Open access peer-reviewed chapter

Pluripotent Stem Cells and Their Dynamic Niche

Written By

Yvonne Reinwald, Jessica Bratt and Alicia El Haj

Submitted: 06 November 2015 Reviewed: 23 February 2016 Published: 20 July 2016

DOI: 10.5772/62671

From the Edited Volume

Pluripotent Stem Cells - From the Bench to the Clinic

Edited by Minoru Tomizawa

Chapter metrics overview

2,750 Chapter Downloads

View Full Metrics

Abstract

Cell-seeded implants are a regenerative medicine strategy that aims to replace injured tissue and restore tissue function. Pluripotent stem cells are promising cell candidates for the development of regenerative medicine therapies as they have the ability to self-renew and commit towards numerous cell types. In vivo, stem cells reside in a dynamic niche, a stem cell-specific microenvironment that possesses chemical, biological and mechanical cues, which drive the stem cell fate and renewal. The connection between stem cells and their niche is a two-way relationship consisting of both cell–cell interaction and cell–extracellular matrix (ECM) interactions. An alternative regenerative medicine approach is the manipulation of the stem cell microenvironment. Hence, novel strategies have been developed including 3D biomaterials and bioreactor technologies providing topographical, chemical and mechanical cues to recreate the stem cell niche. Understanding the mechanisms controlling stem cell fate and the dynamic nature of the stem cell niche will enable researchers to replicate this stem cell-specific microenvironment, and therefore, harness and control the valuable attributes which stem cells possess. This chapter elucidates the importance of pluripotent stem cells and their dynamic niche in regenerative medicine. It further presents novel strategies to replicate chemical, topographical and mechanical stimuli which are essential for the regulation of stem cell fate and hence tissue regeneration.

Keywords

  • Pluripotent stem cells
  • Regenerative medicine
  • Stem cell niche
  • Mechanical cues
  • Bioreactor technologies

1. Introduction

Regenerative medicine is a promising field that aims to develop therapies for currently intractable diseases. These approaches include cell–cell and cell–scaffold implants. Self-renewal and the ability of pluripotent stem cells to commit towards particular cell lineages in response to mechanical, chemical and physical stimuli makes them the ideal building block for such therapies. Embryonic stem cells (ESCs) are isolated during embryological development. Their utilisation in regenerative medicine is controversial, and therefore, adult stem cells have been more thoroughly investigated for their potential use for tissue regeneration. Currently, induced pluripotent stem cells (iPSCs) are being investigated as an alternative source for pluripotent stem cells due to their origin and clinical potential [1].

Stem cells in vivo reside in a dynamic, cell type-specific microenvironment, the so-called niche [25]. This microenvironment is composed of stem cells, supportive stromal cells and surrounding extracellular matrix (ECM) [5, 6]. The niche provides chemical, mechanical and topographical cues facilitating stem cell renewal and controlling stem cell fate [3, 4]. The ECM is an important component of the niche. It is the biological matrix directly surrounding stem cells in vivo and is composed of tissue-specific glycosaminoglycan (GAG), insoluble proteins (e.g. fibronectins, collagens, laminins) and inorganic hydroxyapatites [6, 7]. Research has shown that the biophysical properties of the ECM affect stem cell behaviour. Cells counteract external forces which result from their surroundings, through adjacent cells and ECM stiffness by altering the cytoskeleton tension and through the generation of internal forces [8]. These forces are exerted on and from each individual cell to their environment. The interaction between microenvironment and cells results in the regulation of stem cell behaviour [810]. The cellular response to external mechanical cues is defined as mechanotransduction.

An alternative strategy for tissue repair is the manipulation of the stem cell microenvironment to enable tissue repair through endogenous stem cells [11]. To recreate the stem cell niche, it is important to reproduce the physical and mechanical microenvironment stem cells experience in vivo. Hence, strategies including novel bioreactor technologies have been developed. By replicating the in vivo environment, bioreactors allow the study of mechanical stimuli in combination with chemical and biological signals on cell–cell and cell–biomaterial constructs [12].

This chapter will present the role of pluripotent stem cells and their dynamic niche in regenerative medicine, as well as the importance for the niche replication for the development of novel regenerative therapies. Finally, it will present novel strategies to replicate chemical, biological and mechanical stimuli which are essential for the regulation of stem cell fate and hence tissue regeneration.

Advertisement

2. Stem cell niches

In 1978, the term ‘niche’ was first described by Schofield [13] who theorised that stem cell self-renewal and character is dependent on their environment. Since then, the theory of the stem cell microenvironment has been expanded [14, 15]. Each stem cell population has a unique and specific environment, but there are features which stem cell niches have in common (Figure 1). Tissue-specific (e.g. osteoblasts) and non-specific (stromal cells) heterogeneous cell–cell interactions co-exist in stem cell niches [11]. Secreted-membrane bound factors bind stem cell surface receptors in order to direct stem cell self-renewal and fate (e.g. Wnt, chemokines, Notch, SCF) [1622]. During tissue injury and inflammation, immunological cells regulate the niche [23, 24]. The ECM provides structural support and orientation and serves as storage for soluble factors [25]. It further interacts with stem cells through gap junctions, soluble factors and surface receptors [5, 6]. In addition, physical and physiological parameters of the niche such as shape, elasticity, blood flow and oxygen tension influence stem cell differentiation and self-renewal and regulate the metabolic activity [22, 2632]. The interactions between stem cells and their niche are reciprocal; since stem cells are able to remodel the niche and secrete ECM components in response to the signals they receive from it [3335].

Figure 1.

Components of the stem cell niche. Niches are complex dynamic heterogenous microenvironments containing various cell types, extracellular matrix, soluble factors. The stem cell niche is influenced by a variety of factors including physical and metabolic parameters. The ECM and supportive stromal cells interact with stem cells through gap junction, soluble factors and surface receptors. Systematic signals are carried into the niche by blood vessels to facilitate the recruitment of inflammatory cells and neural signals convey distant physiological cues, such as shear stress, tissue stiffness and oxygen tension, to the stem cell niche. Image adapted with the permission from Lane et al. [11].

2.1. Stem cell niches: role

Niches have specific anatomical locations and form unique stem cells surroundings in vivo. This microenvironment regulates pathophysiological and physiological processes and directs cellular fate and function (Figure 2). The niche provides extracellular signals that maintain a balance between stem cell self-renewal and differentiation enabling stem cells to preserve a dormant and low metabolic state in order to avoid stem cell exhaustion and the accumulation of gene mutations which might result in their transformation into cancer cells [4, 36, 37]. It has been shown that the destabilisation of the stem cell environment is involved in diseases connected to aging, tumorigenesis and degeneration [38]. In adjacent sites within the same tissue, stem cells co-exist in either quiescent or active state [39, 37]. For maintaining stem cell number and to meet the needs for differentiated cells in neighbouring tissues, the balance between asymmetric and symmetric stem cell division is essential [3]. Thus, in order to maintain a healthy stem cell pool, a crosstalk between tissue necessity and state is created by the stem cell niche [40, 41].

Figure 2.

The stem cell microenvironment controls pathophysiological and physiological processes. Image adapted with the permission from Ou et al. [48].

2.2. Extracellular matrix

Due to its dynamic and diverse composition, the ECM provides structural and mechanical support and gives biochemical and physical characteristics to different stem cell niches, which are required for tissue morphogenesis and homeostasis as well as to facilitate stem cell renewal and control stem cell fate [35, 7]. It provides scaffolding to cells and stores soluble growth factors [11, 42, 43]. In addition, secreted or cell surface factors, signalling cascades and gradients, as well as physical factors, such as shear stress, oxygen tension and temperature, contribute to control stem cell behaviour in a well-orchestrated manner [40, 41].

The ECM is mainly composed of water, proteins and polysaccharides with each tissue exhibiting its unique composition and topology [42]. This diversity is caused by a combination of specific molecular interactions between ratios, geometries and isoforms of its components [7]. Two main classes of macromolecules are found in the ECM. These are fibrous proteins (e.g. collagens, elastins, fibronectins and laminins) and proteoglycans (e.g. heparin sulphate, chondroitin sulphate, ketaran sulphate) [44, 45].

The most plentiful fibrous protein in the interstitial ECM is collagen. It contributes to approximately 30% of the total protein amount and provides tensile strength it controls chemotaxis, controls cell attachment and migration as well as influences tissue development [46]. In any given tissue, collagen is present as a non-homogenous mixture of collagen fibre types; however, one type is normally dominant (e.g. in bone: collagen type-I, cartilage: collagen type-II) [42]. In close connection to collagen are elastin fibres which recoil when tissues undergo repeated stretching. In fact, the degree of association with collagen is a limiting factor for tissue elasticity [47]. Besides directing the organisation of the interstitial ECM fibronectin also influences cell adhesion and migration [42, 46]. The majority of the extracellular part within tissues is filled by proteoglycans consisting of GAG chains, which are linked to a protein core and are classified according to the GAG and core protein arrangement [45]. These proteoglycans have numerous different purposes that mirror their hydration, binding, buffering and force-resistance properties [42, 48].

Furthermore, the ECM can be divided into two forms, the basal membrane (BM) and the interstitial matrix [49] both of which are composed of a collagen framework with glycoproteins (non-structural proteins) adhering to it to communicate with adjacent cells via integrins [50].

2.3. Niche cells and cell communication

Stem cell niches in adult tissues are populated by a number of different cell types with each having a particular function. This is shown in the adult hematopoietic stem cell (HSC) niche, which is located along the endosteal surface of trabecular bone, in close vicinity of osteoblasts and the endothelial cells of blood vessels [51]. Osteoblasts in the endosteal niche regulate the HSC number [51, 52] and preserve their quiescence by releasing signalling molecules [5355]. The HSC niche is further inhabited by cell types including stromal cells, bone marrow adipocytes, osteal macrophages, CXCL12-abundant reticular (CAR) cells, nestin-positive mesenchymal stem cells, nestin-positive Schwann cells, endothelial cells [19, 5659].

Besides cells, which reside permanently in the niche such as nerve cells, endothelial cells and connective tissue fibroblasts, cells of the innate and adaptive immune system and cells that are important for the repair of damaged tissues and promote protection against pathogens are also present [11].

When stem cells undergo cell death or apoptosis neighbouring cells belonging to the niche undergo dedifferentiation in order to replace the lost stem cells [11]. For example, the removal of hair follicle stem cells in mice results in the repopulation of the niche by epithelial cells which then sustain the hair renewal [60]. This reprogramming of endogenous differentiated cells into stem cell controlled by the stem cell niche [16], and the maintenance of stem cell sources [61] might hold important clinical promise [11]. The communication between niche cells and stem cells is facilitated either by indirect contact through the secretion of molecules or directly through physical cell–cell interactions. The direct cell–cell contact is mediated by cell–cell adhesion molecules and receptors with membrane-bound ligands [11]. Indirect communication between cells in the HSC niche is clinically used to modify the HSC niche in vivo. For the treatment of bone marrow failure, genetic disorders and haematological malignancy cytokines (G-CSF, GM-CSF) are administered which result in the activation of hematopoietic stem cells, their expansion as well as diminished adhesion to their niche [11].

2.4. Physical and physiological parameters

Stem cell fate is greatly influenced by physical and physiological parameters (Figure 2). Niche topographies modify the stem cell cytoskeletal resulting in the activation of specific signalling pathways and stem cell differentiation [29, 62]. The modification of physical parameters such as substrate stiffness or elasticity, shape and shear forces has been utilised for clinical treatments but also for in vitro hematopoietic stem cell culture [26, 27]. Bone tumour and osteoporosis are treated clinically using drugs that change the balance between physical factors, for example rigid (bone) and elastic (arteriolar, dermal connective tissue). Shear forces and drugs which are used to promote blood flow for the development of embryonic HSC in vivo [28]. Physiological factors such as oxygen tension are important contributors for cell survival and maintenance. Many cells such as HSC inhabit hypoxic microenvironments [30]. Cells in these niches are carried out the glycolysis and express high levels of hypoxia inducible factor 1α. Growing mammalian cells under hypoxic conditions positively influences cell proliferation, survival and function after engraftment [31].

Advertisement

3. Stem cells in regenerative medicine

Medical conditions such as tissue loss, organ failure, cancer abrasion, congenital structural anomalies can already be treated by clinical procedures such as autologous and allogenic organ transplantation and the use of artificial implants. However, these treatments are limited by organ shortages, impairment of healthy tissue during surgery and immune rejections. Breakthroughs in the field of regenerative medicine may enable the utilisation of stem cells and stem cell-based therapies for the restoration of tissue function [63].

Stem cells can be distinguished by their potency (multipotent versus pluripotent) and through their tissue source, that is ESCs, foetal stem cells (FSC), adult stem cells such as mesenchymal stem cells (MSCs) and induced pluripotent stem cells (iPSC) [64, 65] (Figure 3). Stem cells are characterised by their ability to self-renew without senescence for extended culture periods and the potential to differentiate into multiple cell types [66]. Through their pluripotency and multipotency stem cells offer vast cell sources making them ideal for studying degenerative diseases and developing cell-based therapies [64, 65]. They have the capability to generate every tissue type and are essential to human development. However, due to regulatory, ethical and technical considerations involving genetic modification and cell isolation adult stem cells such as MSCs and adipose stem cells have been widely investigated as an alternative cell source to ESCs [63].

Figure 3.

Stem cell types. (a) Embryonic stem cells derived from blastocysts were the first human pluripotent stem cells that could be differentiated to generate various cell types. (b) Induced pluripotent stem cells were first generated by reprogramming somatic adults cells such as skin fibroblasts. (c) Tissue-specific adult stem cells are primed for generating progeny that differentiate into specialized cell types (e.g. stem cells residing in the muscle). Image adapted with the permission from Lutolf et al. [225].

Tissue-specific adult stem cells which can be found in various niches such as the bone marrow or adipose tissue are multipotent and play a crucial role in tissue development, repair and growth [6770]. Regenerative medicine approaches envision that these somatic stem cells could theoretically be harvested from the patient, be differentiated in vitro and injected back into the patient to regenerate impaired tissues without the need to suppress the immune system. Autologous chondrocyte implantation (ACI) for the repair of chondral defects in joints such as the knee is such a regenerative approach. In this two-stage process, the patient’s chondrocytes are expanded in vitro and the subsequently implanted back into the patient [7173].

Yet, low cell numbers, complications isolating adult stem cells from healthy tissues well as limited differentiation potential may require other stem cell types [66]. Latest developments indicate that directed differentiation and the development of differentiation protocols could result in the translation of cell therapies for presently incurable diseases [7477]. The availability of patient-specific cells in required quantities and on demand could revolutionise stem cell therapies and further allow disease modelling and human pluripotent stem cell (hPSC)-based drug discovery [65].

3.1. Pluripotent stem cells

3.1.1. ESCs and FSCs

When ESCs were first isolated, they provided a model for the study of developmental biology but also opened up the possibility to exploit their pluripotency for stem cell-based therapy in order to treat organ damage or dysfunction allowing regenerative medicine to become a reality for the treatment of various diseases [66]. ESCs are harvested from the inner cell mass of the blastocyst-stage embryo, a hollow sphere of cells which is composed of an outer cell layer forming the placenta and an inner cell mass from which the ESCs are derived [66]. ESCs have the ability to proliferate, to maintain an undifferentiated phenotype for extended periods of time in culture and to develop into a large number of somatic cell types [7880]. Early studies utilising ESCs were aimed at treating traumatic injuries of the CNS and degenerative diseases [8184]. Implanting early stage and differentiated ESC into laboratory animals has been shown to improve function, behaviour and morphology, but also to cause teratoma formation and hyperproliferation [85]. The application of ESC has also shown great promise for the regeneration of cartilage, cardiac tissue and peripheral nerves [7689].

The utilisation of tissue-specific stem cell lines isolated from foetal tissue is another possible strategy in regenerative medicine as these cells exhibit higher proliferation potential, more specific differentiation capacity, improved migration and regeneration after implantation [90, 91]. Functional integration of human FSC (hFSC)-derived dopamine neurons in a rodent Parkinson’s disease model is one of the most significant examples for the use of hFSC. It became the basis for clinical trials demonstrating similar effects in patients [9294]. Another study showed that the transplantation of human cortical neuroepithelial stem cells developed from foetal cortical brain did not result in tumour formation and facilitated the recovery of diminished tissue function in a rodent stroke model [95]. Despite associated ethical concerns and the risk for tumour development, these studies have shown great promise for the clinical application of ESC and FSC [64].

3.1.2. Induced pluripotent stem cells

In 2006, Takahashi and Yamanaka [96] discovered that the retroviral expression of pluripotency-specific transcription factors (Oct4, Sox2, Klf4 and c-Myc) reprograms adult somatic cells into a pluripotent state. These iPSCs showed epigenetic and transcriptional similarities to ESCs [9799]. Over the years, the progress has been made in the generation of virus-free/vector-free reprogramming methodologies to avoid vector-induced tumour development [100103]. In addition, iPSCs are now being derived from a variety of different cell types such as blood cells, dermal fibroblasts and keratinocytes at higher efficiencies [104107].

To study the sporadic and genetically inherited diseases, patient-specific iPSCs have been generated (Table 1). These in vitro studies offer a proof-of-concept for the use of iPSCs for disease modelling with the goal to discover novel drugs and disease-specific pathways aiding their treatment. Since several studies were performed on very limited cell numbers, reproducibility of the observed phenotype still needs to be investigated [108].

Disease Derived cells Control cells Result References
Parkinson’s
disease
Dopaminergic
neurons
hiPSC No observed defect, not drug
tested
[103]
Parkinson’s
disease
Dopaminergic
neurons
hiPSC Enhanced chemical sensitivity
causes cell death, drug tested
[309]
Spinal
muscular
atrophy
Motor neurons hiPSC Loss of SMN gene expression
and neuron formation, drug tested
[310]
RETT
syndrome
Neurons hiPSC Reduced soma size and spine
density, loff of synapse, drug
tested
[311]
Familial
dysautonomia
Neural crest cells hiPSC, hESC Loss of neural crest cells, drug
tested
[312]
Long QT 1
syndrome
Cardiomyocytes hiPSC Depolarisation of cardiomyocytes, drug
tested
[313]
Long QT 2
syndrome
Cardiomyocytes hiPSC Depolarisation of cardiomyocytes, drug
tested
[314]
A1-antitrypsin
deficiency
Hepatocytes hiPSC Downregulation of A1-antitrypsin
expression, drug tested
[315]
Timothy
syndrome
Cardiomyocytes hiPSC Depolarisation of cardiomyocytes, drug
tested
[316]

Table 1.

