You are on page 1of 27

Materials

Horizons
View Article Online
REVIEW View Journal | View Issue

Advanced cell culture platforms: a growing quest


for emulating natural tissues
Cite this: Mater. Horiz., 2019,
6, 45
Marziye Mirbagheri,abc Vahid Adibnia,c Bethany R. Hughes,ab
Stephen D. Waldman,ab Xavier Banquyc and Dae Kun Hwang *ab
Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

In the body, cells inhabit within a complex three-dimensional (3D) extracellular matrix that provides
physical and chemical signals to regulate the cell fate. Cultured cells in Petri dishes and tissue culture
flasks (2D) receive completely different environmental cues compared to natural tissues, causing radical
alterations in cell morphology and function. Three-dimensional culture models have been able to
revolutionize biomedical applications by better emulating natural tissues. However, sample handling and
high-throughput screening can be challenging with 3D cell culture. Moreover, most 3D matrices are
unable to quantify intracellular mechanics due to their structurally undefined surface characteristics.
Therefore, highly structured surfaces (221 D) comprising various micro- and nano-patterns were
Received 9th July 2018, introduced to address these limitations. The topographical substrates have also been shown to retain
Accepted 1st November 2018 in vivo cell functionalities, such as proliferative capacity. Here, we review recent advancements in
DOI: 10.1039/c8mh00803e modulation of surface patterns that have been able to control cell adhesion in two or three dimensions,
and their impacts on the cell behavior. Finally, we provide a comparison between 2D, 221 D and 3D
rsc.li/materials-horizons systems and present several clinical applications of non-planar substrates.

a
Department of Chemical Engineering, Faculty of Engineering & Architectural Science, Ryerson University, Toronto, Ontario M5B 2K3, Canada.
E-mail: dkhwang@ryerson.ca
b
Keenan Research Center, Li Ki Shing Knowledge Institute, St. Michael’s Hospital, Toronto, Ontario M5B 1W8, Canada
c
Faculty of Pharmacy, Université de Montréal, C.P. 6128, Succursale Centre Ville, Montreal, Quebec H3C 3J7, Canada

Marziye Mirbagheri obtained her Vahid Adibnia received his PhD in


BS and PhD degrees in Chemical chemical engineering in summer
Engineering from the University 2017 from the Soft Matter &
of Tehran and McGill University, Colloids Laboratory at McGill
respectively. After completing her University. As a postdoctoral
PhD in 2017, as a postdoctoral researcher, he then joined the
fellow, she joined the Advanced Biomaterials and Structured
Material Research Group at Interfaces Laboratory at the
Ryerson University and Li Ka Faculty of Pharmacy of University
Shing Knowledge Institute at St. of Montreal. His research is focused
Michael’s Hospital. She is now on nanoparticle interactions with
continuing her postdoctoral polymers, interfaces and complex
Marziye Mirbagheri training at the Faculty of Vahid Adibnia viscoelastic media for drug delivery
Pharmacy, University of Montreal. purposes. He is also interested in
Dr Mirbagheri has received several awards and fellowships during her biomaterial engineering and tissue-mimetic polymeric structures. Dr
education and career. Her research interests mainly include Adibnia has received several scientific awards and fellowships
microfabrication and microfluidics, tissue engineering, organ-on-a- including the McGill Engineering Doctoral Award (2012), McGill
chip, and nanoparticle interactions with hydrogels and biological Graduate Excellence Fellowship (2015) and FRQNT postdoctoral
entities. fellowship (2018).

This journal is © The Royal Society of Chemistry 2019 Mater. Horiz., 2019, 6, 45--71 | 45
View Article Online

Review Materials Horizons

1 Introduction In their natural environment, nearly all cells reside within a


complex 3D fibrous meshwork, known as the extracellular
Cell culture experiments have been mainly performed on two- matrix (ECM).4,6 The 3D nanostructure of the ECM, which is
dimensional (2D) platforms, such as multi-well plates, Petri specific to each cell type, supports the cells and guides their
dishes, and tissue culture flasks, in which the cells are plated function.7 The loss of 3D cues, disorganized cell–cell inter-
onto a rigid planar surface. These conventional 2D substrates—with actions, high oxygen tension, and high growth factor concentrations
ease of use and high cell viability—have notably improved our in 2D cell culture can considerably alter the inter- and intra-cellular
understanding of basic cell function. However, they are notoriously behavior.8,9
unable to appropriately imitate the cellular microenvironment The critical limitations of 2D cell culture can lead to several
in vivo, leading to physiologically irrelevant cell behaviors in most practical complications. For example, in end-stage destructive
cases. Cells inherently respond to chemical (e.g., surface chemistry joint diseases, such as osteoporosis, osteoarthritis and bone
and material composition) and physical (e.g., mechanical and tumors, the bone quality and bone formation are decreased while
structural properties) signals from their surroundings at all length the bone resorption is increased. Current therapies involve the
scales,1–5 analyze the information and decide to undergo any of replacement of damaged tissue by an implant. The life span of
Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

the cellular fate processes (division, apoptosis, migration and/or these implants and, thus, the patients’ life quality greatly depend
differentiation). on the initial bone tissue response.7 Surface topography is a key
parameter in controlling this initial response10 by modulating the
interactions between tissue and the implant.11 Many researchers
Dr Waldman is a Professor of have already proved the beneficial effects of biomaterials that
Chemical Engineering at Ryerson mimic the bone surface roughness on osteoblast proliferation
University, an Affiliated Scientist and adhesion.12–18 The shortcomings of 2D planar platforms
at the Li Ka Shing Knowledge have been repeatedly demonstrated in other studies as well.19,20
Institute of St. Michael’s According to Lee et al.,20 the cytotoxicity testing of CdTe nano-
Hospital, and the Director of the particles on HepG2 cells is dramatically different between 2D and
Biomedical Engineering Graduate 3D cultures, which can be problematic in preclinical drug
Program (Ryerson University). He screening, where in vitro analyses determine the toxicity of a
was previously a Canada Research target drug.21 Therefore, 2D planar platforms not only fail to
Chair (2003–2013) and recently reproduce the 3D cellular microenvironment in the body, but
named a Fellow of International also can mislead our perception of cell responses.
Orthopaedic Research (2016). His To overcome these limitations, 3D cell culture methods
Stephen D. Waldman research interests are primarily including spheroids, microcarriers, and tissue-engineered models
centered on the development of were introduced.22–33 While microcarriers are mainly used for
tissue engineered cartilages (articular cartilage, auricular cartilage, cellular expansion in vitro, spheroids and 3D scaffold-based
and the intervertebral disc) with specific focus on the regulation of models have diverse applications in tissue engineering, cancer
mechanotransduction pathways, the effect of nutrient metabolism, research and fundamental studies of cellular function.34
and biomaterial-cellular interactions to guide tissue growth.

Xavier Banquy is an associate Dr Dae Kun Hwang, Associate


professor at the Faculty of Professor of Chemical Engineering
Pharmacy, Université de Montréal. at Ryerson University since 2011, is
He has been the Canada Research an applied materials scientist. He
Chair in Bioinspired Materials and is also a Canada Research Chair.
Surfaces since 2013. His research After obtaining his PhD degree in
interests are divided into three main Chemical Engineering from McGill
streams: biomaterials development University in 2006, Dr Hwang did
and characterization of their inter- his postdoctoral training at the
actions with host tissue; Massachusetts Institute of
development of multifunctional Technology (MIT). Dr Hwang has
nanoformulations for cancer over ten years of research and
Xavier Banquy treatment; fundamental research Dae Kun Hwang development experience in
in colloidal and surface science at advanced materials research using
the biointerface. microfluidics. His research focuses on the development of micro-
fabrication methods based on microfluidics, photo-lithography to
generate advanced functional materials of microparticles, membranes,
and wrinkled surfaces for biomedical applications.

46 | Mater. Horiz., 2019, 6, 45--71 This journal is © The Royal Society of Chemistry 2019
View Article Online

Materials Horizons Review

Spheroids are highly dense tissue analogs that are self- and cell transplantation.51 For example, in a recent breakthrough,
assembled due to the tendency of adherent cells to aggregate. European researchers were able to regenerate the entire human
Three-dimensional matrices are highly porous substrates that epidermis using transgenic stem cells,52 providing high hopes for
can support cell function by homogeneously surrounding the treatment of the so-far incurable Junctional Epidermolysis Bullosa
cell with extracellular materials.35 In tissue engineering, primary (JEB) disease. In this research, the keratinocyte cells were cultured
cells from patients adhere and proliferate on the surface of a 3D onto fibrin gels, allowing for preparation of larger grafts from the
scaffold, generating the ECM components of the living tissue. same number of clonogenic cells as would be needed for plastic-
Cellular behavior in 3D cultures is relatively closer to natural cultured grafts.52,53
tissue, specifically, the cell morphology. For example, human Such clinical studies are abundant and what they all have in
fibrochondrocytes have distinct 3D round or oval shapes in vivo common is the importance of cell interactions with the scaf-
and on 3D scaffolds, whereas they exhibit stretched, fibroblast- folding material, which can directly determine the cell response.
like morphologies in 2D culture flasks.36 Cell metabolism and In recent years, it has been extensively documented that
functionality, including proliferation, differentiation and gene surface topography in particular can greatly influence the cell
expression, are also significantly different between 2D and 3D behavior.42,54–56 In this section, first we review topographical
Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

models.19,34,36 This can be substantially important in clinical substrates that lead to planar cell morphologies in which cells
applications, such as mesenchymal stem cell (MSC) transplantation migrate in a 2D plane. Next, advanced 212D structures that
for treating osteogenesis imperfecta,37 where the loss of proliferative induce a crossover from 2D to 3D cell behavior are discussed.
capacity of MSCs in 2D culture can lead to poor engraftment and Here, in addition to physical patterning, surfaces may have been
limited cell survival upon transplantation into the body.38 Therefore, treated chemically to control cell adhesion.44,57,58 Since the
3D systems are mainly favorable to 2D structures for cell culture and focus of this review is on physical cues, the impact of chemical
tissue models. Nonetheless, several challenges remain. surface modifications is discussed only when found relevant.
With 3D cell culture, sample handling and imaging39 as well Moreover, throughout this article, the material types that are
as high-throughput screening35 can be challenging, demanding used to construct the topographical substrates or 3D tissue-
technological innovations to fully benefit from the third engineered models are specified, as they can be important
dimension.40,41 Furthermore, due to structurally undefined surface factors in modulating cell–material interactions.59,60
characteristics, most 3D scaffolds are unable to quantify intra-
cellular mechanics.9 In recent years, highly structured surfaces Physical topographies that induce planar cell morphology
composed of various patterns, also known as 212 dimensional Diverse techniques including photo-,61,62,240 soft,63,64 colloidal65
(212D) objects, have been developed to promote our understanding and electron beam lithography,66 laser patterning,67 and dry
of cell behavior. On these substrates, adhered cells conform to etching66,68 have been used to fabricate topographical surfaces
their specific surface patterns, whether simple, such as grooves, or using various materials, such as polymers, silicon oxide, and
complex, such as plant topographies.42 This can significantly metals.69–71 Surface topography is especially important for
improve drug testing and disease modeling, in which tissue synthetic polymers due to the lack of biological recognition on
models should adequately represent natural cell organizations to their surfaces.72
obtain physiologically relevant information. Schulte et al.73 demonstrated how topographical patterns
This review highlights recent developments in modulation on PEG can solely manipulate cellular behavior without bio-
of surface topography, and its impacts on cell behavior and functionalization. According to their study, cells on smooth
function. While most surface topographies induce 2D planar polymer surfaces showed no pronounced stress fibre network
cell morphologies, we also explain advanced 212D substrates (Fig. 1(a)) with completely round bodies (Fig. 1(b)) and an
that have been able to control cell adhesion and cell shape in inclination to form small clusters (Fig. 1(c)).
three dimensions. Recall that here any topographical pattern However, on patterned surfaces with posts, protrusions were
(e.g., groove, post, pit, and other complex designs) on a flat substrate formed (Fig. 1(d)), and the cell body flattened around the
is recognized as 212D. Moreover, we discuss the advantages and surface structure (Fig. 1(e)) with a noticeably reduced cluster
disadvantages of 212D substrates in comparison with 2D and 3D formation. Similarly, single cells on top of (Fig. 1(f)) or within
models. Finally, we introduce several contemporary applications of (Fig. 1(g)) the grooves showed enhanced spreading and a larger
various non-planar cell culture platforms. cell–surface contact area compared to smooth surfaces (Fig. 1(h)).
These observations indicated that fibroblast cells adhered and
2 Cell culture on structured surfaces spread on imprinted topographies, while the smooth surfaces
were nonadhesive.
Numerous soft and hard materials43–46 are utilized in biomedical The cell morphological and functional responses to surface
and healthcare industries for applications such as tissue patterns greatly depend on the cell type74 as well as the pattern
engineering,47 medical implants,48 and drug delivery.49,50 type and dimensions (several examples are summarized in
Particularly, many types of natural (e.g., collagen, alginate and Table 1). The effects of micro- and nano-grooves on different cells
fibrin) and synthetic polymers (e.g., acrylamide, poly(ethylene have been widely investigated.75–78 On these linear cues, cells
glycol) (PEG), poly(dimethyl siloxane) (PDMS), and poly(methyl usually align and migrate in the groove direction. However, the
methacrylate) (PMMA)) have been used for tissue engineering cell response is strongly governed by the groove width and depth.

This journal is © The Royal Society of Chemistry 2019 Mater. Horiz., 2019, 6, 45--71 | 47
View Article Online

Review Materials Horizons


Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

Fig. 1 Mouse fibroblast cell line (L929) cultured on starPEG hydrogels with smooth surface after 24 h (a–c), post patterns after 4 h (d and e), and line
patterns with groove widths of 5 mm after 48 h (f and g) and 10 mm after 24 h (h) of incubation. Immuno-staining of F-actin (a, d and h) and cell nucleus
(a and h), and electron micrographs of dried cells (b, c and e–g). Scale bars: 5 mm. Reproduced with permission.73

Experiments with fibroblast show that groove width smaller pillars with a height of B2 mm were close to those on a flat
than the cell size provokes a stronger cell response,73 while surface (Fig. 3(a)), while on taller pillars they adopted more
there is a cut-off value below which the cells no longer recognize elongated and branched shapes (Fig. 3(b)). Whereas cells
the patterns.77 For example, rat dermal fibroblasts obtained from migrated on top of the posts for small pillar spacings, they
the ventral skin of male Wistar rats and cultured on nanogrooves encountered alternating flat and rough surfaces for wider
experience a threshold feature size of 35 nm.77 In contrast, cell spacings (Fig. 3(c)). Furthermore, cell migration patterns were
alignment has been shown to be proportional to groove depth changed from a fast and random movement on flat surfaces to
(Fig. 2), with an upper threshold above which no significant a slower and persistent movement on post arrays, allowing
improvements have been observed.62,73 For example, for MDCK their translocation from pillar to pillar. Compared to flat
cell lines cultured on groove structures with groove width surfaces, fewer actin stress fibers and focal adhesions were
B12 mm, the upper threshold for groove depth is reported to observed, and they appeared more aggregated and more stable
be B1.1 mm.62 The groove depth has also been reported to be over time on the pillars. This guidance of focal adhesion
more influential than its width.62,77 formation was also observed in other studies.84,85 More information
On a certain 212D substrate, the cell response may also vary on focal adhesions and their role in transmitting extracellular
with time,77 emphasizing the need for sufficiently prolonged physical or topographic signals can be found in ref. 85. A
culture time to establish the long-term cell behavior in medical potential application of these platforms is for cellular phenotype
applications, such as implants. Guided cell alignment can be discrimination, since various cell types react differently to pillar
crucial for disease modeling and drug testing, where cell topographies.
disruption due to a certain illness is analyzed to identify Studies of cellular mechanotransduction has also greatly
potential drugs. Here, the model tissue should appropriately benefited from post patterns. Analyzing the cell responses to
represent the natural cell organization and function. In many physical cues suggests that when cells adhere to a surface, they
natural tissues, cells are elongated and structurally aligned.79,80 sense their environment by mechanically probing it.88 Adher-
More clinical applications of these platforms are discussed in ent cells exert contractile forces, also known as traction forces,
Section 4. on the substrate using their actin–myosin cytoskeleton to
Another topographical feature that has been used to inves- propel themselves.89–91 Traction forces regulate cell’s motility
tigate cellular behavior is micropillar or micropost. Similar to and morphology,91 thus playing a crucial role in cell signaling,
grooves, the pillar size and spacing influence cell organization. proliferation, differentiation, migration and other biological
Ghibaudo et al.81 studied fibroblast migration onto chemically processes, such as angiogenesis, embryogenesis, inflammation
identical substrates composed of micropillars. In this study, the and wound healing. Specifically, assessing the contractility of human
surfaces were uniformly coated with fibronectin to promote cell stem cell-derived cardiomyocytes83 can help devise better strategies
adhesion, and the pillars were too stiff to be significantly for heart repair,92 disease modeling,93 and drug discovery.94 One of
deformed by cell traction forces.82,83 Cell morphologies on the most important functional characteristics of a cardiomyocyte is

48 | Mater. Horiz., 2019, 6, 45--71 This journal is © The Royal Society of Chemistry 2019
Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

Table 1 Examples of topographical patterns on planar surfaces and their impacts on various cell types

Surface Culture
Pattern Dimension (mm) Method Materials treatment Cell type time Effect Ref.
Grooves 5 o width o 10 Replica molding PEG — Mouse fibroblast 24 h Enhanced adhesion, spreading, and cell 73
Depth: 5 alignment and reduced cluster
Materials Horizons

formation.

