You are on page 1of 14

Available online at www.sciencedirect.

com

Advanced Drug Delivery Reviews 60 (2008) 215 – 228


www.elsevier.com/locate/addr

Biomaterials for stem cell differentiation☆


Eileen Dawson, Gazell Mapili, Kathryn Erickson, Sabia Taqvi, Krishnendu Roy ⁎
Department of Biomedical Engineering, The University of Texas at Austin, Austin, TX 78712, USA
Received 31 July 2007; accepted 11 August 2007
Available online 11 October 2007

Abstract

The promise of cellular therapy lies in the repair of damaged organs and tissues in vivo as well as generating tissue constructs in vitro for
subsequent transplantation. Unfortunately, the lack of available donor cell sources limits its ultimate clinical applicability. Stem cells are a natural
choice for cell therapy due to their pluripotent nature and self-renewal capacity. Creating reserves of undifferentiated stem cells and subsequently
driving their differentiation to a lineage of choice in an efficient and scalable manner is critical for the ultimate clinical success of cellular
therapeutics. In recent years, a variety of biomaterials have been incorporated in stem cell cultures, primarily to provide a conducive
microenvironment for their growth and differentiation and to ultimately mimic the stem cell niche. In this review, we examine applications of
natural and synthetic materials, their modifications as well as various culture conditions for maintenance and lineage-specific differentiation of
embryonic and adult stem cells.
© 2007 Elsevier B.V. All rights reserved.

Keywords: Embryonic stem cells; Adult stem cells; Polymers; Metals; Ceramics; Bioreactors; Differentiation

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
2. Biomaterials used for studying stem cells in 3-D culture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
2.1. Scaffolds in cell culture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
2.2. Natural materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
2.3. Synthetic materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
2.3.1. Polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
2.3.2. Ceramics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
2.3.3. Metals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
3. Modifications to biomaterials for stem cell culture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
3.1. Substrates with altered mechanical properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
3.2. Modifications with chemical and biological stimuli . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
3.3. Patterned biomaterials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
4. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225


This review is part of the Advanced Drug Delivery Reviews theme issue on
“Emerging Trends in Cell-Based Therapeutics”. 1. Introduction
⁎ Corresponding author. Department of Biomedical Engineering, The
University Texas at Austin, ENS 610, C0800, 1 University Station, Austin,
TX 78712, USA. Tel.: +1 512 232 3477. Current therapies in modern medicine mostly involve
E-mail address: kroy@mail.utexas.edu (K. Roy). prevention, manipulation and control of diseases through
0169-409X/$ - see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.addr.2007.08.037
216 E. Dawson et al. / Advanced Drug Delivery Reviews 60 (2008) 215–228

chemical or biological molecules. More recently, the restoration progenitor cells or mesenchymal stem cells (MSCs), hemato-
and direct replacement of diseased cells and tissues are poietic stem cells (HSCs) from both bone marrow and cord
becoming a clinical possibility in large part due to parallel blood as well as adipose-derived stem cells have become
advances in modifying biomaterials and understanding stem attractive cell populations for the field of tissue engineering
cell (SC) behavior [1]. Complete tissue and organ replacement because of their ability to differentiate into variety of cell types
using stem cells is still a distant milestone in which current and relative ease in harvest, isolation and expansion in vitro.
studies are laying the necessary groundwork. Recent review In this review we have focused exclusively on the use of
articles have discussed how certain types of materials are used biomaterials, both natural and synthetic, to maintain and
as substrates to mimic the physico-chemical microenvironments differentiate stem cells. Both adult and ESCs are discussed. A
of cells and tissues [2,3]. Several papers have also carefully major focus of this review is on the chemical and biological
examined materials that have been used to specifically study the modification of materials to better mimic the stem cell niche and
development of bone, cartilage, and skin [4–8]. Others have create microenvironments to control stem cell response.
reported studies on multiple pre-differentiated stem cell
populations and how to combine them together using 2. Biomaterials used for studying stem cells in 3-D culture
biomaterials to form hybrid constructs that closely mimic
native tissue [9,10]. In this review, we focus exclusively on 2.1. Scaffolds in cell culture
biomaterials and how they are revolutionizing studies in tissue
and organ replacements and providing a greater comprehension Although classical 2-D cell cultures on flat surfaces have
of the underlying mechanism behind stem cell behavior and provided us with majority of our knowledge in modern biology,
differentiation within their physiological niche. it is now well accepted that cells (including SCs) reside,
Stem cells are simply defined as a progeny of cells that have proliferate and differentiate inside the body within complex 3-D
the potential to differentiate into a variety of different lineages. microenvironments. Most of the current research in biomaterial-
These cells can be isolated from a variety of sources including directed SC manipulation is focused on such 3-D environments.
embryos, umbilical cord blood, as well as from adult tissues, as Hence this review will primarily focus on materials and
reviewed previously [11]. Since the isolation of mouse concepts that involve 3-D culture of SCs.
embryonic stem cells (mESCs) in 1981 by David and Kaufman, Biomaterial-based scaffolds have been the most important tool
and the subsequent isolation of human embryonic stem cells in providing a 3-D environment to cells, both in culture or inside
(hESCs) in 1998, their exploration in regenerative medicine the body. These 3-D structures provide an ideal platform for cell–
have received great interest. Embryonic stem cells (ESCs) are cell and cell–material communications and their properties can be
isolated from the inner cell mass of the blastocyst during varied to promote differentiation of cells into specific lineages.
embryological development [12,13]. Their pluripotent nature Scaffolds for tissue engineering serve numerous functions and
gives them the ability to differentiate into any one of the three their role during tissue development is dependent upon specific
germ layers: endoderm, ectoderm, and mesoderm. ESCs also properties of the chosen biomaterial. 3-D systems have proven to
have the unique property of indefinite self-renewal i.e. they can enhance osteogenic [16], hematopoietic [17], neural [18], and
be cultured and maintained in an undifferentiated, pluripotent chondrogenic [19,20] differentiation. They serve as biointeractive
state. Thus, ESCs have the potential of providing the biomedical stages promoting cell attachment, proliferation, and organization,
community with a continuous source of all cell types [13]. in addition to acting as delivery vehicles for bioactive molecules
Although ESCs are attractive because of their pluripotency, during tissue formation.
they are also difficult to work with because of this same Properties of biocompatible scaffolds, synthetic or natural, that
characteristic. Maintaining large number of ESCs in an undiffer- must be taken into careful consideration include optimal fluid
entiated state and subsequently directing them to differentiate, in a transport, delivery of bioactive molecules, material degradation,
reliable and reproducible manner, into specific cell types are the cell-recognizable surface chemistries, mechanical integrity and
foremost complications in ESC-based cell therapy [14]. If ESCs the ability to induce signal transduction. The overall success of
remain undifferentiated after implantation in the body, they will tissue organization and development is highly dependent upon
spontaneously differentiate into multiple cell types and form a these properties, since they can ultimately dictate cell adherence,
type of tumor called teratoma [13,15]. To avoid teratoma nutrient/waste transport, matrix synthesis, matrix organization and
formation, ESCs must be guided to differentiate into particular cell differentiation. Most scaffolding materials can be chemically
lineages prior to implantation [13]. On the other hand, ultimate and physically modified to adjust all of these critical parameters,
large scale therapeutic application of these cells necessitates their and a variety of synthetic and natural materials have been used for
maintenance in an undifferentiated state in vitro so that they can be studying SC behavior through specifically manipulating these
differentiated into specific lineages on-demand. properties. Several articles have reviewed the application of
It is now well established that adult tissues carry a variety of scaffolds in tissue engineering in general [21,22].
adult stem cells that are less pluripotent and are more committed
than ESCs. Adult stem cells are often referred to as progenitor 2.2. Natural materials
or multipotent cells since they have limited differentiation
potential. These cells have been found in bone marrow, cord Natural biomaterials used for developing scaffolds can
blood, adipose tissues, neural tissues etc. Bone-marrow derived consist of components found in the extracellular matrix
E. Dawson et al. / Advanced Drug Delivery Reviews 60 (2008) 215–228 217

