.R. Xavier, M. Parani, K.J. Bayless, A.K. Ga- [162] A.J.M. Segeren, J.V. Boskamp, M. Van Den Tempel, Rheological and swelling
harwar, Nanoengineered injectable hydrogels for wound healing application, properties of alginate gels, Faraday Discuss. Chem. Soc. 57 (1974) 255–262,
Acta Biomater. 70 (2018) 35–47, doi:10.1016/j.actbio.2018.01.045. doi:10.1039/DC9745700255.
[140] S. Hong, D. Sycks, H.F. a. Chan, S. Lin, G.P. Lopez, F. Guilak, K.W. Leong, [163] F. Guillemot, A. Souquet, S. Catros, B. Guillotin, J. Lopez, M. Faucon, B. Pip-
X. Zhao, 3D Printing: 3D Printing of Highly Stretchable and Tough Hydro- penger, R. Bareille, M. Rémy, S. Bellance, P. Chabassier, J.C. Fricain, J. Amédée,
gels into Complex, Cellularized Structures, Adv. Mater. 27 (27) (2015) 4034, High-throughput laser printing of cells and biomaterials for tissue engineer-
doi:10.1002/adma.201570182. ing, Acta Biomater. 6 (7) (2010) 2494–2500, doi:10.1016/j.actbio.2009.09.029.
[141] J.W. Nichol, S.T. Koshy, H. Bae, C.M. Hwang, S. Yamanlar, A. Khademhosseini, [164] G. Marino, F. Rosso, G. Cafiero, C. Tortora, M. Moraci, M. Barbarisi, A. Barbarisi,
Cell-laden microengineered gelatin methacrylate hydrogels, Biomaterials 31 β -Tricalcium phosphate 3D scaffold promote alone osteogenic differentiation
(21) (2010) 5536–5544, doi:10.1016/j.biomaterials.2010.03.064. of human adipose stem cells: in vitro study, J. Mater. Sci. Mater. Med. 21 (1)
[142] H. Aubin, J.W. Nichol, C.B. Hutson, H. Bae, A.L. Sieminski, D.M. Cropek, (2010) 353–363, doi:10.10 07/s10856-0 09-3840-z.
P. Akhyari, A. Khademhosseini, Directed 3D cell alignment and elongation [165] R.M. Hodges, N.S. MacDonald, R. Nusbaum, R. Stearns, F. Ezmirlian, P. Spain,
in microengineered hydrogels, Biomaterials 31 (27) (2010) 6941–6951, doi:10. C. McArthur, The strontium content of human bones, J. Biol. Chem. 185 (2)
1016/j.biomaterials.2010.05.056. (1950) 519–524.
[143] M. Nikkhah, N. Eshak, P. Zorlutuna, N. Annabi, M. Castello, K. Kim, [166] E. Bonnelye, A. Chabadel, F. Saltel, P. Jurdic, Dual effect of strontium ranelate:
A. Dolatshahi-pirouz, F. Edalat, H. Bae, Y. Yang, A. Khademhosseini, Directed stimulation of osteoblast differentiation and inhibition of osteoclast formation
endothelial cell morphogenesis in micropatterned gelatin methacrylate hy- and resorption in vitro, Bone 42 (1) (2008) 129–138, doi:10.1016/j.bone.2007.
drogels, Biomaterials 33 (35) (2012) 9009–9018, doi:10.1016/j.biomaterials. 08.043.
2012.08.068. [167] M.T. Vestermark, Strontium in the Bone-Implant Interface, Dan. Med. Bull. 58
[144] S. Lee, J.J. Moon, J.L. West, Three-dimensional micropatterning of bioactive (5) (2011) 1–30.
hydrogels via two-photon laser scanning photolithography for guided 3D [168] P.J. Marie, P. Ammann, G. Boivin, C. Rey, Mechanisms of action and therapeu-
cell migration, Biomaterials 29 (2008) 2962–2968, doi:10.1016/j.biomaterials. tic potential of strontium in bone, Calcif. Tissue Int. 69 (3) (2001) 121–129,
20 08.04.0 04. doi:10.10 07/s0 02230 010 055.
[145] D. Tadic, M. Epple, A thorough physicochemical characterisation of 14 calcium [169] Z. Saidak, P.J. Marie, Strontium signaling: molecular mechanisms and thera-
phosphate-based bone substitution materials in comparison to natural bone, peutic implications in osteoporosis, Pharmacol. Ther. 136 (2) (2012) 216–226,
Biomaterials 25 (6) (2004) 987–994, doi:10.1016/s0142- 9612(03)00621- 5. doi:10.1016/j.pharmthera.2012.07.009.
[146] N.E. Fedorovich, W. Schuurman, H.M. Wijnberg, H.-.J. Prins, P.R. van Weeren, [170] C. Wu, Y. Zhou, C. Lin, J. Chang, Y. Xiao, Strontium-containing mesoporous
J. Malda, J. Alblas, W.J.A. Dhert, Biofabrication of Osteochondral Tissue Equiva- bioactive glass scaffolds with improved osteogenic/cementogenic differenti-
lents by Printing Topologically Defined, Cell-Laden Hydrogel Scaffolds, Tissue ation of periodontal ligament cells for periodontal tissue engineering, Acta
Eng. Part C Methods 18 (1) (2012) 33–44, doi:10.1089/ten.tec.2011.0060. Biomater. 8 (10) (2012) 3805–3815, doi:10.1016/j.actbio.2012.06.023.