Grooves 20 o width Replica molding PEG — Mouse fibroblast 24 h Cells reacted similar to those on a 73
Depth: 5 smooth surface.

Grooves 0.1 o width o1 Replica molding Polystyrene — Rat dermal fibroblast 4 (24) h Enhanced cell alignment. 77
0.075 (0.035)
o depth o0.35

Grooves Width: 6 Photolithography and PMMA — Baby Hamster Kidney — Poor cell alignment. 62
Depth: 0.56 dry etching fibroblast (aka BHK)

This journal is © The Royal Society of Chemistry 2019


Grooves Width: 6 Photolithography and PMMA — MDCK (epithelial cell) — B100% cell alignment. 62
Depth: 0.56 dry etching

Grooves Width: 6 Photolithography and PMMA — Baby Hamster Kidney — B100% cell alignment. 62
Depth: 2 dry etching fibroblast (aka BHK)

Grooves Width: 8 Hot-embossing imprint Polyimide Fibronectin Osteoblast — Cells aligned according to the 76
Depth: 4 lithography mechanical topography rather than
the chemical patterns.

Micropillars Height: 10 Photolithography and PDMS Fibronectin 3T3 fibroblast 6–24 h Cells on top of the posts had more 81
(undeformable) Diameter: 5 replica molding elongated and branched shapes, fewer
Spacing: 5 focal adhesions and slower movements.

Micropillars Height: 0.8 Photolithography and PDMS Fibronectin Malignant fibroblast 1–9 h As oppose to healthy cells, fewer cancer 111
(undeformable) Diameter: 1 replica molding cells were elongated and they exhibited a
Spacing: 1.6 disorganized motility.

Square micropillars Height: 4.5–14 Replica molding PLLA, TCPS, — C2C12 muscle cell 1–4 days Improved initial cell attachment but 112
(undeformable) Length: 5 PEOT-PBT, and reduced proliferation for hydrophobic
Spacing: 2–26 PDMSb materials (PDMS). However, better pro-
liferation on hydrophilic materials.

Nanopillars Diameter: 0.2 Nanoimprint UV resin — Human osteoblast 3 days Long filopodia extensions and 113
(undeformable) Spacing: 0.7 lithography lamellipodia formation.

Sharp-tip nanoposts Height: 0.05–0.6 Interface lithography Silicon — Human foreskin 3 days Elongated cells on medium height 97
Spacing: 0.23 and DRIE fibroblast features, but lower cell viability and
proliferation than on smooth surfaces,
which became more pronounced with
height. The cells also became smaller
and rounder with height.

Sharp-tip Height: 0.05–0.6 Interface lithography Silicon — Human foreskin 3 days Lower cell viability and proliferation 97
nanogrates Width: 0.23 and DRIE fibroblast than on smooth surfaces. Elongated cells
in the grate direction, with enhanced cell
alignment for taller grates.
View Article Online

Review

Mater. Horiz., 2019, 6, 45--71 | 49


Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

Review

Table 1 (continued)

Surface Culture
Pattern Dimension (mm) Method Materials treatment Cell type time Effect Ref.

50 | Mater. Horiz., 2019, 6, 45--71


Nanogrates Height: 0–0.35 Nanoimprint photo- Poly(styrene), — Murine preosteoblast 24 h Physical topography induced higher cell 114
Pitch: 0.42 and 0.8 imprint and PMMA and a alignment than surface chemistry. Cell
lithography cross-linked alignment increased with grate height
dimethacrylate and trough width.

Hexagonal and Diameter: 0.12 Electron beam PMMA — Human osteopro- 21 days Enhanced osteogenesis compared to 98
disordered nanopit Depth: 0.1 lithography genitor and MSC planar surface, specifically on DSQ50
arrays Spacing: 0.3 arrays.a

Square array of Diameter: 0.12 Electron beam litho- PCL — Human MSC 28 days Prolonged maintenance of MSCs 99
nanopits Depth: 0.1 graphy and hot phenotype and multipotency, due to a
Spacing: 0.3 embossing reduced osteogenic differentiation but
improved MSC markers retention.

Orthogonal and Diameter: 0.035, Nanoimprinting and PMMA and PCL — Human fibroblast and — Reduced cell adhesion on ordered arrays 100
hexagonal array 0.075 and 0.12 hot embossing rat epitenon compared to planar surfaces.
of nanopits Spacing: 0.1, 0.2 and 0.3

Micropits Diameter: 7, 15 and 25 Photolithography Quartz — Human fibroblast 24 h Slightly improved cell proliferation, 103
Spacing: 20 and 40 and higher cell motility rates on 7 mm
Depth: 4.8 diameter pits. Cells can enter, divide in
and exit 25 mm diameter pits.

Square micropits Length: 2 Photolithography Quartz and Collagen on Neutrophil — Cell adhesion was stronger on quartz 104
Depth: 0.21 polyimide quartz than polyimide or collagen-coated
Spacing: 6–14 quartz. Cell motility was higher on
collagen-coated-quartz smooth surfaces.
Micropits with 10 mm spacing also
improved the motility.

Membranes with Diameter: 0.1–3 — Polycarbonate — Corneal epithelial cell 21 days Superior stratification on pores with 115
micro- and diameters in the range of 0.1–0.8 mm.
nano-sized
columnar pores
a b
Square arrays with dots displaced randomly by up to 50 nm on x and y axes from their position in a true square. Abbreviations are poly(L-lactic acid) (PLLA), tissue culture polystyrene (TCPS),
and a co-polymer of poly(ethylene oxide) and poly(butylene terephtalate) (PEOT/PBT).
View Article Online

This journal is © The Royal Society of Chemistry 2019


Materials Horizons
View Article Online

Materials Horizons Review

its ability to produce contractile forces, as mechanical contraction


plays a key role in blood circulation throughout the body. For
example, measuring cardiomyocyte contractility can be of interest
in studying cardiac diseases associated with cardiomyopathies
(impaired contractility).
Microfabricated arrays of pillars have been extensively used
to manipulate and measure the mechanical interactions between
cells and their environment.82,83,95,96 Using silicon nanowires, Li
Fig. 2 Scanning electron micrographs (SEM) of BHK cells on PMMA et al.86 compared the mechanical behavior of normal, benign and
microgrooves with 6 mm width and 300 nm (a) or 2 mm (b) depth. Cell malignant mammalian cells (Fig. 3(d)–(f)). According to their
alignment improves with deeper grooves. Scale bars: 120 mm. Reproduced results, cancer cells exhibit a larger traction force than normal
with permission.62 cells by 20 and 50% for the malignant and benign cells,
Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

Fig. 3 (top) SEM images of fibroblast cell response to arrays of undeformable micropillars with height H, diameter D and spacing S. To promote
adhesion, the entire surfaces were coated with fibronectin: (a) H = 2 mm, D = 5 mm, and S = 5 mm; (b) H = 10 mm, D = 5 mm, and S = 5 mm; (c) H = 10 mm,
D = 10 mm, and S = 10 mm.81 (d) Schematic representation of cell adhesion to silicon nanowire arrays and the traction forces generated by cell’s actin
cytoskeleton. SEM images of normal mechanocytes (e) and cancerous L929 (f) cells bending silicon nanowires through their contraction forces. Scale
bars: 2 mm.86 (g) Graphical depiction of a finite-element method analysis of micropost with heights (L) deflected in response to a lateral traction force (F)
of 20 nN. (h) SEM micrographs of human MSCs plated onto PDMS micropost arrays with L = 6.10 (left) and 12.9 mm (right). For this study, only the post tips
were coated with fibronectin. Scale bars: 100 mm.87 Reproduced with permission.

This journal is © The Royal Society of Chemistry 2019 Mater. Horiz., 2019, 6, 45--71 | 51
View Article Online

Review Materials Horizons

respectively. These comparative techniques can be beneficial the square configuration (Fig. 5(a)–(c) and (f)–(h)). On the other
for disease diagnosis. hand, increasing the randomness of nanopit arrangement
Intuitively, taller posts are more susceptible to bending in resulted in denser cell populations with increased levels of
response to a horizontal traction force. Experimentally and OPN and OCN, specifically on disordered square arrays with
numerically, Fu et al.87 demonstrated the relationship between dots displaced randomly by up to 50 nm on x and y axes from
post height and deflection due to traction forces exerted by their position in a true square (DSQ50), as shown in Fig. 5(d),
human MSCs, and its consequent impact on cell behavior (e), (i) and (j).
(Fig. 3(g) and (h)). In contrast to soft (tall) microposts, cells Similarly, MSCs cultured on DSQ50 exhibited intense cell
were well spread with highly organized actin stress fibers and aggregation and early bone nodule formation with both OPN-
large focal adhesions on rigid (short) posts. Therefore, they and OCN-positive regions. In contrast to osteoprogenitors, OPN
established a coupling between cell shape, focal adhesion and and OCN expressions by MSCs were lost on completely random
cytoskeletal tension in response to substrate rigidity. arrays (similar to the far-right, top panel in Fig. 5).
Pillar arrays in these studies had relatively flat tips. Choi Later, McMurray et al.99 demonstrated how culturing cells
et al.97 investigated the behavior of human foreskin fibroblasts on orthogonal arrays of nanopits can allow a long-term maintenance
Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

on sharp-tip nanoposts and nanogrates with various heights. of MSC phenotype and multipotency, in contrast to cells cultured on
They observed a significantly lower cell viability and proliferation displaced arrangements such as DSQ50. In this study, cells on
on sharp-tip nanostructures compared to the smooth planar symmetric orthogonal arrays retained expression of MSC markers,
surfaces, and this effect was more pronounced on nanoposts i.e., STRO-1 and ALCAM (activated leukocyte cell adhesion molecule,
versus nanogrates and with increasing height of both patterns. CD166), while no osteogenic markers (OPN and OCN) were
Whereas cells elongated on medium height nanoposts, they observed. Therefore, following a prolonged culture (28 days), MSCs
became smaller and rounder on taller posts, indicating poor cell could be removed from the square pit arrays using trypsin, plated
adhesion (Fig. 4(a)–(d)). Grate structures also provoked elongated onto a coverslip and treated with differentiation media to promote
cell morphology in the grate direction; however, in contrast to osteogenesis or adipogenesis, indicating that cells not only expressed
posts, cell alignment and elongation were more pronounced on MSC markers but also were multipotent. This again highlights the
taller features (Fig. 4(e)–(h)). This is in agreement with previous impact of surface topography on differentiative capacity of MSCs
observations of enhanced cell alignment with groove depth (ridge with crucial importance in cell transplantation therapies, where
height).62,73 tissue regeneration upon reinfusion depends on MSC differentiation
The last type of ordered surface topography is nano- or into tissue-specific cell types.
micro-pit. Dalby et al.98 studied the interaction of two cell In another study, Curtis et al.100 reported that ordered
types, osteoprogenitors and MSCs, when cultured on nanopits nanopit arrays (orthogonal or hexagonal) in polycaprolactone
of different symmetry, including orthogonal and hexagonal (PCL) or polymethylmethacrylate (PMMA) can considerably
arrays. Compared to cell culture on planar surfaces, human decrease cell adhesion compared to planar surfaces. As illustrated
osteoprogenitor cell density decreased, especially on hexagonal in Fig. 6, cell adhesion on PCL surfaces with pit patterns was
arrays. Measuring the expression of bone-specific ECM proteins very minimal, except for the smallest pits with diameter 35 nm.
by osteoprogenitors, namely osteopontin (OPN) and osteocalcin Perhaps on these small patterns, cells ability to distinguish the
(OCN), a slight increase in production of these proteins was holes diminishes. A similar conclusion was drawn in previous
observed on symmetric pit arrays, with a larger difference for studies.101,102 Moreover, comparing the cells’ orientation on

Fig. 4 SEM micrographs of fibroblast cells cultured on smooth surfaces (a: control for posts, and e: control for grates), needle-like nanoposts (height
increasing from b to d), and blade-like nanogrates (height increasing from f to h). Insets show higher magnification of the sample’s nanotopography, and
arrows (2) in panels (f–h) indicate the nanograte directions. Scale bars: 50 mm. Reproduced with permission.97

52 | Mater. Horiz., 2019, 6, 45--71 This journal is © The Royal Society of Chemistry 2019
View Article Online

Materials Horizons Review


Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

Fig. 5 OPN and OCN staining of osteoprogenitors after 21 days of culture. (top) Images of nanopits with diameter 120 nm, depth 100 nm, and absolute
or average center to center spacing 300 nm in hexagonal, square, displaced (DSQ50) and random configurations (respectively, from left to right).
Osteoprogenitors cultured on the control planar surface with lack of positive OPN and OCN (a and f); hexagonal array with poor cell adhesion (b and g);
square configuration with reduced cell density but slightly increased positive OPN and OCN compared to the control (c and h); DSQ50 showing bone
nodule formation and enhanced OPN and OCN expression (d and i); random arrangement with good cell population and improved OPN and OCN
production (e and j). Actin = red and OPN/OCN = green. Scale bar, as shown in panel b, is the same for panels a–j. Reproduced with permission.98

square or hexagonal pit arrays, they observed that not only were It is widely established that cell response can significantly
the orientations nonrandom but they were also influenced by vary with length scale. While the aforementioned studies
different configurations, thus suggesting that cells are able to analyzed cell behavior on nanopits, Berry et al.103 investigated
distinguish between different symmetries. fibroblast attachment and motility on micropits. They reported
a clear dependence of cell response on pit size and spacing. Cell
proliferation slightly increased only on the smallest pits with
diameter 7 mm (a maximum difference of 6%), with even higher
proliferation on closer pit spacing. The largest pit diameters
(25 mm) allowed the cells to enter the pits while sustaining their
viability. Their results indicated that not only were cells able
to enter these pits, but they could also freely exit them, in
contrast to earlier observations for macrophages that were
trapped inside such pits.2 There was even evidence of cells
actually dividing inside the 25 mm pits prior to the daughter
cell emerging. On the smallest pits, however, a majority
of cells moved over the pits regardless of the spacing.
Furthermore, cells on the smallest pits moved faster, perhaps
by using the pit edges as footholds to gain mechanical
adhesion.104
A possible application of pit patterns is to establish appro-
priate pore sizes for 3D scaffolds in tissue engineering that will
enhance cell penetration (ingrowth)105 whilst maintaining
cell viability and proliferation. Microwells, if classified as pit
patterns with large pit diameters such that cells can colonize
inside the pits, have also been explored for high-throughput
studies of cancer progression and cancer drug discovery106–109
as well as for tissue repair and regeneration applications.110
A number of irregular topographies have also been the
Fig. 6 Rat epitenon cells cultured (24 h) on PCL substrates composed of
subject of several studies. Among the most popular ones is
flat surfaces and pit patterns. In top, left corner of each image, diameter:center-
to-center spacing of nanopits is indicated. Cells are shown in black. Cell
the surface roughness, which is usually quantified by measuring
adhesion on pit arrays is considerably lower than that on flat surfaces, with the protrusions and depressions on a surface.55,116–120 A myriad
relatively larger adhesion for smaller pits. Reproduced with permission.100 of evidence show that surface roughness can play an important