(ECM), such as collagen, fibrinogen, hyaluronic acid, glycosa- Silk fibroin is another versatile natural material that is
minoglycans (GAGs), hydroxyapatite (HA) etc., and therefore isolated from silkworm cocoons. Silk has been developed into
have the advantage of being bioactive, biocompatible, and of porous scaffolds using gas foaming or salt leaching methods
similar mechanical properties as native tissue. Other natural [33–35] and is widely used as suture materials for surgical
materials include those derived from plants, insects or animal applications. The use of silk as a biomaterial for SC culture has
components (e.g. cellulose, chitosan, silk fibroin etc.) with been reviewed recently by Kaplan and colleagues [36].
properties to provide favorable microenvironments for SC Other natural materials used as scaffolds for studies in SC
culture. Disadvantages of using natural materials over synthetic differentiation include chitosan, HA, alginate and coralline.
materials include limited control over physico-chemical prop- Chitosan, isolated and processed from crustacean shells, is
erties, difficult to modify degradation rates, difficulty in hydrophilic in nature, biodegradable, biocompatible, and has
sterilization and purification as well as pathogen/viral issues similar properties to GAGs. One particular study combined
when isolating from different sources. However, in recent years, chitosan with coralline, exoskeletons of marine species or coral,
several natural materials, e.g. chitosan, hyaluronic acid, silk etc. as a composite scaffold to study MSC-osteogenesis since coral
have become commercially available, are well characterized, is composed of calcium carbonate, a component found in bone
and have reproducible, controlled properties. [37]. Another material that has been extensively used as a
Matrigel™, a product currently available commercially, is composite for bone tissue engineering, both from osteoblasts as
comprised of a variety of ECM components including laminin, well as MSCs, is HA, a natural, inorganic component of bone
collagen IV, and heparan sulfate proteoglycans [23,24] and has mineral [38]. HA has the ability to encourage bone in-growth
been used extensively in cell culture [23,25]. In addition to and the inclusion of HA particles in a nanostructured self-
Matrigel's application in normal tissue culture, when injected assembling peptide scaffold, encourages osteogenic differenti-
with cells isolated from the stromal vascular fraction (SVF) of ation of mESCs [39]. Alginate (derived from algae cell walls) is
adipose tissue (i.e. adipose-derived SCs), improved neovascu- a natural polysaccharide that has been evaluated for the
lature formation was observed in an ischemic mouse model encapsulation and differentiation of ESCs [40]. mESCs
[26]. Further studies indicated that cells pre-cultured under encapsulated in alginate poly-L-lysine (PLL) was shown to
proper conditions and then injected with Matrigel could produce support cell proliferation [41]. Moreover, this microencapsu-
tubelike vascular structures [26]. lated structure prevented embryoid body (EB) formation and
Fibrinogen and fibrin are another class of tissue-derived natural promoted differentiation towards a hepatic lineage without the
materials that can be utilized to create three-dimensional scaffold need for EB formation [41].
materials [27]. Fibrin scaffolds have been optimized for neural By crosslinking pullulan, dextran, and fucoidan at a ratio of
differentiation of mESCs [18]. These scaffolds, in conjunction 71:24:5 with a total concentration of 25% (w/v), homogenous
with various growth factors showed significant increase in neuron 3-D hydrogels could be created and stored for up to 4 weeks
and oligodendrocyte production as well as neuronal viability over [42]. CD34+ human umbilical cord blood cells (hUCBCs)
time [28]. Suggs and colleagues have also demonstrated the use of previously differentiated towards the endothelial lineage,
fibrin as a material for mouse ESC culture [29]. Chondrogenesis CD133+ human bone marrow cells (hBMCs) also differenti-
potential of MSCs derived from adipose tissue and bone marrow ated into endothelial cells, and mature endothelial cells that had
has also been evaluated in fibrin gels, and while both cell types been isolated from human saphenous veins, were all cultured
demonstrated cellular functions indicative of chondrogenic on this novel hydrogel [42]. While the scaffolds appear to
differentiation, bone marrow derived MSCs appeared to produce support cell adhesion, and lack cytotoxicity, further experi-
differentiated cells that were more efficient in both collagen II ments need to be performed before in vivo applications in
production as well as proteoglycan synthesis [27]. vascular tissue repair can be determined [42].
Hyaluronic acid is a highly attractive natural biomaterial due to A classic natural material for tissue engineering is collagen
its participation in cell behavior and cell signaling. It is present in and its derivatives. Type II collagen, isolated from bovine
tissue as a gel-like substance but can be chemically modified for cartilage, has been used to culture human mesenchymal stem
efficient processing into fibers, membranes, or microspheres. A cells (hMSCs) in pellet form when combined at the ratio of
modified type of hyaluronic acid is commercially available as 1.25 mg type II collagen per 2.5 ⁎ 105 cells (altering the ratio of
Hyaff® [30]. Hyaff®-based scaffolds are biodegradable and collagen altered the structural integrity of the pellet) [43]. The
combine both the benefit of having a 3-D microenvironment hMSCs cultured under these conditions not only assembled but
comprised of a natural material while allowing for cells to replace were able to reorganize the pellet structure as shown by the
the scaffold with their own ECM. Recently Gerecht et al. reported degree of matrix contraction [43]. In vivo results indicate that
the use of hyaluronic acid hydrogels for maintaining the that this type of 3-D culture can provide a system capable of
pluripotency and undifferentiated state of hESCs [31], and maintaining chondrogenesis [43]. Collagen microbeads have
showed that the addition of soluble growth factors to these also been studied for the expansion of hUCBCs and shown to
hydrogels successfully triggers lineage specific differentiation of not only improve expansion, as compared to a 2-D system, but
these hESCs. In addition, MSCs grown on this type of scaffolds, also improve overall cell viability [44]. Though these results
modeled to resemble a tendon, have been shown in vitro to were not statistically significant, as compared to the 2-D system,
express a variety of ligament proteins while suppressing the cells isolated from these microbeads demonstrated more
indicators of bone and cartilage differentiation [32]. favorable clonogenic ability, suggesting that collagen-beads
218 E. Dawson et al. / Advanced Drug Delivery Reviews 60 (2008) 215–228

maintained cell hematopoietic functions for a longer duration in acrylated polyanhydrides or polyesters have been widely
culture [44]. Additional in vivo results suggested that cells studied for MSC and ESC cultures [57–59]. EBs from
maintained within microbeads had the ability to engraft into the mESCs encapsulated in a poly(ethylene glycol) (PEG) hydro-
bone marrow of NOD/SCID mice even after 12 days of culture, gel, and exposed to transforming growth factor beta (TGF-β),
whereas 2-D cultured hUCBCs failed to engraft in three out of have been shown to up regulate the expression of chondrogenic
four mice [44]. markers [19]. Furthermore, chondrogenic differentiation of
mESCs using PEG-diacrylate hydrogels can be augmented with
2.3. Synthetic materials the addition of glucosamine, an amino monosacchaaride present
in GAGs [20]. Glucosoamine is known to increase the
2.3.1. Polymers production of chondrogenic proteoglycans; therefore its addi-
Polyglycolic acid (PGA), polylactic acid (PLA) and the tion to the hydrogel structure may increase the mechanical
copolymer polylactide-co-glycolide (PLGA) have been exten- properties of the scaffold by synthesizing new ECM [20].
sively used as synthetic 3-D scaffold materials for evaluating cell Recently Healy and colleagues demonstrated that photocros-
behavior [45,46]. These materials are hydrolytically degradable slinked hydrogels incorporating poly(N-isopropylacrylamide-
through bulk erosion due to the presence of ester bonds, and the co-acrylic acid) [p(NIPAAm-co-AAc)] and an acrylated matrix
glycolic/lactic acid byproducts are physiologically removed via metalloproteinase sensitive peptide Gln-Pro-Gln-Gly-Leu-Ala-
metabolic pathways. Polymer molecular weight, copolymeriza- Lys-NH2 (QPQGLAK-NH2) can support short term self
tion ratio, and polydispersity can be easily adjusted to control the renewal and maintenance of hESCs [60].
degradation rate, making these attractive synthetic materials for Though synthetic materials provides the versatility of
tissue engineering. Furthermore, standard methods (e.g., salt creating 3-D microenvironments with tunable features (i.e.,
leaching, sintering, porogen melting, and nanofiber electrospin- mechanical properties, degradation rates, porosities), disadvan-
ning) have been well established to prepare a wide variety of 3-D tages for choosing such materials include poor inherent
scaffolds using these materials [45,47,48]. bioactivity (e.g. PEG), acidic by-products (e.g. PLA or
These scaffolds have been shown to support human ESC PLGA), etc. It is thus critical to modify synthetic materials
growth as well as encourage 3-D tissue-like organization [49]. By with biological or chemical entities to achieve appropriate
using this type of degradable scaffold, it may be possible to allow cellular response (discussed in Section 3).
for the structural support necessary to control growth of ESCs
while allowing the efficient transport of nutrients and other 2.3.2. Ceramics
growth factors. As ESCs form a 3-D tissue structure, the scaffolds Calcium phosphates, bioactive glasses, and other biocera-
degrade, leaving only the differentiated cells. In such a system, mics are desirable materials for studying SC differentiation,
using appropriate growth cues, it may be possible to differentiate especially osteogenesis since these materials have favorable
ESCs into a number of different tissues without altering the initial mechanical properties and can integrate with bone to a higher
matrix material [49]. In our laboratory we have demonstrated that degree than soft biomaterials, thereby enhancing mineralization
the physical properties of the scaffold structure e.g. varying pore and matrix formation. Biphasic calcium phosphate ceramics can
sizes and polymer composition significantly influence mESC be commercially purchased (Triosite™) and MSCs grown on
differentiation into the hematopoietic lineage. Specifically, this material demonstrated continued osteoblastic phenotypic
scaffold porosity was found to be inversely related to ESC properties even after an entire month of culture [61]. The use of
hematopoiesis with an optimal pore size of less then 150 μm [50]. ceramics in tissue engineering has been extensively reviewed in
Hematopoietic progenitor cell (HPC) formation was also shown several articles [62,63].
to be directly proportional to polymer concentration in the
scaffold with optimal concentration being 20% w/v. 2.3.3. Metals
Non-woven fabrics developed from polyethylene tereptha- Titanium, an attractive material due to its inertness and
late (PET), another widely used synthetic polymer, have been biocompatibility, is used widely in orthopedic and dental
studied with MSCs and human cord blood-derived HSCs to surgery [64]. The combination of titanium along with the
evaluate seeding, proliferation, and aggregation for tissue osteogenic differentiation potential of MSCs makes for an
regeneration [51–53]. Nanofibrous scaffolds fabricated using attractive bone regeneration material. MSCs are able to attach
poly(ɛ-caprolactone) (PCL) have also been studied for and proliferate on titanium dishes [64]. Furthermore, after
adipogenesis, chondrogenesis, and osteogenesis [54,55]. appropriate differentiation media is applied to the cultures,
These electrospun polymers produce an interconnected porous titanium scaffolds are able to exhibit signs of bone matrix
matrix capable of encouraging significant cell proliferation and formation [64]. Titanium fiber meshes seeded with rat MSCs
cell–cell interactions in both MSCs and mESCs [56]. The have been cultured in a perfusion bioreactor to study effects on
introduction of adipogenic hormones into the system, resulted osteoblastic differentiation [65]. Levels of mineralization
in specific expression of a transcriptional factor, PPAR-γ, formed by osteoblasts within the scaffolds were compared
associated with adipocyte differentiation [56]. between plain titanium meshes and pre-generated bone ECM-
Acrylated polymers that form hydrogels, such as poly deposited titanium meshes. Results indicated that synergistic
(ethylene glycol) diacrylates (PEGDA) and its derivatives, poly effects of both mechanical stimulation through shear stress and
(6-aminohexyl phosphate acryloyl) ((PPE-HA)-acryl) and the presence of ECM deposition onto the scaffold substrate
E. Dawson et al. / Advanced Drug Delivery Reviews 60 (2008) 215–228 219