[147] H.R. Lin, Y.J. Yen, Porous alginate/hydroxyapatite composite scaffolds for [171] M. Frasnelli, F. Cristofaro, V.M. Sglavo, S. Dirè, E. Callone, R. Ceccato,
bone tissue engineering: preparation, characterization, and in vitro stud- G. Bruni, A.I. Cornaglia, L. Visai, Synthesis and characterization of strontium-
ies, J. Biomed. Mater. Res. - Part B Appl. Biomater. 71 (1) (2004) 52–65, substituted hydroxyapatite nanoparticles for bone regeneration, Mater. Sci.
doi:10.10 02/jbm.b.30 065. Eng. C 71 (2017) 653–662, doi:10.1016/j.msec.2016.10.047.
[148] S. Bose, S. Tarafder, Calcium phosphate ceramic systems in growth factor and [172] U. Thormann, S. Ray, U. Sommer, T. ElKhassawna, T. Rehling, M. Hundge-
drug delivery for bone tissue engineering: a review, Acta Biomater. 8 (4) burth, A. Henß, M. Rohnke, J. Janek, K.S. Lips, C. Heiss, G. Schlewitz, G. Sza-
(2012) 1401–1421, doi:10.1016/j.actbio.2011.11.017. lay, M. Schumacher, M. Gelinsky, R. Schnettler, V. Alt, Bone formation induced
[149] J. Cui, C. Ma, Z. Li, L. Wu, W. Wei, M. Chen, B. Peng, Z. Deng, Polydopamine- by strontium modified calcium phosphate cement in critical-size metaphyseal
functionalized polymer particles as templates for mineralization of hydroxya- fracture defects in ovariectomized rats, Biomaterials 34 (34) (2013) 8589–
patite: biomimetic and in vitro bioactivity, RSC Adv. 6 (8) (2016) 6747–6755, 8598, doi:10.1016/j.biomaterials.2013.07.036.
doi:10.1039/c5ra24821c. [173] M. Schumacher, A.S. Wagner, J. Kokesch-Himmelreich, A. Bernhardt,
[150] K. Rezwan, Q.Z. Chen, J.J. Blaker, A.R. Boccaccini, Biodegradable and bioac- M. Rohnke, S. Wenisch, M. Gelinsky, Strontium substitution in apatitic CaP
tive porous polymer/inorganic composite scaffolds for bone tissue engineer- cements effectively attenuates osteoclastic resorption but does not inhibit os-
ing, Biomaterials 27 (18) (2006) 3413–3431, doi:10.1016/j.biomaterials.2006. teoclastogenesis, Acta Biomater. 37 (2016) 184–194, doi:10.1016/j.actbio.2016.
01.039. 04.016.
[151] A.J. Engler, S. Sen, H.L. Sweeney, D.E. Discher, Matrix elasticity directs stem [174] L. Maïmoun, T.C. Brennan, I. Badoud, V. Dubois-Ferriere, R. Rizzoli, P. Am-
cell lineage specification, Cell 126 (4) (2006) 677–689, doi:10.1016/j.cell.2006. mann, Strontium ranelate improves implant osseointegration, Bone 46 (5)
06.044. (2010) 1436–1441, doi:10.1016/j.bone.2010.01.379.
[152] Y. He, Y. Dong, F. Cui, X. Chen, R. Lin, Ectopic osteogenesis and scaffold [175] S. Kargozar, M. Montazerian, E. Fiume, F. Baino, Multiple and promising ap-
biodegradation of nano-hydroxyapatite-Chitosan in a rat model, PLoS ONE 10 plications of strontium (Sr)-containing bioactive glasses in bone tissue engi-
(8) (2015) 1–15, doi:10.1371/journal.pone.0135366. neering, Front. Bioeng. Biotechnol. 7 (2019), doi:10.3389/fbioe.2019.00161.
22 S. Heid and A.R. Boccaccini / Acta Biomaterialia 113 (2020) 1–22

[176] E. Fiume, J. Barberi, E. Verné, F. Baino, Bioactive glasses: from parent 45S5 production, Biomaterials 25 (15) (2004) 2941–2948, doi:10.1016/j.
Composition to Scaffold-Assisted Tissue-Healing Therapies, J. Funct. Biomater. biomaterials.2003.09.086.
9 (1) (2018) 24, doi:10.3390/jfb9010024. [179] A. Abalymov, L. Van der Meeren, M. Saveleva, E. Prikhozhdenko, K. Dewet-
[177] C. Gao, S. Peng, P. Feng, C. Shuai, Bone biomaterials and interactions with tinck, B.V. Parakhonskiy, A.G. Skirtach, Cells-grab-on particles – a novel ap-
stem cells, Bone Res. 5 (2017) 17059, doi:10.1038/boneres.2017.59. proach to control cell focal adhesion on hybrid hydrogels, ACS Biomater. Sci.
[178] P. Valerio, M.M. Pereira, A.M. Goes, M.F. Leite, The effect of ionic products Eng. (2020), doi:10.1021/acsbiomaterials.0c00119.
from bioactive glass dissolution on osteoblast proliferation and collagen

You might also like