This journal is © The Royal Society of Chemistry 2019 Mater. Horiz., 2019, 6, 45--71 | 53
View Article Online

Review Materials Horizons

role in cell function, for example increasing cell adhesion and


growth.121 Informative reviews are available on this subject.9,44
In an interesting work by Unadkat et al.,122 mathematical
algorithms were used to design a library of nonbiased and
random surface features on poly(lactic acid) with more than
2000 different topographies. As demonstrated in Fig. 7, a
multitude of MSC cell morphologies, including elongated,
branched, and round, was obtained. Using high-throughput
screening, they illustrated MSC proliferation and osteogenic
differentiation on the surface patterns.
The 212D platforms reviewed in this section induce different Fig. 8 SEM images of human corneal epithelial cells cultured on (a) nano-
cell morphologies, and furnish valuable information about cell grooved (width: 400 nm, depth: 600 nm, and ridge width: 70 nm) and (b) flat
metabolism and function compared to the oversimplified 2D control silicon oxide substrates,123 and (c) fibroblasts on a micropit-patterned
quartz surface (diameter: 25 mm and spacing: 40 mm).103 Reproduced with
planar surfaces (see Table 1 for more examples). Nonetheless, permission.
Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

cells on these substrates generally lay flat and migrate in a 2D


plane similarly to those on a smooth surface. Fig. 8(a) and (b)
demonstrates an example of such behavior by comparing the Advanced topographical substrates to induce 3D cell behavior
human corneal epithelial cell shapes on a flat control and a Various physical and chemical treatments have been employed
nanogrooved surface. The only exception is for micropits that to create 212D scaffolds with more resemblance to topographical
allow a certain level of 3D migration when cells enter the pits103 complexity in natural ECM,124 and to manipulate cell adhesion
(Fig. 8(c)). and shape in 3 dimensions.68,125,126
Devices that can capture more of the 3D complexity in Using direct laser writing (DLW), Klein et al.127 fabricated a
natural tissues are more desirable for understanding the cell substrate consisting of pillars that were connected by beams at
behavior in vivo. In the following, we will discuss novel 212D two different heights by 10 mm offset (Fig. 9(a)).
platforms that have been able to induce cell movement and Poly(ethylene glycol) diacrylate (PEG-DA) crosslinked with
morphology in 3 dimensions. pentaerythritol tetraacrylate (PETA) was selected for the main

Fig. 7 Fluorescent microscopic images of spread and elongated MSCs, with alignments according to various topographical features (a–d). Actin stained
with Alexa Fluor 488 phalloidin = green, and nucleus stained with TOTO-3 = red. SEM images of cell morphologies on different topographies (e–h), and
round cells on two distinct platforms exhibiting different cell membrane textures (i and j). Scale bars: 90 mm (a–h) and 10 mm (i and j). Reproduced with
permission.122

54 | Mater. Horiz., 2019, 6, 45--71 This journal is © The Royal Society of Chemistry 2019
View Article Online

Materials Horizons Review

In the aforementioned studies, an ECM protein (i.e., fibronectin)


was used to control binding sites and, therefore, manipulate cell
adhesion to induce 3D morphologies. Using a similar strategy,
Ochesner et al.68 were able to provide a 3D shape control of single
cells in microwell arrays. In their study, cell adhesion was limited
within the microwells by passivation of the upper flat surface, while
backfilling the wells with adhesive proteins or lipid bilayers.
Another creative use of adhesive proteins for cell adhesion
was demonstrated by Richter et al.,130 who produced 212D cell
instructive platforms by incorporating two distinct ECM binding
proteins in a passivating framework. They illustrated that cells
with different sets of integrin receptors preferentially adhered to
one of the proteins. However, since the binding sites were in the
same height, cells spread in a 2D plane. Positioning the distinct
Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

binding proteins at different levels can possibly be a strategic


approach to obtain 3D cell shapes.
In a recent work by our group,131 surface wrinkling was
implemented as a physical technique to precisely distribute
adhesion sites and induce 3D cell morphologies. Using photo-
lithography and microfluidics, arrays of PEG-DA posts were
Fig. 9 (a) SEM image of a 221 D polymer scaffold composed of a protein- created. By controlling the UV exposure time, a thin layer of
repellent PEG-DA framework, which has embedded photoresist Ormo- partially cured polymer was formed on the posts, which wrinkled
comp cubes on the PEG-DA beams (one cube is highlighted in red). when exposed to plasma. The wrinkling mechanism is extensively
Primary chicken fibroblasts cultivated on the scaffold are immunostained
explained by Kim et al.132 Whereas fibroblast cells adhered to the
for fibronectin (red) and f-actin (green). Top view (b) and 3D construction
(c and d) of confocal image stacks illustrating adherent cells to one or flat surface when the posts were smooth, they preferentially
more fibronectin-positive Ormocomp cubes at the same or different conformed to the wrinkled posts exhibiting 3D morphologies
heights. Reproduced with permission.127 (Fig. 11(a) and (b)). It has also been shown in previous studies
that cells prefer the flat surface in arrays of smooth pillars.81
Interestingly, some cells on adjacent wrinkled posts appeared to
skeleton of the scaffold, as PEG-DA is known for its protein- form bridges, as shown in Fig. 12(c).
repellent properties that prevent cell attachment. To control In contrast to 212D platforms reviewed in the previous section,
cell adhesion and ECM distribution, protein-binding Ormo- studies on advanced surface topographies that are capable of
comp cubes were embedded on PEG-DA beams. Ormocomp is a producing 3D cell geometries have just begun in the past decade.
biocompatible photoresist. Upon incubation of these scaffolds Nonetheless, they have provided exciting opportunities to elucidate
with fibronectin, this ECM protein bound preferentially to the the cell metabolism and function in more detail, and help
Ormocomp cubes, as shown by immunofluorescence labelling advance engineered microenvironments to better emulate in vivo
in Fig. 9. Therefore, chicken fibroblasts cultured on these conditions.
platforms adhered and spread on a single or multiple Ormo-
comp cubes, particularly revealing a true 3D growth pattern
when spreading between cubes on beams at different heights 3 A comparison with three-dimensional
(Fig. 9(c) and (d)). cell culture systems
Greiner et al.128 studied the growth of several cell types on
fibronectin-coated 212D platforms fabricated also using DLW In this section, we first briefly introduce different 3D cell
(Fig. 10 (left)). As illustrated in Fig. 10, the cells adapted true 3D culture techniques and explain their differences with 2D mono-
geometries on these platforms. It was also demonstrated that layer models. Next, the advantages and disadvantages of 3D cultures
cell and nuclear volumes may change between 2D and 3D cell are weighed against 212D platforms for human bone-marrow-derived
morphology, depending on the cell type. Whereas epithelial MSCs, as one of the most investigated cell types using these
cells retained similar cell and nuclear volumes regardless of the techniques in the literature.
cell morphology, fibroblasts displayed significantly increased
cell and nuclear volumes for the 3D cell shapes compared to the 3.1 3D vs. 2D (planar surfaces)
2D. They attributed these differences to the tissue-specific cell Three-dimensional ex vivo cell culture models can be categorized
arrangements in vivo, where fibroblasts reside within loose as spheroids or microtissues, microcarriers, and tissue-engineered
connective tissues and epithelial cells within densely-packed models. Microcarriers are mainly used to expand anchorage-
flat layers. Nevertheless, the nucleus-to-cell (N/C) volume ratios dependent cells in vitro, providing several advantages compared
remained constant for all cell types and culture conditions. N/C to 2D models such as enhancing in vitro differentiation30 with high
ratio alteration is a widely used metric in cancer detection.129 reproducibility and scale-up potential.31 However, due to their

This journal is © The Royal Society of Chemistry 2019 Mater. Horiz., 2019, 6, 45--71 | 55
View Article Online

Review Materials Horizons


Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

Fig. 10 Buffalo rat liver cell growth on a 2D control surface (a), and a stiff (b) and a soft (c) 221 D scaffold, which were all coated with fibronectin to control
cell adhesion. Left panels indicate the SEM images of culturing structures, the middle panels present the maximum projections of confocal image stacks,
and the right panels are 3D constructions of confocal images. F-Actin = green and nucleus = blue. Reproduced with permission.128

The term spheroid is more commonly used because most tissue


aggregates are spherical, whereas a more general term, namely
multi-cellular microtissues or simply microtissues, describes
other shapes as well.133,134 Several single culture or co-culture
techniques, such as pellet culture, spinner culture, hanging
drop, rotating wall vessel (RWV) and microfluidics, have been
used to create microtissues.32,33 Regardless of the fabrication
method, spheroids with a high cell density create more in vivo-
like microenvironments by forming complex cell–cell and cell–
ECM interactions in addition to establishing diffusion barriers
to soluble components (e.g., nutrients, oxygen, growth factors
and wastes). These biomimetic characteristics are particularly
crucial for cancer drug discovery applications. The molecular
gradient of nutrients and oxygen can cause necrosis and
hypoxia toward the center of the spheroids, and the enhanced
Fig. 11 SEM images of bovine fibroblast cell growth on smooth (a) and intercellular communications can regulate cell function and
wrinkled (b and c) posts. Comparing (a and b), cells preferred the flat surface fate.32 Similarly, solid tumors often show hypoxia/necrosis,
when posts were smooth, whereas with the wrinkled posts they attached to and slow proliferation and drug diffusion barriers that contribute
spread on the posts, conforming their 3D curvature. Some cells even formed
bridges between the wrinkled posts, creating an overhanging connection
to cancer drug resistance.135 These features are not properly
between them (c). Scale bars: 30 mm. Reproduced with permission.131 reflected in monolayer cell cultures in which cells proliferate
rapidly, show high drug activity and have equal access to high
concentrations of nutrients and oxygen.
relatively less diverse applications, we will mainly focus on spheroids Several other studies have demonstrated the benefits of
and 3D tissue-engineered models in this review. spheroids over 2D monolayers.136–138 For example, as illustrated
Cellular spheroids are simple 3D tissue analogs that are self- in Fig. 12, hepatocyte spheroids exhibit liver-like functionalities
assembled due to the tendency of adherent cells to aggregate. with improved differentiation and reduced proliferation in

56 | Mater. Horiz., 2019, 6, 45--71 This journal is © The Royal Society of Chemistry 2019
View Article Online

Materials Horizons Review


Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

Fig. 12 Histological examination of human liver and HepG2 spheroids. Scale bar: 100 mm. Reproduced with permission.139

contrast to 2D monolayers of hepatocytes, which rapidly used to grow murine neural progenitor cells. The self-assembly
dedifferentiate and lose critical hepatocyte functions.139 was triggered by injection of cell culture medium, causing the
Nonetheless, various challenges in spheroid culture procedures peptides to aggregate into nanofibers with high aspect ratios,
still remain. Conventional methods for spheroid production, such thereby forming a solid hydrogel that mechanically supported
as spinner cultures, hanging drop, and rotating systems, can be the medium (Fig. 13(a)). The molecular design of the scaffolds
labor-intensive, time-consuming, and are unable to maintain promoted neurite sprouting and direct neurite growth, inducing
spheroid size uniformity on large scales,140,141 whereas more very rapid cell differentiation into neurons.
sophisticated platforms142 can be more complicated and Alginate-based scaffolds have been used to promote directed
require specially trained users for fabrication and operation. axon regeneration in vitro and in adult rat spinal cord lesions
Table 2 provides a list of advantages and disadvantages of in vivo.155 The hydrogel microstructure contained self-assembled
spheroids compared to 2D models. anisotropic capillaries that were formed during the gel synthesis
In 3D tissue-engineered models, cells are grown in 3D matrices (Fig. 13(b)). After seeding with adult neural progenitor cells,
or scaffolds that are prepared ex vivo using natural and/or synthetic which are known to improve axon regeneration, these gels were
materials. Generally, it is preferable to use the materials extracted implanted into acute cervical spinal cord lesions in rats and
from living tissue ECM that are specific to the cells under study. integrated into the spinal cord parenchyma without inflammatory
Commercially available gelatinous protein mixtures, such as responses. In accordance with in vitro findings, the anisotropic
Matrigels, and decellularized tissues used for treating cardio- structure of these implants induced directed axon regeneration
vascular diseases165 and diabetes166 as well as for reconstruction across the artificial scaffold, as shown in Fig. 13(b)(V) and (VI).
medicine167 are practical examples of this method. However, Moreover, the combined cell- and artificial scaffold-based therapy
such matrices are often difficult to customize and may vary from further improved neuron restoration. The alginate-based hydrogels
batch to batch.158 Alternatively, attention has been paid to were presented as a promising strategy for nerve regrowth
synthetic polymers that can form 3D networks with microscopic following spinal cord injury, when the targeted neuron rein-
pores. Polymer selection depends on cell adhesion, biocompatibility, nervation depends on longitudinally directed regrowth of
wettability and structural properties such as porosity, pore geometry, transected axons.
transport and degradation kinetics. Detailed reviews on artificial 3D Tumor microenvironments are temporally and spatially
cell culture scaffolds and their characteristics can be found heterogeneous, and it is unclear how the gradations in biophysical
elsewhere.4,39,158,168–174 Spatial organization of cells in the media cues can impact malignant phenotype and response to therapy.
directly depends on their microstructural geometry. Therefore, Three dimensional platforms capable of mimicking this complexity
perfecting the material structure to mimic the tissue is a key step can provide valuable information about tumor etiology, growth and
in regulating cell growth and promoting the functionality of specific spreading. In an attempt to replicate the heterogeneity of tumor
cell types to produce targeted outcomes that are unachievable in 2D ECMs, Pedron et al.156 generated 3D hydrogels with controlled
planar models. spatial gradient of matrix and cellular content (Fig. 13(c)).
An example of artificial 3D scaffolds that direct cell pro- Investigating glioblastoma multiforme (GBM), as a common
liferation and differentiation is a 3D self-assembled network of form of brain tumor in adults, they demonstrated how gradients
peptide amphiphiles introduced by Silva et al.,175 which was in cell density can lead to heterogeneous gene expression.

This journal is © The Royal Society of Chemistry 2019 Mater. Horiz., 2019, 6, 45--71 | 57
View Article Online

Review Materials Horizons

Table 2 Key differences in cellular characteristics and processes between 2D and 3D cell culture models

Cellular
characteristics 2D monolayer 3D cell spheroids 3D tissue-engineered models Ref.
Morphology Flat and stretched Aggregates of natural 3D shapes at Can assume stellate, round or elongated 135, 139,
monolayer the interior of spheroid morphologies. Regardless of the shape they 143 and
are homogeneously surrounded by the ECM 144

Adhesion Cells adhere to the Cell-to-cell adhesion is dominant Both cell–cell and cell–ECM adhesion can be 144 and
supporting surface observed 145

Viability Cell viability is acceptable Cell viability is improved or is 3D cultures can significantly increase cell 32, 136,
similar to 2D viability 137, 139,
146 and
147

Stage of cell cycle Cells are more likely to be May contain proliferating, — 139
at the same cycle due to quiescent, hypoxic and necrotic
Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

being evenly exposed to the cells. Larger spheroids show more


medium necrosis in the center due to
hypoxia

Proliferation Often proliferate faster Lower or comparable proliferation Lower, comparable or higher proliferation 139, 143,
than in vivo rate compared to 2D rate compared to 2D 148–150

Differentiation Often reduces the Can enhance the differentiation Can preserve the differentiative capacities 34, 136,
differentiative capacities potential of multipotent cells and provide a more reliable understanding 151–153
of stem cells of differentiation mechanisms

Gene and protein Gene and protein Gene and protein expression Can furnish similar gene and protein 34, 139
expression expression levels generally profiles are closer to natural expressions as in in vivo conditions and 154
differ from in vivo models tissues

Cellular Cell–surface interactions Cellular and cell–ECM Interactions of the cell cytoskeleton and the 34, 138
interactions prevail rather than cell–cell interactions are promoted ECM can significantly vary between 2D and and 139
and cell–ECM interactions 3D scaffold-based models, impacting
various cell activities, such as apoptosis

Directed growth Cells grow randomly Cellular aggregates are mainly Can be customized to direct cellular growth 133, 134,
spherical, whereas other shapes in various patterns, such as elongated or 155 and
were obtained in limited studies heterogeneous cell distributions 156

Sample handling Simple and well suited for Sample handling is more Sample handling and high-throughput 35, 39 and
and imaging high-throughput screening complicated while large number screening is more challenging 157
of spheroids can be provided for
high-throughput drug testing.