profoundly enhanced osteoblastic differentiation. Layer by contractile forces when integrated onto substrates and can
layer (LbL) nano-assembly, a technique which allows for the change the overall microstructure of the scaffold during tissue
surface modification of materials can fabricate layered coatings development depending on the mechanical properties of the
to modify any substrate material. Using nanoparticles of TiO2 it material [69,70]. Recently Discher and colleagues elegantly
is possible to produce a thin surface film which increases the demonstrated the importance of substrate mechanical properties
attachment of MSCs by creating a rougher surface [66]. on SC differentiation using MSCs as a model [71]. Therefore
Titanium nitride (TiN) can also be deposited on a surface in the overall mechanical integrity of scaffolding materials is a key
the form of a thin film by using a DC magnetron sputtering element that needs to be addressed when evaluating material
technique [67]. The TiN coating encouraged hMSC adherence properties that effect differentiation pathways of SCs.
as compared with traditional implant material (TiAlV) [67]. Bone marrow-derived MSCs are highly sensitive and
Another metal based material suitable for SC differentiation is responsive to mechanical stimulation in vitro [71,72]. It is
Cytomatrix™, a tantalum based scaffold material. In studies speculated that mechanical stimuli activates cell surface
performed in our laboratory, Cytomatrix™ has been shown to be receptors and focal adhesion sites, which in turn triggers
able to effectively induce EB formation in mESCs in both static intracellular signaling cascades. This leads to specific gene
conditions and dynamic culture conditions [17]. As compared to a activation that modulates ECM secretion. MSC differentiation
2-D culture, the EBs formed on Cytomatrix™ appeared to have is influenced both by their physical microenvironment e.g. by
an ECM-like structure that covered not only the EB itself, but mechanical cyclic stresses applied directly to the cells [73] as
extended onto the surface of the scaffold indicating both cell–cell well as by the inherent biomaterial properties (i.e. material
and cell–matrix interactions [17]. Additionally, EBs formed on integrity, crystallinity, crosslinking density, overall micro- and
the Cytomatrix™ scaffolds appeared to be of a smaller size, macro-porosity etc.).
avoiding the formation of aggregates usually associated with 2-D Substrates such as natural or synthetic hydrogels, closely
cultures [17]. HSC generation in 3-D cultures, especially those resemble the consistency of soft, native tissues, making them
cultured under dynamic conditions, was significantly higher as attractive scaffold materials for soft tissue engineering. On the
compared to 2-D culture and revealed a higher potential for other hand, MSC differentiation into connective tissue lineages
further differentiation into the myeloid lineage [17]. Extensive (i.e. bone, cartilage, ligaments, and tendons) require materials
cDNA microarray analysis of mESCs differentiated in 2-D as with higher mechanical strength to closely mimic the tissue
well as 3-D static and dynamic conditions demonstrated mechanical properties. However, hydrogel-like materials, can
significant differences in gene expression [68]. Cells differenti- be modified to have increased modulus of elasticity, making
ated in the different culture conditions expressed genes unique to them more suitable for applications in connective tissue
their scaffold environment as well as to their culture conditions engineering. For example, collagen gels can be adjusted to
[68]. have a higher modulus by adding HA, thereby mimicking the
composition of bone which is mostly composed of collagen
3. Modifications to biomaterials for stem cell culture fibers and phosphate minerals. Adding HA to collagen at a 1:1
ratio increases the modulus from 0.392 MPa to 0.422 MPa,
Chemical and biological modifications to biomaterials can which is comparable to trabecular bone (E = 0.443 MPa) [74].
directly influence SC behavior by altering substrate properties, Other studies have fabricated collagen composites to contain
surface interactions, scaffold degradation rate, microenviron- PLA and chitin fibers to provide increased mechanical integrity
ment architecture and ultimately manipulating the signal and have demonstrated higher human MSC attachment [75].
transduction pathways in SCs. Biomaterials can be designed Silk-based materials have also been commonly used for
to fine-tune their degradation kinetics, present specific ligand- MSC culture [76–78]. Silk exhibits higher modulus of elasticity
based signals and/or release of biological molecules in response over other natural materials, such as collagen. However de-
to their microenvironment. These influence cell–matrix inter- contamination and purification methods of silk, prior to their
actions and should lead to altered gene expression and lineage use in vivo, are extremely critical in order to avoid inflammatory
specificity. Numerous studies have demonstrated how modified and immunogenic reactions. Silk-based scaffolds seeded with
biomaterials and scaffold surfaces introduce specific biological hMSCs were shown to induce bone formation in critical-sized,
responses in SCs. Ultimately, the goal of biomaterial-directed cranial defects (larger than 4 mm) of nude mice, indicated by the
SC culture is to mimic the properties, both physical and presence of bone sialoprotein, osteopontin, and osteocalcin
biochemical, of the physiological SC niche. This section [76]. Furthermore, efficient cartilage formation was also seen
provides a comprehensive review of the various concepts in when differentiating MSCs into the chondrogenic pathway
modifying biomaterials for maintenance and differentiation of within silk scaffolds [35].
SCs. In addition to natural materials, synthetic materials can also be
chemically modified to enhance mechanical properties. For
3.1. Substrates with altered mechanical properties example, simply increasing the macromer concentration in photo-
crosslinked hydrogel scaffolds has been shown to increase the
The intrinsic mechanical properties of a biomaterial are of modulus of elasticity, such as within (PPE-HA)-acryl hydrogels
significant interest since they influence the forces exerted by [57]. A four times increase in the amount of acrylated-PEG
cells on their substrates. For instance, chondrocytes exert reacted with PPE-HA showed an almost 10-fold increase in shear
220 E. Dawson et al. / Advanced Drug Delivery Reviews 60 (2008) 215–228

modulus (3 to 26 kPa) [57]. Additionally this biodegradable, These growth factors include retinoic acid (RA), activin-A,
phosphate-based synthetic material is highly conducive to bone TGF-β, and insulin-like growth factor 1 (IGF-I). Activin-A has
tissue engineering due to the phosphate degradation product, been shown to promote the differentiation of mESC derived
which could aid in overall scaffold mineralization during EBs towards the endoderm germ layer [91]. Introducing activin-
osteogenesis [57]. A into EB cultures has been used in creating lung epithelial
Scaffolds have been increasingly applied to study the progenitor cells [91]. Both RA and TGF-β are involved in
differentiation of ESCs. While most studies focus on fluid pancreatic formation [92]. The addition of RA and activin-A
transport and other material properties of the scaffold, recent have demonstrated success in promoting in vitro differentiation
studies suggest that the mechanical stiffness may also play a key of mESCs into α, β, γ, and δ cells, all of which are pancreatic
role in controlling differentiation [49,50,82]. Studies reported by endocrine cells [92].
Battista et al. using scaffolds of collagen, fibronectin (FN) and Growth factors, hormones, and chemicals have classically
laminin, have indicated that alteration of mechanical properties been directly added into the culture medium. More recently
of scaffolds can drastically alter EB formation. When the elastic these bio-molecules have been directly incorporated within the
modulus was increased from 16 to 34 Pa the formation of EBs scaffold structure or into the scaffold biomaterial in a variety of
was severely inhibited suggesting that the increase in elastic ways. Fig. 1 provides a schematic representation of some of the
modulus resulted in an inhibition of apoptosis [82]. A second methods used in incorporating these molecules into a 3-D
possible explanation may be that the denser network of these scaffold. Soluble growth factors can be directly encapsulated or
stiffer gels may have altered ESC growth. In our lab, however, incorporated during the scaffold fabrication process [93,94] and
we were able to show similar trends with HSC differentiation has been used widely. However, in order to mimic the patterned
[50]. Our results indicated that scaffolds with higher modulus distribution of growth factors and ECM molecules in the SC
were more conductive towards HSC generation. niche, it is necessary to develop methods to sequester these bio-
Although not a focus of this review, MSC and ESC factors in a localized microenvironment to prevent their
differentiation induced by mechanical stimulation has also diffusion into other regions of the scaffolds. Heparan sulfates
been extensively investigated using bioreactors [65,79–81], or have been known to bind and protect growth factors, especially
in vitro systems that provide dynamic culturing conditions. FGF-2, in the ECM. In our studies, we have successfully
Cells in vivo consistently undergo fluid shear stress and demonstrated the effective binding of FGF-2 within photo-
mechanical strain, thereby influencing cellular interactions and polymerizable PEGDA scaffolds by covalently conjugating
responses. In our lab, the significance of mimicking this in vivo acrylated-PEG moieties to heparan sulfate [95]. Using immuno-
environment was demonstrated with the differentiation of histochemistry we effectively demonstrated that FGF-2 was
mESCs into HSCs. Cytomatrix, a porous, tantalum-based only localized or sequestered in a region of a multi-layered
biocompatible scaffold was used to culture mESCs [68]. It was scaffold that had heparin-PEG-acrylate.
shown that in comparison to 2-D culture, the 3-D tantalum A widely used method of growth factor delivery in tissue
scaffold was more efficient at promoting HSC formation [68]. engineering has been simple physical adsorption of biomole-
Furthermore, when combining 3-D environment with a dynamic cules on the biomaterial or scaffold surface. Carstens et al.,
culture in a spinner flask, the efficiency of HSC formation was using absorbable collagen sponges as delivery vehicles for
even greater (1.4–2.2 times) than that of 3-D culture alone [68]. recombinant human BMP-2 (rhBMP-2), showed in situ
osteogenesis in a craniofacial mandibular defect [96]. The
3.2. Modifications with chemical and biological stimuli chemotactic effects of rhBMP-2, attracted MSCs to the vicinity
of the implants. These cells were characterized to be spindle-
SC differentiation can be directly mediated by presenting shaped and having a pre-osteoblastic phenotype. In another
appropriate biological or chemical signals in their microenvi- study titanium fiber mesh scaffolds were coated with arginine-
ronment. It is well established that specific growth factors, glycine-aspartic acid (RGD) [97], a cell adhesive, integrin-
hormones and cytokines can enhance proliferation and lineage- binding peptide found in FN and laminin. MSCs were shown to
specific differentiation of SCs. For example, fibroblast growth attach more strongly to these RGD-coated scaffolds, however
factor-2 (FGF-2) has been shown to increase self-renewal of no change was observed in ECM secretion.
MSCs and maintenance of their multi-lineage differentiation Although protein adsorption to scaffold surfaces can be an
potential [83–86]. Additionally, bone morphogenetic proteins effective route for presentation of bioactive molecules [98,99],
(BMPs), have shown to have significant role in the regeneration desorption of the protein during culture period and the inherent
of skeletal tissues, especially bone [87,88]. poor reproducibility of adsorption processes limits its applica-
Propagation of ESCs without a feeder layer can be bility. Covalent conjugation of bioactive molecules to the
maintained in the presence of a variety of cytokines including biomaterial surface should provide a more reproducible and
leukemia inhibitory factor (LIF), stem cell factor (SCF), and controlled method of presentation since both ligand amount and
FMS-like tyrosine kinase 3 ligand (Flt3-ligand). Recently it was ligand density can be controlled. By incorporating methacrylic
also demonstrated that FGF-2 could allow long term self acid within a PEGDA macromer solution prior to photo-
renewal of hESCs and maintain their pluripotent status [89,90]. polymerizing, we have shown that the surface of 3-D hydrogel
Addition of other growth factors can induce ES differentiation scaffolds can be functionalized with carboxylic acid groups.
and ultimately formation of a variety of different tissue types. These free carboxyl groups can be subsequently activated to
E. Dawson et al. / Advanced Drug Delivery Reviews 60 (2008) 215–228 221

Fig. 1. Schematic of biochemical modification and spatial patterning of scaffolds for stem cell culture.