Patient-specific disease models utilising human iPSC-derived cells.

Table adapted from Wu and Hochedlinger [108]

The generation of autologous cells for cell therapy is another possible application for the iPSC technology as it minimises the challenges associated with human ESC-based therapies [108]. Researchers have described that iPSCs facilitate the reduction of the blood cell phenotype in a sickle cell anaemia mouse model [109]. iPSCs were derived from a transgenic mouse exhibiting a mutation in the human haemoglobin sequence. These iPSCs were genetically corrected and differentiated into haematopoietic progenitor cells. Subsequent implantation of these cells into the mouse model resulted in a normal haemoglobin level and restored phenotype [109]. Despite these remarkable results, haematopoietic stem cells with the ability for multilineage differentiation have not yet been generated. In addition, this study utilised retrovirus-derived iPSCs and it remains unclear if similar results can be achieved in a retrovirus-free approach. Comparable studies implanting iPSC-derived progenitors for mesodermal and ectodermal cells into animal models have been described (e.g. neurons, cardiomyocytes, blood) [110112].

The conversion of somatic cells into iPSCs and the advances that have been made in their generation has enabled researchers to utilise disease-specific cells for disease modelling as well as drug screening. Moreover, it opens up opportunities for the generation of custom-made iPSCs for cell therapies [108] (Figure 4). Patient-specific iPSCs are interesting autologous cell sources as they would eliminate the need to suppress the patient’s immune system and could be generated in larger quantities. However, future studies are required to investigate the effect of genetic modifications of host and donor cells when injecting iPSC into patients [64].

Figure 4.

The potential of iPSC technology in regenerative medicine. Reprogramming a patient’s somatic cells, for example blood cells or skin cells generates patient-specific iPSCs. These iPSCs can be differentiated into specific cell types which are subsequently used for either cell therapy or disease modelling. Cell therapy aims to fabricate autologous differentiated cells for implantation into a single patient. Disease modelling on the other hand is based on reproducing a cell phenotype from iPSC-derived specialist cells as present in the patient and utilising these cells for large-scale drug screening for the treatment of any patient with the same disease. Image adapted with the permission from Wu and Hochedlinger [108].

3.1.3. Comparison of ESCs and iPSCs

ESC and iPSC cell lines display biological differences between one another and how comparable the both cell types are might affect their functionality and safety. Researchers have shown that hiPSC and hESCs are highly similar [113116]. However, variations in gene expression [116], DNA methylation [117, 118], differentiation potential [119, 120] and teratoma-forming propensity [121] have been reported. Genetic background [122], passage number [116, 123], lab-to-lab differences and the use of vectors in their generation have extensive influence on the function and gene expression of PSC. Moreover, hiPSC and hESC have been shown to carry copy number variations (deletions and duplications) [124126] and point mutations [127], which are consequences of their culture. Some mutations might also occur during reprogramming [108]. Since differences in transcriptional and DNA methylation profiles between 12 hiPSC and 20 hESC lines have been described, large sample numbers are required to robustly investigate potential differences and their influence on pluripotent stem cells [128].

3.2. Clinical potential of pluripotent stem cells

Translation of pluripotent stem cell-based therapies into the clinic will depend on several factors such as cell purification, the efficiency of cell-lineage-specific differentiation leading to functional cells, eliminating tumour generation and finally the generation of novel organs [108]. Even though iPSCs have been differentiated into a number of cell types including cardiomyocytes [129133], neurons [134, 135], blood [133, 136, 137] and pancreatic cells [137, 138], purities over 95% have not been reported and isolating these cells from a heterogeneous cell population is difficult. Furthermore, ESC-/iPSC-derived cells are mostly immature and whether their stage will affect their clinical performance remains to be investigated in a cell- and disease-specific manner [108].

Another important issue for the application of PSC in regenerative medicine is their integration into the host tissue. Organs exhibit a balance between the numbers of each cell type, their geometrical arrangement and their developmental stages. Whether the injection of cell suspension will auto-regulate cell type numbers to generate an endogenous tissue and whether the transplanted cells will function in synergy with existing cells still needs to be investigated [108]. In addition, regulatory requirements for pluripotent stem cell–derived therapies remain high as shown by a clinical trial performed by Geron, a biopharmaceutical company that tested human ESC–derived oligodendrocytes for spinal cord injury [139]. Microscopic cysts found in cell-transplanted mice resulted in extensive studies for the batch-to-batch assessment of cyst formation as well as follow-up safety study on these cyst containing grafts were requested by the FDA [139].

Concerns over the use of hiPSCs for cell therapy are also arsing due to studies reporting on genetic and epigenetic modifications during the reprogramming process such as protein coding and DNA methylation [127, 140]. To-date, progress in the development of iPSC protocols has eliminated the need for vectors reducing the risk of tumour formation [113, 141, 142]. Furthermore, whether iPSC-derived cell types retain their epigenetic ‘memory’ remains under investigation. Studies have shown that both mouse and human iPSCs preserve an epigenetic profile which is related to the donor cell source and this may affect subsequent differentiation [143146]. Drawbacks for the application of patient-specific iPSCs also include regulatory, cost and time requirements for the generation of patient-specific iPSC-based treatments as well as scalability and good manufacturing practice (GMP)-compliant cell therapies.

Despite all this, in recent years, the focus has been on translating hESC research into the clinic such as the trial sponsored by Geron which has highlighted the obstacles but also promises of pluripotent stem cell therapy. Furthermore, Japanese researchers were applying for regulatory approval to use patient-specific iPSC-derived RPE cells to treat macular degeneration [147]. Pluripotent stem cell research has evolved from the isolation of hPSC to the development of differentiation protocols and early clinical trials. Less than two decades into hPSC research, the utilisation of human pluripotent stem cell–derived treatments are under clinical investigation [65].

3.3. Multipotent stem cells as alternative cell sources

Limited availability of embryonic and foetal tissue as well as ongoing research into the development of improved protocols yielding matured differentiated cells may be limiting factors for the use of ESC, FSC and iPSC. Alternatives to pluripotent stem cells are multipotent stem cells such as MSCs and adipose-derived stem cells which have been extensively studied for regenerative medicine therapies.

3.3.1. Mesenchymal stem cells

Non-hematopoietic multipotent MSCs are derived from the mesoderm and are present in a variety of connective tissues and postnatal organs [63]. Their discovery was reported over 130 years ago when research suggested that bone marrow is the source collagen depositing cells that can develop into various cell types including bone and cartilage [148]. Caplan [149] then named these marrow cells MSCs after their capability to differentiate into cells of mesenchyme origin such as bone, cartilage, tendon, muscle and adipose tissue. The clinical routine to treat haematological diseases with bone marrow makes MSCs interesting contenders for cell-based therapies. MSCs isolated from bone marrow are probably among the most characterised and clinical utilised stem cell types [150]. The multipotency of MSC allows them to differentiate into bone, cartilage, muscle and neural cells [151153]. They offer the possibility for autologous cell transplantation and at the same time eliminating the risk for graft-versus-host diseases. MSCs are non-immunogenic, affect the maturation and response of immune cells and do not cause tumour formation when implanted in allogeneic hosts [64]. Finally, the injection of MSC leads to the secretion of growth factors by host cells including brain-derived neurotrophic factor (BDNF), vascular endothelial growth factor (VEGF) and nerve growth factor (NGF) among others. In summary, the capacity of MSC to change their default state and their potential to promote tissue regeneration may surpass their application for haematopoietic diseases and might make them a suitable tool to treat degenerative diseases such as neurological and neurodegenerative disorders [154]. Employing MSC in combination with biomaterials or alone resulted in functional regeneration of paralysed limbs, diminished cavity formation in the spinal cord and axonal regrowth [155158]. MSCs have also been utilised in first clinical trials to treat myocardial infarcts, stroke and diseases of the CNS [159, 160]. These studies have demonstrated that bone marrow–derived MSCs are a promising stem cell type for clinical application and stem cell therapies [64]. However, limited integration into the host tissue was observed in clinical trials indicating that the detected benefit in these trials were most likely due to the secretion of cytokines and soluble factors [161].

MSCs with similar properties to bone marrow–derived MSC have been isolated from trabecular bone [162], periosteum [163], synovial membrane [164], skeletal muscle [165], skin [166], pericytes [167], peripheral blood [168] and umbilical cord [169, 170]. Due to their low number upon isolation, adult stem cells require in vitro expansion and modification before being employed therapeutically [63].

3.3.2. Adipose-Derived stem cells

Adipose-derived stem cells (ASCs) are a promising cell type as they can be easily isolated in large numbers from adipose tissue without causing severe donor site morbidity and discomfort [171, 172]. Over 80% of adipose tissue is composed of mature adipocytes and the stromal vascular fraction consisting of vascular smooth muscle cells, fibroblasts, preadipocytes, endothelial cells, resident immune cells and ASCs [173, 174]. Many of the ASC properties and density differ according to the harvest location of the adipose tissue [63]. Studies have shown that the ASCs proliferation rate is dependent on age, tissue location and type, the culture conditions and isolation procedure, but is in general higher compared to the doubling rate of bone marrow–derived MSCs [175178]. Recent studies have shown that their application is not restricted to mesodermal tissues instead they can be employed for endodermal and ectodermal tissue regeneration [63]. ASCs have been successfully differentiated into adipogenic [179181], osteogenic [182], cardiomyogenic [183, 184], chondrogenic [182, 185, 186], angiogenic [187, 188], tenogenic [189], hepatogenic [190, 191] lineages. Moreover, studies utilising scaffolds in combination with growth factors have been carried out to assess their tissue regenerative potential [192, 193]. In summary, ASCs are a valuable cell source for the development of cell-based therapies and have been shown to be safe and effective in clinical and preclinical studies [171, 172]. However, since ASCs are characterised as multipotent adult stem cells, their differentiation potential is restricted compared to ESC and iPSC. Even though limited numbers of clinical studies have investigated the therapeutic potential of ASCs and their lineage-specific differentiation depends on the site of harvest, gender and age of donor [194, 195], there are practical advantages of using ASCs in regenerative medicine [63].

Advertisement

4. Regenerative medicine strategies: engineering artificial niches for the control of stem cell fate

Adams and Scadden [196] elucidated the concept that the stem cell niche is ‘dynamic’ and its properties change during development and with varying physiological conditions. These changes affect stem cell fate, but could also be utilised as potential therapeutic tool in regenerative medicine. Consequently, the control of nanotopography, mechanical and chemical properties of the ECM among others in engineered constructs as well as their mechanical loading is essential for directing stem cell fate in bioartificial systems and for the development of regenerative therapies (Figure 5) [197].

Figure 5.

Engineering stem cell niches for the development of regenerative medicine strategies. To replicate niches in order to control stem cell fate chemical, topographical and mechanical properties are being mimicked using engineering techniques. Image adapted with the permission from Gazzatto et al. 2014 [5].

4.1. Biochemical signals as stem cell fate regulators

Chemical signals that cells are exposed to in the ECM activate signalling cascades determining cell proliferation, differentiation, migration and apoptosis [198, 199]. Hence, it is essential to mimic chemical cues of the ECM to control stem cell behaviour. Growth factors, which regulate cell adhesion, proliferation and lineage development, are an important type of chemical cue [200]. Growth factors are either added to the culture medium or secreted by niche cells and stem cells. They are crucial for stem cell fate decision and are spatiotemporally regulated during embryonic development [201]. A chemically defined protocol for the directed differentiation of hESC towards chondrocytes in 2D culture was developed where the pluripotent stem cells undergo intermediate developmental stages by supplementing the culture medium with exogenous growth factors and culturing cells on substrates of known matrix proteins over a culture period of 14 days [202, 203] (Figure 6A). The cartilage repair capacity of these chondroprogenitors was assessed by incorporation into fibrin hydrogels and implantation using an osteochondral defect model in the patellar groove of athymic RNU rats. Results showed that chondrogenic cells derived from hESC using a chemically defined differentiation protocol promoted cartilage repair [202] (Figure 6B).

Figure 6.

Differentiation of hESC and in vivo cartilage formation of hESC-derived chondroprogenitors. (A) hESC underwent a 14-day defined protocol for promoting chondrogenic differentiation. (B) Characterisation of chondroprogenitor cells by quantifying the upregulation of pluripotency, chondrogenic and cartilage ECM markers using qPCR. Results are shown as means ±SD (n − 3). Chondrogenic cells derived from MAN7 showed high expression of SOX9 (inset shows IgG control) and Safranin-O staining (inset shows pre-treated with chondroitinase ABC) at the end of the protocol. Scale bars = 100 mm. (C) Assessment of in vivo cartilage repair capacity of pre-differentiated hESC after 12 weeks by macroscopic observation of the gross appearance of the RNU rat patella groove. Histological sections through the knee were stained for HE, Saf-O immunological assessment through Col-I and COL-II staining. Cartilage repair was scored using Pineda’s system (0-worst to 14-best). Human cells were detected 12 weeks after implantation by immunohistochemistry staining. Scale bars black = defect area; white = 500 mm. Abbreviations: alkaline phosphatase (ALP); extracellular matrix (ECM); fibronectin (FN); glyceraldehyde-3-phosphate dehydrogenase (GAPDH); gelatine (GEL); Safranin-O (Saf-O); collagen type-I (Col-I); collagen type-II (Col-II); haematoxylin (HE). Figures were modified with the permission from Cheng et al. [202] and Oldershaw et al. [203].

Furthermore, the importance of cell–cell interactions in soluble factor signalling has revealed that certain cell types will respond only to locally secreted growth factors when in direct contact with adjacent cells [204]. Growth factors affect secreting cells (autocrine) and other cells (paracrine) in vivo. Soluble factors are often bound to the ECM limiting their diffusion and enhancing their efficiency. This can be replicated by tethering these to a biomaterial substrate [205]. Kuhl et al. [200] demonstrated that growth factors in their matrix-bound state were more effective than their soluble form. When proteins are incorporated or linked to biomaterials, they are commonly denatured or degraded. Consequently, short peptide sequences mimicking growth factors and chemokines are integrated to control stem cell fate [206]. The RGD sequence present in ECM proteins has been widely linked to biomaterial surfaces demonstrating enhanced osteogenesis and chondrogenesis compared to non-modified surfaces [206208]. Bone morphogenic proteins have been shown to play a key role in stem cell activity and osteoblastic differentiation [209, 210]. Another approach is a surface modification of natural and synthetic biomaterials with specific functional chemical groups [211]. Examples include the functionalisation of PEG hydrogels with phosphate groups for bone mineralisation, carboxyl groups resembling GAG for cartilage and tert-butyl groups to mimic lipids for adipose tissue [212]. Moreover, self-assembled monolayers (SAM) functionalised with methyl-, hydroxyl-, amino- and carboxyl groups have been shown to promote osteogenic differentiation of MSC with the amino group being the most effective chemical group [213].

Growth factors and reactive chemical groups have been successfully used to guide stem cell differentiation. It was shown that immobilised chemical cues are more effective than soluble factors. However, improved control over bioactivity, spacing and orientation of the immobilised growth factors is required to guide-/direct-specific stem cell differentiation [214]. Further studies investigating the effect of chemical signals on stem cell fate are summarised in Table 2.

Cell type Material Chemical group Result References
Soluble factors
hMSC PLGA BMP-2 Enhanced osteogenic
differentiation
[317]
hMSC PLLA BMP-2 Enhanced osteogenic
differentiation
[318]
hMSC Chitosan/collagen
IV
VEGF Endothelia differentiation [319321]
hESC TCP coated
with FN,
gelatine and
both
Sequential addition of
Wnt3a, Activin-A, BMP-
4, Follistatin, GDF5,
FGF2, NT4
Chondrogenic
differentiation
[202, 203]
Peptides
hMSC Alginate Osteopontin peptide Osteogenic differentiation [322]
hMSC PLGA Osteocalcin peptide Osteogenic differentiation [323]
hMSC BCP/PLA RGD osteogenic differentiation [324]
rMSC HA-PLG BMP-2 peptide Increased ALP
expression, ectopic bone
formation in vivo
[325]
Chemically reactive groups
hMSC Silk fibroin –COO–, =C=O, SO3H,
NH2, CH3
Enhanced osteogenic
differentiation
[326]
hMSC PEG PO3 Increase of osteogenic
markers at protein
and gene level
[212, 327, 328]
hMSC Glass COOH, CH3, OH, NH2,
SO3H, SH
–NH2 and –SH group
promoted and maintained
osteogenesis, −OH and –
COOH promoted
chondrogenesis
[329331]

Table 2.

Effect of chemical signals on stem cell differentiation.

Table adapted with the permission from Griffin et al. [214]

hMSC: human mesenchymal stem cells; hESC: human embryonic stem cells; PLGA: poly (lactic-co-glycolic acid); PLLA: poly(L-lactic acid); PEG: polyethylene glycol; HA-PLG: hydroxyapatite (HA)/poly(lactic-co-glycolic acid); BMP: bone morphogenetic protein; PMMA-g-PEG: poly(methyl methacrylate)-graft-poly(ethylene glycol; BCP: biphasic calcium phosphate; TCP: tissue culture plastic; EGF: epidermal growth factor; PDGF-AA: VEGF: vascular endothelial growth factor.