Scale up and Less expensive for large- Depending on the fabrication Reproducibility can be challenging and 135, 139,
reproducibility scale studies and easier to technique, spheroid properties scaffold’s properties may vary from batch to 141, 158
reproduce can be reproducible or greatly batch for naturally derived scaffolds and 159
variable

Complexity Involve the simplest Often labor-intensive, More expensive and more complicated 141
procedures time-consuming, more compared to 2D
expensive and more complicated

Drug testing Drugs appear effective Cells often show more resistance Cells responses are more reliable, showing 135 and
but the results are to therapeutics and better more resistance to drugs or providing better 160–162
physiologically irrelevant represent in vivo conditions prediction of drug cytotoxicity and efficacy

Wound healing Lower effectiveness Spheroids have been shown to These models can allow a more reliable 163 and
compared to 3D models, promote angiogenesis and wound studying of wound healing by better 164
as simulating natural healing compared to single cells evaluating the cell contractility and matrix
conditions is almost recovered from 2D cultures compaction as well as simulating the
impossible mechanical forces in tissue in vivo

Moreover, they reported that a spatially graded matrix stiffness furnishes a thorough comparison between 3D scaffold-based
can significantly alter cell proliferation and morphology. and 2D monolayer models. Ultimately, an ideal environment
Several other novel hydrogel designs for cell culture have for 3D cell culture requires a control of structural and physico-
been reported in the literature176–182 that demonstrate the chemical properties on length scales from nano to macro, and
superiority of 3D cultures to 2D planar substrates. Table 2 also time scales from seconds to weeks. Therefore, the next step in

58 | Mater. Horiz., 2019, 6, 45--71 This journal is © The Royal Society of Chemistry 2019
View Article Online

Materials Horizons Review


Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

Fig. 13 (a) (I) Schematic illustration of the self-assembly of peptide amphiphile molecules into nanofibers. (II) SEM micrograph of the nanofiber network
after adding the cell medium. (III) Macroscopic image of the gel formed by adding peptide solution to the cell culture medium.175 (b) (I) Illustration of
anisotropic capillary alginate gel formation. (II) Macroscopic image of the gel. (III) Cross section of the gel. (IV) Longitudinal section of the gel. (V and VI)
Confocal fluorescent micrographs show regenerating axons in the alginate capillaries. Scale bars: 1 cm (b, II), 100 mm (b, III and IV), and 50 mm (b, V and
VI).155 (c) (I) Schematic illustration of gelatin hydrogels designed by microfluidics that contain a gradient in cell number. (II) Increase in hypoxia inducible
factor 1 gene expression (green) with increasing cell (blue) density across the gel. Scale bar: 50 mm.156 Reproduced with permission.

effectively mimicking the ECM and guiding cell behavior is to Organ printing, which uses this technology to prototype living
develop materials whose properties can be tuned according to human organs (Fig. 14), offers new promise for solving organ
the time and length scale of the cell development. The conventional transplantation crisis.191–194
fabrication approaches, such as lithography, molding and solvent Bioprinted structures offer various desirable features to the
casting, are unable to sufficiently control the scaffold properties. field of tissue engineering. Spatial patterning of complex 3D
These include nanoscale surface modifications, such as biofunctio- environments can be relatively easy using 3D printing.183,198
nalization; microstructural details, such as pore size and shape, Bioprinting also allows for co-culturing of multiple cell-types,
porosity, and pore distribution; macroscale architecture, such as the which can be attractive for 3D modeling of neural tissues185,199
external appearances of artificial devices. Hence, a user-controlled that contain various glial cell types.183 Moreover, it is known for
and replicable technique seems essential. 3D bioprinting has its ability to control the scaffold porosity and produce inter-
recently emerged to address these shortcomings. connected pores,200 which can greatly enhance cell attachment and
Bioprinting is a rapidly growing computer-aided technology ingrowth,201 both desirable in tissue engineering applications.202,203
that can create complex tissue architectures by patterning a cell- For more information about different aspects of bioprinting, such
encapsulating bioink.183 Owing to its significantly improved reso- as the choice of material and bioprinter, challenges, and future
lution and cheaper cost in the recent years, it is implemented in perspectives of this technology, the reader is encouraged to see
tissue engineering to create 3D cell-laden materials in which the detailed reviews available elsewhere.183,185,194,204–206
bioink contains the cells,184,185 or to produce scaffolds onto which
cells are seeded.186 Due to its versatile patterning of cells and 3.2 3D vs. 212D (non-planar surfaces)
scaffolding materials, bioprinting has been able to revolutionize Two-dimensional monolayer cultures are usually used as control
biomedical applications, such as organ-on-a-chip devices,187 experiments when developing 3D culture systems and investigating
cancer research studies,188,189 and pharmaceutical analyses.190 their impacts on cellular functions. Therefore, direct comparison

This journal is © The Royal Society of Chemistry 2019 Mater. Horiz., 2019, 6, 45--71 | 59
View Article Online

Review Materials Horizons


Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

Fig. 14 A 3D bioprinter (a) was used to prototype a human ear (b).195 (c) Image of a 3D printed bionic ear during in vitro culture. Scale bar: 1 cm.196 Panels
(d, e and f) show a heart valve model designed by Solidworks, the bioprinted valve conduit, and the cross-sectional view of live/dead cells from surface to
4300 mm depth, respectively.197 Reproduced with permission.

between 2D and 3D models, reported in Section 3.1, could be readily Comparing 212D and 3D systems, spheroids might accom-
found in the literature. However, topographical substrates are not modate higher cell culture capacity in cell aggregates than 212D
usually directly compared with 3D systems in a single study, substrates and 3D scaffolds with limited culture spaces.212
requiring a broader and more careful survey to acquire this However, the biomaterial-based culture methods often improve
comparative information. In addition, an accurate comparison cell proliferation (Table 3), while spheroids usually yield similar/
can be conducted only between similar cell types that are lower proliferation rate compared to 2D monolayers.136,139,211
derived from the same tissue origin. For example, Baksh On the other hand, MSC differentiation can be improved using
et al.207 demonstrated that human MSCs derived from umbilical all three methods, thus, the optimal selection should be based
cord and bone marrow may exhibit different proliferative and on the final application. Due to their structurally controlled
differentiative capacities. Therefore, we will only focus on bone- surface characteristics, topographical substrates can provide
marrow-derived human MSCs in this section, since they have more detailed information about intercellular mechanics by
been more extensively investigated using all culture systems. tailoring and manipulating cell–cell interactions. Similarly, they
Bone-marrow-derived MSCs are multipotential cells which may be more desirable for high-throughput screening and
are capable of self-renewing and differentiating to a number of imaging, since cells are far more accessible for optical char-
cell types, including osteocytes, chondrocytes, adipocytes, acterization when they are plated onto a surface than confined
hepatocytes, myocytes, neurons, and cardiomyocytes. For tissue within the pores of a 3D scaffold with high thickness.
engineering applications, MSCs are often expanded and/or While both spheroids and scaffold-based models can be
differentiated in vitro prior to in vivo transplantation. Conventional good sources for obtaining a supply of specifically differentiated
2D monolayer methods normally fail to maintain and/or support MSCs or providing valuable information about MSC differentiation,
efficient cellular functions such as differentiation and replication. one might be more desirable than the other depending on the
On the other hand, topographical substrates, spheroids and 3D purpose of the culture. In tissue engineering applications, such as
tissue constructs have been shown to improve MSC differentiation cartilage repair, spheroids usually suffer from small size and
while providing proliferation and viability similar to or higher than uniformly weak mechanical properties,213 and these limitations
2D models, according to several studies summarized in Table 3. can be addressed using 3D scaffold-based models. Spheroids,
However, the cell characteristics as well as biomaterial properties however, have been rather more attractive for other applications
can greatly impact the outcome. For example, 3D-polycaprolactone/ such as cancer research by better mimicking the human solid
hydroxyapatite (PCL/HAp) scaffolds have been shown to increase tumor microenvironments, where necrotic, quiescent and proliferat-
the proliferation of bone-marrow-derived MSCs compared to 2D ing cells are present.34 In some cases, a combination of both
monolayers, whereas adipose-derived MSCs exhibited similar methods can be more effective. For example, hydrogel encapsulated
replication rate on 3D scaffolds and monolayer cultures.208 In spheroids have been shown to exhibit greater resistance to apoptosis
another study, MSC viability on nanofibrous topographies with than entrapped dissociated cells.136 Nonetheless, a judicious
random fibre orientations, which was assumed to better mimic selection of the cell culture method is a trade-off between
the structure of the native bone ECM, was higher than that on the efficacy and performance of the model, sample handling,
aligned fibres.209 complexity, reproducibility and economical factors.

60 | Mater. Horiz., 2019, 6, 45--71 This journal is © The Royal Society of Chemistry 2019
View Article Online

Materials Horizons Review

Table 3 Bone-marrow-derived MSC cellular functions in 221 D and 3D cell culture methods in comparison with 2D models

Structure/fabrication Cell morphology Viability Proliferation Differentiation Ref.


212D substrates
Square array of PMMA Flat and polygonal — Higher proliferation Higher osteoblastic 98 and
nanopits morphology and cell phenotype/ differentiation on displaced 99
multipotency retention arrays than 2D flat and
on ordered arrays than ordered nanopits
2D planar control and
displaced nanopits

Silicon nanowires More spherical and less Cell survivability — Shortest nanowires significantly 210
elongated morphologies was close to promoted osteogenic
with sturdy protrusions monolayers differentiation than 2D mono-
layers and longer nanowires

PCL nanofiber Aligned according to Higher viability on Improved proliferation on Osteochondrogenic 209
substrates fibre orientation. Cells random fibres than random fibres compared differentiation was significantly
Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

on aligned fibers were aligned fibres and with aligned orientations increased on random fibres
elongated and spindle- 2D culture plates and monolayers
shaped, whereas they
were mostly round on
random-oriented fibers

3D cell spheroids
Spinner flasks and Continuous layer of Comparable 4.3% or 3.5% proliferating Increased osteogenic and 211
RWV elongated MSCs on viability to cells for spinner and RWV, adipogenic differentiation
the outer surface with monolayer cultures which was comparable to in spheroids, where the latter
irregularly shaped cells monolayers was more significant in RWV
in the spheroid interior

Hydrogel encapsulated — Entrapped spheroids Proliferation did not differ Similar osteogenic potential 136
spheroids formed using exhibit greater between entrapped for spheroids and their
the hanging drop resistance to spheroids and dissociated monolayer counterparts, when
technique apoptosis than cells encapsulated in fibrin gels
entrapped cells from
monolayer cultures

Photolithography and 3D cell aggregates High viability was — Adipogenic and osteogenic 212
micropatterning obtained for at least differentiation was more
techniques 30 days efficient in spheroids than
monolayers

3D tissue-engineered models
3D PCL nanofibrousa Flat fibroblast-like cells on — Cell proliferation was Chondrogenic differentiation 213
scaffolds the surface with round, higher than in cell pellets was similar to cell pellets
chondrocyte-like cells in
middle, surrounded by
abundant cartilaginous
matrix

b-Tricalcium phosphate — Mean values of Proliferation was 3D culture provided higher 150
(TCP) ceramics metabolic activity proportional to porosity osteogenic differentiation
was larger with in 3D scaffolds, and than 2D monolayer, but it was
porosity, however, significantly higher independent of porosity
the comparison was than monolayer culture
elusive due to high
standard deviations

3D-Polycaprolactone/ — Cells were highly Higher proliferation rate ALP activity and mineral 208
hydroxyapatite (PCL/ viable than 2D monolayer deposition were higher on the
HAp) scaffold 3D scaffold than 2D culture
plates, indicating higher
osteogenic differentiation
a
Cell functions in this example are in comparison with their spheroid counterparts.

4 Applications 4.1 Cell therapy


Cell-based devices, analyses and therapies are increasingly being Cell therapy has been widely used in a variety of clinical
used in clinical applications. Here, we describe some recent contexts. Stem cell transplantation is a promising therapeutic
examples, although many others can be found in the literature. strategy for several illnesses, such as neurological214 and

This journal is © The Royal Society of Chemistry 2019 Mater. Horiz., 2019, 6, 45--71 | 61
View Article Online

Review Materials Horizons

cardiovascular215 diseases, type 1 diabetes,216 and osteogenesis that biomimetic culture platforms comprising physical topo-
imperfecta.37,217 graphies, such as wrinkles227 and grooves,228 can stimulate
Prior to implantation, the stem cells isolated from a donor cardiomyocyte cell alignment and, therefore, lead to physio-
are expanded in vitro using conventional 2D tissue culture logically relevant responses.228
techniques.218 However, it has been demonstrated that certain The dynamic cell alignment can be an important factor
cell-specific properties of stem cells are lost over time in 2D when modeling the progression of various diseases. For example,
culture to the point they no longer reflect in vivo cell behavior.38,219 during heart failure, the highly aligned cardiac muscle fibres
These cells are usually unable to maintain long-term tissue become disorganized, causing progressive fibrosis and ventri-
regeneration upon reinfusion into the body, highlighting the cular dilatation. While many of the in vitro models using surface
necessity of improved methods for in vitro cell culture. For topographies are static, a tunable culture platform that can
example, 212D99 and 3D211 cultures have been shown to enhance dynamically manipulate the temporal and spatial cell align-
osteogenic and adipogenic differentiation potential in MSCs ments can be attractive. Lam et al.229 were the first to introduce
when compared with 2D monolayer cultures. reversible cell alignment by compression-induced dynamic
Recently, 3D scaffolds were used to regenerate an entire, lamellar patterns on PDMS substrates. However, their pattern
Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

fully functional epidermis on a seven-year-old child suffering versatility was limited to one direction. To address this limitation,
from a life-threatening form of JEB.52 For this treatment, the Guvendiren and Burdick230 reported the fabrication of strain-
keratinocyte cells were cultured on 3D fibrin scaffolds, allowing responsive buckling patterns on PDMS substrates that could
the preparation of larger grafts from the same number of cells control the human MSC organization with a higher complexity
as would be needed for 2D plastic-cultured grafts.52,53 and versatility. In their study, multidirectional lamellar patterns
appeared and disappeared when relaxing and stretching the
4.2 Vaccine production substrates, respectively, in various directions and angles. They
In 1949, the key scientific discovery by Enders et al.220 that observed that the preferential alignment of MSCs in the presence
poliovirus could be grown in cell culture led to the development of lamellar topography was completely eliminated by stretching
of poliovirus vaccine, and to significant advances in producing the platform, producing a disorganized cell arrangement. Such
other virus vaccines in the following years. Prior to this, the few platforms can have considerable potential for targeted disease
available vaccines were produced in animal systems, such as modeling and drug discovery.
embryonated eggs for influenza and yellow fever viruses.221
Currently, many viruses are still being produced in animal
systems due to reliably high yields. However, these methods are 4.4 Drug screening
time consuming and expensive. In addition, in some cases, Drug development and pharmacokinetics studies using animal
such as embryonated eggs for seasonal influenza, the vaccines models are costly, time consuming and arguably unethical
are developed more than 6 months in advance of the flu season when using an extensive number of animals.231 Therefore,
to allow for adequate time to produce the required quantities. the early stages of drug discovery rely on toxicity testing
This lag can be problematic if the selected strains do not match using in vitro cell-based assays, to at least exclude poor ther-
the predominant ones during the winter. Unfortunately, people apeutic candidates before animal testing.94,232 To increase their
with egg allergies cannot use vaccines that are produced in eggs. predictive accuracy, the cell-based systems are expected to
Therefore, finding an alternative approach, such as in vitro cell effectively represent essential aspects of in vivo conditions.
culture, can be attractive.222 In this regard, Tree et al.223 introduced However, a substantial number of new therapeutic agents fail
the 3D culture of MDCK cells using microcarriers224 as a potential in late-stage human drug testing,233 calling for better in vivo-
technique for producing influenza virus A vaccine strains. mimetic culture platforms.
Although they suggested that a comparable yield of influenza virus The cell behavior is greatly regulated by the cell’s genetic
A production could be obtained between scaled-up 3D cell cultures codes and its communications with neighboring cells.234 Therefore,
and embryonated eggs, the differences may be larger for a different the spatial position of cells is crucial for their functionality235 and
strain of influenza virus. Thus, the search for an efficient approach their response to a therapeutic agent. For example, multicellular
is still ongoing. tumor spheroids have been shown to provide more physiologically
relevant models for anticancer drug screening compared to
4.3 Disease modeling monolayer cultures, because the intercellular adhesion in
Engineered tissues can furnish a powerful tool to probe the tumor spheroids can assist tumor cells in evading the cytotoxic
disruption in cell organization and function in response to a effects of anticancer drugs.236 Cell patterning on engineered
disease.225 However, to obtain biologically relevant models, the cell extracellular environments and, then replicating the cultured
arrangement should be closely related to the tissue organization cells in miniaturized regular arrays can improve cell-based drug
in vivo. For example, human embryonic stem cells can be directed testing.232,237 For example, tumor-on-a-chip systems compris-
into cardiac lineages, such as cardiomyocytes,226 for disease ing arrays of tumor spheroids have been presented by several
modeling. Whereas ventricular cardiomyocytes are naturally researchers to promote cancer drug discovery238 by providing a
aligned, human embryonic stem cell-derived cardiomyocytes are large number of uniform tumor spheroids for enhanced drug
heterogenous and randomly organized. It has been demonstrated screening.239