attach FN or other ligands [100]. Murine MSCs seeded onto Semi-interpenetrating polymer networks (sIPN) which form a
such FN-functionalized scaffolds created by an LbL micro- hydrogel material can be chemically altered to mimic ECM in a
fabrication system, and cells effectively adhered and trans- variety of ways, notably, the mechanical properties can be altered
formed into osteoblasts. as well as adding functional cell-adhesion ligands (at varied
RGD can be covalently conjugated to PEG-macromers using densities). p(NIPAAm-co-AAc) was crosslinked with Gln-Pro-
the widely used N-hydroxysuccinimidyl-ester (NHS) chemistry Gln-GLY-Leu-Ala-Lys-NH2, which can be cleaved by matrix
[101–104]. This has been reported for functionalization of metalloproteinase-13 [60]. The sIPN was further functionalized
PEGDA scaffolds for hMSC differentiation into osteoblasts with a polyacrylic acid-graft-Ac-CGGNGEPRGD-TYRAY-NH2
[105]. Cell–matrix interactions were enhanced in RGD- [p(AAc)-g-RGD] [60]. The p(AAc) chains were modified with
functionalized hydrogels leading to increased MSCs viability. (Ac-CGGNGEPRGDTYRAY-NH2) which provides an active
Nuttelman and colleagues showed that the viability of the RGD site to promote cellular adhesion [60]. hESCs cultured on
encapsulated hMSCs increases from 15% to 75% when RGD is this fully synthetic ECM replacement, functionalized with 0 to
incorporated into PEGDA hydrogels [105]. It is also possible to 150 μM of RGD complexes, were shown to be morphologically
differentiate hESCs into chondrocytic like cells in an RGD similar to hESCs cultured on an embryonic fibroblast feeder layer
modified PEGDA hydrogel [106]. The hESC-derived cells not and continued to produce markers characteristic of undifferenti-
only morphologically resembled chondrocytes but also RT-PCR ated hESCs [60].
analysis indicated that the cells expressed a number of In another study, incorporation of RGD in oligo(PEG-
chondrocytic markers and demonstrated the ability to produce fumarate) hydrogels enhanced alkaline phosphatase (ALP)
7% w/v of GAG after three weeks in culture [106]. MSCs activity of osteoblasts when compared to non-functionalized
cultured under similar conditions without RGD produced 3.5% hydrogels [107]. Shin et al. compared RGD with an
w/v of GAG accumulation [59]. osteopontin-derived peptide, and determined that hydrogels
222 E. Dawson et al. / Advanced Drug Delivery Reviews 60 (2008) 215–228

modified with the osteopontin-derived peptide triggered minimal cell toxicity, and showed gene expression of bone-
significantly higher osteoblast migration than RGD-modified specific markers [117].
hydrogels. However, RGD remained the dominating peptide for In addition to the supplementation of bio-chemical stimuli by
cell-attachment [108]. growth factors and other cytokines, it may be possible to
A study completed by Feng et al. compared the effect of stimulate alterations in the micro-environment by modifying the
adsorbed, conjugated, and soluble FN on hHSC expansion in material surface chemistry. By altering hydrophobicity of
PET non-woven scaffolds. Their study indicated that conjuga- peptides used to create scaffolds it may be possible to encourage
tion of FN directly to the scaffolds resulted in a higher a variety of cellular interactions [118]. Hydrophilicty of poly
expansion percentage of CD34+ cells [51]. Because the FN was (DL-lactide) (PDLLA), PLA, PGA and PLGA can be altered by
directly conjugated to the scaffold material, detachment of FN submerging the scaffolds in various concentrations of potassium
from the scaffold could be prevented, and perhaps aid in hydroxide [118]. As compared to non-surface treated scaffolds,
stabilizing FN in the culture system [51]. all poly(α-hydroxyl ester) scaffolds treated with potassium
In addition to presenting ligands to control differentiation via hydroxide (0.1 M) were able to support a larger number of
a covalently linked hydrogel, it may be possible to utilize other mESCs [118] indicating that for each polymer used, there may
materials to mimic the cellular environment. HSCs located in exist an optimized hydrophobicity that will best promote
the thymus receive specific microenvironmental signals direct- cellular growth.
ing them toward the T lineage [111]. Traditionally, studies Konno et al. studied the effects of electrostatic charge on
focusing on T cell differentiation employ the usage of mESCs by culturing mESCs on photoimmobilized polymers
retrovirally transfected stromal cells to present the necessary with LIF [109]. It was found that only one polymer surface,
cues to cause T lineage commitment. In our lab we have gelatin coupled with azidophenyl groups, improved the growth
employed a biomaterial based magnetic microbead system to and maintenance of mESCs [109]. The other polymers (coupled
effectively present ligands necessary to promote T cell with azidophenyl groups), poly(acrylic acid) and poly(2-
formation (e.g. the notch ligand DLL4) [111]. Mouse methacryloyl-oxyethyl phosphorylcholine-co-methacrylic
hematopoietic stem cells (mHSCs) isolated from the bone acid) (PMAc50) did not advance the growth of mESCs. In
marrow of mice and cultured with this microbead system fact, these polymers prevented cell adhesion resulting in EB
effectively produced Thy1.2+ early T cells [111]. This presents a formation, and induced cell differentiation [109]. Carbonated
versatile biomaterial-based method to quantitatively present apatite surfaces may also have the potential to induce
ligands to SCs in order to direct lineage-specific differentiation differentiation as mESCs cultured on this material proliferated
and study cellular processes during cell differentiation. in a differentiated state [110].
Scaffold mineralization during osteogenic differentiation of In addition to proteins, scaffolds can also operate as effective
MSCs can also be enhanced through chemical modification of delivery vehicles for small bioactive molecules that direct SC
the biomaterial structure. For example, adding a phosphoester growth and organization. Covalent conjugation and controlled-
group to photo-polymerizable PEG-based hydrogels not only release of dexamethasone, a synthetic corticosteroid crucial for
provides biodegradability but has also been shown to promote MSC osteogenesis in vitro, has been achieved through a
mineralization of encapsulated MSCs. The use of such hydrolytically-labile lactide component incorporated within
phosphoester-containing hydrogels significantly increases PEG hydrogels [119]. Human MSCs were encapsulated within
ALP and osteocalcin levels in differentiated cells [112,113]. It these gels, which lead to their efficient osteogenic differenti-
was shown that cell–matrix interactions and the viability of ation. The gene expression levels for two common osteogenic
encapsulated MSCs increase in the presence of phosphate markers, ALP and core binding factor alpha 1 (Cbfa1),
molecules within these hydrogel scaffolds [105]. Phosphate significantly increased in MSCs cultured in dexamethasone-
moieties contribute to the adsorption of osteopontin, a functionalized scaffolds.
sialoprotein that binds to bone mineralization and mediates Degradable thin films may also act as an efficient delivery
cell adhesion. vehicle of SCs [120]. Films coated with FN enabled adipose
Natural materials can also be chemically altered to improve derived SCs to adhere to poly(L-lactide-co-ɛ-caprolactone)
specific properties for tissue engineering. For example, Zhang films as efficiently as cells adhering to tissue culture plastic
and colleagues have created PEGylated fibrin patches to study surfaces, with no significant difference in number of cells
in vitro differentiation of MSCs into endothelial cell lineage attached even up to 24 hours after seeding [120]. The polymer
for potential use in myocardial repair [114]. Chitosan, a film itself degraded over a prolonged period of time, with a
polysaccharide, is another example of a natural biomaterial sharp increase in degradation rate at week 6 [120].
that has been modified to be thermo-responsive [115,116] by Delivery of growth factors and chemicals can also be
incorporating hydroxybutyl groups onto the polymer backbone. mediated using degradable particles, entrapped within the
The modified water soluble polymer undergoes sol–gel biomaterial scaffold, which could provide temporal release
transition when exposed to 37 °C [117]. This modified chitosan kinetics for signaling biomolecules over a prolonged period of
has been used in encapsulation and in vitro culture of hMSCs time [121–124]. Osteogenic studies of rat MSCs using
with the ultimate goal of developing an injectable cell- recombinant human TGF-β1 encapsulated in polymer blends
biomaterial composite for degenerative disk diseases. MSCs of PEG-PLGA particles (sized at an average of 158 μm) has
were effectively encapsulated within these chitosan gels, with been reported by Peter et al. [125]. Here they show that a
E. Dawson et al. / Advanced Drug Delivery Reviews 60 (2008) 215–228 223

Table 1
Application of natural biomaterials in stem cell culture
Biomaterial Chemical modifications Application in stem cell culture Cell source References
Matrigel™ Cell culture applications [22,23,25]
Vascularization mADSC [26]
Fibrin Addition of growth factors Neural differentiation mESC [18,29]
Cell culture mESC [27]
Chondrogenesis hMSC
Hyaff® (hyaluron) Ligament formation Sheep MSC [32]
Hyaluron Maintenance of pluripotency hESC [31]
Addition of soluble growth factors Lineage specific differentiation
Modified with photoreactive groups Cell proliferation hESC [31]
Application of hyaluronidase Cell removal
Silk fibroin Medical suture material N/A [33–35]
Chitosan with coralline Osteogenesis Murine MSC [37]
Hydroxyapatite (HA) Osteogenesis mESC [39]
Alginate Cell encapsulation and differentiation ESC [40]
Pullulan, Dextran, and Addition of poly-L-lysine cross-linked with sodium Hepatic possible applications to vascular repair mESC [41]
Fucoidan trimetaphosphate hESC [42]
Collagen Type II Chondrogenesis hMSC [43]
Collagen microbeads Cell expansion and viability, hUCBC [44]
Hematopoiesis
Collagen Addition of HA MSC seeding and proliferation Rat MSC [74]
Addition of crosslinked chitin and PLA Cell attachment hMSC [75]
Addition of recombinant human BMP-2 (rhBMP-2) Osteogenesis Porcine MSC [96]
Silk Osteogenesis hMSC [76]
Chondrogenesis hMSC [35]
Chitosan Conjugation of hydroxybutyl groups Potential degenerative disk therapies hMSC [117]