4.2. Topographical signals as regulators of stem cell fate

Cells in their native in vivo environment engage with a variety of macro (tissues)-, micro (cells)- and nano-sized (proteins and ligands) topographical features. The basement membrane with its nanoscale pores, ridges and fibres is the most crucial ECM structure providing tissue organisation and support [215]. Micro- and nanopatterning techniques including soft lithography, electrospinning, layer-by-layer microfluidic patterning, three-dimensional printing, reactive ion etching and ion milling have resulted in the fabrication of scaffolds with controlled porosity, geometry and rigidity and texture [216218]. The production of specific surface topography scales (nano, micro), types (ridges, pit, pillar, grooves) and distributions (random, regular) has enabled researchers to study the influence of topographical signals on stem cell differentiation [214]. Structural cues greatly affect gene expression as they determine cell shape, elongation, positioning of focal adhesion and cell–cell interactions [29, 219221]. Cell shape is a crucial regulator for cell physiology and function as well as dictator for stem cell differentiation [222]. Manasek et al. [223] reported that cell shape regulates myocardial development, whereas ECM-induced cell shape changes are responsible for the proliferation and differentiation of capillary endothelial cells [224]. Numerous studies have demonstrated that stem cell fate can be guided through controlling their shape by artificial extracellular matrices [197]. Traditionally, stem cells are expanded two-dimensionally on rigid, flat-coated or non-coated tissue culture plastic and are exposed to soluble factors in the liquid growth medium. These plastic culture substrates are usually coated with collagen or laminin, feeder cell layers or with hydrogels such as Matrigel [225]. However, this culture approach is very different from the native microenvironment cells are experiencing in vivo, where they are residing in the stem cell niche anchored to ECM though adhesion molecules [225]. Two-dimensional culture approaches provide simplified methods to expand stem cells and to study individual cues influencing their fate decisions in vitro. This has allowed the identification, patterning and concentration of soluble to tethered ECM molecules regulating stem cell niche and its inhabitants. ECM arrays have been utilised to screen for molecules individually or in combination that induce fate changes [226, 227]. The development of 3D-culture systems compared to traditional 2D culture systems results in a more rounded cell morphology. Culturing chondrocytes in 2D, for example resulted in a dedifferentiation and change from chondrogenic to fibroblastic phenotype [228], whereas pellet culture or incorporation of chondrocytes in hydrogels maintained their native phenotype [229, 230]. It was further demonstrated that for bone marrow–derived MSCs to undergo chondrogenic differentiation cells would have to be cultured with a spherical shape either as spheroid culture or encapsulated in hydrogels [221, 231233]. However, when MSCs were cultured in hydrogels facilitating their adhesion and spreading stem cells exhibited a fibroblastic phenotype [234]. Table 3 provides a summary of studies examining the correlation between surface topography and stem cell differentiation. Oh et al. [235] found that MSCs cultured on ~100-nm nanotubes differentiated towards the osteogenic lineage, whereas MSCs seeded on 30-nm nanotubes did not differentiate [235]. Moreover, surface topography can also cause cell elongation on material surfaces resulting in the distortion of the nuclei shape as the nuclei are mechanically integrated into the cells, which has been shown to influence osteoblastic stem cell differentiation [237]. When ESCs are aggregated into embryoid bodies (EB), which mimic early stages of embryonic development, it becomes clear that geometrical signals also affect cell–cell interaction and hence guide stem cell differentiation [66]. The investigation of hexagonal patterns, random nanopits, disordered and ordered squares utilising on hMSCs electron beam lithography (EBL) showed that highly asymmetric structures resulted in highest osteogenic expression in the absence of growth factor supplemented medium [237] indicating that nano-sized topographical induction of stem cell differentiation is as effective as chemical induction [238]. Micropatterns have also been shown to affect cell shape. Changes from rounded (on small islands) to flattened (on large islands) shapes result in modifications of the actin cytoskeleton and focal adhesions and hence controlling the lineage commitment of MSC into a variety of cell phenotypes [29].

Cell type Material Topographical feature Result References
hMSC TiO2 Nanotubes (15, 100 nm) Cells grown on 15 nm
exhibited enhanced
integrin clustering, cell
spreading, osteogenesis
[332]
hMSC TiO2 Nanotubes (30, 50, 70
and 100 nm)
Osteogenic
differentiation promoted
on 70–100 nm nanotubes
[235]
hMSC PDMS Islands 1000 μm2 facilitated
osteogenesis
[29]
hMSC PDMS Grafting (350 nm) Neurogenesis [333]
hMSC PDMS Micropattern, stripped
grooves collagen type-
I coated
Neuronal differentiation
enhanced
[334]
hMSC PMMA Hexagonal pattern,
nanopits, disordered
and ordered squares
Enhanced osteogenesis
on disordered squares
[237]
hBMSC Hydrogenated
amorphous
carbon
Grooves (80/40, 40/30,
30/20 μm-width/
spacing; 24 nm depth)
Neurogenesis, absence
of growth factor
supplemented medium
[335]
hESC PDMS Square shaped
fibronectin surrounded
by pluronic-F127,
micropattern
Myogenesis and
chondrogenesis
[336]
hESC PDMS Grooves Neuronal [337]
mESC PLLA Fibrous grating (50–500
nm), TCP
Enhanced osteogenesis
on fibrous gratings
[338]

Table 3.

Effect of surface nanotopography and stem cell fate.

Table adapted with the permission from Griffin et al. [214]

hMCS: human mesenchymal stem cell; hBMSC: human bone marrow–derived stem cells; hESC: human embryonic stem cell; mESC: mouse embryonic stem cell; PDMS: polydimethylsiloxane; PMMA: polymethyl methacrylate; PLLA: poly(L-lactide); TiO2: titanium dioxide, TCP: tissue culture plastic.

4.3. Biomaterials affecting stem cell fate

The control of stem cell fate through biomaterial nanotopography is promising as these approaches are not subjected to short-term degradation as are chemical cues. In addition, an individual biomaterial could exhibit various nanotopographies and gradients which would allow for further clinical applications [214]. Biomaterials are being designed and utilised to mimic the microenvironment and its cues stem cells experience in vivo. Nano- and microtopography replicating the ECM signals have been shown to direct stem cell differentiation. ECM components linked to the biomaterials surface in form of small peptides or through the incorporation of growth factors have been successful approaches for guided stem cell differentiation and provide advanced stem cell-based clinical approaches [214, 225] To overcome limiting factors for the clinical use of pluripotent stem cells, new culture systems based on advanced biomaterials are required which more closely mimic the native in vivo milieu and support application related stem cell fate decisions [225]. In summary, simplifying a complex three-dimensional stem cell niche into a two-dimensional biomaterial approach is a potent tool to study control machineries regulating stem cell biology. In order to reconstruct the complexity and interplay of the stem cell niche and its components it is, however, necessary to utilise 3D culture approaches [225]. Hence, there have been numerous strategies to create 3D biomaterial matrices with a variety of structural, chemical and mechanical properties as artificial growth environments for cells each with its advantages and disadvantages [239]. Since the investigation of niche components and their effect on stem cell fate is a complex undertaking, 3D high-throughput approaches similar to 2D ECM protein arrays are being developed [240242]. These could be produced by printing or liquid-dispensing technologies [263]. For example, 3D PEG gel arrays were prepared to study the effect of gel degradation and cell adhesion ligand concentration [264]. Once ideal biomaterial candidates have been identified, it is essential to examine their in vivo performance by implantation into hosts, some of which have been reviewed elsewhere [239].

Biomaterials have also been utilised to study the interplay between stem cells and support cells such as vascular cells, neural cells and stromal cells [225]. Electropatterning of living cells within PEG hydrogels resulted in the deposition of multicellular aggregates of known size and shape, which was shown to regulate biosynthesis of chondrocytes by increase of sulphated GAG in larger cell aggregates [245]. Moreover, 3D angiogenesis was studied combining microfluidics with gel patterning [246]. Gradients are known to be crucial for the regulation of dynamic processes during development and during tissue regeneration and homeostasis. Hence, the development of biomaterial gradients has found wide interest in regenerative medicine [225, 247249]. It has been shown for bone/marrow [52, 250], prostate [251] and breast [252] that cell–cell interactions in those stem cell niches are influenced by paracrine hormone signalling [197].

Furthermore, cell–cell interactions have been studied in co-culture approaches, which on the other hand, do not allow for the identification of the particular function of soluble or immobilised molecules. In vivo, cytokines and growth factors are mostly immobilised to the ECM suggesting that tethered chemical factors are more stable and their signalling is longstanding. So has for example, the immobilisation of FGF2 to a synthetic polymer resulted in an increase in its potency and stability and subsequent enhancement in ECS proliferation and ERKI activation [253]. Additionally, the tethering of EGF to a biomaterial scaffold has demonstrated an increased effectiveness compared to its soluble equivalent [254]. These studies have demonstrated that investigating chemical cues individually is informative, but testing growth factors in a high-throughput manner on polymer arrays is more desirable for identifying pluripotent stem cell regulators [255].

4.4. ECM elasticity as stem cell fate regulator

ECM subjects cells to multiple physical signals including mechanical signals due to its stiffness [256258]. Cells adhering to their surrounding employ contractile forces resulting in tensile stresses within their cytoskeleton [259]. The connection between these forces and ECM stiffness has a huge influence on cell migration [260, 261], proliferation [262] and apoptosis [263] (Table 4). Stem cells have shown to respond to mechanical properties of their surrounding microenvironment [214, 264]. Engler et al. [264] observed that when MSCs were cultured on collagen-coated polyacrylamide hydrogels with varying stiffness, cells differentiation without the use of supplemented medium. On soft gels mimicking brain tissue (0.1–1.0 kPa), MSCs underwent neurogenic differentiation. Stiffer gels resembling muscle tissue enabled myogenic development and very stiff gels (25–40 kPa) replicating the bone properties resulted in osteogenic differentiation of MSC [264]. Similar observations were made by other researchers for two-dimensional and three-dimensional culture [265269]. MSC incorporated into collagen-GAG scaffolds with stiffness of 0.5, 1 and 1.5 kPa exhibited different chondrogenic differentiation. Softer gels allowed for chondrogenic and stiffer gels for osteogenic differentiation [269]. It is assumed that the substrate stiffness induces alterations in the activity of focal adhesion and causes remodelling, which triggers a cascade if signalling pathways enabling cell differentiation [270, 271]. In addition, integrins are thought to be a central cell structure for sensing mechanical stimulation [270, 271]. A limitation to these studies is the fact that different tissues may have similar stiffness, and hence, it might not be possible to direct stem cell fate by a single mechanical property of the surrounding. Instead, it is necessary to consider a more complex interplay of extrinsic and intrinsic factors influencing stem cell differentiation [197].

Cell type Material Mechanical cues Result References
ECM stiffness
MSC Polyacrylamide
gel, collagen
coated
Stiffness-0.1–1.0,
25–40 kPa
Lineage commitment according to
substrate stiffness; softer gel-
neuronal, stiffer gel-osteoblastic
differentiation
[264]
ANSC Interfacial
hydrogel
Substrate moduli
0.01–10 kPa
Self-renewal, cell spreading and
differentiation inhibited on soft gels
(0.01 kPa); cell proliferation and
neuronal differentiation maintained
on ≥0.1 kPa; glial differentiation on
gels 1–10 kPa
[339]
hMSC Polyacrylamide
gel, collagen
and fibronectin coated
Marrow and
adipose tissue
0.25 kPa
Cells were quiescent but
maintained multilineage potential,
[35]
MSC Collagen-
glycosaminogly
can
0.5, 1 and 1.5 kPa Softer gels triggered chondrogenic
differentiation; stiffer gels resulted
in osteogenic differentiation
[269]
Mechanical stimulation
MSC Protein-coated
membranes
1, 5, 10, 15%
cyclic uniaxial
stretch
Myogenic differentiation for 5 and
10%; 1 and 15% failure of
myogenic lineage
[340342]
AMSC 10% uniaxial cyclic
strain at 1 Hz for
7 days
Decreased expression of myogenic
markers
[343]
MSC Uniform biaxial
strain
Enhanced osteogenesis and
calcium deposition
[344347]
AMSC Pulsatile fluid flow Enhanced osteogenesis [348]
mESC 4–12% strain,
1 Hz, 24 h
Differentiation into vascular smooth
muscle cells; increase in
proliferation
[349]
hESC Cyclic strain promoting of self-renewal, inhibition
of differentiation,
[350]
MSC Agarose Cyclic confined
compression
Chondrogenic differentiation [351]
MSC Spheroids IHP Chondrogenesis [352]
hBMSC PCL IHP, 270 kPa, 1 Hz,
over 21 days
Enhanced metabolic activity,
upregulation of osteogenic markers
and calcium deposition
[295]
FCSC Ex vivo cultured
chick femurs
IHP, 270 kPa, 1 Hz,
over 21 days
Enhanced mineralisation [287]

Table 4.

Effect of ECM stiffness and mechanical stimulation on stem cell differentiation.

MSC: mesenchymal stem cells; ANSC: adult neuronal stem cells, AMSC: adipose-derived mesenchymal stem cells, hBMSC: human bone marrow–derived mesenchymal stem cells, FCSC: fetal chick stem cells, kPa: kilo Pascal, Hz: hertz, IHP: intermittent hydrostatic pressure.

4.5. Biomechanical regulation of stem cell fate

The body’s cells are constantly subjected to a variety of mechanical stimuli through muscle action, blood flow, gravity and other physical and physiological processes [197].To ensure tissue health and function, the interplay between cells and mechanical cues is essential. It is assumed that these mechanical factors are involved in diseases including osteoarthritis, osteoporosis and atherosclerosis [277]. Furthermore, there is mounting evidence that mechanical forces are crucial for development [278, 279] stem cell lineage commitment and fate decisions [197]. Despite early in vivo studies by Glückmann et al. [280], Fang and Hall [281], Hall and Herring [282] and Murray and Drachman [283] on the effect of mechanical factors on development little was known on how biomechanical cues affect gene expression and stem cell fate [280283]. One difficulty in studying these interactions is the complexity by which cells sense mechanical signals. Mechanical loading of tissue can result in a variety of stresses, strains, pressures, fluid flows, osmotic pressures and electric charges on biological molecules [284]. These modifications of the cell’s microenvironment result in structural changes of ECM proteins, as well as the activity of immobilised or soluble growth factors. For this reason it is challenging to differentiate the effect of direct mechanical forces in vivo from indirect mechanically driven effects in adhesive and paracrine signals and resulting changes such as cell shape [197, 285]. Nevertheless, mechanical forces are influencing cellular processes directly and the mechanism by which cells perceive external mechanical stimuli have been described by Liedtke and Kim [285]. Cellular responses to physical stimuli are not simply a reaction to the input stimuli, instead they are also linked to cytoskeletal changes, ECM interactions and the production of cellular forces [29, 264]. Stem cell response to controlled physical forces as well as the biomechanical mechanism and signalling pathways that direct stem cell lineage commitment are being investigated [286]. Several research groups have attempted to isolate the effect of applied mechanical stimuli such as fluid shear, strain and compression. Table 4 summarises studies on the effect of externally applied mechanical stimuli and ECM stiffness on stem cell fate. These studies demonstrate that the effect of physical stimulation is dependent on stem cell source, type and state of pre-differentiation. Dynamic mechanical compression, for example increases chondrogenic marker expression of BhMSC. However, ESC-derived EB exhibit downregulation of chondrogenic markers when subjected to the same stimulation regime [197]. It is also under investigation whether mechanical cues provided by bioreactors alone or in combination with other cues can induce lineage commitment in PSC [287, 288].

Advertisement

5. Bioreactor technologies promoting mechanical stimuli

As elucidated, earlier physical forces in vivo are essential cues during development and for regulating stem cell fate. Furthermore, it has been widely acknowledged that in vitro mechanical cues (e.g. stress, strain, shear, compression, hydrostatic pressure) are greatly influencing the cell morphology, cell adhesion, proliferation and gene regulation [214, 289]. In order to provide physiological relevant growth environments for cells and tissue-engineered constructs, various types of bioreactors have been developed [287, 290]. Bioreactors are devices that utilise mechanical forces to influence biological processes under closely controlled and monitored conditions [291, 292]. They offer biophysical cues encouraging cells to differentiate or/and produce ECM prior to implantation in vivo [293]. In general, bioreactors meet the following applications in regenerative medicine-1: they provide spatially uniform cell distribution, 2: they deliver physiological relevant concentrations of oxygen and carbon dioxide as well as nutrients in the culture medium, 3: they support mass transport to the core of the tissue engineered construct facilitating cell survival throughout tissue-engineered constructs, 4: they provide physical stimuli to regulate stem cell differentiation and proliferation, 5: they facilitate tissue development and accelerate construct maturation [287, 294297]. Studies utilising mechanical forces for the direction of stem cell fate and tissue development are summarised in Table 4.

5.1. Bioreactor designs and their application for tissue regeneration

Many attempts in regenerative medicine employ traditional cell culture approaches. These, however, are associated with numerous disadvantages. To overcome limitations in cell culture techniques and to provide mechanical stimulus for tissue development, various types of custom-made and commercially available bioreactors have been manufactured [293] (Figure 7).

Figure 7.

Static and dynamic cell culture systems. Traditional cell culture systems include well plate formats, petri dishes, T-flasks and roller bottles (left). To overcome disadvantages of static culture dynamic culture systems have been developed such as spinner flasks, rotating vessels, hollow fibre systems (right). Image adapted with the permission from Pörtner et al. 2005.

5.1.1. Rotating wall bioreactors

The rotating wall vessel bioreactor was originally developed by Schwarz and colleagues at Nasa’s Johnson Space Centre to protect cell culture experiments from high forces during space shuttle take-off and landing [290, 298]. This bioreactor consists of a cylindrical chamber in which the scaffolds are free to move within the cell culture media. A permeable membrane is inserted in the vessel wall to allow for sufficient gas exchange. The vessel is rotated so that the downward gravitational force and the upward hydrodynamic force are balanced so that the scaffolds remain suspended in the media experiencing microgravity. This bioreactor has been utilised for regenerative studies on bone, cartilage, human ovarian surface epithelial cells among others [299301].

5.1.2. Flow perfusion

A fluid pump is used to pump media through a sample chamber which is designed so that fluid flow is directed through the centre of cellular scaffolds [293]. Superior fluid transport was obtained for flow perfusion bioreactors when comparing spinner flasks, rotating wall bioreactors and flow perfusion bioreactors. For the comparison of the three bioreactor types, consistent flow rate, scaffold type and cell densities were utilised resulting in homogenous cell distribution throughout the scaffold only for the flow perfusion bioreactor [302, 303]. In order to obtain optimum experimental outcome, fluid flow rate needs to be adjusted to facilitate a balance between mass transfer of waste products and nutrients and to retain newly produced ECM [293].

5.1.3. Compression bioreactors

Compression bioreactors are commonly used for cartilage and bone tissue engineering and can be adjusted for both dynamic and static loading as studies have exhibited shown that dynamic loading is more suitable for certain tissues such as cartilage, but negative for others [291, 292]. A compression bioreactor commonly consists of a motor, a controlling mechanism providing various magnitudes of displacements and frequencies as well as a tool to provide linear motion [304]. Evenly distributed load is applied to cellular scaffolds via flat plates [305].

5.1.4. Strain bioreactors

Strain bioreactors are normally utilised for the development of tendon and ligament tissues, but also for bone, cartilage and cardiovascular tissues [293]. The bioreactor design is similar to compression bioreactors whereas instead of applying force through flat plates, scaffolds are hold in place via clamps when tensile force is employed.