62 | Mater. Horiz., 2019, 6, 45--71 This journal is © The Royal Society of Chemistry 2019
View Article Online

Materials Horizons Review

5 Summary and outlook References


Cells in their biological ECM reside within a complex 3D 1 M. E. Lukashev and Z. Werb, ECM signalling: Orchestrating
network, in which every physical and chemical detail is essential cell behaviour and misbehaviour, Trends Cell Biol., 1998, 8,
for cell functionality. The oversimplified monolayer cell cultures 437–441.
have appreciably promoted our understanding of cell function. 2 A. S. G. Curtis and C. D. W. Wilkinson, Reactions of cells to
Nonetheless, they are notoriously inadequate for successful clinical topography, J. Biomater. Sci., Polym. Ed., 1998, 9, 1313–1329.
applications. Adapting to their new environment, cultured cells 3 M. J. Bissell, A. Rizki and I. S. Mian, Tissue architecture:
lose specific properties over time and deviate from their in vivo The ultimate regulator of breast epithelial function, Curr.
behaviors. Thus, there has been an increasing demand for culture Opin. Cell Biol., 2003, 15, 753–763.
platforms that better emulate biological conditions. 4 J. Lee, M. J. Cuddihy and N. A. Kotov, Three-dimensional
A wide range of ordered and random topographical surfaces cell culture matrices: State of the art, Tissue Eng., Part B,
(identified as 212D substrates), capable of better mimicking the 2008, 14, 61–86.
natural ECM, and their impact on cell behavior were detailed in 5 B. Geiger, J. P. Spatz and A. D. Bershadsky, Environmental
this review. With a relatively simple fabrication procedure, 212D
Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

sensing through focal adhesions, Nat. Rev. Mol. Cell Biol.,


platforms show great potential in advancing biomedical appli- 2009, 10, 21–33.
cations by retaining in vivo cell functionalities, such as their 6 L. Wang and R. L. Carrier, in Advances in Biomimetics, ed.
proliferative capacity. This can be significantly important for A. George, InTech, 2011, ch. Biomimetic topography:
cell transplantation and tissue engineering, where the cells Bioinspired cell culture substrates and scaffolds.
should display a long-term regeneration upon reinfusion into 7 M. M. Stevens and J. H. George, Exploring and engineering
the body. In addition, topographical structures with tunable the cell surface interface, Science, 2005, 310, 1135–1138.
characteristics that induce 3D cell morphology furnish exciting 8 J. El-Ali, P. K. Sorger and K. F. Jensen, Cells on chips,
opportunities for better understanding cellular response to Nature, 2006, 442, 403–411.
environmental cues. Stimuli-responsive platforms are another 9 A. M. Ross, Z. Jiang, M. Bastmeyer and J. Lahann, Physical
type of novel artificial ECMs that are expected to contrive aspects of cell culture substrates: Topography, roughness,
the dynamic investigation of cellular responses. Despite these and elasticity, Small, 2012, 8, 336–355.
advances, major challenges remain as innumerate factors 10 C. J. Bettinger, R. Langer and J. T. Borenstein, Engineering
simultaneously contribute to cell behavior. Nonetheless, recent substrate topography at the micro- and nanoscale to control
advances in micro- and nanofabrication techniques have paved cell function, Angew. Chem., Int. Ed., 2009, 48, 5406–5415.
the route to studying a single cell population and to address 11 B. Kasemo and J. Gold, Implant surfaces and interface
these challenges. processes, Adv. Dent. Res., 1999, 13, 8–20.
Topographical substrates are also more desirable for cell 12 T. J. Webster, C. Ergun, R. H. Doremus, R. W. Siegel and
imaging and high-throughput screening in comparison with R. Bizios, Enhanced functions of osteoblasts on nanophase
3D models, in which optical-readouts can be particularly ceramics, Biomaterials, 2000, 21, 1803–1810.
cumbersome. Nonetheless, in certain applications, such as 13 Y. Wan, Y. Wang, Z. Liu, X. Qu, B. Han, J. Bei and S. Wang,
cancer research and producing implantable tissue constructs, Adhesion and proliferation of OCT-1 osteoblast-like cells
spheroids and 3D porous scaffolds can be better options as they on micro- and nano-scale topography structured poly(L-lactide),
better emulate in vivo ECMs. However, a judicious selection Biomaterials, 2005, 26, 4453–4459.
between topographical surfaces and 3D models depends on the 14 E. Palin, N. H. Liu and T. J. Webster, Mimicking the
efficiency and performance of the models to achieve a targeted nanofeatures of bone increases bone-forming cell adhesion
outcome as well as on sample handling, complexity, reprodu- and proliferation, Nanotechnology, 2005, 16, 1828–1835.
cibility and economical factors. 15 M. J. Dalby, D. McCloy, M. Robertson, H. Agheli, D. Sutherland,
S. Affrossman and R. O. C. Oreffo, Osteoprogenitor response to
semi-ordered and random nanotopographies, Biomaterials,
Conflicts of interest 2006, 27, 2980–2987.
16 M. J. Dalby, D. McCloy, M. Robertson, C. D. Wilkinson and
There are no conflicts of interest to declare.
R. O. Oreffo, Osteoprogenitor response to defined topographies
with nanoscale depths, Biomaterials, 2006, 27, 1306–1315.
17 P. T. de Oliveira, S. F. Zalzal, M. M. Beloti, A. L. Rosa and
Acknowledgements
A. Nanci, Enhancement of in vitro osteogenesis on titanium
Support from Bourses d’excellence TransMedTech (to M. M. by chemically produced nanotopography, J. Biomed. Mater.
and V. A.), FRQNT (The Fonds de recherche du Québec – Nature Res., Part A, 2007, 80, 554–564.
et technologies to V. A.), CRC (Canada Research Chairs program 18 A. C. De Luca, M. Zink, A. Weidt, S. G. Mayr and A. E. Markaki,
to X. B. and D. K. H.) and NSERC (Natural Sciences and Effect of microgrooved surface topography on osteoblast
Engineering Research Council of Canada to X. B. and D. K. H. maturation and protein adsorption, J. Biomed. Mater. Res.,
(Discovery Grant 2017-04489)) is gratefully acknowledged. Part A, 2015, 103, 2689–2700.

This journal is © The Royal Society of Chemistry 2019 Mater. Horiz., 2019, 6, 45--71 | 63
View Article Online

Review Materials Horizons

19 J. E. Barralet, L. Wang, M. Lawson, J. T. Triffitt, P. R. Cooper 35 B. Derby, Printing and prototyping of tissues and scaffolds,
and R. M. Shelton, Comparison of bone marrow cell growth Science, 2012, 338, 921–926.
on 2D and 3D alginate hydrogels, J. Mater. Sci.: Mater. Med., 36 G. K. Tan, D. L. M. Dinnes, P. T. Myers and J. J. Cooper-
2005, 16, 515–519. White, Effects of biomimetic surfaces and oxygen tension
20 J. Lee, G. D. Lilly, R. C. Doty, P. Podsiadlo and N. A. Kotov, on redifferentiation of passaged human fibrochondrocytes
In vitro toxicity testing of nanoparticles in 3D cell culture, in 2D and 3D cultures, Biomaterials, 2011, 32, 5600–5614.
Small, 2009, 5, 1213–1221. 37 E. M. Horwitz, D. J. Prockop, P. L. Gordon, W. W. K. Koo,
21 S. Breslin and L. O’Driscoll, Three-dimensional cell L. A. Fitzpatrick, M. D. Neel, M. E. McCarville, P. J. Orchard,
culture: the missing link in drug discovery, Drug Discovery R. E. Pyeritz and M. K. Brenner, Clinical responses to bone
Today, 2013, 18, 240–249. marrow transplantation in children with severe osteogenesis
22 S. Even-Ram and K. M. Yamada, Cell migration in 3D imperfecta, Blood, 2001, 97, 1227–1231.
matrix, Curr. Opin. Cell Biol., 2005, 17, 524–532. 38 M. A. Baxter, R. F. Wynn, S. N. Jowitt, J. E. Wraith,
23 T. Elkayam, S. Amitay-Shaprut, M. Dvir-Ginzberg, T. Harel L. J. Fairbairn and I. Bellantuono, Study of telomere length
and S. Cohen, Enhancing the drug metabolism activities of reveals rapid aging of human marrow stromal cells follow-
Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

C3A–A human hepatocyte cell line–by tissue engineering ing in vitro expansion, Stem Cells, 2004, 22, 675–682.
within alginate scaffolds, Tissue Eng., 2006, 12, 1357–1368. 39 F. Pampaloni, E. G. Reynaud and E. H. K. Stelzer, The third
24 J. Barrila, A. L. Radtke, A. Crabbé, S. F. Sarker, M. M. Herbst- dimension bridges the gap between cell culture and live
Kralovetz, C. M. Ott and C. A. Nickerson, Organotypic 3D cell tissue, Nat. Rev. Mol. Cell Biol., 2007, 8, 839–845.
culture models: using the rotating wall vessel to study host- 40 M. Y. Lee, R. A. Kumar, S. M. Sukumaran, M. G. Hogg,
pathogen interactions, Nat. Rev. Microbiol., 2010, 8, 791–801. D. S. Clark and J. S. Dordick, Three-dimensional cellular
25 C. R. Thoma, M. Zimmermann, I. Agarkova, J. M. Kelm and microarray for high-throughput toxicology assays, Proc.
W. Krek, 3D cell culture systems modeling tumor growth Natl. Acad. Sci. U. S. A., 2008, 105, 59–63.
determinants in cancer target discovery, Adv. Drug Delivery 41 T. G. Fernandes, S. J. Kwon, S. S. Bale, M. Y. Lee, M. M. Diogo,
Rev., 2014, 69–70, 29–41. D. S. Clark, J. M. S. Cabral and J. S. Dordick, Three-dimensional
26 D. D. McKinnon, D. W. Domaille, J. N. Cha and cell culture microarray for high-throughput studies of stem cell
K. S. Anseth, Biophysically defined and cytocompatible fate, Biotechnol. Bioeng., 2010, 106, 106–118.
covalently adaptable networks as viscoelastic 3D cell culture 42 G. Fontana, J. Gershlak, M. Adamski, J. S. Lee, S. Matsumoto,
systems, Adv. Mater., 2014, 26, 865–872. H. D. Le, B. Binder, J. Wirth, G. Gaudette and W. L. Murphy,
27 Y. C. Lu, W. Song, D. An, B. J. Kim, R. Schwartz, M. Wu and Biofunctionalized plants as diverse biomaterials for human cell
M. Ma, Designing compartmentalized hydrogel micro- culture, Adv. Healthcare Mater., 2017, 6, 1601225.
particles for cell encapsulation and scalable 3D cell culture, 43 J. E. Puskas and Y. Chen, Biomedical application of
J. Mater. Chem. B, 2015, 3, 353–360. commercial polymers and novel polyisobutylene-based
28 S. Utech, R. Prodanovic, A. S. Mao, R. Ostafe, D. J. Mooney and thermoplastic elastomers for soft tissue replacement, Bio-
D. A. Weitz, Microfluidic generation of monodisperse, structu- macromolecules, 2004, 5, 1141–1154.
rally homogeneous alginate microgels for cell encapsulation and 44 N. M. Alves, I. Pashkuleva, R. L. Reis and J. F. Mano,
3D cell culture, Adv. Healthcare Mater., 2015, 4, 1628–1633. Controlling cell behavior through the design of polymer
29 V. van Duinen, S. J. Trietsch, J. Joore, P. Vulto and surfaces, Small, 2010, 6, 2208–2220.
T. Hankemeier, Microfluidic 3D cell culture: from tools 45 I. Milos̃ev, Metallic materials for biomedical applications:
to tissue models, Curr. Opin. Biotechnol., 2015, 35, 118–126. Laboratory and clinical studies, Pure Appl. Chem., 2010, 83,
30 L. Y. Sun, D. K. Hsieh, W. S. Syu, Y. S. Li, H. T. Chiu and 309–324.
T. W. Chiou, Cell proliferation of human bone marrow 46 I. Milos̃ev, in Surface treatments of titanium with antibacterial
mesenchymal stem cells on biodegradable microcarriers agents for implant applications, ed. S. Djokić, Springer Inter-
enhances in vitro differentiation potential, Cell Proliferation, national Publishing, Cham, 2016, pp. 1–87.
2010, 43, 445–456. 47 C. M. Agrawal and R. B. Ray, Biodegradable polymeric
31 S. Sart, S. N. Agathos and Y. Li, Engineering stem cell fate scaffolds for musculoskeletal tissue engineering, J. Biomed.
with biochemical and biomechanical properties of micro- Mater. Res., Part A, 2001, 55, 141–150.
carriers, Biotechnol. Prog., 2013, 29, 1354–1366. 48 J. A. Shibli, S. Grassi, L. Cristina de Figueiredo, M. Feres,
32 R. Z. Lin and H. Y. Chang, Recent advances in three- E. Marcantonio, G. Iezzi and A. Piattelli, Influence of implant
dimensional multicellular spheroid culture for biomedical surface topography on early osseointegration: A histological study
research, Biotechnol. J., 2008, 3, 1172–1184. in human jaws, J. Biomed. Mater. Res., Part B, 2007, 80, 377–385.
33 T.-M. Achilli, J. Meyer and J. R. Morgan, Advances in the 49 M. Goldberg, R. Langer and X. Jia, Nanostructured materials
formation, use and understanding of multi-cellular spheroids, for applications in drug delivery and tissue engineering,
Expert Opin. Biol. Ther., 2012, 12, 1347–1360. J. Biomater. Sci., Polym. Ed., 2007, 18, 241–268.
34 M. Ravi, V. Paramesh, S. R. Kaviya, E. Anuradha and 50 H. Hillaireau and P. Couvreur, Nanocarriers’ entry into the
F. D. Paul Solomon, 3D cell culture systems: Advantages cell: relevance to drug delivery, Cell. Mol. Life Sci., 2009, 66,
and applications, J. Cell. Physiol., 2015, 230, 16–26. 2873–2896.