loading of 6.0 ng TGF-β1/ mg of microparticle provided an By creating a microwell array system using photolithography
optimal dose for growth factor delivery for enhancing MSC and plasma etching techniques, Mohr and colleagues were able
proliferation and transformation into osteoblasts [125]. Tables 1 to create spatially uniform aggregates of undifferentiated hESCs
and 2 summarizes the various natural and synthetic biomaterials without the use of MEFs [128]. After culturing in microwells,
used in SC culture, their chemical modifications (as discussed in the hESCs could be removed from the microwells and further
Section 3) as well as their specific applications. cultured under standard conditions (plates coated with Matrigel)
showing little sign of prior differentiation [128]. EBs formed
3.3. Patterned biomaterials from cells isolated from this microwell system ranged in size
from 200 to 359 μm, with a majority of EBs between 280 and
Recent developments in micro and nanofabrication techni- 359 μm (78% of the total cell population) [128]. In comparison,
ques have opened up a myriad of possibilities in studying and EBs that were derived on tissue culture polystyrene were more
controlling the differentiation of SCs. These techniques allow variable in size, with only 31% of the total cell population
for the controlled design of highly reproducible features on a falling in a range of size between 280 and 359 μm in diameter
cellular level as well as the possibility of creating spatially and [128]. This system can be further altered by creating microwells
temporally patterned scaffold structures that might ultimately of different sizes and depths in order to form a variety of EB
lead to generation of complex, hybrid tissue structures. structures allowing for the optimization of EB size [128].
ESC lineage commitment of individual cells depends on a Khademhosseini et al. reported micropatterned hydrogels of
number of variables. One important factor is the EB shape hyaluronic acid, with photoreactive methacrylates [129]. A
and size [126]. The ability to effectively control the size of composite of hyaluronic acid microwells was formed and
EBs in a reproducible manner may have a large impact on the mESCs were either “docked” in these wells or encapsulated in
ability to control and scale up ESC differentiation. One of the the hydrogel [129]. Both conditions promoted cell viability and
approaches to culturing hESCs is to create a co-culture of also illustrated how cells can be manipulated to colonize in
hESCs on top of inactivated murine embryonic fibroblasts specific patterns and shapes based on the microarchitecture of
(MEFs). Using polydimethylsiloxane (PDMS) molds, Kha- the wells [129].
demhosseini et al. were able to create microwells on a glass The microenvironment comprising complex tissues includes
surface [127]. MEFs were then seeded onto the PDMS a variety of cellular signaling moieties presented in 3-D spatial
surface and subsequently inactivated to allow for hESC seed- patterns. While many scaffolds incorporate various biomole-
ing. Results indicated that hESCs grew in highly homoge- cules into the scaffold construct, they do so in bulk, i.e. a
neous aggregates and demonstrated both high cell viability as random distribution of factors throughout the scaffold. Using
well as markers indicative of an undifferentiated cell state LbL microfabrication and a digital micro-mirror (DMD)-based
[127]. system we have incorporated precise, pre-designed spatial
224 E. Dawson et al. / Advanced Drug Delivery Reviews 60 (2008) 215–228

Table 2
Application of synthetic biomaterials in stem cell culture
Biomaterial Chemical modifications Application in stem cell Cell References
culture source
poly(ɛ-caprolactone) (PCL) Cell propagation mESC [56]
Addition of adipogenic promoting factors Adipogenesis mESC [56]
Adipogenesis hMSC [54]
Chondrogenesis and Rat [55]
Osteogenesis MSC
Poly(L-lactic acid) (PLLA) Hematopoiesis mESC [50]
Polyglycolic acid (PGA), polylactic acid, Cell proliferation and 3-D hESC [49]
(PLA), polylactide-co-glycolide (PLGA) organization
Poly(ethylene glycol) diacrylates Cell and bio-chemical Goat [59]
(PEGDA) hydrogels MSC
molecule encapsulation hMSC [58]
Addition of glucosamine Chondrogenesis mESC [20]
Incorporation of RGD-PEG-acrylates Cell viability, Osteogenesis hMSC [105]
Modified with RGD Chondrogenesis hESC [106]
Release of dexamethasone Osteogenesis hMSC [119]
poly(N-isopropylacrylamide-co-acrylic acid) Incorporated into a photocrosslinked Cell self-renewal and maintenance hESC [60]
[p(NIPAAm-co-AAc)] hydrogel with a metalloproteinase sensitive peptide
PEGDA scaffolds Incorporation of methacrylic acid Osteogenesis Murine [100]
MSC
poly(6-aminohexyl phosphate acryloyl) Cell and bio-chemical Goat [57]
(PPE-HA-acryl) molecule encapsulation MSC
Polyethylene terephthalate (PET) Cell seeding, proliferation, and aggregation Rat [53]
MSC
hMSC [52]
hHSC [51]
Conjugated with FN CD34+ proliferation hHSC [51]
Poly(ethylene glycol) (PEG) hydrogel Exposed to TGF-β Chondrogenesis mESC [19]
Addition of a phosphoester Osteogenesis Goat [112]
MSC
Biphasic calcium phosphate(Triosite™) Osteogenesis hMSC [61]
Titanium Cell attachment and Rat [64]
proliferation MSC
Differentiation media Osteogenesis Rat [64,65]
MSC
Coated with RGD MSC attachment Rat [97]
MSC
Surface modification with TiO2 Cell adherence Murine [66]
MSC
Surface modification with TiO2 Cell adherence hMSC [64,67]
Tantalum (Cytomatrix) Hematopoiesis mESC [17,66,68]
PPE-HA-acrl Increase acrylated-PEG Osteogenesis Goat [57]
MSC
sIPN p(NIPAAm-coI-AAc) crosslinked with Cell propagation hESC [60]
Gln-Pro-Gln-GLY-Leu-Ala-Lys-NH2 and
functionalized with p(AAc) and RGD complexes
Oligo(PEG-fumarate) hydrogels Modified with osteopontin-derived peptide Osteoblast migration Rat [108]
MSC
Modified with RGD Cell attachment Rat [108]
MSC
Gelatin Coupled with azidophenyl groups Cell growth mESC [109]
Magnetic microbeads T cell formation mHSC [111]
Poly(α-hydroxyl ester) scaffolds Treated with potassium hydroxide Cell growth mESC [118]
Poly(L-lactide-co-ɛ-caprolactone) films Coated with FN ESC adherence hADSC [120]
PEG-PLGA polymer blends Encapsulation of recombinant human TGF-β1 MSC proliferation and Rat [125]
osteogenesis MSC
PDMS molds Seeded with MEFs Cell viability and hESC [127]
proliferation
Microwell array system Create spatially uniform hESC [128]
aggregates of undifferentiated cells
HA microwell system Cell viability, controlled cell patterning mESC [129]
and shaping
Biomaterial microarray Time efficient cell viability and hESC [130]
differentiation studies
E. Dawson et al. / Advanced Drug Delivery Reviews 60 (2008) 215–228 225

patterning of molecules into 3-D scaffolds to create complex [6] J.Y. Lim, H.J. Donahue, Biomaterial characteristics important to skeletal
microenvironments more indicative of native tissues [95,100]. tissue engineering, J. Musculoskelet. Neuronal. Interact. 4 (2004) 396–398.
[7] J. Raghunath, H.J. Salacinski, K.M. Sales, P.E. Butler, A.M. Seifalian,
Using these techniques it may be possible to create a scaffold Advancing cartilage tissue engineering: the application of stem cell
containing regions conducive to directing a single stem technology, Curr. Opin. Biotechnol. 16 (2005) 503–509.
population to multiple, spatially organized tissues. By utilizing [8] H. Yoshikawa, A. Myoui, Bone tissue engineering with porous
ECM components capable of sequestering growth factors hydroxyapatite ceramics, J. Artif. Organs 8 (2005) 131–136.
[9] J. Elisseeff, C. Puleo, F. Yang, B. Sharma, Advances in skeletal tissue
(heparan), we demonstrated that the spatial pattern can be
engineering with hydrogels, Orthod. Craniofac. Res. 8 (2005) 150–161.
maintained during the culture period. [95]. Fig. 1 provides a [10] M.N. Rahaman, J.J. Mao, Stem cell-based composite tissue constructs for
schematic representation of such spatial patterning concept. regenerative medicine, Biotechnol. Bioeng. 91 (2005) 261–284.
[11] M. Mimeault, S.K. Batra, Concise review: recent advances on the
4. Concluding remarks significance of stem cells in tissue regeneration and cancer therapies,
Stem Cells 24 (2006) 2319–2345.
[12] N.D. Evans, E. Gentleman, J.M. Polak, Scaffolds for stem cells, Materials
The use of SCs for cellular therapy offers enormous prospect Today 9 (2006) 26–33.
in replacing damaged or lost tissue function. By utilizing SCs, [13] M.T. Mitjavila-Garcia, C. Simonin, M. Peschanski, Embryonic stem
the creation of an unlimited supply of fully differentiated cells cells: meeting the needs for cell therapy, Adv. Drug. Deliv. Rev. 57 (2005)
can be obtained, thus making cellular therapy a reality. 1935–1943.
Ultimately there are a number of obstacles researchers must [14] Y. Li, D.A. Kniss, L.C. Lasky, S.T. Yang, Culturing and differentiation of
murine embryonic stem cells in a three-dimensional fibrous matrix,
overcome before this potential is reached. In addition to Cytotechnology 41 (2003) 23–35.
overcoming the difficulty of obtaining a therapeutically [15] C. Chai, K.W. Leong, Biomaterials approach to expand and direct
meaningful number of cells, a large variety of in vivo and in differentiation of stem cells, Molec. Ther. 15 (2007) 467–480.
vitro work is left to be done prior to the application of these cells [16] E. Garreta, E. Genove, S. Borros, C.E. Semino, Osteogenic differentiation
on a clinical level. Specifically, the safety of using SCs, of mouse embryonic stem cells and mouse embryonic fibroblasts in a
three-dimensional self-assembling peptide scaffold, Tissue Eng. 12
especially when considering the usage of allogenic or even (2006) 2215–2227.
xenogenic donor cells must be carefully established before [17] H. Liu, K. Roy, Biomimetic three-dimensional cultures significantly
using SC therapy in a clinical setting [11]. increase hematopoietic differentiation efficacy of embryonic stem cells,
The cell microenvironment is known to play a significant Tissue Eng. 11 (2005) 319–330.
role in determining progenitor cell fate and function. The [18] S.M. Willerth, K.J. Arendas, D.I. Gottlieb, S.E. Sakiyama-Elbert,
Optimization of fibrin scaffolds for differentiation of murine embryonic
precise coordination of interpreting spatial and temporal cues stem cells into neural lineage cells, Biomaterials 27 (2006) 5990–6003.
from their microenvironment is highly essential for SCs to [19] N.S. Hwang, M.S. Kim, S. Sampattavanich, J.H. Baek, Z. Zhang, J.
create complex, functional tissues. Advanced and high- Elisseeff, Effects of three-dimensional culture and growth factors on the
throughput assays, such as extracellular microarrays and other chondrogenic differentiation of murine embryonic stem cells, Stem Cells
24 (2006) 284–291.
technologies extensively reviewed by Khademhosseini et al.,
[20] N.S. Hwang, S. Varghese, P. Theprungsirikul, A. Canver, J. Elisseeff,
[131–133], have uncovered specific cell interactions with ECM Enhanced chondrogenic differentiation of murine embryonic stem cells in
components and polymers that can directly stimulate SC hydrogels with glucosamine, Biomaterials 27 (2006) 6015–6023.
signaling and response. As biomaterial research advances, [21] H. Shin, S. Jo, A.G. Mikos, Biomimetic materials for tissue engineering,
new materials as well as innovations in their manipulation and Biomaterials 24 (2003) 4353–4364.
usage continue. By utilizing high throughput arrays to [22] J.L. Drury, D.J. Mooney, Hydrogels for tissue engineering: scaffold
design variables and applications, Biomaterials 24 (2003) 4337–4351.
distinguish biomaterial functionality the assessment of these [23] D.M. Bissell, D.M. Arenson, J.J. Maher, F.J. Roll, Support of cultured
materials for ultimate usage can occur in short periods of time, hepatocytes by a laminin-rich gel. Evidence for a functionally significant
wasting few materials and ultimately allowing for a more rapid subendothelial matrix in normal rat liver, J. Clin. Invest. 79 (1987) 801–812.
end product [130]. For SC based cellular therapy to be a viable [24] H.K. Kleinman, M.L. McGarvey, L.A. Liotta, P.G. Robey, K.
therapeutic option, major advances in both the understanding of Tryggvason, G.R. Martin, Isolation and characterization of type IV
procollagen, laminin, and heparan sulfate proteoglycan from the EHS
the local cues necessary for lineage commitment and the sarcoma, Biochemistry 21 (1982) 6188–6193.
biomaterials necessary to promote differentiation is necessary. [25] C. Xu, M.S. Inokuma, J. Denham, K. Golds, P. Kundu, J.D. Gold, M.K.
Carpenter, Feeder-free growth of undifferentiated human embryonic stem
References cells, Nat. Biotechnol. 19 (2001) 971–974.
[26] V. Planat-Benard, J.S. Silvestre, B. Cousin, M. Andre, M. Nibbelink, R.
[1] J.P. Vacanti, R. Langer, Tissue engineering: the design and fabrication of Tamarat, M. Clergue, C. Manneville, C. Saillan-Barreau, M. Duriez, A.
living replacement devices for surgical reconstruction and transplantation, Tedgui, B. Levy, L. Penicaud, L. Casteilla, Plasticity of human adipose
Lancet 354 (1) (1999) SI32–SI34. lineage cells toward endothelial cells: physiological and therapeutic
[2] P.K. Yarlagadda, M. Chandrasekharan, J.Y. Shyan, Recent advances and perspectives, Circulation 109 (2004) 656–663.
current developments in tissue scaffolding, Biomed. Mater. Eng. 15 (2005) [27] G.I. Im, Chondrogenesis from mesenchymal stem cells derived from
159–177. adipose tissue on the fibrin scaffold, Curr. Appl. Phys. 5 (2005) 438–443.
[3] T. Ahsan, R.M. Nerem, Bioengineered tissues: the science, the [28] S.M. Willerth, T.E. Faxel, D.I. Gottlieb, S.E. Sakiyama-Elbert, The
technology, and the industry, Orthod. Craniofac. Res. 8 (2005) 134–140. effects of soluble growth factors on embryonic stem cell differentiation
[4] A. El-Ghannam, Bone reconstruction: from bioceramics to tissue inside of fibrin scaffolds, Stem Cells 25 (2007) 2235–2244.
engineering, Expert. Rev. Med. Devices 2 (2005) 87–101. [29] H. Liu, S.F. Collins, L.J. Suggs, Three-dimensional culture for expansion
[5] R.E. Horch, J. Kopp, U. Kneser, J. Beier, A.D. Bach, Tissue engineering and differentiation of mouse embryonic stem cells, Biomaterials 27 (2006)
of cultured skin substitutes, J. Cell. Mol. Med. 9 (2005) 592–608. 6004–6014.
226 E. Dawson et al. / Advanced Drug Delivery Reviews 60 (2008) 215–228