5.1.5. Hydrostatic pressure bioreactor

Hydrostatic pressure bioreactors have been utilised for the development of bone and cartilage tissues among others. Professor El Haj’s group has developed a novel hydrostatic bioreactor allowing for ease of handling and scale up of sample numbers that can be mechanically stimulated simultaneously [287, 288] (Figure 8). This bioreactor system is composed of a sealed aluminium chamber suitable for standard tissue culture well plates. By compressing incubator air and subsequently pumping it into the bioreactor chamber, hydrostatic pressure is applied to samples creating a gas–liquid interface between cell culture medium and the air. A temperature controller maintains the temperature of the compressed air at 37°C. The gas phase is removed from the bioreactor chamber, pumped back into the incubator and reused for the next stimulation cycle. A system accompanying software controls continuous and sinusoidal waveforms at various pressures (0–280 kPa) and frequencies (0.0001–2 Hz).

Figure 8.

Hydrostatic force bioreactor. The bioreactor system (A) consists of a compressor, computer for monitoring of applied pressures and a bioreactor chamber (B) fitting standard tissue culture well plates, which is connected to a stimulator within a standard cell culture incubator (C). Image was adapted with the permission from Reinwald et al. 2015 [287].

Pluripotent stem cells are interesting cell types due to their capability to differentiate into numerous cell types. A limitation, however, is that hPSCs develop their functionality at later stages in development. At present, protocols yield in differentiated cells that match the embryonic stages of development. Protocols are needed that result in cells with adult-like functionality [65]. The investigation of signalling pathways controlling cell development has led to successful strategies to direct stem cell fate [306], and small molecules have been utilised for the creation of differentiation protocols [202, 203, 307, 308]. In addition, bioreactors have been successfully employed to direct stem cell fate [288, 295]. Recently, a collaborative approach between Manchester University and Keele University started to investigate the effect of hydrostatic pressure on the maturation of hESC-derived chondroprogenitors and hMSC. Chondroprogenitors were obtained following a defined 2 weeks of differentiation protocol (Figure 6A) [202, 203]. To enhance maturation of these cells and to differentiate hMSC towards chondrogenic lineage, they were cultured in growth factor supplemented medium as either spheroids or embedded in fibrin hydrogels and subjected to intermittent hydrostatic pressure at 270 kPa, 1Hz for 1 h daily. Results suggested that a combination of mechanical and chemical cues resulted in the production of matrix proteins collagen and GAG as well as the upregulation and maintained expression of chondrogenic markers Aggrecan and SOX9 (Figure 9).

Figure 9.

Stem cell differentiation was promoted by mechanical and chemical cues. hMSC spheroids (A) were cultured in growth factor supplemented medium either under static conditions or when subjected to hydrostatic pressure at 270 kPa, 1 Hz and 1 h daily over a period of 14 days. Positive immunocytochemistry staining for Aggrecan was observed for both culture conditions (B). To quantify changes in the expression of chondrogenic markers, Aggrecan and SOX9 quantitative polymerase chain reactor was performed revealing 2.5-fold Aggrecan upregulation and maintained SOX9 expression compared to static controls (n = 5). Increased levels of matrix proteins collagen and GAG and total protein were detected from biochemical assays (D) (n = 5). Scale bars = 150 μm (light microscopic images) and 100 μm (immunocytochemistry stains).

Advertisement

6. Conclusion

This chapter highlights the valuable role of multipotent and pluripotent stem cells and the importance of replicating their niche in vitro to develop novel regenerative therapies. Despite the ethical and safety concerns associated with PSC, they offer valuable differentiation and proliferation potential and could possibly offer a valuable cell source for clinical applications.

The stem cell niche influences stem cell fate in many ways, including mechanical support, elasticity, topography, biochemical signals, oxygen tension and cell communication. By artificially engineering the dynamic stem cell niche, stem cell fate can be directed. Advanced 3D biomaterials are being harnessed to mimic the in vivo environment by providing physical and chemical support and by allowing for cell–cell interactions. Bioreactor technologies are able to replicate the mechanical stimuli and mimic the physiologically relevant environment.

Whilst promising advances have been made in regenerative medicine, significant obstacles have been identified and these must be overcome before novel cell and tissue engineering therapies are clinically established.

Advertisement

Acknowledgments

This work was supported by the UKRMP Hub. The authors would also like to thank Mr. Diogo Mosqueira-Alves-Moreira-Silva, Mr. Luis Costa-Marques and Mr. Shah Mijaan Ali.