64 | Mater. Horiz., 2019, 6, 45--71 This journal is © The Royal Society of Chemistry 2019
View Article Online

Materials Horizons Review

51 B. Dhandayuthapani, Y. Yoshida, T. Maekawa and D. S. Kumar, 67 E. A. Bremus-Koebberling and S. Beckemper, Nano structures
Polymeric Scaffolds in Tissue Engineering Application: A via laser interference patterning for guided cell growth of
Review, Int. J. Polym. Sci., 2011, 2011, 290602. neuronal cells, J. Laser Appl., 2012, 24, 042013.
52 T. Hirsch, et al., Regeneration of the entire human epidermis 68 M. Ochesner, M. R. Dusseiller, H. M. Grandin, S. Luna-
using transgenic stem cells, Nature, 2017, 551, 327–332. Morris, M. Textor, V. Vogel and M. L. Smith, Micro-well
53 G. Pellegrini, R. Ranno, G. Stracuzzi, S. Bondanza, L. Guerra, arrays for 3D shape control and high resolution analysis of
G. Zambruno, G. Micali and M. De Luca, The control of single cells, Lab Chip, 2007, 7, 1074–1077.
epidermal stem cells (holoclones) in the treatment of massive 69 D. Falconnet, G. Csucs, H. M. Grandin and M. Textor,
full-thickness burns with autologous keratinocytes cultured on Surface engineering approaches to micropattern surfaces
fibrin, Transplantation, 1999, 68, 868–879. for cell-based assays, Biomaterials, 2006, 27, 3044–3063.
54 A. S. Andersson, F. Bäckhed, A. von Euler, A. Richter- 70 V. N. Truskett and M. P. C. Watts, Trends in imprint
Dahlfors, D. Sutherland and B. Kasemo, Nanoscale features lithography for biological applications, Trends Biotechnol.,
influence epithelial cell morphology and cytokine production, 2006, 24, 312–317.
Biomaterials, 2003, 24, 3427–3436. 71 A. M. Ross and J. Lahann, Surface engineering the cellular
Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

55 R. L. Price, K. Ellison, K. M. Haberstroh and T. J. Webster, microenvironment via patterning and gradients, J. Polym.
Nanometer surface roughness increases select osteoblast Sci., Part B: Polym. Phys., 2013, 51, 775–794.
adhesion on carbon nanofiber compacts, J. Biomed. Mater. 72 X. Liu and P. X. Ma, Polymeric scaffolds for bone tissue
Res., Part A, 2004, 70, 129–138. engineering, Ann. Biomed. Eng., 2004, 32, 477–486.
56 L. Altomare, N. Gadegaard, L. Visai, M. C. Tanzi and 73 V. A. Schulte, M. Dı́ez, M. Möller and M. C. Lensen, Surface
S. Farè, Biodegradable microgrooved polymeric surfaces topography induces fibroblast adhesion on intrinsically
obtained by photolithography for skeletal muscle cell nonadhesive poly(ethylene glycol) substrates, Biomacromo-
orientation and myotube development, Acta Biomater., lecules, 2009, 10, 2795–2801.
2010, 6, 1948–1957. 74 A. I. Teixeira, P. F. Nealey and C. J. Murphy, Responses of
57 K. Kolind, K. W. Leong, F. Besenbacher and M. Foss, human keratocytes to micro- and nanostructured sub-
Guidance of stem cell fate on 2D patterned surfaces, strates, J. Biomed. Mater. Res., Part A, 2004, 71, 369–376.
Biomaterials, 2012, 33, 6626–6633. 75 X. F. Walboomers, H. J. E. Croes, L. A. Ginsel and J. A. Jansen,
58 S. Turunen, A. M. Haaparanta, R. Äänismaa and M. Kellomäki, Growth behavior of fibroblasts on microgrooved polystyrene,
Chemical and topographical patterning of hydrogels for neural Biomaterials, 1998, 19, 1861–1868.
cell guidance in vitro, J. Tissue Eng. Regener. Med., 2013, 7, 76 J. L. Charest, M. T. Eliason, A. J. Garcı́a and W. P. King,
253–270. Combined microscale mechanical topography and chemical
59 K. S. Masters and K. S. Anseth, Cell–material interactions, patterns on polymer cell culture substrates, Biomaterials,
Adv. Chem. Eng., 2004, 29, 7–46. 2006, 27, 2487–2494.
60 H. S. Dhowre, S. Rajput, N. A. Russell and M. Zelzer, 77 W. A. Loesberg, J. te Riet, F. C. M. J. M. van Delft, P. Schön,
Responsive cell–material interfaces, Nanomedicine, 2015, C. G. Figdor, S. Speller, J. J. W. A. van Loon, X. F. Walboomers
10, 849–871. and J. A. Jansen, The threshold at which substrate nanogroove
61 P. Clark, P. Connolly, A. S. G. Curtis, J. A. T. Dow and dimensions may influence fibroblast alignment and adhesion,
C. D. W. Wilkinson, Topographical control of cell behaviour: I. Biomaterials, 2007, 28, 3944–3951.
Simple step cues, Development, 1987, 99, 439–448. 78 E. Lamers, X. F. Walboomers, M. Domanski, J. te Riet, F. C. M. J.
62 P. Clark, P. Connolly, A. S. G. Curtis, J. A. T. Dow and M. van Delft, R. Luttge, L. A. J. A. Winnubst, H. J. G. E.
C. D. W. Wilkinson, Topographical control of cell behaviour: II. Gardeniers and J. A. Jansen, The influence of nanoscale grooved
multiple grooved substrata, Development, 1990, 108, 635–644. substrates on osteoblast behavior and extracellular matrix
63 J. Mai, C. S. Sun, S. Li and X. Zhang, A microfabricated platform deposition, Biomaterials, 2010, 31, 3307–3316.
probing cytoskeleton dynamics using multidirectional 79 J. Chen, Q. Wang, S. Chen, S. A. Wickline and S. K. Song,
topographical cues, Biomed. Microdevices, 2007, 9, 523–531. In vivo diffusion tensor MRI of the mouse retina: a non-
64 K. M. Beussman, M. L. Rodriguez, A. Leonard, N. Taparia, invasive visualization of tissue organization, NMR Biomed.,
C. R. Thompson and N. J. Sniadecki, Micropost arrays for 2011, 24, 447–451.
measuring stem cell-derived cardiomyocyte contractility, 80 T. Banerjee, S. Mukherjee, S. Ghosh, M. Biswas, S. Dutta,
Methods, 2016, 94, 43–50. S. Pattari, S. Chatterjee and A. Bandyopadhyay, Clinical
65 D. S. Sutherland, M. Broberg, H. Nygren and B. Kasemo, significance of markers of collagen metabolism in rheumatic
Influence of nanoscale surface topography and chemistry mitral valve disease, PLoS One, 2014, 9, e90527.
on the functional behaviour of an adsorbed model macro- 81 M. Ghibaudo, L. Trichet, J. Le Digabel, A. Richert, P. Hersen
molecule, Macromol. Biosci., 2001, 1, 270–273. and B. Ladoux, Substrate topography induces a crossover
66 W. B. Tsai, Y. C. Ting, J. Y. Yang, J. Y. Lai and H. L. Liu, from 2D to 3D behavior in fibroblast migration, Biophys. J.,
Fibronectin modulates the morphology of osteoblast-like 2009, 97, 357–368.
cells (MG-63) on nano-grooved substrates, J. Mater. Sci.: 82 B. Li, L. Xie, Z. C. Starr, Z. Yang, J. S. Lin and J. H. Wang,
Mater. Med., 2009, 20, 1367–1378. Development of micropost force sensor array with culture

This journal is © The Royal Society of Chemistry 2019 Mater. Horiz., 2019, 6, 45--71 | 65
View Article Online

Review Materials Horizons

experiments for determination of cell traction forces, Cell 97 C. H. Choi, S. H. Hagvall, B. M. Wu, J. C. Y. Dunn, R. E.
Motil. Cytoskeleton, 2007, 64, 509–518. Beygui and C. J. Kim, Cell interaction with three-dimensional
83 M. L. Rodriguez, B. T. Graham, L. M. Pabon, S. J. Han, sharp-tip nanotopography, Biomaterials, 2007, 28, 1672–1679.
C. E. Murry and N. J. Sniadecki, Measuring the contractile 98 M. J. Dalby, N. Gadegaard, R. Tare, A. Andar, M. O. Riehle,
forces of human induced pluripotent stem cell-derived P. Herzyk, C. D. W. Wilkinson and R. O. C. Oreffo, The
cardiomyocytes with arrays of microposts, J. Biomech. Eng., control of human mesenchymal cell differentiation using
2014, 136, 051005. nanoscale symmetry and disorder, Nat. Mater., 2007, 6,
84 X. F. Walboomers, W. Monaghan, A. S. G. Curtis and 997–1003.
J. A. Jansen, Attachment of fibroblasts on smooth and 99 R. J. McMurray, N. Gadegaard, P. M. Tsimbouri, K. V. Burgess,
microgrooved polystyrene, J. Biomed. Mater. Res., 1999, 46, L. E. McNamara, R. Tare, K. Murawski, E. Kingham,
212–220. R. O. C. Oreffo and M. J. Dalby, Nanoscale surfaces for the
85 M. T. Frey, I. Y. Tsai, T. P. Russell, S. K. Hanks and long-term maintenance of mesenchymal stem cell phenotype
Y. L. Wang, Cellular responses to substrate topography: and multipotency, Nat. Mater., 2011, 10, 637–644.
Role of myosin II and focal adhesion kinase, Biophys. J., 100 A. S. G. Curtis, N. Gadegaard, M. J. Dalby, M. O. Riehle,
Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

2006, 90, 3774–3782. C. D. W. Wilkinson and G. Aitchison, Cells react to


86 Z. Li, J. Song, G. Mantini, M. Y. Lu, H. Fang, C. Falconi, nanoscale order and symmetry in their surroundings, IEEE
L. J. Chen and Z. L. Wang, Quantifying the traction force of Trans Nanobioscience, 2004, 3, 61–65.
a single cell by aligned silicon nanowire array, Nano Lett., 101 A. S. G. Curtis, B. Casey, J. O. Gallagher, D. Pasqui,
2009, 9, 3575–3580. M. A. Wood and C. D. W. Wilkinson, Substratum nano-
87 J. Fu, Y. K. Wang, M. T. Yang, R. A. Desai, X. Yu, Z. Liu and topography and the adhesion of biological cells. Are
C. S. Chen, Mechanical regulation of cell function with symmetry or regularity of nanotopography important?,
geometrically modulated elastomeric substrates, Nat. Biophys. Chem., 2001, 94, 275–283.
Methods, 2010, 7, 733–736. 102 J. O. Gallagher, K. F. McGhee, C. D. W. Wilkinson and
88 S. Huang, C. S. Chen and D. E. Ingber, Control of cyclin D1, M. O. Riehle, Interaction of animal cells with ordered
p27Kip1, and cell cycle progression in human capillary nanotopography, IEEE Trans Nanobioscience, 2002, 1, 24–28.
endothelial cells by cell shape and cytoskeletal tension, 103 C. C. Berry, G. Campbell, A. Spadiccino, M. Robertson and
Mol. Biol. Cell, 1998, 9, 3179–3193. A. S. G. Curtis, The influence of microscale topography on
89 A. K. Harris, P. Wild and D. Stopak, Silicone rubber fibroblast attachment and motility, Biomaterials, 2004, 25,
substrata: a new wrinkle in the study of cell locomotion, 5781–5788.
Science, 1980, 208, 177–179. 104 J. Tan, H. Shen and W. M. Saltzman, Micron-scale positioning
90 K. Burton, J. H. Park and D. L. Taylor, Keratocytes generate of features influences the rate of polymorphonuclear leukocyte
traction forces in two phases, Mol. Biol. Cell, 1999, 10, migration, Biophys. J., 2001, 81, 2569–2579.
3745–3769. 105 M. Werner, S. B. G. Blanquer, S. P. Haimi, G. Korus,
91 A. D. Bershadsky, N. Q. Balaban and B. Geiger, Adhesion- J. W. C. Dunlop, G. N. Duda, D. W. Grijpma and A. Petersen,
dependent cell mechanosensitivity, Annu. Rev. Cell Dev. Surface curvature differentially regulates stem cell migration
Biol., 2003, 19, 677–695. and differentiation via altered attachment morphology and
92 V. Karantalis, W. Balkan, I. H. Schulman, K. E. Hatzistergos nuclear deformation, Adv. Sci., 2017, 4, 1600347.
and J. M. Hare, Cell-based therapy for prevention and 106 M. Singh, D. A. Close, S. Mukundan, P. A. Johnston and
reversal of myocardial remodeling, Am. J. Physiol. Heart S. Sant, Production of uniform 3D microtumors in hydrogel
Circ. Physiol., 2012, 303, H256–H270. microwell arrays for measurement of viability, morphology,
93 C. Dambrot, R. Passier, D. Atsma and C. L. Mummery, and signaling pathway activation, Assay Drug Dev. Technol.,
Cardiomyocyte differentiation of pluripotent stem cells 2015, 13, 570–583.
and their use as cardiac disease models, Biochem. J., 107 M. Singh, S. Mukundan, M. Jaramillo, S. Oesterreich and S. Sant,
2011, 434, 25–35. Three-dimensional breast cancer models mimic hallmarks of
94 M. Grskovic, A. Javaherian, B. Strulovici and G. Q. Daley, size-induced tumor progression, Cancer Res., 2016, 76, 1–12.
Induced pluripotent stem cells-opportunities for disease 108 J. Casey, X. Yue, T. D. Nguyen, A. Acun, V. R. Zellmer,
modelling and drug discovery, Nat. Rev. Drug Discovery, S. Zhang and P. Zorlutuna, 3D hydrogel-based microwell
2011, 10, 915–929. arrays as a tumor microenvironment model to study breast
95 C. S. Chen, J. L. Alonso, E. Ostuni, G. M. Whitesides and cancer growth, Biomed. Mater., 2017, 12, 025009.
D. E. Ingber, Cell shape provides global control of focal 109 S. Sant and P. A. a. Johnston, The production of 3D tumor
adhesion assembly, Biochem. Biophys. Res. Commun., 2003, spheroids for cancer drug discovery, Drug Discovery Today:
307, 355–361. Technol., 2017, 23, 27–36.
96 J. L. Tan, J. Tien, D. M. Pirone, D. S. Gray, K. Bhadriraju 110 N. I. Martins, M. P. Sousa, C. A. Custódio, V. C. Pinto,
and C. S. Chen, Cells lying on a bed of microneedles: An P. J. Sousa, G. Minas, F. Cleymand and J. F. Mano, Multilayered
approach to isolate mechanical force, Proc. Natl. Acad. Sci. membranes with tuned well arrays to be used as regenerative
U. S. A., 2003, 100, 1484–1489. patches, Acta Biomater., 2017, 57, 313–323.