[30] P. Brun, G. Abatangelo, M. Radice, V. Zacchi, D. Guidolin, D. Daga [51] Q. Feng, C. Chai, X.S. Jiang, K.W. Leong, H.Q. Mao, Expansion of
Gordini, R. Cortivo, Chondrocyte aggregation and reorganization into engrafting human hematopoietic stem/progenitor cells in three-dimen-
three-dimensional scaffolds, J. Biomed. Mater. Res. 46 (1999) 337–346. sional scaffolds with surface-immobilized fibronectin, J. Biomed. Mater.
[31] S. Gerecht, J.A. Burdick, L.S. Ferreira, S.A. Townsend, R. Langer, G. Res. A 78 (2006) 781–791.
Vunjak-Novakovic, Hyaluronic acid hydrogel for controlled self-renewal [52] W.L. Grayson, T. Ma, B. Bunnell, Human mesenchymal stem cells tissue
and differentiation of human embryonic stem cells, Proc. Natl. Acad. Sci. development in 3D PET matrices, Biotechnol. Prog. 20 (2004) 905–912.
U.S.A. 104 (2007) 11298–11303. [53] Y. Takahashi, Y. Tabata, Homogeneous seeding of mesenchymal stem cells
[32] S. Cristino, F. Grassi, S. Toneguzzi, A. Piacentini, B. Grigolo, S. Santi, M. into nonwoven fabric for tissue engineering, Tissue Eng. 9 (2003) 931–938.
Riccio, E. Tognana, A. Facchini, G. Lisignoli, Analysis of mesenchymal [54] W.J. Li, R. Tuli, X. Huang, P. Laquerriere, R.S. Tuan, Multilineage
stem cells grown on a three-dimensional HYAFF 11-based prototype differentiation of human mesenchymal stem cells in a three-dimensional
ligament scaffold, J. Biomed. Mater. Res. A 73 (2005) 275–283. nanofibrous scaffold, Biomaterials 26 (2005) 5158–5166.
[33] S. Hofmann, H. Hagenmuller, A.M. Koch, R. Muller, G. Vunjak- [55] M. Shin, H. Yoshimoto, J.P. Vacanti, In vivo bone tissue engineering
Novakovic, D.L. Kaplan, H.P. Merkle, L. Meinel, Control of in vitro using mesenchymal stem cells on a novel electrospun nanofibrous
tissue-engineered bone-like structures using human mesenchymal stem scaffold, Tissue Eng. 10 (2004) 33–41.
cells and porous silk scaffolds, Biomaterials 28 (2007) 1152–1162. [56] X. Kang, Y. Xie, H.M. Powell, L. James Lee, M.A. Belury, J.J. Lannutti, D.A.
[34] R. Nazarov, H.J. Jin, D.L. Kaplan, Porous 3-D scaffolds from regenerated Kniss, Adipogenesis of murine embryonic stem cells in a three-dimensional
silk fibroin, Biomacromolecules 5 (2004) 718–726. culture system using electrospun polymer scaffolds, Biomaterials 28 (2007)
[35] Y. Wang, U.J. Kim, D.J. Blasioli, H.J. Kim, D.L. Kaplan, In vitro cartilage 450–458.
tissue engineering with 3D porous aqueous-derived silk scaffolds and [57] Q. Li, J. Wang, S. Shahani, D.D. Sun, B. Sharma, J.H. Elisseeff, K.W.
mesenchymal stem cells, Biomaterials 26 (2005) 7082–7094. Leong, Biodegradable and photocrosslinkable polyphosphoester hydro-
[36] Y. Wang, H.J. Kim, G. Vunjak-Novakovic, D.L. Kaplan, Stem cell-based gel, Biomaterials 27 (2006) 1027–1034.
tissue engineering with silk biomaterials, Biomaterials 27 (2006) [58] C.R. Nuttelman, M.C. Tripodi, K.S. Anseth, In vitro osteogenic
6064–6082. differentiation of human mesenchymal stem cells photoencapsulated in
[37] M. Gravel, T. Gross, R. Vago, M. Tabrizian, Responses of mesenchymal PEG hydrogels, J. Biomed. Mater. Res. A 68 (2004) 773–782.
stem cell to chitosan-coralline composites microstructured using coralline [59] C.G. Williams, T.K. Kim, A. Taboas, A. Malik, P. Manson, J. Elisseeff,
as gas forming agent, Biomaterials 27 (2006) 1899–1906. In vitro chondrogenesis of bone marrow-derived mesenchymal stem
[38] F. Zhao, W.L. Grayson, T. Ma, B. Bunnell, W.W. Lu, Effects of hydroxyapatite cells in a photopolymerizing hydrogel, Tissue Eng. 9 (2003) 679–688.
in 3-D chitosan-gelatin polymer network on human mesenchymal stem cell [60] Y.J. Li, E.H. Chung, R.T. Rodriguez, M.T. Firpo, K.E. Healy, Hydrogels
construct development, Biomaterials 27 (2006) 1859–1867. as artificial matrices for human embryonic stem cell self-renewal,
[39] E. Garreta, D. Gasset, C. Semino, S. Borros, Fabrication of a three- J. Biomed. Mater. Res. A 79 (2006) 1–5.
dimensional nanostructured biomaterial for tissue engineering of bone, [61] J. Toquet, R. Rohanizadeh, J. Guicheux, S. Couillaud, N. Passuti, G.
Biomol. Eng. 24 (2007) 75–80. Daculsi, D. Heymann, Osteogenic potential in vitro of human bone
[40] S.K. Dean, Y. Yulyana, G. Williams, K.S. Sidhu, B.E. Tuch, marrow cells cultured on macroporous biphasic calcium phosphate
Differentiation of encapsulated embryonic stem cells after transplanta- ceramic, J. Biomed. Mater. Res. 44 (1999) 98–108.
tion, Transplantation 82 (2006) 1175–1184. [62] H. Ohgushi, A.I. Caplan, Stem cell technology and bioceramics: from cell
[41] T. Maguire, E. Novik, R. Schloss, M. Yarmush, Alginate-PLL to gene engineering, J. Biomed. Mater. Res. 48 (1999) 913–927.
microencapsulation: effect on the differentiation of embryonic stem [63] K. Rezwan, Q.Z. Chen, J.J. Blaker, A.R. Boccaccini, Biodegradable and
cells into hepatocytes, Biotechnol. Bioeng. 93 (2006) 581–591. bioactive porous polymer/inorganic composite scaffolds for bone tissue
[42] N.B. Thebaud, D. Pierron, R. Bareille, C. Le Visage, D. Letourneur, L. engineering, Biomaterials 27 (2006) 3413–3431.
Bordenave, Human endothelial progenitor cell attachment to polysac- [64] M. Maeda, M. Hirose, H. Ohgushi, T. Kirita, In vitro mineralization by
charide-based hydrogels: a pre-requisite for vascular tissue engineering, mesenchymal stem cells cultured on titanium scaffolds, J. Biochem.
J. Mater. Sci., Mater. Med. 18 (2007) 339–345. (Tokyo) 141 (2007) 729–736.
[43] C.F. Chang, M.W. Lee, P.Y. Kuo, Y.J. Wang, Y.H. Tu, S.C. Hung, Three- [65] N. Datta, Q.P. Pham, U. Sharma, V.I. Sikavitsas, J.A. Jansen, A.G. Mikos,
dimensional collagen fiber remodeling by mesenchymal stem cells requires In vitro generated extracellular matrix and fluid shear stress synergistically
the integrin–matrix interaction, J. Biomed. Mater. Res. A. 80 (2007) enhance 3D osteoblastic differentiation, Proc. Natl. Acad. Sci. U. S. A.
466–474. 103 (2006) 2488–2493.
[44] H.S. Kim, J.B. Lim, Y.H. Min, S.T. Lee, C.J. Lyu, E.S. Kim, H.O. Kim, [66] D.S. Kommireddy, S.M. Sriram, Y.M. Lvov, D.K. Mills, Stem cell
Ex vivo expansion of human umbilical cord blood CD34+ cells in a attachment to layer-by-layer assembled TiO2 nanoparticle thin films,
collagen bead-containing 3-dimensional culture system, Int. J. Hematol. Biomaterials 27 (2006) 4296–4303.
78 (2003) 126–132. [67] M. Manso-Silvan, J.M. Martinez-Duart, S. Ogueta, P. Garcia-Ruiz, J. Perez-
[45] M.J. Mondrinos, S. Koutzaki, E. Jiwanmall, M. Li, J.P. Dechadarevian, P.I. Rigueiro, Development of human mesenchymal stem cells on DC sputtered
Lelkes, C.M. Finck, Engineering three-dimensional pulmonary tissue titanium nitride thin films, J. Mater. Sci., Mater. Med. 13 (2002) 289–293.
constructs, Tissue Eng. 12 (2006) 717–728. [68] H. Liu, J. Lin, K. Roy, Effect of 3D scaffold and dynamic culture
[46] C.S. Young, H. Abukawa, R. Asrican, M. Ravens, M.J. Troulis, L.B. condition on the global gene expression profile of mouse embryonic stem
Kaban, J.P. Vacanti, P.C. Yelick, Tissue-engineered hybrid tooth and cells, Biomaterials 27 (2006) 5978–5989.
bone, Tissue Eng. 11 (2005) 1599–1610. [69] A. Subramanian, H.Y. Lin, Crosslinked chitosan: its physical properties
[47] M. Borden, M. Attawia, C.T. Laurencin, The sintered microsphere matrix and the effects of matrix stiffness on chondrocyte cell morphology and
for bone tissue engineering: in vitro osteoconductivity studies, J. Biomed. proliferation, J. Biomed. Mater. Res. A 75 (2005) 742–753.
Mater. Res. 61 (2002) 421–429. [70] J.M. Zaleskas, B. Kinner, T.M. Freyman, I.V. Yannas, L.J. Gibson, M.
[48] A.S. Lin, T.H. Barrows, S.H. Cartmell, R.E. Guldberg, Microarchitectural Spector, Contractile forces generated by articular chondrocytes in collagen-
and mechanical characterization of oriented porous polymer scaffolds, glycosaminoglycan matrices, Biomaterials 25 (2004) 1299–1308.
Biomaterials 24 (2003) 481–489. [71] A.J. Engler, S. Sen, H.L. Sweeney, D.E. Discher, Matrix elasticity directs
[49] S. Levenberg, N.F. Huang, E. Lavik, A.B. Rogers, J. Itskovitz-Eldor, R. stem cell lineage specification, Cell 126 (2006) 677–689.
Langer, Differentiation of human embryonic stem cells on three-dimensional [72] G.H. Altman, R.L. Horan, I. Martin, J. Farhadi, P.R. Stark, V. Volloch, J.C.
polymer scaffolds, Proc. Natl. Acad. Sci. U. S. A. 100 (2003) 12741–12746. Richmond, G. Vunjak-Novakovic, D.L. Kaplan, Cell differentiation by
[50] S. Taqvi, K. Roy, Influence of scaffold physical properties and stromal mechanical stress, FASEB J. 16 (2002) 270–272.
cell coculture on hematopoietic differentiation of mouse embryonic stem [73] C.A. Simmons, S. Matlis, A.J. Thornton, S. Chen, C.Y. Wang, D.J.
cells, Biomaterials 27 (2006) 6024–6031. Mooney, Cyclic strain enhances matrix mineralization by adult human
E. Dawson et al. / Advanced Drug Delivery Reviews 60 (2008) 215–228 227