References

  1. 1. Shohreh Mashayekhan, Maryam Hajiabbas and Ali Fallah (2013). Stem Cells in Tissue Engineering, Pluripotent Stem Cells, Dr. Deepa Bhartiya (Ed.), InTech, DOI: 10.5772/54371. Available from: http://www.intechopen.com/books/pluripotent-stem-cells/stem-cells-in-tissue-engineering
  2. 2. Moore K.A.,Lemischka I.R. Stem cells and their niches. Science. 2006; 311:1880–1885.
  3. 3. Fuchs E.,Tumbar T., Guasch G. Socializing with the neighbors: stem cells and their niche. Cell. 2004; 116:769–778.
  4. 4. Morrison S.J.,Spradling A.C. Stem cells and niches: mechanisms that promote stem cell maintenance throughout life. Cell.2008; 132(4)598–611.
  5. 5. Gattazzo F.,Urciuolo A., Bonaldo P. Extracellular matrix: a dynamic microenvironment for stem cell niche. Biochim Biophys Acta. 2014; 1840(8):2506–2519.
  6. 6. Place E.S.,Evans N.D., Stevens M.M. Complexity in biomaterials for tissue engineering. Nat. Mater. 2009; 8:457–470.
  7. 7. Stevens M.M.,George J.H. Exploring and engineering the cell surface interface. Science. 2005; 310:1135–8.
  8. 8. Halder G.,Dupont S., Piccolo S. Transduction of mechanical and cytoskeletal cues by YAP and TAZ. Nat. Rev. Mol. Cell Biol. 2012; 13:591–600.
  9. 9. Mammoto A.,Mammoto T., Ingber D.E. Mechanosensitive mechanisms in transcriptional regulation. J. Cell Sci. 2012; 125:3061–3073.
  10. 10. Mammoto T.,Ingber D.E. Mechanical control of tissue and organ development. Development. 2010; 137:1407–1420.
  11. 11. Lane S.W.,Williams D.A., Watt F.M. Modulating the stem cell niche for tissue regeneration. Nat. Biotechnol. 2014; 32:795–803.
  12. 12. Reinwald Y.,Leonard K.H.L., Henstock J.R.,. Evaluation of the growth environment of a hydrostatic force bioreactor for preconditioning of tissue-engineered constructs. Tissue Eng. Part C Methods. 2015; 21(1):1–14.
  13. 13. Schofield R. The relationship between the spleen colony-forming cell and the haemopoietic stem cell. Blood Cells. 1978; 4:7–25.
  14. 14. Hall P.A.,Watt F.M. Stem cells: the generation and maintenance of cellular diversity. Development. 1989; 106:619–633.
  15. 15. Watt F.M.,Hogan B.L. Out of Eden: stem cells and their niches. Science. 2000; 287:1427–1430.
  16. 16. van Es J.H., et al. Dll1+ secretory progenitor cells revert to stem cells upon crypt damage. Nat. Cell Biol. 2012; 14:1099–1104.
  17. 17. Katayama Y., et al. Signals from the sympathetic nervous system regulate hematopoietic stem cell regress from bone marrow. Cell. 2006; 124:407–421.
  18. 18. Barker N., et al. Identification of stem cells in small intestine and colon by marker gene Lgr5. Nature. 2007; 449:1003–1007.
  19. 19. Greenbaum A., et al. CXCL12 in early mesenchymal progenitors is required for haematopoietic stem-cell maintenance. Nature. 2013; 495:227–230.
  20. 20. Ding L.,Morrison S.J. Haematopoietic stem cells and early lymphoid progenitors occupy distinct bone marrow niches. Nature. 2013; 495:231–235.
  21. 21. Reya T.,Clevers H. Wnt signalling in stem cells and cancer. Nature. 2005; 434:843–850.
  22. 22. Méndez-Ferrer S.,Lucas D., Battista M., Frenette P.S. Haematopoietic stem cell release is regulated by circadian oscillations. Nature. 2008; 452:442–447.
  23. 23. Fujisaki J., et al. In vivo imaging of Treg cells providing immune privilege to the haematopoietic stem-cell niche. Nature. 2011; 474:216–219.
  24. 24. Jaiswal S., et al. CD47 is upregulated on circulating hematopoietic stem cells and leukemia cells to avoid phagocytosis. Cell. 2009; 138:271–285.
  25. 25. Williams D.A.,Rios M., Stephens C., Patel V.P. Fibronectin and VLA-4 in haematopoietic stem cell-microenvironment interactions. Nature. 1991; 352:438–441.
  26. 26. Holst J., et al. Substrate elasticity provides mechanical signals for the expansion of hemopoietic stem and progenitor cells. Nat. Biotechnol. 2010; 28:1123–1128.
  27. 27. Gilbert P.M., et al. Substrate elasticity regulates skeletal muscle stem cell self-renewal in culture. Science. 2010; 329:1078–1081.
  28. 28. North T.E., et al. Hematopoietic stem cell development is dependent on blood flow. Cell. 2009; 137:736–748.
  29. 29. McBeath R.,Pirone D.M., Nelson C.M., Bhadriraju K., Chen C.S. Cell shape, cytoskeletal tension, and RhoA regulate stem cell lineage commitment. Dev. Cell. 2004; 6:483–495.
  30. 30. Kimura W.,Sadek H.A. The cardiac hypoxic niche: emerging role of hypoxic microenvironment in cardiac progenitors. Cardiovasc. Diagn. Ther. 2012; 2:278–289.
  31. 31. Muscari C., et al. Priming adult stem cells by hypoxic pre-treatments for applications in regenerative medicine. J. Biomed. Sci. 2013; 20:63.
  32. 32. Forristal C.E., et al. Pharmacologic stabilization of HIF-1alpha increases hematopoietic stem cell quiescence in vivo and accelerates blood recovery after severe irradiation. Blood. 2013; 121:759–769.
  33. 33. Watt F.M., Driskell R.R. The therapeutic potential of stem cells. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2010; 365:155–163.
  34. 34. Trumpp A., Essers M., Wilson A. Awakening dormant haematopoietic stem cells. Nat. Rev. Immunol. 2010; 10:201–209.
  35. 35. Kurtz A., Oh S. Age related changes of the extracellular matrix and stem cell maintenance. Prev. Med. Baltimore. 2012; 54:S50–S56 (Suppl.).
  36. 36. Orford K.W., Scadden D.T. Deconstructing stem cell self-renewal: genetic insights into cell-cycle regulation. Nat. Rev. Genet. 2008; 9:115–128.
  37. 37. Greco V., Guo S. Compartmentalized organization: a common and required feature of stem cell niches? Development. 2010; 137:1586–1594.
  38. 38. Jones D.L., Wagers A.J. No place like home: anatomy and function of the stem cell niche. Nat. Rev. Mol. Cell Biol. 2008; 9:11–21.
  39. 39. Li L., Clevers H. Coexistence of quiescent and active adult stem cells in mammals. Science. 2010; 327:542–545.
  40. 40. Lander A.D., Kimble J., Clevers H., Fuchs E., Montarras D., Buckingham M., Calof A.L., Trumpp A., Oskarsson T. What does the concept of the stem cell niche really mean today? BMC Biol. 2012; 10:19.
  41. 41. Wagers A.J. The stem cell niche in regenerative medicine. Cell Stem Cell. 2012; 10:362–369.
  42. 42. Frantz C., Stewart K.M., Weaver V.M. The extracellular matrix at a glance. J. Cell Sci. 2011; 123:4195–4200.
  43. 43. Brizzi M.F., Tarone G., Defilippi P. Extracellular matrix, integrins, and growth factors as tailors of the stem cell niche. Curr. Opin. Cell Biol. 2012; 24:645–651.
  44. 44. Jarvelainen H., Sainio A., Koulu M., Wight T.N., Penttinen R. Extracellular matrix molecules: potential targets in pharmacotherapy. Pharmacol. Rev. 2009; 61:198–223.
  45. 45. Schaefer L., Schaefer R.M. Proteoglycans: from structural compounds to signaling molecules. Cell Tissue Res. 2010; 339:237–246.
  46. 46. Rozario T., DeSimone D.W. The extracellular matrix in development and orphogenesis: a dynamic view. Dev. Biol. 2010; 341:126–140.
  47. 47. Wise S.G., Weiss A.S. Tropoelastin. Int. J. Biochem. Cell Biol. 2009; 41:494–497.
  48. 48. Ou K.L., Hosseinkhani H. Development of 3D in vitro technology for medical applications. Int. J. Mol. Sci. 2014; 15:17938–17962.
  49. 49. Kular J.K., Basu S., Sharma R.I. The extracellular matrix: Structure, composition, age-related differences, tools for analysis and applications for tissue engineering. J. Tissue Eng. 2014; 5:1–17.
  50. 50. Bosman F.T., Stamenkovic I. Functional structure and composition of the extracellular matrix. J. Pathol. 2003; 200:423–428.
  51. 51. Zhang J., Niu C., Ye L., Huang H., He X., Tong W.G., Ross J., Haug J., Johnson T., Feng J.Q., Harris S., Wiedemann L.M., Mishina Y., Li L. Identification of the haematopoietic stem cell niche and control of the niche size. Nature. 2003; 425:836–841.
  52. 52. Calvi L.M., Adams G.B., Weibrecht K.W., Weber J.M., Olson D.P., Knight M.C., Martin R.P., Schipani E., Divieti P., Bringhurst F.R., Milner L.A., Kronenberg H.M., Scadden D.T. Osteoblastic cells regulate the haematopoietic stem cell niche. Nature. 2003; 425:841–846.
  53. 53. Yoshihara H., Arai F., Hosokawa K., Hagiwara T., Takubo K., Nakamura Y., Gomei Y., Iwasaki H., Matsuoka S., Miyamoto K., Miyazaki H., Takahashi T., Suda T. Thrombopoietin/MPL signaling regulates hematopoietic stem cell quiescence and interaction with the osteoblastic niche. Cell Stem Cell. 2007; 1:685–697.
  54. 54. Arai F., Hirao A., Ohmura M., Sato H. Tie2/angiopoietin-1 signaling regulates hematopoietic stem cell quiescence in the bone marrow niche. Cell. 2004; 118:149–161.
  55. 55. Weber J.M., Calvi L.M. Notch signaling and the bone marrow hematopoietic stem cell niche. Bone. 2010; 46:281–285.
  56. 56. Nakamura-Ishizu A., Suda T., Hematopoietic stem cell niche: an interplay among a repertoire of multiple functional niches. Biochim. Biophys. Acta. 2013; 1830:2404–2409.
  57. 57. Kunisaki Y., Frenette P.S. The secrets of the bone marrow niche: enigmatic niche brings challenge for HSC expansion. Nat. Med. 2012; 18:864–865.
  58. 58. Shen Y., Nilsson S.K. Bone, microenvironment and hematopoiesis. Curr. Opin. Hematol. 2012; 19:250–255.
  59. 59. Smith J.N.P., Calvi L.M. Concise review: current concepts in bone marrow microenvironmental regulation of hematopoietic stem and progenitor cells. Stem Cells. 2013; 31:1044–1050.
  60. 60. Rompolas P., Mesa K.R., Greco V. Spatial organization within a niche as a determinant of stem-cell fate. Nature. 2013; 502:513–518.
  61. 61. Tian H., Biehs B., Warming S., Leong K.G., Rangell L., Klein O.D., de Sauvage F.J. A reserve stem cell population in small intestine renders Lgr5-positive cells dispensable. Nature. 2011; 478:255–259.
  62. 62. Kilian K.A., Bugarija B., Lahn B.T., Mrksich M. Geometric cues for directing the differentiation of mesenchymal stem cells. Proc. Natl. Acad. Sci. USA. 2010; 107:4872–4877.
  63. 63. Mizuno H., Tobita M., Uysal A.C. Concise review: adipose-derived stem cells as a novel tool for future regenerative medicine. Stem Cells. 2012; 30:804–810.
  64. 64. Sykova E., Forostyak S. Stem cells in regenerative medicine. Laser Ther. 2013; 22(2):87–92.
  65. 65. Tabar V., Studer L. Pluripotent stem cells in regenerative medicine: challenges and recent progress. Nat. Rev. Genet. 2014; 15(2):82–92.
  66. 66. Rippon H.J., Bishop A.E. Embryonic stem cells. Cell Prolif. 2004; 37:23–34.
  67. 67. Lavker R.R., Sun T.T. Epidermal stem cells: properties, markers and location. Proc. Natl. Acad. Sci. USA. 2000; 97:13473.
  68. 68. Uchida N., Buck D.W., He D., Reitsma M.J., Masek M., Phan T.V., Tsukamoto A.S., Gage F.H., Weissman I.L. Direct isolation of human central nervous system stem cells. Proc. Natl. Acad. Sci. USA. 2000; 97:14720.
  69. 69. Vessey C.J., de la Hall P.M. Hepatic stem cells: a review. Pathology (Phila). 2001; 33:130.
  70. 70. Wagers A.J., Christensen J.L., Weissman I.L. Cell fate determination from stem cells. Gene Ther. 2002; 9:606.
  71. 71. McCarthy H.S., Roberts S. A histological comparison of the repair tissue formed when using either Chondrogide® or periosteum during autologous chondrocyte implantation. Osteoarthr. Cartil. 2013; 21:2048–2057.
  72. 72. Bailey A., Goodstone N., Roberts S., Hughes J., Roberts S., van Niekerk L., Richardson J., Rees D. Rehabilitation after oswestry autologous-chondrocyte implantation: the OsCell protocol. J. Sport Rehabil. 2003; 12:104–118.
  73. 73. Roberts S., Menage J., Sandell L.J., Evans E.H., Richardson J.B. Immunohistochemical study of collagen types I and II and procollagen IIA in human cartilage repair tissue following autologous chondrocyte implantation. Knee. 2009; 16:398–404.
  74. 74. Kriks S., Shim J.W., Piao J., Ganat Y.M., Wakeman D.R., Xie Z., Carrillo-Reid L., Auyeung G., Antonacci C, Buch A., Yang L., Beal M.F., Surmeier D.J., Kordower J.H., Tabar V., Studer L., . Dopamine neurons derived from human ES cells efficiently engraft in animal models of Parkinson’s disease. Nature. 2011; 480:547–551.
  75. 75. Ma L., Hu B., Liu Y., Vermilyea S.C., Liu H., Gao L., Sun Y., Zhang X., Zhang S.C. Human embryonic stem cell-derived GABA neurons correct locomotion deficits in quinolinic acid-lesioned mice. Cell Stem Cell. 2012; 10:455–464.
  76. 76. Wang S., Bates J., Li X., Schanz S., Chandler-Militello D., Levine C., Maherali N., Studer L., Hochedlinger K., Windrem M., Goldman S.A.Human iPSC-derived oligodendrocyte progenitor cells can myelinate and rescue a mouse model of congenital hypomyelination. Cell Stem Cell. 2013; 12:252–264.
  77. 77. Shiba Y., Fernandes S., Zhu W., Filice D., Veronica Muskheli V., Kim J., Palpant N.J., Gantz J., White Moyes K., Reinecke H., Van Biber B., Dardas T., Mignone J.L., Izawa A., Hanna R., Viswanathan M., Gold J.D., Kotlikoff M.I., Sarvazyan N., Kay M.W., Murry C.E., Laflamme M.A. Human ES-cell-derived cardiomyocytes electrically couple and suppress arrhythmias in injured hearts. Nature. 2012; 489:322–325.
  78. 78. Smith A., Heath J.K., Donaldson D.D., Wong G.G., Moreau J., Stahl M., Rogers, D. Inhibition of pluripotential embryonic stem cell differentiation by purified polypeptides. Nature. 1988; 336:688.
  79. 79. Williams R.L., Hilton D.J., Pease S., Willson T.A., Stewart C.I., Gearing D.P., Wagner E.F., Metcalf D., Nicola N.A., Gough N.M. Myeloid leukaemia inhibitory factor maintains the developmental potential of embryonic stem cells. Nature. 1988; 36:684.
  80. 80. Thomson J.A., Itskovitz-Eldor J., Shapiro S.S., Waknitz M.A., Swiergiel J.J., Marshall V.S., Jones J.M. Embryonic stem cell lines derived from human blastocysts. Science. 1998; 282(5391):1145–1147.
  81. 81. Hofstetter C.P., Schwarz E.J., Hess D., Widenfalk J., El Manira A., Prockop D.J., Olson L. Marrow stromal cells form guiding strands in injured spinal cord and promote recovery. Proc. Natl. Acad. Sci. USA. 2002; 99(4):2199–2204.
  82. 82. Xu L., Yan J., Chen D., Welsh A.M., Hazel T., Johe K., Hatfield G., Koliatsos V.E. Human neural stem cell grafts ameliorate motor neuron disease in SOD-1 transgenic rats. Transplantation. 2006; 82(7):865–875.
  83. 83. McDonald J.W., Liu X.Z., Qu Y., Liu S., Mickey S.K., Turetsky D., Gottlieb D.I., Choi D.W. Transplanted embryonic stem cells survive, differentiate and promote recovery in injured rat spinal cord. Nat. Med. 1999; 5(12):1410–1412.
  84. 84. Nistor G.I., Totoiu M.O., Haque N., Carpenter M.K., Keirstead H.S. Human embryonic stem cells differentiate into oligodendrocytes in high purity and myelinate after spinal cord transplantation. Glia. 2005; 49(3):385–396.
  85. 85. Seminatore C., Polentes J., Ellman D., Kozubenko N., Itier V., Tine S., Tritschler L., Brenot M., Guidou E., Blondeau J., Lhuillier M., Bugi A., Aubry L., Jendelova P., Sykova E., Perrier A.L., Finsen B., Onteniente B. The postischemic environment differentially impacts teratoma or tumor formation after transplantation of human embryonic stem cell-derived neural progenitors. Stroke J. Cereb. Circ. 2010; 41 (1), 153–159.
  86. 86. Toh W.S., Lee, E.H, Guo X.M., Chan J.K., Yeow C.H., Choo A.B., Cao T. Cartilage repair using hyaluronan hydrogel-encapsulated human embryonic stem cell-derived chondrogenic cells. Biomaterials. 2010; 31(27):6968–6980.
  87. 87. Kehat I., Khimovich L., Caspi O., Gepstein A., Shofti R., Arbel G., Huber I., Satin J., Itskovitz-Eldor J., Gepstein L. Electromechanical integration of cardiomyocytes derived from human embryonic stem cells. Nat. Biotechnol. 2004; 22(10):1282–1289 .
  88. 88. Hwang N.S., Varghese S., Lee H.J., Zhang Z., Ye Z., Bae J., Cheng L., Elisseeff J. In vivo commitment and functional tissue regeneration using human embryonic stem cell-derived mesenchymal cells. Proc. Natl. Acad. Sci. USA. 2008; 105(52):20641–20646.
  89. 89. Lee E.J., Xu L., Kim G.H., Kang S.K., Lee S.W., Park S.H., Kim S., Choi T.H., Kim H.S. Regeneration of peripheral nerves by transplanted sphere of human mesenchymal stem cells derived from embryonic stem cells. Biomaterials. 2012; 33:7039–7046.
  90. 90. Horiguchi S., Takahashi J., Kishi Y., Morizane A., Okamoto Y., Koyanagi M., Tsuji M., Tashiro K., Honjo T., Fujii S., Hashimoto N. Neural precursor cells derived from human embryonic brain retain regional specificity. J. Neurosci. Res. 2004; 75(6):817–824.
  91. 91. Burnstein R.M., Foltynie T., He X., Menon D.K., Svendsen C.N., Caldwell M.A. Differentiation and migration of long term expanded human neural progenitors in a partial lesion model of Parkinson’s disease. Intl. J. Biochem. Cell Biol. 2004; 36(4): 702–713.
  92. 92. Brundin P., Strecker R.E., Gage F.H., Lindvall O., Bjorklund A. Intracerebral transplantation of dopamine neurons: understanding the functional role of the mesolimbocortical dopamine system and developing a therapy for Parkinson’s disease. Ann. NY Acad. Sci. 1988; 537:148–160.
  93. 93. Clarke D.J., Brundin P., Strecker R.E., Nilsson O.G., Bjorklund A., Lindvall O. Experimental brain research. Exp. Cereb. 1988; 73(1):115–126.
  94. 94. Lindvall O., Brundin P., Widner H., Rehncrona S., Gustavii B., Frackowiak R., Leenders, K.L, Sawle G., Rothwell J.C., Marsden C.D., Björklund A. Grafts of fetal dopamine neurons survive and improve motor function in Parkinson’s disease. Science. 1990; 247(4942):574–577.
  95. 95. Pollock K., Stroemer P., Patel S., Stevanato L., Hope A., Miljan E., Dong Z., Hodges H., Price J., Sinden J.D. A conditionally immortal clonal stem cell line from human cortical neuroepithelium for the treatment of ischemic stroke. Exp. Neurol. 2006; 199(1):143–155.
  96. 96. Takahashi K., Yamanaka S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell. 2006; 126:663–676.
  97. 97. Maherali N., Sridharan R., Xie W., Utikal J., Eminli S., Arnold K., Stadtfeld M., Yachechko R., Tchieu J., Jaenisch R., Plath K., Hochedlinger K. Directly reprogrammed fibroblasts show global epigenetic remodeling and widespread tissue contribution. Cell Stem Cell. 2007; 1:55–70.
  98. 98. Okita K., Ichisaka T., Yamanaka S. Generation of germline-competent induced pluripotent stem cells. Nature. 2007; 448:313–317.
  99. 99. Wernig M., vMeissner, A., Foreman, R., Brambrink, T., Ku, M., Hochedlinger, K., Bradley E., Bernstein, B.E., Jaenisch, R. In vitro reprogramming of fibroblasts into a pluripotent ES-cell-like state. Nature. 2007; 448:318–324.