66 | Mater. Horiz., 2019, 6, 45--71 This journal is © The Royal Society of Chemistry 2019
View Article Online

Materials Horizons Review

111 T. Tzvetkova-Chevolleau, A. Stéphanou, D. Fuard, J. Ohayon, micro- and nanostructured substrates, J. Cell Sci., 2003,
P. Schiavone and P. Tracqui, The motility of normal and 116, 1881–1892.
cancer cells in response to the combined influence of the 124 F. Klein, T. Striebel, J. Fischer, Z. Jiang, C. M. Franz, G. von
substrate rigidity and anisotropic microstructure, Biomaterials, Freymann, M. Wegener and M. Bastmeyer, Elastic fully
2008, 29, 1541–1551. three-dimensional microstructure scaffolds for cell force
112 B. J. Papenburg, E. D. Rodrigues, M. Wessling and measurements, Adv. Mater., 2010, 22, 868–871.
D. Stamatialis, Insights into the role of material surface 125 M. R. Dusseiller, M. L. Smith, V. Vogel and M. Textor,
topography and wettability on cell–material interactions, Microfabricated three-dimensional environments for sin-
Soft Matter, 2010, 6, 4377–4388. gle cell studies, Biointerphases, 2006, 1, P1–P4.
113 E. Kim; J. Lee; S. Ahn; H. Jeon and K. Lee Cell culture over 126 J. N. Hanson Shepherd, S. T. Parker, R. F. Shepherd,
nanopatterned surface fabricated by holographic lithography M. U. Gillette, J. A. Lewis and R. G. Nuzzo, 3D micro-
and nanoimprint lithography, 2008 3rd IEEE International periodic hydrogel scaffolds for robust neuronal cultures,
Conference on Nano/Micro Engineered and Molecular Adv. Funct. Mater., 2011, 21, 47–54.
Systems, 2008, pp. 725–728. 127 F. Klein, B. Richter, T. Stribel, C. M. Franz, G. von Freymann,
Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

114 J. Sun, Y. Ding, N. J. Lin, J. Zhou, H. Ro, C. L. Soles, M. Wegener and M. Bastmeyer, Two-component polymer
M. T. Cicerone and S. Lin-Gibson, Exploring cellular contact scaffolds for controlled three-dimensional cell culture, Adv.
guidance using gradient nanogratings, Biomacromolecules, Mater., 2011, 23, 1341–1345.
2010, 11, 3067–3072. 128 A. M. Greiner, F. Klein, T. Gudzenko, B. Richter, T. Striebel,
115 B. A. Dalton, M. D. M. Evans, G. A. McFarland and B. G. Wundari, T. J. Autenrieth, M. Wegener, C. M. Franz
J. G. Steele, Modulation of corneal epithelial stratification and M. Bastmeyer, Cell type-specific adaptation of cellular
by polymer surface topography, J. Biomed. Mater. Res., and nuclear volume in micro-engineered 3D environ-
1999, 45, 384–394. ments, Biomaterials, 2015, 69, 121–132.
116 D. D. Deligianni, N. D. Katsala, P. G. Koutsoukos and 129 V. Nandakumar, L. Kelbauskas, K. F. Hernandez, K. M.
Y. F. Missirlis, Effect of surface roughness of hydroxyapatite Lintecum, P. Senechal, K. J. Bussey, P. C. W. Davies, R. H.
on human bone marrow cell adhesion, proliferation, Johnson and D. R. Meldrum, Isotropic 3D nuclear morphometry
differentiation and detachment strength, Biomaterials, of normal, fibrocystic and malignant breast epithelial cells
2000, 22, 87–96. reveals new structural alterations, PLoS One, 2012, 7, e29230.
117 D. D. Deligianni, N. Katsala, S. Ladas, D. Sotiropoulou, 130 B. Richter, V. Hahn, S. Bertels, T. K. Claus, M. Wegener,
J. Amedee and Y. F. Missirlis, Effect of surface roughness G. Delaittre, C. Barner-Kowollik and M. Bastmeyer, Guiding
of the titanium alloy Ti-6Al-4V on human bone marrow cell cell attachment in 3D microscaffolds selectively functionalized
response and on protein adsorption, Biomaterials, 2001, with two distinct adhesion proteins, Adv. Mater., 2017,
22, 1241–1251. 29, 1604342.
118 A. Ranella, M. Barberoglou, S. Bakogianni, C. Fotakis and 131 M. Li, N. Hakimi, R. Perez, S. Waldman, J. A. Kozinski and
E. Stratakis, Tuning cell adhesion by controlling the roughness D. K. Hwang, Microarchitecture for a three-dimensional
and wettability of 3D micro/nano silicon structures, Acta wrinkled surface platform, Adv. Mater., 2015, 27, 1880–1886.
Biomater., 2010, 6, 2711–2720. 132 P. Kim, M. Abkarian and H. A. Stone, Hierarchical folding
119 R. A. Gittens, T. McLachlan, R. Olivares-Navarrete, Y. Cai, of elastic membranes under biaxial compressive stress,
S. Berner, R. Tannenbaum, Z. Schwartz, K. H. Sandhage and Nat. Mater., 2011, 10, 952–957.
B. D. Boyan, The effects of combined micron-/submicron-scale 133 D. M. Dean, A. P. Napolitano, J. Youssef and J. R. Morgan,
surface roughness and nanoscale features on cell proliferation Rods, tori, and honeycombs: The directed self-assembly of
and differentiation, Biomaterials, 2011, 32, 3395–3403. microtissues with prescribed microscale geometries, FASEB
120 L. G. Villa-Diaz, A. M. Ross, J. Lahann and P. H. Krebsbach, J., 2007, 21, 4005–4012.
Concise review: the evolution of human pluripotent stem 134 A. P. Napolitano, P. Chai, D. M. Dean and J. R. Morgan,
cell culture: from feeder cells to synthetic coatings, Stem Dynamics of the self-assembly of complex cellular aggregates
Cells, 2013, 31, 1–7. on micromolded nonadhesive hydrogels, Tissue Eng., 2007, 13,
121 H. Shadpour and N. L. Allbritton, In situ roughening of 2087–2094.
polymeric microstructures, ACS Appl. Mater. Interfaces, 135 H. Karlsson, M. Fryknäs, R. Larsson and P. Nygren, Loss of
2010, 2, 1086–1093. cancer drug activity in colon cancer HCT-116 cells during
122 H. V. Unadkat, M. Hulsman, K. Cornelissen, B. J. Papenburg, spheroid formation in a new 3D spheroid cell culture
R. K. Truckenmüller, A. E. Carpenter, M. Wessling, G. F. Post, system, Exp. Cell Res., 2012, 318, 1577–1585.
M. Uetz, M. J. T. Reinders, D. Stamatialis, C. A. van Blitterswijk 136 K. C. Murphy, S. Y. Fang and J. K. Leach, Human mesenchymal
and J. de Boer, An algorithm-based topographical biomaterials stem cell spheroids in fibrin hydrogels exhibit improved cell
library to instruct cell fate, Proc. Natl. Acad. Sci. U. S. A., 2011, survival and potential for bone healing, Cell Tissue Res., 2014,
108, 16565–16570. 357, 91–99.
123 A. I. Teixeira, G. A. Abrams, P. J. Bertics, C. J. Murphy and 137 S. S. Ho, K. C. Murphy, B. Y. K. Binder, C. B. Vissers and
P. F. Nealey, Epithelial contact guidance on well-defined J. K. Leach, Increased survival and function of mesenchymal

This journal is © The Royal Society of Chemistry 2019 Mater. Horiz., 2019, 6, 45--71 | 67
View Article Online

Review Materials Horizons

stem cell spheroids entrapped in instructive alginate hydrogels, cells: An in vitro and in vivo study, Acta Biomater., 2008, 4,
Stem Cells Transl. Med., 2016, 5, 773–781. 1904–1915.
138 Y. B. Lee, E. M. Kim, H. Byun, H.-k. Chang, K. Jeong, 151 P. R. Baraniak and T. C. McDevitt, Scaffold-free culture of
Z. M. Aman, Y. S. Choi, J. Park and H. Shin, Engineering mesenchymal stem cell spheroids in suspension preserves
spheroids potentiating cell–cell and cell–ECM interactions multilineage potential, Cell Tissue Res., 2012, 347, 701–711.
by self-assembly of stem cell microlayer, Biomaterials, 152 Y. Yamaguchi, J. Ohno, A. Sato, H. Kido and T. Fukushima,
2018, 165, 105–120. Mesenchymal stem cell spheroids exhibit enhanced in vitro
139 S. C. Ramaiahgari, M. W. den Braver, B. Herpers, V. Terpstra, and in vivo osteoregenerative potential, BMC Biotechnol.,
J. N. M. Commandeur, B. van de Water and L. S. Price, A 3D 2014, 14, 105.
in vitro model of differentiated HepG2 cell spheroids with 153 T. S. Ramasamy, J. S. L. Yu, C. Selden, H. Hodgson and
improved liver-like properties for repeated dose high- W. Cui, Application of three-dimensional culture conditions to
throughput toxicity studies, Arch. Toxicol., 2014, 88, 1083–1095. human embryonic stem cell-derived definitive endoderm cells
140 G. Mehta, A. Y. Hsiao, M. Ingram, G. D. Luker and S. Takayama, enhances hepatocyte differentiation and functionality, Tissue
Opportunities and challenges for use of tumor spheroids as Eng., Part A, 2013, 19, 360–367.
Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

models to test drug delivery and efficacy, J. Controlled 154 H. Tanaka, S. Tanaka, K. Sekine, S. Kita, A. Okamura,
Release, 2012, 164, 192–204. T. Takebe, Y. W. Zheng, Y. Ueno, J. Tanaka and H. Taniguchi,
141 R. Edmondson, J. J. Broglie, A. F. Adcock and L. Yang, The generation of pancreatic b-cell spheroids in a simulated
Three-dimensional cell culture systems and their applications microgravity culture system, Biomaterials, 2013, 34, 5785–5791.
in drug discovery and cell-based biosensors, Assay Drug Dev. 155 P. Prang, R. Muller, A. Eljaouhari, K. Heckmann, W. Kunz,
Technol., 2014, 12, 207–218. T. Weber, C. Faber, M. Vroemen, U. Bogdahn and
142 K. Moshksayan, N. Kashaninejad, M. Ebrahimi Warkiani, N. Weidner, The promotion of oriented axonal regrowth
J. G. Lock, H. Moghadas, B. Firoozabadi, M. S. Saidi and in the injured spinal cord by alginate-based anisotropic
N. T. Nguyen, Spheroids-on-a-chip: Recent advances and capillary hydrogels, Biomaterials, 2006, 27, 3560–3569.
design considerations in microfluidic platforms for spheroid 156 S. Pedron, E. Becka and B. A. Harley, Spatially gradated
formation and culture, Sens. Actuators, B, 2018, 263, 151–176. hydrogel platform as a 3D engineered tumor microenviron-
143 H. Huang, Y. Ding, X. S. Sun and T. A. Nguyen, Peptide ment, Adv. Mater., 2015, 27, 1567–1572.
hydrogelation and cell encapsulation for 3D culture of 157 Y. C. Tung, A. Y. Hsiao, S. G. Allen, Y.-s. Torisawa, M. Ho
MCF-7 breast cancer cells, PLoS One, 2013, 8, e59482. and S. Takayama, High-throughput 3D spheroid culture
144 B. M. Baker and C. S. Chen, Deconstructing the third and drug testing using a 384 hanging drop array, Analyst,
dimension – how 3D culture microenvironments alter 2011, 136, 473–478.
cellular cues, J. Cell Sci., 2012, 125, 3015–3024. 158 A. Abbott, Biology’s new dimension, Nature, 2003, 424, 870–872.
145 T. Weiß, R. Schade, T. Laube, A. Berg, G. Hildebrand, 159 L. A. Gurski, N. J. Petrelli, X. Jia and M. C. Farach-Carson,
R. Wyrwa, M. Schnabelrauch and K. Liefeith, Two-photon 3D matrices for anti-cancer drug testing and development,
polymerization of biocompatible photopolymers for micro- Oncol. Issues, 2010, 25, 20–25.
structured 3D biointerfaces, Adv. Mater., 2011, 13, B264–B273. 160 S. J. Fey and K. Wrzesinski, Determination of drug toxicity
146 M. Wolun-Cholewa, K. Langer, K. Szymanowski, A. Glodek, using 3D spheroids constructed from an immortal human
A. Jankowska, W. Warchol and J. Langer, An efficient 3D hepatocyte cell line, Toxicol. Sci, 2012, 127, 403–411.
cell culture method on biomimetic nanostructured grids, 161 Q. Meng, Three-dimensional culture of hepatocytes for
PLoS One, 2013, 8, e72936. prediction of drug-induced hepatotoxicity, Expert Opin.
147 T. Xu, P. Molnar, C. Gregory, M. Das, T. Boland and Drug Metab. Toxicol., 2010, 6, 733–746.
J. J. Hickman, Electrophysiological characterization of 162 J. Yin, Q. Meng, G. Zhang and Y. Sun, Differential methotrexate
embryonic hippocampal neurons cultured in a 3D collagen hepatotoxicity on rat hepatocytes in 2D monolayer culture and
hydrogel, Biomaterials, 2009, 30, 4377–4383. 3D gel entrapment culture, Chem. – Biol. Interact., 2009, 180,
148 S. Li, J. Lao, B. P. C. Chen, Y. S. Li, Y. Zhao, J. Chu, 368–375.
K. D. Chen, T. C. Tsou, K. Peck and S. Chien, Genomic 163 S.-h. Hsu and P. S. Hsieh, Self-assembled adult adipose-
analysis of smooth muscle cells in 3-dimensional collagen derived stem cell spheroids combined with biomaterials
matrix, FASEB J., 2003, 17, 97–99. promote wound healing in a rat skin repair model, Wound
149 E. M. Chandler, C. M. Berglund, J. S. Lee, W. J. Polacheck, Repair Regen., 2015, 23, 57–64.
J. P. Gleghorn, B. J. Kirby and C. Fischbach, Stiffness 164 T. P. Amadeu and B. Coulomb, Cutaneous wound healing:
of photocrosslinked RGD-alginate gels regulates adipose Myofibroblastic differentiation and in vitro models, Int.
progenitor cell behavior, Biotechnol. Bioeng., 2011, 108, J. Lower Extremity Wounds, 2003, 2, 60–68.
1683–1692. 165 Y. Duan, Z. Liu, J. O’Neill, L. Q. Wan, D. O. Freytes and
150 P. Kasten, I. Beyen, P. Niemeyer, R. Luginbühl, M. Bohner G. Vunjak-Novakovic, Hybrid gel composed of native heart
and W. Richter, Porosity and pore size of b-tricalcium matrix and collagen induces cardiac differentiation of
phosphate scaffold can influence protein production and human embryonic stem cells without supplemental growth
osteogenic differentiation of human mesenchymal stem factors, J. Cardiovasc. Transl. Res., 2011, 4, 605–615.

68 | Mater. Horiz., 2019, 6, 45--71 This journal is © The Royal Society of Chemistry 2019
View Article Online

Materials Horizons Review

166 S. K. Goh, S. Bertera, P. Olsen, J. E. Candiello, W. Halfter, host-guest interactions assist cell infiltration and in situ
G. Uechi, M. Balasubramani, S. A. Jahnson, B. M. Sicari, tissue regeneration, Biomaterials, 2016, 101, 217–228.
E. Kollar, S. F. Badylak and I. Banerjee, Perfusion- 182 X. Dou and C. Feng, Amino acids and peptide-based
decellularized pancreas as a natural 3D scaffold for pan- supramolecular hydrogels for three-dimensional cell culture,
creatic tissue and whole organ engineering, Biomaterials, Adv. Mater., 2017, 29, 1604062.
2013, 34, 6760–6772. 183 S. Knowlton, S. Anand, T. Shah and S. Tasoglu, Bioprinting
167 S. L. Spear, P. M. Parikh, E. Reisin and N. G. Menon, for neural tissue engineering, Trends Neurosci., 2018, 41,
Acellular dermis-assisted breast reconstruction, Aesthetic 31–46.
Plast. Surg., 2008, 32, 418–425. 184 T. Billiet, E. Gevaert, T. De Schryver, M. Cornelissen and
168 R. Langer and D. A. Tirrell, Designing materials for biology P. Dubruel, The 3D printing of gelatin methacrylamide
and medicine, Nature, 2004, 428, 487–492. cell–laden tissue-engineered constructs with high cell via-
169 N. Huebsch and D. J. Mooney, Inspiration and application bility, Biomaterials, 2014, 35, 49–62.
in the evolution of biomaterials, Nature, 2009, 462, 426–432. 185 W. Aljohani, M. W. Ullah, X. Zhang and G. Yang, Bioprinting
170 M. W. Tibbitt and K. S. Anseth, Hydrogels as extracellular and its applications in tissue engineering and regenerative
Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