mesenchymal stem cells via the extracellular signal-regulated kinase [93] J.A. Jansen, J.W. Vehof, P.Q. Ruhe, H. Kroeze-Deutman, Y. Kuboki, H.
(ERK1/2) signaling pathway, J. Biomech. 36 (2003) 1087–1096. Takita, E.L. Hedberg, A.G. Mikos, Growth factor-loaded scaffolds for
[74] L. Liu, L. Zhang, B. Ren, F. Wang, Q. Zhang, Preparation and charac- bone engineering, J. Control. Release 101 (2005) 127–136.
terization of collagen-hydroxyapatite composite used for bone tissue [94] T.P. Richardson, M.C. Peters, A.B. Ennett, D.J. Mooney, Polymeric system
engineering scaffold, Artif. Cells Blood Substit. Immobil. Biotechnol. 31 for dual growth factor delivery, Nat. Biotechnol. 19 (2001) 1029–1034.
(2003) 435–448. [95] G. Mapili, Y. Lu, S. Chen, K. Roy, Laser-layered microfabrication of
[75] X. Li, Q. Feng, W. Wang, F. Cui, Chemical characteristics and spatially patterned functionalized tissue-engineering scaffolds, J. Biomed.
cytocompatibility of collagen-based scaffold reinforced by chitin fibers for Mater. Res. B Appl. Biomater. 75 (2005) 414–424.
bone tissue engineering, J. Biomed. Mater. Res. B Appl. Biomater. 77 (2006) [96] M.H. Carstens, M. Chin, X.J. Li, In situ osteogenesis: regeneration of 10-cm
219–226. mandibular defect in porcine model using recombinant human bone
[76] L. Meinel, R. Fajardo, S. Hofmann, R. Langer, J. Chen, B. Snyder, G. morphogenetic protein-2 (rhBMP-2) and Helistat absorbable collagen
Vunjak-Novakovic, D. Kaplan, Silk implants for the healing of critical sponge, J. Craniofac. Surg. 16 (2005) 1033–1042.
size bone defects, Bone 37 (2005) 688–698. [97] H.L. Holtorf, J.A. Jansen, A.G. Mikos, Ectopic bone formation in rat
[77] L. Meinel, S. Hofmann, V. Karageorgiou, L. Zichner, R. Langer, D. marrow stromal cell/titanium fiber mesh scaffold constructs: effect of
Kaplan, G. Vunjak-Novakovic, Engineering cartilage-like tissue using initial cell phenotype, Biomaterials 26 (2005) 6208–6216.
human mesenchymal stem cells and silk protein scaffolds, Biotechnol. [98] Y. Takahashi, M. Yamamoto, Y. Tabata, Enhanced osteoinduction by
Bioeng. 88 (2004) 379–391. controlled release of bone morphogenetic protein-2 from biodegradable
[78] L. Meinel, V. Karageorgiou, S. Hofmann, R. Fajardo, B. Snyder, C. Li, L. sponge composed of gelatin and beta-tricalcium phosphate, Biomaterials
Zichner, R. Langer, G. Vunjak-Novakovic, D.L. Kaplan, Engineering 26 (2005) 4856–4865.
bone-like tissue in vitro using human bone marrow stem cells and silk [99] X.B. Yang, H.I. Roach, N.M. Clarke, S.M. Howdle, R. Quirk, K.M.
scaffolds, J. Biomed. Mater. Res. A 71 (2004) 25–34. Shakesheff, R.O. Oreffo, Human osteoprogenitor growth and differen-
[79] M.E. Gomes, C.M. Bossano, C.M. Johnston, R.L. Reis, A.G. Mikos, tiation on synthetic biodegradable structures after surface modification,
In vitro localization of bone growth factors in constructs of bio- Bone 29 (2001) 523–531.
degradable scaffolds seeded with marrow stromal cells and cultured in [100] Y. Lu, G. Mapili, G. Suhali, S. Chen, K. Roy, A digital micro-mirror device-
a flow perfusion bioreactor, Tissue Eng. 12 (2006) 177–188. based system for the microfabrication of complex, spatially patterned tissue
[80] H.L. Holtorf, N. Datta, J.A. Jansen, A.G. Mikos, Scaffold mesh size affects engineering scaffolds, J. Biomed. Mater. Res. A 77 (2006) 396–405.
the osteoblastic differentiation of seeded marrow stromal cells cultured in a [101] B.K. Brandley, R.L. Schnaar, Covalent attachment of an Arg-Gly-Asp
flow perfusion bioreactor, J. Biomed. Mater. Res. A 74 (2005) 171–180. sequence peptide to derivatizable polyacrylamide surfaces: support of fibro-
[81] H.L. Holtorf, J.A. Jansen, A.G. Mikos, Flow perfusion culture induces blast adhesion and long-term growth, Anal. Biochem. 172 (1988) 270–278.
the osteoblastic differentiation of marrow stroma cell-scaffold constructs [102] D.L. Hern, J.A. Hubbell, Incorporation of adhesion peptides into
in the absence of dexamethasone, J. Biomed. Mater. Res. A 72 (2005) nonadhesive hydrogels useful for tissue resurfacing, J. Biomed. Mater.
326–334. Res. 39 (1998) 266–276.
[82] S. Battista, D. Guarnieri, C. Borselli, S. Zeppetelli, A. Borzacchiello, L. [103] U. Hersel, C. Dahmen, H. Kessler, RGD modified polymers: biomaterials for
Mayol, D. Gerbasio, D.R. Keene, L. Ambrosio, P.A. Netti, The effect of stimulated cell adhesion and beyond, Biomaterials 24 (2003) 4385–4415.
matrix composition of 3D constructs on embryonic stem cell differen- [104] R.L. Schnaar, B.G. Langer, B.K. Brandley, Reversible covalent
tiation, Biomaterials 26 (2005) 6194–6207. immobilization of ligands and proteins on polyacrylamide gels, Anal.
[83] G. Bianchi, A. Banfi, M. Mastrogiacomo, R. Notaro, L. Luzzatto, R. Biochem. 151 (1985) 268–281.
Cancedda, R. Quarto, Ex vivo enrichment of mesenchymal cell progenitors [105] C.R. Nuttelman, M.C. Tripodi, K.S. Anseth, Synthetic hydrogel niches
by fibroblast growth factor 2, Exp. Cell Res. 287 (2003) 98–105. that promote hMSC viability, Matrix Biol. 24 (2005) 208–218.
[84] I. Martin, A. Muraglia, G. Campanile, R. Cancedda, R. Quarto, Fibroblast [106] N.S. Hwang, S. Varghese, Z. Zhang, J. Elisseeff, Chondrogenic
growth factor-2 supports ex vivo expansion and maintenance of osteogenic differentiation of human embryonic stem cell-derived cells in arginine-
precursors from human bone marrow, Endocrinology 138 (1997) glycine-aspartate-modified hydrogels, Tissue Eng. 12 (2006) 2695–2706.
4456–4462. [107] H. Shin, J.S. Temenoff, G.C. Bowden, K. Zygourakis, M.C. Farach-
[85] L.A. Solchaga, K. Penick, J.D. Porter, V.M. Goldberg, A.I. Caplan, J.F. Carson, M.J. Yaszemski, A.G. Mikos, Osteogenic differentiation of rat
Welter, FGF-2 enhances the mitotic and chondrogenic potentials of bone marrow stromal cells cultured on Arg-Gly-Asp modified hy-
human adult bone marrow-derived mesenchymal stem cells, J. Cell. drogels without dexamethasone and beta-glycerol phosphate, Bioma-
Physiol. 203 (2005) 398–409. terials 26 (2005) 3645–3654.
[86] S. Tsutsumi, A. Shimazu, K. Miyazaki, H. Pan, C. Koike, E. Yoshida, K. [108] H. Shin, K. Zygourakis, M.C. Farach-Carson, M.J. Yaszemski, A.G. Mikos,
Takagishi, Y. Kato, Retention of multilineage differentiation potential of Attachment, proliferation, and migration of marrow stromal osteoblasts
mesenchymal cells during proliferation in response to FGF, Biochem. cultured on biomimetic hydrogels modified with an osteopontin-derived
Biophys. Res. Commun. 288 (2001) 413–419. peptide, Biomaterials 25 (2004) 895–906.
[87] A.H. Reddi, N.S. Cunningham, Initiation and promotion of bone [109] T. Konno, N. Kawazoe, G. Chen, Y. Ito, Culture of mouse embryonic stem
differentiation by bone morphogenetic proteins, J. Bone Miner. Res. cells on photoimmobilized polymers, J. Biosci. Bioeng. 102 (2006) 304–310.
8 (2) (1993) S499–S502. [110] A.J. Melville, J. Harrison, K.A. Gross, J.S. Forsythe, A.O. Trounson, R.
[88] J.M. Wozney, The bone morphogenetic protein family and osteogenesis, Mollard, Mouse embryonic stem cell colonisation of carbonated apatite
Mol. Reprod. Dev. 32 (1992) 160–167. surfaces, Biomaterials 27 (2006) 615–622.
[89] M.E. Levenstein, T.E. Ludwig, R.H. Xu, R.A. Llanas, K. VanDenHeuvel- [111] S. Taqvi, L. Dixit, K. Roy, Biomaterial-based notch signaling for the
Kramer, D. Manning, J.A. Thomson, Basic fibroblast growth factor support differentiation of hematopoietic stem cells into T cells, J. Biomed. Mater.
of human embryonic stem cell self-renewal, Stem Cells 24 (2006) 568–574. Res. A 79 (2006) 689–697.
[90] R.H. Xu, R.M. Peck, D.S. Li, X. Feng, T. Ludwig, J.A. Thomson, Basic [112] D.A. Wang, C.G. Williams, F. Yang, N. Cher, H. Lee, J.H. Elisseeff,
FGF and suppression of BMP signaling sustain undifferentiated Bioresponsive phosphoester hydrogels for bone tissue engineering,
proliferation of human ES cells, Nat. Methods 2 (2005) 185–190. Tissue Eng. 11 (2005) 201–213.
[91] H.J. Rippon, J.M. Polak, M. Qin, A.E. Bishop, Derivation of distal lung [113] J. Wang, H.Q. Mao, K.W. Leong, A novel biodegradable gene carrier
epithelial progenitors from murine embryonic stem cells using a novel based on polyphosphoester, J. Am. Chem. Soc. 123 (2001) 9480–9481.
three-step differentiation protocol, Stem Cells 24 (2006) 1389–1398. [114] G. Zhang, X. Wang, Z. Wang, J. Zhang, L. Suggs, A PEGylated fibrin
[92] M. Nakanishi, T.S. Hamazaki, S. Komazaki, H. Okochi, M. Asashima, patch for mesenchymal stem cell delivery, Tissue Eng. 12 (2006) 9–19.
Pancreatic tissue formation from murine embryonic stem cells in vitro, [115] J.H. Cho, S.H. Kim, K.D. Park, M.C. Jung, W.I. Yang, S.W. Han, J.Y.
Differentiation 75 (2007) 1–11. Noh, J.W. Lee, Chondrogenic differentiation of human mesenchymal
228 E. Dawson et al. / Advanced Drug Delivery Reviews 60 (2008) 215–228