  100. 100. Okita K., Nakagawa M., Hyenjong H., Ichisaka T., Yamanaka S. Generation of mouse induced pluripotent stem cells without viral vectors. Science. 2008; 322:949–953.
  101. 101. Yu J., Hu K., Smuga-Otto K., Tian S., Stewart R., Slukvin I.I., Thomson J.A. Human induced pluripotent stem cells free of vector and transgene sequences. Science. 2009; 324:797–801.
  102. 102. Stadtfeld M., Nagaya M., Utikal J., Weir G., Hochedlinger K. Induced pluripotent stem cells generated without viral integration. Science. 2008; 322:945–949.
  103. 103. Soldner F., Hockemeyer D., Beard C., Gao Q., Bell G.W., Cook E.G., Hargus G., Blak A., Cooper O., Mitalipova M., Isacson O., Jaenisch R. Parkinson’s disease patient-derived induced pluripotent stem cells free of viral reprogramming factors. Cell. 2009; 136:964–977.
  104. 104. Lowry W.E., Richter L., Yachechko R., Pyle A.D., Tchieu J., Sridharan R., Clark A.T., Plath K. Generation of human induced pluripotent stem cells from dermal fibroblasts. Proc. Natl. Acad. Sci. USA. 2008; 105:2883–2888.
  105. 105. Loh Y.H., Agarwal S., Park I., Urbach A., Huo H., Heffner G.C., Kim K., Miller J.D., Ng K., Daley G.Q. Generation of induced pluripotent stem cells from human blood. Blood. 2009; 113:5476–5479.
  106. 106. Kunisato A., Wakatsuki M., Shinba H., Ota T., Ishida I., Nagao K. Direct generation of induced pluripotent stem cells from human non-mobilized blood. Stem Cells Dev. 2011; 20:159–168.
  107. 107. Aasen T., Raya A., Barrero M.J., Garreta E., Consiglio A., Gonzalez F., Vassena R., Bilić J., Pekarik V., Tiscornia G., Edel M., Boué S., Izpisúa Belmonte J.C. Efficient and rapid generation of induced pluripotent stem cells from human keratinocytes. Nat. Biotechnol. 2008; 26:1276–1284.
  108. 108. Wu S.M., Hochedlinger K. Harnessing the potential of induced pluripotent stem cells for regenerative medicine. Nat. Cell Biol. 2011; 13(5):497–505.
  109. 109. Hanna J., Wernig M., Markoulaki S., Sun C.W., Meissner A., Cassady J.P., Beard C., Brambrink T., Wu L.C., Townes T.M., Jaenisch R. Treatment of sickle cell anemia mouse model with iPS cells generated from autologous skin. Science. 2007; 318:1920–1923.
  110. 110. Wernig M., Zhao J.P., Pruszak J., Hedlund E., Fu D., Soldner F., Broccoli V., Constantine-Paton M., Isacson O., Jaenisch R. Neurons derived from reprogrammed fibroblasts functionally integrate into the fetal brain and improve symptoms of rats with Parkinson’s disease. Proc. Natl. Acad. Sci. USA. 2008; 105:5856–5861.
  111. 111. Zhang J., Wilson G.F., Soerens A.G., Koonce C.H., Yu J., Palecek S.P., Thomson J.A., Kamp T.J. Functional cardiomyocytes derived from human induced pluripotent stem cells. Circ. Res. 2009; 104:e30–e41.
  112. 112. Nelson T.J., Martinez-Fernandez A., Yamada S., Perez-Terzic C., Ikeda Y., Terzic A. Repair of acute myocardial infarction by human stemness factors induced pluripotent stem cells. Circulation. 2009; 120:408–416.
  113. 113. Yu J., Vodyanik M.A., Smuga-Otto K., Antosiewicz-Bourget J., Frane J.L., Tian S., Nie J., Jonsdottir G.A., Ruotti V., Stewart R., Slukvin I.I., Thomson J.A. Induced pluripotent stem cell lines derived from human somatic cells. Science. 2007; 318:1917–1920.
  114. 114. Park I.H., Zhao R., West J.A., Yabuuchi A., Huo H., Ince T.A., Lerou P.H., Lensch M.W., Daley G.Q. Reprogramming of human somatic cells to pluripotency with defined factors. Nature. 2008; 451:141–146.
  115. 115. Guenther M.G., Frmapton G.M., Soldner F., Hockemeyer D., Mitalipova M., Jaenisch R., Young R.A. Chromatin structure and gene expression programs of human embryonic and induced pluripotent stem cells. Cell Stem Cell. 2010; 7:249–257.
  116. 116. Chin M.H., Mason M.J., Xie W., Volinia S., Singer M., Peterson C., Ambartsumyan G., Aimiuwu O., Richter L., Zhang J., Khvorostov I., Ott V., Grunstein M., Lavon N., Benvenisty N., Croce C.M., Clark A.T., Baxter T., Pyle A.D., Teitell M.A., Pelegrini M., Plath K., Lowry W.E. Induced pluripotent stem cells and embryonic stem cells are distinguished by gene expression signatures. Cell Stem Cell. 2009; 5:111–123.
  117. 117. Doi A., Park I.H., Wen B., Murakami P., Aryee M.J., Irizarry R., Herb B., Ladd-Acosta C., Rho J., Loewer S., Miller J., Schlaeger T., Daley G.Q., Feinberg A.P. Differential methylation of tissue- and cancer-specific CpG island shores distinguishes human induced pluripotent stem cells, embryonic stem cells and fibroblasts. Nat. Genet. 2009; 41:1350–1353.
  118. 118. Deng J., Shoemaker R., Xie B., Gore A., LeProust E.M., Antosiewicz-Bourget J., Egli D., Maherali N., Park I.H., Yu J., Daley G.Q., Eggan K., Hochedlinger K., Thomson J., Wang W., Gao Y., Zhang K. Targeted bisulfite sequencing reveals changes in DNA methylation associated with nuclear reprogramming. Nat. Biotechnol. 2009; 27:353–360.
  119. 119. Hu B.Y., Weick J.P., Yu J., Ma L.X., Zhang X.Q., Thomson J.A., Zhang S.C. Neural differentiation of human induced pluripotent stem cells follows developmental principles but with variable potency. Proc. Natl. Acad. Sci. USA. 2010; 107:4335–4340.
  120. 120. Feng Q., Lu S.J., Klimanskaya I., Gomes I., Kim D., Chung Y., Honig G.R., Kim K.S., Lanza R. Hemangioblastic derivatives from human induced pluripotent stem cells exhibit limited expansion and early senescence. Stem Cells. 2010; 28:704–712.
  121. 121. Miura K., Okada Y., Aoi T., Okada A., Takahashi K., Okita K., Nakagawa M., Koyanagi M., Tanabe K., Ohnuki M., Ogawa D., Ikeda E., Okano H., Yamanaka S. Variation in the safety of induced pluripotent stem cell lines. Nat. Biotechnol. 2009; 27:743–745.
  122. 122. Stadtfeld M., Apostolou E., Akutsu H., Fukuda A., Follett P., Natesan S., Kono T., Shioda T., Hochedlinger K. Aberrant silencing of imprinted genes on chromosome 12qF1 in mouse induced pluripotent stem cells. Nature. 2010; 465:175–181.
  123. 123. Polo J.M., Liu S., Figueroa M.E., Kulalert W., Eminli S., Tan K.Y., Apostolou E., Stadtfeld M., Li Y., Shioda T., Natesan S., Wagers A.J., Melnick A., Evans T., Hochedlinger K. Cell type of origin influences the molecular and functional properties of mouse induced pluripotent stem cells. Nat. Biotechnol. 2010; 28:848–855.
  124. 124. Laurent, L.C., Ulitsky, I., Slavin, I., Tran, H., Schork, A., Morey, R., Lynch, C., Harness, J.V., Lee, S., Barrero, M.J., Ku, S., Martynova, M., Semechkin, R., Galat, V., Gottesfeld, J., Izpisua Belmonte, J.C., Murry, C., Keirstead, H.S., Park, H.S., Schmidt, U., Laslett, A.L., Muller, F.J., Nievergelt, C.M., Shamir, R., Loring, J.F. Dynamic changes in the copy number of pluripotency and cell proliferation genes in human ESCs and iPSCs during reprogramming and time in culture. Cell Stem Cell. 2011; 8:106–118.
  125. 125. Hussein S.M., Batada N.N., Vuoristo S., Ching R.W., Autio R., Närvä E., Ng S., Sourour M., Hämäläinen R., Olsson C., Lundin K., Mikkola M., Trokovic R., Peitz M., Brüstle O., Bazett-Jones D.P., Alitalo K., Lahesmaa R., Nagy A., Otonkoski T. Copy number variation and selection during reprogramming to pluripotency. Nature. 2011; 471:58–62.
  126. 126. Mayshar Y., Ben-David U., Lavon N., Biancotti J.C., Yakir B., Clark A.T., Plath K., Lowry W.E., Benvenisty N. Identification and classification of chromosomal aberrations in human induced pluripotent stem cells. Cell Stem Cell. 2010; 7:521–531.
  127. 127. Gore A., Li Z., Fung H.L., Young J.E., Agarwal S., Antosiewicz-Bourget J., Canto I., Giorgetti A., Israel M.A., Kiskinis E., Lee J.H., Loh Y.H., Manos P.D., Montserrat N., Panopoulos A.D., Ruiz S., Wilbert M.L., Yu J., Kirkness E.F., Izpisua Belmonte J.C., Rossi D.J., Thomson J.A., Eggan K., Daley G.Q., Goldstein L.S., Zhang K. Somatic coding mutations in human induced pluripotent stem cells. Nature. 2011; 471:63–67.
  128. 128. Bock C., Kiskinis E., Verstappen G., Gu H., Boulting G., Smith Z.D., Ziller M., Croft G.F., Amoroso M.W., Oakley D.H., Gnirke A., Eggan K., Meissner A. Reference Maps of human ES and iPS cell variation enable high-throughput characterization of pluripotent cell lines. Cell. 2011; 144:439–452.
  129. 129. Yang L., Soonpaa M.H., Adler E.D., Roepke T.K., Kattman S.J., Kennedy M., Henckaerts E., Bonham K., Abbott G.W., Linden R.M., Field L.J., Keller G.M. Human cardiovascular progenitor cells develop from a KDR+ embryonic-stem-cell derived population. Nature. 2008; 453:524–528.
  130. 130. van Laake L.W., Passier R., Monshouwer-Kloots J., Verkleij A.J., Lips D.J., Freund C., den Ouden K., Ward-van Oostwaard D., Korving J., Tertoolen L.G., van Echteld C.J., Doevendans P.A., Mummery C.L. Human embryonic stem cell-derived cardiomyocytes survive and mature in the mouse heart and transiently improve function after myocardial infarction. Stem Cell Res. 2007; 1:9–24.
  131. 131. Laflamme M.A., Chen K., Naumova A.V., Muskheli V., Fugate J.A., Dupras S.K., Hassanipour M., Police S., O’Sullivan, C., Collins L., Chen Y., Minami E., Gill E.A., Ueno S., Yuan C., Gold J., Murry C.E. Cardiomyocytes derived from human embryonic stem cells in pro-survival factors enhance function of infarcted rat hearts. Nat. Biotechnol. 2007; 25:1015–1024.
  132. 132. Kattman S.J., Witty A.D., Gagliardi M., Dubois N.C., Niapour M., Hotta A., Ellis J., Keller G. Stage-specific optimization of activin/nodal and bmp signaling promotes cardiac differentiation of mouse and human pluripotent stem cell lines. Cell Stem Cell. 2011; 8:228–240.
  133. 133. Irion S., Clarke R.L., Luche H., Kim I., Morrison S.J., Fehling H.J., Keller G.M. Temporal specification of blood progenitors from mouse embryonic stem cells and induced pluripotent stem cells. Development. 2010; 137:2829–2839.
  134. 134. Swistowski A., Peng J., Liu Q., Mali P., Rao M.S., Cheng L., Zeng, X. Efficient generation of functional dopaminergic neurons from human induced pluripotent stem cells under defined conditions. Stem Cells. 2010; 28:1893904.
  135. 135. Nizzardo M., Simone C., Falcone M., Locatelli F., Riboldi G., Comi G.P., Corti S. Human motor neuron generation from embryonic stem cells and induced pluripotent stem cells. Cell. Mol. Life Sci. 2010; 67:3837–3847.
  136. 136. Ma YD, Lugus JJ, Park C, Choi K. Differentiation of Mouse Embryonic Stem Cells into Blood. Current protocols in stem cell biology. 2008;CHAPTER:Unit-1F.4. doi:10.1002/9780470151808.sc01f04s6.
  137. 137. D’Amour, K.A, Agulnick A.D., Eliazer S., Kelly O.G., Kroon E., Baetge E.E. Efficient differentiation of human embryonic stem cells to definitive endoderm. Nat. Biotechnol. 2005; 23:1534–1541.
  138. 138. Kroon E., Martinson L.A., Kadoya K., Bang A.G., Kelly O.G., Eliazer S., Young H., Richardson M., Smart N.G., Cunningham J., Agulnick A.D., D’Amour K.A., Carpenter M.K., Baetge E.E. Pancreatic endoderm derived from human embryonic stem cells generates glucose responsive insulin-secreting cells in vivo. Nat. Biotechnol. 2008; 26:443–452.
  139. 139. Strauss S. Geron trial resumes, but standards for stem cell trials remain elusive. Nat. Biotechnol. 2010; 28:989–990.
  140. 140. Lister R., Kida Y.S., Hawkins R.D., Nery J.R., Hon G., Antosiewicz-Bourget J., O’Malley, R., Castanon R., Klugman S., Downes M., Yu., R., Stewart, R., Ren, B., Thomson, J.A., Evans, R.M., Ecker, J.R. Hotspots of aberrant epigenomic reprogramming in human induced pluripotent stem cells. Nature. 2011; 471:68–73.
  141. 141. Warren L., Manos P.D., Ahfeldt T., Loh Y.H., Li H., Lau F., Ebina W., Mandal P.K., Smith Z.D., Meissner A., Daley G.Q., Brack A.S., Collins J.J., Cowan C., Schlaeger T.M., Rossi D.J. Highly efficient reprogramming to pluripotency and directed differentiation of human cells with synthetic modified mRNA. Cell Stem Cell. 2010; 7:618–630.
  142. 142. Fusaki N., Ban H., Nishiyama A., Saeki K., Hasegawa M. Efficient induction of transgene-free human pluripotent stem cells using a vector based on Sendai virus, an RNA virus that does not integrate into the host genome. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 2009; 85:348–362.
  143. 143. Polo J.M., Liu S., Figueroa M.E., Kulalert W., Eminli S., Tan K.Y., Apostolou E., Stadtfeld M., Li Y., Shioda T., Natesan S., Wagers A.J., Melnick A., Evans T., Hochedlinger K. Cell type of origin influences the molecular and functional properties of mouse induced pluripotent stem cells. Nat. Biotechnol. 2010; 28:848–855.
  144. 144. Kim K., Zhao R., Doi A., Ng K., Unternaehrer J., Cahan P., Huo H., Loh Y.H., Aryee M.J., Lensch M.W., Li H., Collins J.J., Feinberg A.P., Daley G.Q. Donor cell type can influence the epigenome and differentiation potential of human induced pluripotent stem cells. Nat. Biotechnol. 2011; 29:1117–1119.
  145. 145. Kim K., Doi A., Wen B., Ng K., Zhao R., Cahan P., Kim J., Aryee M.J., Ji H., Ehrlich L.I.R., Yabuuchi A., Takeuchi A., Cunniff K.C., Hongguang H., Mckinney-Freeman S., Naveiras O., Yoon T.J., Irizarry R.A., Jung N., Seita J., Hanna J., Murakami P., Jaenisch R., Weissleder R., Orkin S.H. Epigenetic memory in induced pluripotent stem cells. Nature. 2010; 467:285–290.
  146. 146. Bar-Nur O., Russ H.A., Efrat S., Benvenisty N. Epigenetic memory and preferential lineage-specific differentiation in induced pluripotent stem cells derived from human pancreatic islet β cells. Cell Stem Cell. 2011; 9:17–23.
  147. 147. Cyranoski D. Stem cells cruise to clinic. Nature. 2013; 494:413.
  148. 148. Friedenstein A.J., Gorskaja J.F., Kulagina N.N. Fibroblast precursors in normal and irradiated mouse hematopoietic organs. Exp. Hematol. 1976; 4:267–274.
  149. 149. Caplan A.I. Mesenchymal stem cells. J. Orthop. Res. 1991; 9(5):641–50.
  150. 150. Arboleda D., Forostyak S., Jendelova P., Marekova D., Amemori T., Pivonkova H., Masinova K., Sykova E. Transplantation of predifferentiated adipose-derived stromal cells for the treatment of spinal cord injury. Cell. Mol. Neurobiol. 2011; 31(7):1113–1122.
  151. 151. Krause D.S., Plasticity of marrow-derived stem cells. Gene Ther. 2002; 9(11):754–758.
  152. 152. Mezey E., Chandross K.J., Harta G., Maki R.A., McKercher S.R. Turning blood into brain: cells neuronal antigens generated in vivo from bone marrow. Science. 2000; 290(5497):1779–1782.
  153. 153. Prockop D.J. Marrow stromal cells as stem cells for nonhematopoietic tissues. Science. 1997; 276(5309):71–74.
  154. 154. La Spada A., Ranum L.P. Molecular genetic advances in neurological disease:special review issue. Hum. Mol. Genet. 2010; 19(R1):R1–3.
  155. 155. Urdzikova L., Jendelova P., Glogarova K., Burian M., Hajek M., Sykova E. Transplantation of bone marrow stem cells as well as mobilization by granulocyte-colony stimulating factor promotes recovery after spinal cord injury in rats. J. Neurotrauma. 2006; 23(9):1379–1391.
  156. 156. Sykova E., Jendelova P. Migration, fate and in vivo imaging of adult stem cells in the CNS. Cell Death Differ. 2007; 14(7):1336–1342.
  157. 157. Hejcl A., Sedy J., Kapcalova M., Toro D.A., Amemori T., Lesny P., Likavcanova-Masinova K., Krumbholcova E., Pradny M., Michalek J., Burian M., Hajek M., Jendelova P., Sykova E. HPMA-RGD hydrogels seeded with mesenchymal stem cells improve functional outcome in chronic spinal cord injury. Stem cells and development 2010; 19(10):1535–1546.
  158. 158. Ohta M., Suzuki Y., Noda T., Ejiri Y., Dezawa M., Kataoka K., Chou H., Ishikawa N., Matsumoto N., Iwashita Y., Mizuta E., Kuno S., Ide C. Bone marrow stromal cells infused into the cerebrospinal fluid promote functional recovery of the injured rat spinal cord with reduced cavity formation. Exp. Neurol. 2004; 187 (2), 266–278.
  159. 159. Daley G.Q. The promise and perils of stem cell therapeutics. Cell Stem Cell. 2012; 10:740–749.
  160. 160. Bang O.Y., Lee J.S., Lee P.H., Lee G. Autologous mesenchymal stem cell transplantation in stroke patients. Ann. Neurol. 2005; 57 (6), 874–882.
  161. 161. McNiece I. Stem cells and regenerative medicine. J. Regen. Med. 2012; 1:1.
  162. 162. Song L., Young N.J., Webb N.E., Tuan R.S. Origin and characterization of multipotential mesenchymal stem cells derived from adult human trabecular bone. Stem Cells Dev. 2005; 14:712–721.
  163. 163. Choi Y.S., Noh S.E., Lim S.M., Lee C.W., Kim C.S., Im M.W., Lee M.H., Kim D.I. Multipotency and growth characteristic of periosteum-derived progenitor cells for chondrogenic, osteogenic, and adipogenic differentiation. Biotechnol. Lett. 2008; 30:593–601.
  164. 164. De Bari C., Dell’Accio F., Tylzanowski P., Luyten F.P. Multipotent mesenchymal stem cells from adult human synovial membrane. Arthritis Rheum. 2001; 44:1928–1942.
  165. 165. Dodson M.V., Hausman G.J., Guan L., Du M., Rasmussen T.P., Poulos S.P., Mir P., Bergen W.G., Fernyhough M.E., McFarland D.C., Rhoads R.P., Soret B., Reecy J.M., Velleman S.G., Jiang Z. Skeletal muscle stem cells from animals I. Basic cell biology. Int. J. Biol. Sci. 2010; 6:465–474.
  166. 166. Belicchi M., Pisati F., Lopa R., Porretti L., Fortunato F., Sironi M., Scalamogna M., Parati E.A., Bresolin N., Torrente Y. Human skin-derived stem cells migrate throughout forebrain and differentiate into astrocytes after injection into adult mouse brain. J. Neurosci. Res. 2004; 77:475–486.
  167. 167. Feng J., Mantesso A., Sharpe P.T. Perivascular cells as mesenchymal stem cells. Exp. Opin. Biol. Ther. 2010; 10:1441–1451.
  168. 168. Shi M., Ishikawa M., Kamei N., Nakasa T., Adachi N., Deie M., Asahara T., Ochi M. Acceleration of skeletal muscle regeneration in a rat skeletal muscle injury model by local injection of human peripheral blood-derived cd133-positive cells. Stem Cells. 2009; 27:949–960.
  169. 169. Baksh D., Yao R., Tuan R.S. Comparison of proliferative and multilineage differentiation potential of human mesenchymal stem cells derived from umbilical cord and bone marrow. Stem Cells. 2007; 25:1384–1392.
  170. 170. Musina R.A., Bekchanova E.S., Sukhikh G.T. Comparison of mesenchymal stem cells obtained from different human tissues. Bull. Exp. Biol. Med. 2005; 139:504–509.
  171. 171. Gimble J.M., Katz A.J., Bunnell B.A. Adipose-derived stem cells for regenerative medicine. Circ. Res. 2007; 100:1249–1260.