matrix mimics for 3D cell culture, Biotechnol. Bioeng., 2009, medicine, Int. J. Biol. Macromol., 2018, 107, 261–275.
103, 655–663. 186 T. Q. Huang, X. Qu, J. Liu and S. Chen, 3D printing of
171 M. Matsusaki, C. P. Case and M. Akashi, Three-dimensional biomimetic microstructures for cancer cell migration,
cell culture technique and pathophysiology, Adv. Drug Delivery Biomed. Microdevices, 2014, 16, 127–132.
Rev., 2014, 74, 95–103. 187 J. U. Lind, T. A. Busbee, A. D. Valentine, F. S. Pasqualini,
172 H. Wang and S. C. Heilshorn, Adaptable hydrogel networks H. Yuan, M. Yadid, S. J. Park, A. Kotikian, A. P. Nesmith,
with reversible linkages for tissue engineering, Adv. Mater., P. H. Campbell, J. J. Vlassak, J. A. Lewis and K. K. Parker,
2015, 27, 3717–3736. Instrumented cardiac microphysiological devices via multi-
173 S. R. Caliari and J. A. Burdick, A practical guide to hydro- material three-dimensional printing, Nat. Mater., 2017, 16,
gels for cell culture, Nat. Methods, 2016, 13, 405–414. 303–308.
174 A. M. Rosales and K. S. Anseth, The design of reversible 188 S. Knowlton, A. Joshi, B. Yenilmez, I. T. Ozbolat, C. K. Chua,
hydrogels to capture extracellular matrix dynamics, Nat. A. Khademhosseini and S. Tasoglu, Advancing cancer
Rev. Mater., 2016, 1, 15012. research using bioprinting for tumor-on-a-chip platforms,
175 G. A. Silva, C. Czeisler, K. L. Niece, E. Beniash, D. A. Harrington, Int. J. Bioprint., 2016, 2, 3–8.
J. A. Kessler and S. L. Stupp, Selective differentiation of neural 189 V. E. Santo, S. P. Rebelo, M. F. Estrada, P. M. Alves, E. Boghaert
progenitor cells by high-epitope density nanofibers, Science, and C. Brito, Drug screening in 3D in vitro tumor models:
2004, 303, 1352–1355. Overcoming current pitfalls of efficacy read-outs, Biotechnol. J.,
176 V. Jayawarna, M. Ali, T. A. Jowitt, A. F. Miller, A. Saiani, 2017, 12, 1600505.
J. E. Gough and R. V. Ulijn, Nanostructured hydrogels for 190 J. Norman, R. D. Madurawe, C. M. V. Moore, M. A. Khan
three-dimensional cell culture through self-assembly of and A. Khairuzzaman, A new chapter in pharmaceutical
fluorenylmethoxycarbonyl-dipeptides, Adv. Mater., 2006, manufacturing: 3D-printed drug products, Adv. Drug Delivery
18, 611–614. Rev., 2017, 108, 39–50.
177 N. Li, Q. Zhang, S. Gao, Q. Song, R. Huang, L. Wang, L. Liu, 191 V. Mironov, T. Boland, T. Trusk, G. Forgacs and R. R. Markwald,
J. Dai, M. Tang and G. Cheng, Three-dimensional graphene Organ printing: computer-aided jet-based 3D tissue engineering,
foam as a biocompatible and conductive scaffold for neural Trends Biotechnol., 2003, 21, 157–161.
stem cells, Sci. Rep., 2013, 3, 1604. 192 T. Boland, V. Mironov, A. Gutowska, E. A. Roth and R. R.
178 L. Jin, Z. Feng, T. Wang, Z. Ren, S. Ma, J. Wu and D. Sun, A Markwald, Cell and organ printing 2: Fusion of cell aggregates
novel fluffy hydroxylapatite fiber scaffold with deep in three-dimensional gels, Anat. Rec., Part A, 2003, 272,
interconnected pores designed for three dimensional cell 497–502.
culture, J. Mater. Chem. B, 2014, 2, 129–136. 193 N. E. Fedorovich, J. Alblas, W. E. Hennink, F. C. Öner and
179 R. Jacob, D. Ghosh, P. K. Singh, S. K. Basu, N. Nath Jha, W. J. A. Dhert, Organ printing: the future of bone regeneration?
S. Das, P. K. Sukul, S. Patil, S. Sathaye, A. Kumar, Trends Biotechnol., 2011, 29, 601–606.
A. Chowdhury, S. Malik, S. Sen and S. K. Maji, Self healing 194 S. V. Murphy and A. Atala, 3D bioprinting of tissues and
hydrogels composed of amyloid nano fibrils for cell culture organs, Nat. Biotechnol., 2016, 32, 773–785.
and stem cell differentiation, Biomaterials, 2015, 54, 97–105. 195 K. Markstedt, A. Mantas, I. Tournier, H. M. Ávila, D. Hägg
180 B. Zhang, M. Mongomery, L. Davenport-Huyer, A. Korolj and P. Gatenholm, 3D bioprinting human chondrocytes
and M. Radisic, Platform technology for scalable assembly with nanocellulose-alginate bioink for cartilage tissue
of instantaneously functional mosaic tissues, Sci. Adv., engineering applications, Biomacromolecules, 2015, 16,
2015, 1, e1500423. 1489–1496.
181 Q. Feng, K. Wei, S. Lin, Z. Xu, Y. Sun, P. Shi, G. Li and 196 M. S. Mannoor, Z. Jiang, T. James, Y. L. Kong, K. A. Malatesta,
L. Bian, Mechanically resilient, injectable, and bioadhesive W. O. Soboyejo, N. Verma, D. H. Gracias and M. C. McAlpine,
supramolecular gelatin hydrogels crosslinked by weak 3D printed bionic ears, Nano Lett., 2013, 13, 2634–2639.

This journal is © The Royal Society of Chemistry 2019 Mater. Horiz., 2019, 6, 45--71 | 69
View Article Online

Review Materials Horizons

197 B. Duan, E. Kapetanovic, L. A. Hockaday and J. T. Butcher, 212 W. Wang, K. Itaka, S. Ohba, N. Nishiyama, U.-i. Chung,
Three-dimensional printed trileaflet valve conduits using Y. Yamasaki and K. Kataoka, 3D spheroid culture system
biological hydrogels and human valve interstitial cells, on micropatterned substrates for improved differentiation
Acta Biomater., 2014, 10, 1836–1846. efficiency of multipotent mesenchymal stem cells, Biomaterials,
198 X. Cui and T. Boland, Human microvasculature fabrication 2009, 30, 2705–2715.
using thermal inkjet printing technology, Biomaterials, 213 W.-J. Li, R. Tuli, C. Okafor, A. Derfoul, K. G. Danielson,
2009, 30, 6221–6227. D. J. Hall and R. S. Tuan, A three-dimensional nanofibrous
199 R. Lozano, L. Stevens, B. C. Thompson, K. J. Gilmore, scaffold for cartilage tissue engineering using human
R. Gorkin, E. M. Stewart, M. in het Panhuis, M. Romero- mesenchymal stem cells, Biomaterials, 2005, 26, 599–609.
Ortega and G. G. Wallace, 3D printing of layered brain-like 214 S. U. Kim and J. de Vellis, Stem cell-based cell therapy in
structures using peptide modified gellan gum substrates, neurological diseases: A review, J. Neurosci. Res., 2009, 87,
Biomaterials, 2015, 67, 264–273. 2183–2200.
200 C. X. F. Lam, X. M. Mo, S. H. Teoh and D. W. Hutmacher, 215 V. Jeevanantham, M. Butler, A. Saad, A. Abdel-Latif, E. K.
Scaffold development using 3D printing with a starch- Zuba-Surma and B. Dawn, Adult bone marrow cell therapy
Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

based polymer, Mater. Sci. Eng., C, 2002, 20, 49–56. improves survival and induces long-term improvement in
201 B. Leukers, H. Gülkan, S. H. Irsen, S. Milz, C. Tille, cardiac parameters: A systematic review and meta-analysis,
M. Schieker and H. Seitz, Hydroxyapatite scaffolds for bone Circulation, 2012, 126, 551–568.
tissue engineering made by 3D printing, J. Mater. Sci.: 216 L. Li, H. Hui, X. Jia, J. Zhang, Y. Liu, Q. Xu and D. Zhu,
Mater. Med., 2005, 16, 1121–1124. Infusion with human bone marrow-derived mesenchymal
202 S. Yang, K. F. Leong, Z. Du and C. K. Chua, The design of stem cells improves b-cell function in patients and non-
scaffolds for use in tissue engineering. Part I. Traditional obese mice with severe diabetes, Sci. Rep., 2016, 6, 37894.
factors, Tissue Eng., 2001, 7, 679–689. 217 E. M. Horwitz, P. L. Gordon, W. K. K. Koo, J. C. Marx,
203 S. Yang, K. F. Leong, Z. Du and C. K. Chua, The design of M. D. Neel, R. Y. McNall, L. Muul and T. Hofmann,
scaffolds for use in tissue engineering. Part II. Rapid Isolated allogeneic bone marrow-derived mesenchymal
prototyping techniques., Tissue Eng., 2002, 8, 1–11. cells engraft and stimulate growth in children with osteo-
204 H. N. Chia and B. M. Wu, Recent advances in 3D printing genesis imperfecta: Implications for cell therapy of bone,
of biomaterials, J. Biol. Eng., 2015, 9, 4. Proc. Natl. Acad. Sci. U. S. A., 2002, 99, 8932–8937.
205 C. Mandrycky, Z. Wang, K. Kim and D. H. Kim, 3D 218 O. N. Koç, J. Day, M. Nieder, S. L. Gerson, H. M. Lazarus
bioprinting for engineering complex tissues, Biotechnol. and W. Krivit, Allogeneic mesenchymal stem cell infusion
Adv., 2016, 34, 422–434. for treatment of metachromatic leukodystrophy (MLD) and
206 M. C. Echave, P. Sánchez, J. L. Pedraz and G. Orive, Progress of Hurler syndrome (MPS-IH), Bone Marrow Transplant., 2002,
gelatin-based 3D approaches for bone regeneration, J. Drug 30, 215–222.
Delivery Sci. Technol., 2017, 42, 63–74. 219 J. Reiser, X. Y. Zhang, C. S. Hemenway, D. Mondal,
207 D. Baksh, R. Yao and R. S. Tuan, Comparison of proliferative L. Pradhan and V. F. La Russa, Potential of mesenchymal
and multilineage differentiation potential of human stem cells in gene therapy approaches for inherited and
mesenchymal stem cells derived from umbilical cord and acquired diseases, Expert Opin. Biol. Ther., 2005, 5, 1571–1584.
bone marrow, Stem Cells, 2007, 25, 1384–1392. 220 J. F. Enders, T. H. Weller and F. C. Robbins, Cultivation of
208 B. Chuenjitkuntaworn, T. Osathanon, N. Nowwarote, the lansing strain of poliomyelitis virus in cultures of
P. Supaphol and P. Pavasant, The efficacy of polycaprolac- various human embryonic tissues, Science, 1949, 109, 85–87.
tone/hydroxyapatite scaffold in combination with mesen- 221 S. L. Plotkin and S. A. Plotkin in Vaccines, ed. S. A. Plotkin,
chymal stem cells for bone tissue engineering, J. Biomed. W. A. Orenstein and P. A. Offit, Elsevier, Saunders, 5th edn,
Mater. Res., Part A, 2016, 104, 264–271. 2008, ch. A short history of vaccination.
209 Y. Reinwald and A. J. El Haj, Hydrostatic pressure in 222 P. N. Barrett, W. Mundt, O. Kistner and M. K. Howard,
combination with topographical cues affects the fate of Vero cell platform in vaccine production: moving towards
bone marrow-derived human mesenchymal stem cells for cell culture-based viral vaccines, Expert Rev. Vaccines, 2009,
bone tissue regeneration, J. Biomed. Mater. Res., Part A, 2018, 8, 607–618.
106, 629–640. 223 J. A. Tree, C. Richardson, A. R. Fook, J. C. Clegg and D. Looby,
210 S.-W. Kuo, H.-I. Lin, J. H.-C. Ho, Y.-R. V. Shih, H.-F. Chen, Comparison of large-scale mammalian cell culture systems
T.-J. Yen and O. K. Lee, Regulation of the fate of human with egg culture for the production of influenza virus A vaccine
mesenchymal stem cells by mechanical and stereo- strains, Vaccine, 2001, 19, 3444–3450.
topographical cues provided by silicon nanowires, Bioma- 224 G. Andrei, Three-dimensional culture models for human
terials, 2012, 33, 5013–5022. viral diseases and antiviral drug development, Antiviral
211 J. E. Frith, B. Thomson and P. G. Genever, Dynamic three- Res., 2006, 71, 96–107.
dimensional culture methods enhance mesenchymal stem 225 G. Lee, E. P. Papapetrou, H. Kim, S. M. Chambers,
cell properties and increase therapeutic potential, Tissue M. J. Tomishima, C. A. Fasano, Y. M. Ganat, J. Menon,
Eng., Part C, 2010, 16, 735–749. F. Shimizu, A. Viale, V. Tabar, M. Sadelain and L. Studer,

70 | Mater. Horiz., 2019, 6, 45--71 This journal is © The Royal Society of Chemistry 2019
View Article Online

Materials Horizons Review

Modelling pathogenesis and treatment of familial dysauto- 233 A. Astashkina, B. Mann and D. W. Grainger, A critical
nomia using patient-specific iPSCs, Nature, 2009, 461, 402–406. evaluation of in vitro cell culture models for high-throughput
226 L. Yang, M. H. Soonpaa, E. D. Adler, T. K. Roepke, S. J. drug screening and toxicity, Pharmacol. Ther., 2012, 134,
Kattman, M. Kennedy, E. Henckaerts, K. Bonham, G. W. 82–106.
Abbott, R. M. Linden, L. J. Field and G. M. Keller, Human 234 T. H. Qazi, D. J. Mooney, G. N. Duda and S. Geissler,
cardiovascular progenitor cells develop from a KDR+ embryonic- Biomaterials that promote cell–cell interactions enhance
stem-cell-derived population, Nature, 2008, 453, 524–528. the paracrine function of MSCs, Biomaterials, 2017, 140,
227 J. I. Luna, J. Ciriza, M. E. Garcia-Ojeda, M. Kong, A. Herren, 103–114.
D. K. Lieu, R. A. Li, C. C. Fowlkes, M. Khine and K. E. 235 F. Xu, J. Wu, S. Wang, N. G. Durmus, U. A. Gurkan and
McCloskey, Multiscale biomimetic topography for the U. Demirci, Microengineering methods for cell-based
alignment of neonatal and embryonic stem cell-derived microarrays and high-throughput drug-screening applications,
heart cells, Tissue Eng., Part C, 2011, 17, 579–588. Biofabrication, 2011, 3, 034101.
228 J. Wang, A. Chen, D. K. Lieu, I. Karakikes, G. Chen, 236 ditM. A. Faute, L. Laurent, D. Ploton, M.-F. Poupon,
W. Keung, C. W. Chan, R. J. Hajjar, K. D. Costa, M. Khine J.-C. Jardillier and H. Bobichon, Distinctive alterations of
Published on 01 November 2018. Downloaded on 1/20/2019 9:00:43 PM.

and R. A. Li, Effect of engineered anisotropy on the invasiveness, drug resistance and cell–cell organization in
susceptibility of human pluripotent stem cell-derived ven- 3D-cultures of MCF-7, a human breast cancer cell line, and
tricular cardiomyocytes to arrhythmias, Biomaterials, 2013, its multidrug resistant variant, Clin. Exp. Metastasis, 2002,
34, 8878–8886. 19, 161–167.
229 M. T. Lam, W. C. Clem and S. Takayama, Reversible on- 237 K. Bhadriraju and C. S. Chen, Engineering cellular micro-
demand cell alignment using reconfigurable microtopography, environments to improve cell-based drug testing, Drug
Biomaterials, 2008, 29, 1705–1712. Discovery Today, 2002, 7, 612–620.
230 M. Guvendiren and J. A. Burdick, Stem cell response to 238 A. Albanese, A. K. Lam, E. A. Sykes, J. V. Rocheleau and
spatially and temporally displayed and reversible surface W. C. W. Chan, Tumour-on-a-chip provides an optical
topography, Adv. Healthcare Mater., 2013, 2, 155–164. window into nanoparticle tissue transport, Nat. Commun.,
231 J. Webster, P. Bollen, H. Grimm and M. Jennings, Ethical 2013, 4, 2718.
implications of using the minipig in regulatory toxicology 239 L. Y. Wu, D. Di Carlo and L. P. Lee, Microfluidic self-
studies, J. Pharmacol. Toxicol. Methods, 2010, 62, 160–166. assembly of tumor spheroids for anticancer drug discovery,
232 D. R. Jung, R. Kapur, T. Adams, K. A. Giuliano, M. Mrksich, Biomed. Microdevices, 2008, 10, 197–202.
H. G. Craighead and D. L. Taylor, Topographical and 240 B. R. Hughes, M. Mirbagheri, S. D. Waldman and
physicochemical modification of material surface to enable D. K. Hwang, Direct cell–cell communication with three-
patterning of living cells, Crit. Rev. Biotechnol., 2001, 21, dimensional cell morphology on wrinkled microposts,
111–154. Acta Biomater., 2018, 78, 89–97.

This journal is © The Royal Society of Chemistry 2019 Mater. Horiz., 2019, 6, 45--71 | 71

You might also like