stem cells using a thermosensitive poly(N-isopropylacrylamide) and [124] M.C. Wake, P.D. Gerecht, L. Lu, A.G. Mikos, Effects of biodegradable
water-soluble chitosan copolymer, Biomaterials 25 (2004) 5743–5751. polymer particles on rat marrow-derived stromal osteoblasts in vitro,
[116] E. Ruel-Gariepy, M. Shive, A. Bichara, M. Berrada, D. Le Garrec, A. Biomaterials 19 (1998) 1255–1268.
Chenite, J.C. Leroux, A thermosensitive chitosan-based hydrogel for the [125] S.J. Peter, L. Lu, D.J. Kim, G.N. Stamatas, M.J. Miller, M.J. Yaszemski,
local delivery of paclitaxel, Eur. J. Pharm. Biopharm. 57 (2004) 53–63. A.G. Mikos, Effects of transforming growth factor beta1 released from
[117] J.M. Dang, D.D. Sun, Y. Shin-Ya, A.N. Sieber, J.P. Kostuik, K.W. Leong, biodegradable polymer microparticles on marrow stromal osteoblasts
Temperature-responsive hydroxybutyl chitosan for the culture of mesen- cultured on poly(propylene fumarate) substrates, J. Biomed. Mater. Res.
chymal stem cells and intervertebral disk cells, Biomaterials 27 (2006) 50 (2000) 452–462.
406–418. [126] E.S. Ng, R.P. Davis, L. Azzola, E.G. Stanley, A.G. Elefanty, Forced
[118] J. Harrison, S. Pattanawong, J.S. Forsythe, K.A. Gross, D.R. Nisbet, H. aggregation of defined numbers of human embryonic stem cells into
Beh, T.F. Scott, A.O. Trounson, R. Mollard, Colonization and embryoid bodies fosters robust, reproducible hematopoietic differentia-
maintenance of murine embryonic stem cells on poly(alpha-hydroxy tion, Blood 106 (2005) 1601–1603.
esters), Biomaterials 25 (2004) 4963–4970. [127] A. Khademhosseini, L. Ferreira, J. Blumling III, J. Yeh, J.M. Karp, J.
[119] C.R. Nuttelman, M.C. Tripodi, K.S. Anseth, Dexamethasone-functiona- Fukuda, R. Langer, Co-culture of human embryonic stem cells with
lized gels induce osteogenic differentiation of encapsulated hMSCs, J. murine embryonic fibroblasts on microwell-patterned substrates, Bioma-
Biomed. Mater. Res. A 76 (2006) 183–195. terials 27 (2006) 5968–5977.
[120] C.A. Burks, K. Bundy, P. Fotuhi, E. Alt, Characterization of 75:25 poly [128] J.C. Mohr, J.J. de Pablo, S.P. Palecek, 3-D microwell culture of human
(L-lactide-co-epsilon-caprolactone) thin films for the endoluminal embryonic stem cells, Biomaterials 27 (2006) 6032–6042.
delivery of adipose-derived stem cells to abdominal aortic aneurysms, [129] A. Khademhosseini, G. Eng, J. Yeh, J. Fukuda, J. Blumling III, R. Langer, J.
Tissue Eng. 12 (2006) 2591–2600. A. Burdick, Micromolding of photocrosslinkable hyaluronic acid for cell
[121] T.A. Holland, Y. Tabata, A.G. Mikos, Dual growth factor delivery from encapsulation and entrapment, J. Biomed. Mater. Res. A 79 (2006) 522–532.
degradable oligo(poly(ethylene glycol) fumarate) hydrogel scaffolds for [130] D.G. Anderson, S. Levenberg, R. Langer, Nanoliter-scale synthesis of
cartilage tissue engineering, J. Control. Release 101 (2005) 111–125. arrayed biomaterials and application to human embryonic stem cells, Nat.
[122] Z. Lalani, M. Wong, E.M. Brey, A.G. Mikos, P.J. Duke, Spatial and Biotechnol. 22 (2004) 863–866.
temporal localization of transforming growth factor-beta1, bone [131] D.G. Anderson, D. Putnam, E.B. Lavik, T.A. Mahmood, R. Langer,
morphogenetic protein-2, and platelet-derived growth factor-A in Biomaterial microarrays: rapid, microscale screening of polymer–cell
healing tooth extraction sockets in a rabbit model, J. Oral Maxillofac. interaction, Biomaterials 26 (2005) 4892–4897.
Surg. 61 (2003) 1061–1072. [132] C.J. Flaim, S. Chien, S.N. Bhatia, An extracellular matrix microarray for
[123] H. Park, J.S. Temenoff, T.A. Holland, Y. Tabata, A.G. Mikos, Delivery of probing cellular differentiation, Nat. Methods 2 (2005) 119–125.
TGF-beta1 and chondrocytes via injectable, biodegradable hydrogels for [133] A. Khademhosseini, R. Langer, J. Borenstein, J.P. Vacanti, Microscale
cartilage tissue engineering applications, Biomaterials 26 (2005) technologies for tissue engineering and biology, Proc. Natl. Acad. Sci. U. S. A.
7095–7103. 103 (2006) 2480–2487.

You might also like