  172. 172. Tobita M., Orbay H., Mizuno H. Adipose-derived stem cells: Current findings and future perspectives. Discov. Med. 2011; 11:160–170.
  173. 173. Weisberg S.P., McCann D., Desai M., Rosenbaum M., Leibel R.L., Ferrante A.W., Jr. Obesity is associated with macrophage accumulation in adipose tissue. J. Clin. Investig. 2003; 112:1796–1808.
  174. 174. Xu H., Barnes G.T., Yang Q., Tan G., Yang D., Chou C.J., Sole J., Nichols A., Ross J.S., Tartaglia L.A., Chen H. Chronic inflammation in fat plays a crucial role in the development of obesity-related insulin resistance. J. Clin. Investig. 2003; 112:1821–1830.
  175. 175. Zuk P.A., Zhu M., Mizuno H., Huang J., Futrell J.W., Katz A.J., Benhaim P., Lorenz H.P., Hedrick M.H. Multilineage cells from human adipose tissue: implications for cell-based therapies. Tissue Eng. 2001; 7:211–228.
  176. 176. De Ugarte D.A., Morizono K., Elbarbary A., Alfonso Z., Zuk P.A., Zhu M., Dragoo J.L., Ashjian P., Thomas B., Benhaim P., Chen I., Fraser J., Hedrick M.H. Comparison of multi-lineage cells from human adipose tissue and bone marrow. Cells Tissues Organs. 2003; 174:101–109.
  177. 177. Izadpanah R., Trygg C., Patel B., Kriedt C., Dufour J., Gimble J.M., Bunnell B.A. Biologic properties of mesenchymal stem cells derived from bone marrow and adipose tissue. J. Cell Biochem. 2006; 99:1285–1297.
  178. 178. Mitchell J.B., McIntosh K., Zvonic S., Garrett S., Floyd Z.E., Kloster A., Di Halvorsen Y., Storms R.W., Goh B., Kilroy G., Wu X., Gimble J.M. Immunophenotype of human adipose-derived cells: temporal changes in stromal-associated and stem cell-associated markers. Stem Cells. 2006; 24:376–385.
  179. 179. Brayfield C.A., Marra K.G., Rubin J.P. Adipose tissue regeneration. Curr. Stem Cell Res. Ther. 2010; 5:116–121.
  180. 180. Cherubino M., Marra K.G. Adipose-derived stem cells for soft tissue reconstruction. Regen. Med. 2009; 4:109–117.
  181. 181. Rubin J.P., Marra K.G. Soft tissue reconstruction. Methods Mol. Biol. 2011; 702:395–400.
  182. 182. Dragoo JL, Samimi B, Zhu M . Tissue-engineered cartilage and bone using stem cells from human infrapatellar fat pads. J. Bone Jt. Surg. Br. 2003; 85:740–747.
  183. 183. Lee W.C., Sepulveda J.L., Rubin J.P., Marra K.G. Cardiomyogenic differentiation potential of human adipose precursor cells. Int. J. Cardiol. 2009; 133:399–401.
  184. 184. Planat-Bénard V., Menard C., André M., Puceat M., Perez A., Garcia-Verdugo J.M., Pénicaud L., Casteilla L. Spontaneous cardiomyocyte differentiation from adipose tissue stroma cells. Circ. Res. 2004; 94:223–229.
  185. 185. Estes B.T., Diekman B.O., Gimble J.M., Guilak F. Isolation of adipose derived stem cells and their induction to a chondrogenic phenotype. Nat. Protoc. 2010; 5:1294–1311.
  186. 186. Estes B.T., Guilak F. Three-dimensional culture systems to induce chondrogenesis of adipose-derived stem cells. Methods Mol. Biol. 2011; 702:201–217.
  187. 187. Rehman J., Traktuev D., Li J., Merfeld-Clauss S., Temm-Grove C.J., Bovenkerk J.E., Pell C.L., Johnstone B.H., Considine R.V., March K.L. Secretion of angiogenic and anti-apoptotic factors by human adipose stromal cells. Circulation. 2004; 109:1292–1298.
  188. 188. Cherubino M., Rubin J.P., Miljkovic N., Kelmendi-Doko A., Marra K.G. Adipose-derived stem cells for wound healing applications. Ann. Plast. Surg. 2011; 66:210–215.
  189. 189. Uysal A.C., Mizuno H. Tendon regeneration and repair with adipose derived stem cells. Curr. Stem Cell Res. Ther. 2010; 5:161–167.
  190. 190. Aurich H., Sgodda M., Kaltwasser P., Vetter M., Weise A., Liehr T., Brulport M., Hengstler J.G., Dollinger M.M., Fleig W.E., Christ B. Hepatocyte differentiation of mesenchymal stem cells from human adipose tissue in vitro promotes hepatic integration in vivo. Gut. 2009; 58:570–581.
  191. 191. Banas A., Teratani T., Yamamoto Y., Tokuhara M., Takeshita F., Osaki M., Kato T., Okochi H., Ochiya T. Rapid hepatic fate specification of adipose-derived stem cells and their therapeutic potential for liver failure. J. Gastroenterol. Hepatol. 2009; 24:70–77.
  192. 192. Mesimäki K., Lindroos B., Törnwall J., Mauno J., Lindqvist C., Kontio R., Miettinen S., Suuronen R. Novel maxillary reconstruction with ectopic bone formation by gmp adipose stem cells. Int. J. Oral. Maxillofac. Surg. 2009; 38:201–209.
  193. 193. Lendeckel S., Jödicke A., Christophis P., Heidinger K., Wolff J., Fraser J.K., Hedrick M.H., Berthold L., Howaldt H.P. Autologous stem cells (adipose) and fibrin glue used to treat widespread traumatic calvarial defects: case report. J. Craniomaxillofac. Surg. 2004; 32:370–373.
  194. 194. Gimble J.M., Guilak F., Bunnell B.A. Clinical and preclinical translation of cell-based therapies using adipose tissue-derived cells. Stem Cell Res. Ther. 2010; 1:19.
  195. 195. Bailey A.M., Kapur S., Katz A.J. Characterization of adipose-derived stem cells: an update. Curr. Stem Cell Res. Ther. 2010; 5:95–102.
  196. 196. Adams G.B., Scadden D.T. A niche opportunity for stem cell therapeutics. Gene Ther. 2008; 15:96–99.
  197. 197. Guilak F., Cohen D.M., Estes B.T., Gimble J.M., Liedtke W., Chen C.S. Control of stem cell fate by physical interactions with the extracellular matrix. Cell Stem Cell. 2009;5:17–26.
  198. 198. Ratne B.D. Biomaterials Science: an Introduction to Materials in Medicine, 2nd ed. Amsterdam, The Netherlands: Elsevier Academic Press, 2004.
  199. 199. Lafrenie, R.M, Yamada K.M. Integrin-dependent signal transduction. J Cell. Biochem. 1996; 61:543–553.
  200. 200. Kuhl P.R., Griffith-Cima L.G. Tethered epidermal growth factor as a paradigm for growth factor-induced stimulation from the solid phase. Nat. Med. 1996; 2:1022e7.
  201. 201. Kelly P.N., Dakic A., Adams J.M., Nutt S.L., Strasser, A. Tumor growth need not be driven by rare cancer stem cells. Science. 2007; 317:337.
  202. 202. Cheng A., Kapacee Z., Peng J., Lucas R.J., Hardingham T.E., Kimber S.J. Cartilage repair using human embryonic stem cell-derived chondroprogenitors. Stem Cells Transl. Med. 2014; 3:1287–1294.
  203. 203. Oldershaw R.A., Baxter M.A., Lowe E.T., Bates N., Grady L.M., Soncin F., Brison D.R., Hardingham T.E., Kimber S.J. Directed differentiation of human embryonic stem cells toward chondrocytes. Nat. Biotechnol. 2010; 28:1187–1194.
  204. 204. Guo W., Lasky J.L., Chang C.J., Mosessian S., Lewis X., Xiao Y., Yeh J.E., Chen J.Y., Iruela-Arispe M.L., Varella-Garcia M., Wu H. Nature. 2008; 453:529.
  205. 205. Neering S.J., Bushnell T., Sozer S., Ashton J., Rossi R.M., Wang P.Y., Bell D.R., Heinrich D., Bottaro A., Jordan C.T. Leukemia stem cells in a genetically defined murine model of blast-crisis CML. Blood. 2007; 110:2578.
  206. 206. Hern D.L., Hubbell J.A. Incorporation of adhesion peptides into non-adhesive hydrogels useful for tissue resurfacing. J. Biomed. Mater. Res. 1998; 39:266–276.
  207. 207. Chien C.Y., Tsai W.B. Poly(dopamine)-assisted immobilization of Arg-Gly-Asp peptides, hydroxyapatite, and bone morphogenic protein-2 on titanium to improve the osteogenesis of bone marrow stem cells. ACS Appl. Mater. Interfaces. 2013; 5:6975–6983.
  208. 208. Hsiong S.X., Huebsch N., Fischbach C., Kong H.J., Mooney D.J. Integrin-adhesion ligand bond formation of pre-osteoblasts and stem cells in three-dimensional RGD presenting matrices. Biomacromolecules. 2008; 9:1843–1851.
  209. 209. Luu H.H., Song W.X., Luo X., Manning D., Luo J., Deng Z.L., Sharff K.A., Montag A.G., Haydon R.C., He T.C. Distinct roles of bone morphogenetic proteins in osteogenic differentiation of mesenchymal stem cells. J Orthop. Res. 2007; 25:665–677.
  210. 210. Zouani O.F., Kalisky J., Ibarboure E., Durrieu M.C. Effect of BMP-2 from matrices of different stiffnesses for the modulation of stem cell fate. Biomaterials. 2013; 34:2157–2166.
  211. 211. Shin H., Zygourakis K., Farach-Carson M.C., Yaszemski M.J., Mikos A.G. Attachment, proliferation, and migration of marrow stromal osteoblasts cultured on biomimetic hydrogels modified with an osteopontin-derived peptide. Biomaterials. 2004; 25:895–906.
  212. 212. Benoit D.S., Schwartz M.P., Durney A.R., Anseth K.S. Small functional groups for controlled differentiation of hydrogel encapsulated human mesenchymal stem cells. Nat. Mater. 2008; 7:816–823.
  213. 213. Phillips J.E., Petrie T.A., Creighton F.P., García A.J. Human mesenchymal stem cell differentiation on self-assembled monolayers presenting different surface chemistries. Acta Biomater. 2010; 6:12–20.
  214. 214. Griffin M.F., Butler P.E., Seifalian A.M., Kalaskar D.M. Control of stem cell fate by engineering their micro and nanoenvironment. World J Stem Cells. 2015; 7(1):37–50.
  215. 215. Abrams G.A., Goodman S.L., Nealey P.F., Franco M., Murphy C.J. Nanoscale topography of the basement membrane underlying the corneal epithelium of the rhesus macaque. Cell Tissue Res. 2000; 299:39–46.
  216. 216. van Dorp W.F., Zhang X., Feringa B.L., Hansen T.W., Wagner J.B., De Hosson J.T. Molecule-by-molecule writing using a focused electron beam. ACS Nano. 2012; 6:10076–10081.
  217. 217. Lai G.J., Shalumon K.T., Chen S.H., Chen J.P. Composite chitosan/silk fibroin nanofibers for modulation of osteogenic differentiation and proliferation of human mesenchymal stem cells. Carbohydr. Polym. 2014; 111:288–297.
  218. 218. Xu T., Zhao W., Zhu J.M., Albanna M.Z., Yoo J.J., Atala A. Complex heterogeneous tissue constructs containing multiple cell types prepared by inkjet printing technology. Biomaterials. 2013; 34:130–139.
  219. 219. Dalby M.J., McCloy D., Robertson M., Wilkinson C.D., Oreffo R.O. Osteoprogenitor response to defined topographies with nanoscale depths. Biomaterials. 2006; 27:1306–1315.
  220. 220. Zanetti N.C., Solursh M. Induction of chondrogenesis in limb mesenchymal cultures by disruption of the actin cytoskeleton. J. Cell Biol. 1984; 99(1 Pt 1):115–123.
  221. 221. McBride S.H., Knothe Tate M.L. Modulation of stem cell shape and fate A: the role of density and seeding protocol on nucleus shape and gene expression. Tissue Eng. Part A. 2008; 14:1561–1572.
  222. 222. Folkman J., Moscona A. Role of cell shape in growth control. Nature. 1978; 273:345–349.
  223. 223. Manasek F.J., Burnside M.B., Waterman R.E. Myocardial cell shape change as a mechanism of embryonic heart looping. Dev. Biol. 1972; 29:349–371.
  224. 224. Ingber D. Extracellular matrix and cell shape: potential control points for inhibition of angiogenesis. J. Cell. Biochem. 1991; 47:236–241.
  225. 225. Lutolf M.P., Gilbert P.M., Blau H.M. Designing materiasl to direct stem-cell fate. Nature. 2009; 46(26):433–441.
  226. 226. Flaim C.J., Chien S., Bhatia S.N. An extracellular matrix microarray for probing cellular differentiation. Nat. Methods. 2005; 2:119–125.
  227. 227. Soen Y., Mori A., Palmer T.D., Brown P.O. Exploring the regulation of human neural precursor cell differentiation using arrays of signaling microenvironments. Mol. Syst. Biol. 2006; 2:37.
  228. 228. Holtzer H., Abbott J., Lash J., Holtzer S. The loss of phenotypic traits by differentiated cells in vitro, I. Dedifferentiation of cartilage cells. Proc. Natl. Acad. Sci. USA. 1960; 46:1533–1542.
  229. 229. Abbott J., Holtzer H. The loss of phenotypic traits by differentiated cells. 3. The reversible behavior of chondrocytes in primary cultures. J. Cell Biol. 1966; 28:473–487.
  230. 230. Benya P.D., Shaffer J.D. Dedifferentiated chondrocytes re-express the differentiated collagen phenotype when cultured in agarose gels. Cell. 1982; 30:215–224.
  231. 231. Erickson G.R., Gimble J.M., Franklin D.M., Rice H.E., Awad H., Guilak F. Chondrogenic potential of adipose tissue-derived stromal cells in vitro and in vivo. Biochem. Biophys. Res. Commun. 2002; 290:763–769.
  232. 232. Hoben G.M., Koay E.J., Athanasiou K.A. Fibrochondrogenesis in two embryonic stem cell lines: effects of differentiation timelines. Stem Cells. 2008; 26:422–430.
  233. 233. Johnstone B., Hering T.M., Caplan A.I., Goldberg V.M., Yoo J.U. In vitro chondrogenesis of bone marrow-derived mesenchymal progenitor cells. Exp. Cell Res. 1998; 238:265–272.
  234. 234. Awad H.A., Wickham M.Q., Leddy H.A., Gimble J.M., Guilak F. Chondrogenic differentiation of adipose-derived adult stem cells in agarose, alginate, and gelatin scaffolds. Biomaterials. 2004; 25:3211–3222.
  235. 235. Oh S., Brammer K.S., Li Y.S., Teng D., Engler A.J., Chien S., Jin S. Stem cell fate dictated solely by altered nanotube dimension. Proc. Natl. Acad. Sci. USA. 2009; 106:2130–2135.
  236. 236. Brammer K.S., Oh S., Cobb C.J., Bjursten L.M., van der Heyde H., Jin S. Improved bone-forming functionality on diametercontrolled TiO(2) nanotube surface. Acta Biomater. 2009; 5:3215–3223.
  237. 237. Dalby M.J., Gadegaard N., Tare R., Andar A., Riehle M.O., Herzyk P., Wilkinson C.D., Oreffo R.O. The control of human mesenchymal cell differentiation using nanoscale symmetry and disorder. Nat. Mater. 2007; 6:997–1003.
  238. 238. Zouani O.F., Chanseau C., Brouillaud B., Bareille R., Deliane F., Foulc M.P., Mehdi A., Durrieu M.C. Altered nanofeature size dictates stem cell differentiation. J. Cell Sci. 2012; 125:1217–1224.
  239. 239. Reinwald, Y., Shakesheff, K. and Howdle, S. (2011) Biomedical Devices, in Porous Polymers (eds M. S. Silverstein, N. R. Cameron and M. A. Hillmyer), John Wiley & Sons, Inc., Hoboken, NJ, USA. doi: 10.1002/9780470929445.ch9
  240. 240. Underhill G.H., Bhatia S.N. High-throughput analysis of signals regulating stem cell fate and function. Curr. Opin. Chem. Biol. 2007; 11:357–366.
  241. 241. Gidrol X., Fouqué B., Ghenim L., Haguet V., Picollet-D’hahan N., Schaack B. 2D and 3D cell microarrays in pharmacology. Curr. Opin. Pharmacol. 2009; 9:664–668.
  242. 242. Fernandes T.G., Diogo M.M., Clark D.S., Dordick J.S., Cabral J.M.S. High-throughput cellular microarray platforms: applications in drug discovery, toxicology and stem cell research. Trends Biotechnol. 2009; 27:342–349.
  243. 243. Lee, M.Y., Kumar, R.A., Sukumaran, S.M., Hogg, M.G., Clark, D.S., Dordick, J.S. Three-dimensional cellular microarray for high-throughput toxicology assays. Proc. Natl. Acad. Sci. USA. 2008; 105:59–63.
  244. 244. Jongpaiboonkit L., King W.J., Murphy W.L. Screening for 3D environments that support human mesenchymal stem cell viability using hydrogel arrays. Tissue Eng. A. 2009; 15:343–353.
  245. 245. Albrecht D.R., Underhill G.H., Wassermann T.B., Sah R.L., Bhatia S.N. Probing the role of multicellular organization in three-dimensional microenvironments. Nat. Methods. 2006; 3:369–375.
  246. 246. Sudo R., Chung S., Zervantonakis I.K., Vickerman V., Toshimitsu Y., Griffith L.G., Kamm R.D. Transport-mediated angiogenesis in 3D epithelial coculture. FASEB J. 2009; 23:2155–2164.
  247. 247. Whitesides G.M. The origins and the future of microfluidics. Nature. 2006; 442:368–373.
  248. 248. Chung B.G., Flanagan L.A., Rhee S.W., Schwartz P.H., Lee A.P., Monuki E.S., Jeon N.L. Human neural stem cell growth and differentiation in a gradient generating microfluidic device. Lab Chip. 2005; 5:401–406.
  249. 249. Choi N.W., Cabodi M., Held B., Gleghorn J.P., Bonassar L.J., Stroock A.D. Microfluidic scaffolds for tissue engineering. Nat. Mater. 2007; 6:908–915.
  250. 250. Kronenberg H.M. PTH regulates the hematopoietic stem cell niche in bone. Adv. Exp. Med. Biol. 2007; 602:57–60.
  251. 251. Kasper S. Exploring the origins of the normal prostate and prostate cancer stem cell. Stem Cell Rev. 2008; 4:193–201.
  252. 252. Brisken C., Duss S. Stem cells and the stem cell niche in the breast:an integrated hormonal and developmental perspective. Stem Cell Rev. 2007; 3:147–156.
  253. 253. Nur-E-Kamal A., Ahmed I., Kamal J., Babu A.N., Schindler M., Meiners S. Covalently attached FGF-2 to three-dimensional polyamide nanofibrillar surfaces demonstrates enhanced biological stability and activity. Mol. Cell. Biochem. 2008; 309:157–166.
  254. 254. Fan V.H., Tamama K., Au A., Littrell R., Richardson L.B., Wright J.W., Wells A., Griffith L.G. Tethered epidermal growth factor provides a survival advantage to mesenchymal stem cells. Stem Cells. 2007; 25:1241–1251.
  255. 255. Anderson D.G., Levenberg S., Langer R. Nanoliter-scale synthesis of arrayed biomaterials and application to human embryonic stem cells. Nat. Biotechnol. 2004; 22:863–866.
  256. 256. Kshitiz D.M.G., Hubbi M.E., Ahn E.H., Downey J., Afzal J., Kim D.H., Rey S., Chang C., Kundu A., Semenza G.L., Abraham, R.M., Levchenko, A. Matrix rigidity controls endothelial differentiation and morphogenesis of cardiac precursors. Sci. Signal. 2012; 5:ra41.
  257. 257. Dela Paz N.G., Walshe T.E., Leach L.L., Saint-Geniez M., D’Amore, P.A. Role of shear-stress-induced VEGF expression in endothelial cell survival. J. Cell Sci. 2012; 125:831–843.
  258. 258. Bakeine G.J., Ban J., Grenci G., Pozzato A., Zilio S.D., Prasciolu M., Businaro L., Tormen M., Ruaro M.E. Design, fabrication and evaluation of nanoscale surface topography as a tool in directing differentiation and organisation of embryonic stem-cell-derived neural precursors. Microelect. Eng. 2009; 86:1435–1438.
  259. 259. Ingber D.E. The mechanochemical basis of cell and tissue regulation. Mech. Chem. Biosyst. 2004; 1:53–68.
  260. 260. Guo W.H., Frey M.T., Burnham N.A., Wang Y.L. Substrate rigidity regulates the formation and maintenance of tissues. Biophys. J. 2006; 90:2213–2220.
  261. 261. Pelham R.J. Jr., Wang Y. Cell locomotion and focal adhesions are regulated by substrate flexibility. Proc. Natl. Acad. Sci. USA. 1997; 94:13661–13665.
  262. 262. Hadjipanayi E., Mudera V., Brown R.A. Close dependence of fibroblast proliferation on collagen scaffold matrix stiffness. J. Tissue Eng. Regen. Med. 2008; 3:77–84.
  263. 263. Wang H.B., Dembo M., Wang Y.L. Substrate flexibility regulates growth and apoptosis of normal but not transformed cells. Am. J. Physiol. Cell Physiol. 2000; 279:C1345–C1350.
  264. 264. Engler A.J., Sen S., Sweeney H.L., Discher D.E. Matrix elasticity directs stem cell lineage specification. Cell. 2006; 126:677–689.
  265. 265. Wingate K., Bonani W., Tan Y., Bryant S.J., Tan W. Compressive elasticity of three-dimensional nanofiber matrix directs mesenchymal stem cell differentiation to vascular cells with endothelial or smooth muscle cell markers. Acta Biomater. 2012; 8:1440–1449.
  266. 266. Nieponice A., Soletti L., Guan J., Deasy B.M., Huard J., Wagner W.R., Vorp D.A. Development of a tissue-engineered vascular graft combining a biodegradable scaffold, muscle-derived stem cells and a rotational vacuum seeding technique. Biomaterials. 2008; 29:825–833.
  267. 267. Zhang G., Drinnan C.T., Geuss L.R., Suggs L.J. Vascular differentiation of bone marrow stem cells is directed by a tunable three-dimensional matrix. Acta Biomater. 2010; 6:3395–3403.
  268. 268. Banerjee A., Arha M., Choudhary S., Ashton R.S., Bhatia S.R., Schaffer D.V., Kane R.S. The influence of hydrogel modulus on the proliferation and differentiation of encapsulated neural stem cells. Biomaterials. 2009; 30:4695–4699.
  269. 269. Murphy C.M., Matsiko A., Haugh M.G., Gleeson J.P., O’Brien, F.J. Mesenchymal stem cell fate is regulated by the composition and mechanical properties of collagen-glycosaminoglycan scaffolds. J. Mech. Behav. Biomed. Mater. 2012; 11:53–62.
  270. 270. Janmey P.A., McCulloch C.A. Cell mechanics: integrating cell responses to mechanical stimuli. Annu. Rev. Biomed. Eng. 2007; 9:1–34.
  271. 271. Chen W., Villa-Diaz L.G., Sun Y., Weng S., Kim J.K., Lam R.H., Han L., Fan R., Krebsbach P.H., Fu J. Nanotopography influences adhesion, spreading, and self-renewal of human embryonic stem cells. ACS Nano. 2012; 6:4094–4103.
  272. 272. Chuang C.K., Sung L.Y., Hwang S.M., Lo W.H., Chen H.C., Hu Y.C. Baculovirus as a new gene delivery vector for stem cell engineering and bone tissue engineering. Gene Ther. 2007; 14:1417–1424.
  273. 273. Lee S.J., Kang S.W., Do H.J., Han I., Shin D.A., Kim J.H., Lee S.H. Enhancement of bone regeneration by gene delivery of BMP2/Runx2 bicistronic vector into adipose-derived stromal cells. Biomaterials. 2010; 31:5652–5659.
  274. 274. Morito A., Kida Y., Suzuki K., Inoue K., Kuroda N., Gomi K., Arai T., Sato T. Effects of basic fibroblast growth factor on the development of the stem cell properties of human dental pulp cells. Arch. Histol. Cytol. 2009; 72:51–64.
  275. 275. Zhang J., Wilson G.F., Soerens A.G., Koonce C.H., Yu J., Palecek S.P., Thomson J.A., Kamp T.J. Functional cardiomyocytes derived from human induced pluripotent stem cells. Circ. Res. 2009; 104:e30–e41.
  276. 276. Levenberg S., Ferreira L.S., Chen-Konak L., Kraehenbuehl T.P., Langer R. Isolation, differentiation and characterization of vascular cells derived from human embryonic stem cells. Nat. Protocol. 2010; 5:1115–1126.
  277. 277. Ingber D.E. Mechanobiology and diseases of mechanotransduction. Ann. Med. 2003; 35:564–577.
  278. 278. Estes B.T., Gimble J.M., Guilak F. Mechanical signals as regulators of stem cell fate. Curr. Top. Dev. Biol. 2004; 60:91–126.
  279. 279. Kshitiz Park, J., Kim, P., Helen, W., Engler, A.J., Levchenko, A., Kim, D.H. Control of stem cell fate and function by engineering physical microenvironments. Integr. Biol. 2012; 4(9):1008–1018.
  280. 280. Glücksmann A. The role of mechanical stresses in bone formation in vitro. J. Anat. 1942; 76:132–139.
  281. 281. Fang J., Hall B.K. Differential expression of neural cell adhesion molecule (NCAM) during osteogenesis and secondary chondrogenesis in the embryonic chick. Int. J. Dev. Biol. 1995; 39:519–528.
  282. 282. Hall B.K., Herring S.W. Paralysis and growth of the musculoskeletal system in the embryonic chick. J. Morphol. 1990; 206:45–56.
  283. 283. Murray P.D.F., Drachman B.D. The role of movement in the development of joints and related structures: the head and neck in the chick embryo. J. Embryol. Exp. Morphol. 1969; 22:349–371.
  284. 284. Guilak F., Sah R.L., Setton L.A. Physical regulation of cartilage metabolism. In: Mow V.C., Hayes W.C. (eds.). Basic Orthopaedic Biomechanics. (Philadelphia: Lippincott-Raven), 1997; pp. 179–207.
  285. 285. Liedtke W., Kim C. Functionality of the TRPV subfamily of TRP ion channels: add mechano-TRP and osmo-TRP to the lexicon! Cell. Mol. Life Sci. 2005; 62:2985–3001.
  286. 286. Wang J.H., Thampatty B.P. Mechanobiology of adult and stem cells. Int. Rev. Cell Mol. Biol. 2008; 271:301–346.
  287. 287. Reinwald Y., Leonard K.H.L., Henstock J.R., Whiteley J.P., Osborne J.M., Waters S.L., Levesque P., El Haj A.J. Evaluation of the growth environment of a hydrostatic force bioreactor for preconditioning of tissue-engineered constructs. Tissue Eng. Part C. 2015; 21(1):1–14.
  288. 288. Henstock J.R., Rotherham M., Rose J.B., El Haj A.J. Cyclic hydrostatic pressure stimulates enhanced bone development in the foetal chick femur in vitro. Bone. 2013; 53:468.
  289. 289. Luo W., Jones S.R., Yousaf M.N. Geometric control of stem cell differentiation rate on surfaces. Langmuir. 2008; 24:12129–12133.
  290. 290. El Haj A.J., Cartmell S.H. Bioreactors for bone tissue engineering. Proc. Inst. Mech. Eng. 2010H; 224:1523.
  291. 291. Darling E.M., Athanasiou K.A. Biomechanical strategies for articular cartilage regeneration. Ann. Biomed. Eng. 2003; 31:1114–1124.
  292. 292. Darling E.M., Athanasiou K.A. Articular cartilage bioreactors and bioprocesses. Tissue Eng. 2003; 9:9–26.
  293. 293. Partap S., Plunkett N.A., O’Brien, F.J. Bioreactors in Tissue Engineering. In: Eberli D. (ed.).Tissue Engineering. 2010; ISBN:978-953-307-079-7.
  294. 294. Freed L.E., Guilak F., Guo X.E., Gray M.L., Tranquillo R., Holmes J.W., Radisic M., Sefton M.V., Kaplan D., Vunjak-Novakovic G. Advanced tools for tissue engineering: scaffolds, bioreactors, and signaling. Tissue Eng. 2006; 12:3285–3305.
  295. 295. Reinwald Y., El Haj A.J. Mechanical and topographical cues affecting the mesenchymal stem cell fate for bone tissue regeneration. 2016 (manuscript under revision).
  296. 296. Vunjak-Novakovic G., Obradovic B., Martin I., Bursac P.M., Langer R., Freed L.E. Dynamic cell seeding of polymer scaffolds for cartilage tissue engineering. Biotechnol. Prog. 1998; 14:193–202.
  297. 297. Bancroft G.N., Sikavitsas V.I., Mikos A.G. Design of a flow perfusion bioreactor system for bone tissue-engineering applications. Tissue Eng. 2003; 9:549–554.
  298. 298. Schwarz R.P., Goodwin T.J., Wolf D.A. Cell culture for three-dimensional modeling in rotating-wall vessels: an application of simulated microgravity. J. Tissue Cult. Methods. 1992; 14:51–57.
  299. 299. Li W.J., Jiang Y.J., Tuan R.S., Cell-nanofiber-based cartilage tissue engineering using improved cell seeding, growth factor, and bioreactor technologies. Tissue Eng. Part A. 2008; 14(5):69–48.
  300. 300. Lawrenson K., Benjamin E., Turmaine M., Jacobs I., Gayther S., Dafou D. In vitro three-dimensional modelling of human ovarian surface epithelial cells. Cell Prolif 2009; 42(3):385–93.
  301. 301. Consolo F., Bariani C., Mantalaris F., Redaelli A., Morbiducci U., Computional modelling for the optimization of a cardiogenic 3D bioprocess of encapsulated embryonic stem cells. Biomech. Model. Mechanobiol. 2012; 11(1–2):261–77.
  302. 302. Goldstein A.S., Juarez T.M., Helmke C.D., Gustin M.C., Mikos A.G. Effect of convection on osteoblastic cell growth and function in biodegradable polymer foam scaffolds. Biomaterials. 2001; 22:1279–1288.
  303. 303. Yu X., Botchwey E.A., Levine E.M., Pollack S.R., Laurencin C.T. Bioreactor-based bone tissue engineering: the influence of dynamic flow on osteoblast phenotypic expression and matrix mineralization. Proc. Natl. Acad. Sci. USA. 2004; 101:11203–11208.
  304. 304. Huang C.Y., Hagar K.L., Frost L.E., Sun Y., Cheung H.S. Effects of cyclic compressive loading on chondrogenesis of rabbit bone-marrow derived mesenchymal stem cells. Stem Cells. 2004; 22:313–323.
  305. 305. Thorpe S.D., Buckley C.T., Vinardell T., O’Brien F.J., Campbell V.A., Kelly D.J. Dynamic compression can inhibit chondrogenesis of mesenchymal stem cells. Biochem. Biophys. Res. Commun. 2008; 377:458–462.
  306. 306. Barberi T., Bradbury M., Dincer Z., Panagiotakos G., Socci N.D., Studer L. Derivation of engraftable skeletal myoblasts from human embryonic stem cells. Nat. Med. 2007; 13:642–648.
  307. 307. Li W., Li K., Wei W., Ding S. Chemical approaches to stem cell biology and therapeutics. Cell Stem Cell. 2013; 13:270–283.
  308. 308. Li W., Ding S. Small molecules that modulate embryonic stem cell fate and somatic cell reprogramming. Trends Pharmacol. Sci. 2010; 31:36–45.
  309. 309. Nguyen H.N., Byers B., Cord B., Shcheglovitov A., Byrne J., Gujar P., Kee K., Schüle B., Dolmetsch R.E., Langston W., Palmer T.D., Pera R.R. LRRK2 mutant iPSC-derived DA neurons demonstrate increased susceptibility to oxidative stress. Cell Stem Cell. 2011; 8:267–280.
  310. 310. Ebert A.D., Yu J., Rose F.F.,Jr., Mattis V.B., Lorson C.L., Thomson J.A., Svendsen C.N. Induced pluripotent stem cells from a spinal muscular atrophy patient. Nature. 2009; 457:277–280.
  311. 311. Marchetto M.C., Carromeu C., Acab A., Yu D., Yeo G.W., Mu Y., Chen G., Gage F.H., Muotri A.R. A model for neural development and treatment of Rett syndrome using human induced pluripotent stem cells. Cell. 2010; 143:527–539.
  312. 312. Lee G., Papapetrou E.P., Kim H., Chambers S.M., Tomishima M.J., Fasano C.A., Ganat Y.M., Menon J., Shimizu F., Viale A., Tabar V., Sadelain M., Studer L. Modelling pathogenesis and treatment of familial dysautonomia using patient-specific iPSCs. Nature. 2009; 461:402–406.
  313. 313. Moretti A., Bellin M., Welling A., Jung C.B., Lam J.T., Bott-Flügel L., Dorn T., Goedel A., Höhnke C., Hofmann F., Seyfarth M., Sinnecker D., Schömig A., Laugwitz K.L. Patient-specific induced pluripotent stem-cell models for long-QT syndrome. N. Engl. J. Med. 2010; 363:1397–1409.
  314. 314. Itzhaki I., Maizels L., Huber I., Zwi-Dantsis L., Caspi O., Winterstern A., Feldman O., Gepstein A., Arbel G., Hammerman H., Boulos M., Gepstein L. Modelling the long QT syndrome with induced pluripotent stem cells. Nature. 2011; 471:225–229.
  315. 315. Rashid S.T., Corbineau S., Hannan N., Marciniak S.J., Miranda E., Alexander G., Huang-Doran I., Griffin J., Ahrlund-Richter L., Skepper J., Semple R., Weber A., Lomas D.A., Vallier L. Modeling inherited metabolic disorders of the liver using human induced pluripotent stem cells. J. Clin. Investig. 2010; 120:3127–3136.
  316. 316. Yazawa M., Hsueh B., Jia X., Pasca A.M., Bernstein J.A., Hallmayer J., Dolmetsch R.E. Using induced pluripotent stem cells to investigate cardiac phenotypes in Timothy syndrome. Nature. 2011; 471:230–234.
  317. 317. Ko E., Yang K., Shin J., Cho S.W. Polydopamine-assisted osteoinductive peptide immobilization of polymer scaffolds for enhanced bone regeneration by human adipose-derived stem cells. Biomacromolecules. 2013; 14:3202–3213.
  318. 318. Beazley K.E., Nurminskaya M. BMP2 cross-linked by transglutaminase 2 to collagen-plla scaffold promotes osteogenic differentiation in mesenchymal stem cells. Biotechnol. Lett. 2014; 36:1901–1907.
  319. 319. Chiang C.K., Chowdhury M.F., Iyer R.K., Stanford W.L., Radisic M. Engineering surfaces for site-specific vascular differentiation of mouse embryonic stem cells. Acta Biomater. 2010; 6:1904–1916.
  320. 320. Poh C.K., Shi Z., Lim T.Y., Neoh K.G., Wang W. The effect of VEGF functionalization of titanium on endothelial cells in vitro. Biomaterials. 2010; 31:1578–85.
  321. 321. Rahman N., Purpura K.A., Wylie R.G., Zandstra P.W., Shoichet M.S. The use of vascular endothelial growth factor functionalized agarose to guide pluripotent stem cell aggregates toward blood progenitor cells. Biomaterials. 2010; 31:8262–8270.
  322. 322. Lee J.Y., Choo J.E., Park H.J., Park J.B., Lee S.C., Jo I., Lee S.J., Chung C.P., Park Y.J. Injectable gel with synthetic collagen-binding peptide for enhanced osteogenesis in vitro and in vivo. Biochem. Biophys. Res. Commun. 2007; 357:68–74.
  323. 323. Lee J.S., Lee J.S., Murphy W.L. Modular peptides promote human mesenchymal stem cell differentiation on biomaterial surfaces. Acta Biomater. 2010; 6:21–28.
  324. 324. Shin Y.M., Jo S.Y., Park J.S., Gwon H.J., Jeong S.I., Lim Y.M. Synergistic effect of dual-functionalized fibrous scaffold with BCP and RGD containing peptide for improved osteogenic differentiation. Macromol. Biosci. 2014; 14:1190–1198.
  325. 325. Lin Z.Y., Duan Z.X., Guo X.D., Li J.F., Lu H.W., Zheng Q.X., Quan D.P., Yang S.H. Bone induction by biomimetic PLGA-(PEGASP) n copolymer loaded with a novel synthetic BMP-2-related peptide in vitro and in vivo. J. Control Release. 2010; 144:190–195.
  326. 326. Murphy A.R., St John P., Kaplan D.L. Modification of silk fibroin using diazonium coupling chemistry and the effects on hMSC proliferation and differentiation. Biomaterials. 2008; 29:2829–2838.
  327. 327. Nuttelman C.R., Tripodi M.C., Anseth K.S. Synthetic hydrogel niches that promote hMSC viability. Matrix Biol. 2005; 24:208–218.
  328. 328. Nuttelman C.R., Benoit D.S., Tripodi M.C., Anseth K.S. The effect of ethylene glycol methacrylate phosphate in PEG hydrogels on mineralization and viability of encapsulated hMSCs. Biomaterials. 2006; 27:1377–1386.
  329. 329. Liu X., Feng Q., Bachhuka A., Vasilev K. Surface modification by allylamine plasma polymerization promotes osteogenic differentiation of human adipose-derived stem cells. ACS Appl. Mater. Interfaces. 2014; 6:9733–9741.
  330. 330. Ren Y.J., Zhang H., Huang H., Wang X.M., Zhou Z.Y., Cui F.Z., An Y.H. In vitro behavior of neural stem cells in response to different chemical functional groups. Biomaterials. 2009; 30:1036–1044.
  331. 331. Curran J.M., Chen R., Hunt J.A. The guidance of human mesenchymal stem cell differentiation in vitro by controlled modifications to the cell substrate. Biomaterials. 2006; 27:4783–4793.
  332. 332. Park J., Bauer S., von der Mark K., Schmuki P. Nanosize and vitality: TiO2 nanotube diameter directs cell fate. Nano Lett. 2007; 7:1686–1691.
  333. 333. Yim E.K., Pang S.W., Leong K.W. Synthetic nanostructures inducing differentiation of human mesenchymal stem cells into neuronal lineage. Exp. Cell Res. 2007; 313:1820–1829.
  334. 334. Biehl J.K., Yamanaka S., Desai T.A., Boheler K.R., Russell B. Proliferation of mouse embryonic stem cell progeny and the spontaneous contractile activity of cardiomyocytes are affected by microtopography. Dev. Dyn. 2009; 238:1964–1973.
  335. 335. D’Angelo F., Armentano I., Mattioli S., Crispoltoni L., Tiribuzi R., Cerulli G.G., Palmerini C.A., Kenny J.M., Martino S., Orlacchio A. Micropatterned hydrogenated amorphous carbon guides mesenchymal stem cells towards neuronal differentiation. Eur. Cell Mater. 2010; 20:231–244.
  336. 336. Gao L., McBeath R., Chen C.S. Stem cell shape regulates a chondrogenic versus myogenic fate through Rac1 and N-cadherin. Stem Cells. 2010; 28:564–572.
  337. 337. Lu D., Chen C.S., Lai C.S., Soni S., Lam T., Le C., Chen E.Y., Nguyen T., Chin W.C. Microgrooved surface modulates neuron differentiation in human embryonic stem cells. Methods Mol. Biol. 2016; 1307:281–7.
  338. 338. Smith L.A., Liu X., Hu J., Wang P., Ma P.X. Enhancing osteogenic differentiation of mouse embryonic stem cells by nanofibers. Tissue Eng. Part A. 2009; 15:1855–1864.
  339. 339. Saha K., Keung A.J., Irwin E.F., Li Y., Little L., Schaffer D.V., Healy K.E. Substrate modulus directs neural stem cell behavior. Biophys. J. 2008; 95:4426–4438.
  340. 340. Yang Y., Beqaj S., Kemp P., Ariel I., Schuger L. Stretch-induced alternative splicing of serum response factor promotes bronchial myogenesis and is defective in lung hypoplasia. J. Clin. Investig. 2000; 106:1321–1330.
  341. 341. Hamilton D.W., Maul T.M., Vorp D.A. Characterization of the response of bone marrow-derived progenitor cells to cyclic strain: implications for vascular tissue-engineering applications. Tissue Eng. 2004; 10:361–369.
  342. 342. Gong Z., Niklason L.E. Small-diameter human vessel wall engineered from bone marrow-derived mesenchymal stem cells (hMSCs). FASEB J. 2008; 22:1635–1648.
  343. 343. Lee W.C., Maul T.M., Vorp D.A., Rubin J.P., Marra K.G. Effects of uniaxial cyclic strain on adipose-derived stem cell morphology, proliferation, and differentiation. Biomech. Model. Mechanobiol. 2007; 6:265–273.
  344. 344. Sen B., Xie Z., Case N., Ma M., Rubin C., Rubin J. Mechanical strain inhibits adipogenesis in mesenchymal stem cells by stimulating a durable beta-catenin signal. Endocrinology. 2008; 149:6065–6075.
  345. 345. Simmons C.A., Matlis S., Thornton A.J., Chen S., Wang C.Y., Mooney D.J. Cyclic strain enhances matrix mineralization by adult human mesenchymal stem cells via the extracellular signal-regulated kinase (ERK1/2) signaling pathway. J. Biomech. 2003; 36:1087–1096.
  346. 346. Thomas G.P., El Haj A.J. Bone marrow stromal cells are load responsive in vitro. Calcif. Tissue Int. 1996; 58:101–108.
  347. 347. Yoshikawa T., Peel S.A., Gladstone J.R., Davies J.E. Biochemical analysis of the response in rat bone marrow cell cultures to mechanical stimulation. Biomed. Mater. Eng. 1997; 7:369–377.
  348. 348. Knippenberg M., Helder M.N., Doulabi B.Z., Semeins C.M., Wuisman P.I., Klein-Nulend J. Adipose tissue-derived mesenchymal stem cells acquire bone cell-like responsiveness to fluid shear stress on osteogenic stimulation. Tissue Eng. 2005; 11:1780–1788.
  349. 349. Shimizu N., Yamamoto K., Obi S., Kumagaya S., Masumura T., Shimano Y., Naruse K., Yamashita J.K., Igarashi T., Ando J. Cyclic strain induces mouse embryonic stem cell differentiation into vascular smooth muscle cells by activating PDGF receptor beta. J. Appl. Physiol. 2008; 104:766–772.
  350. 350. Saha S., Ji L., de Pablo J.J., Palecek S.P. Inhibition of human embryonic stem cell differentiation by mechanical strain. J. Cell. Physiol. 2006; 206:126–137.
  351. 351. Elder S.H., Kimura J.H., Soslowsky L.J., Lavagnino M., Goldstein S.A. Effect of compressive loading on chondrocyte differentiation in agarose cultures of chick limb-bud cells. J. Orthop. Res. 2000; 18:78–86.
  352. 352. Angele P., Yoo J.U., Smith C., Mansour J., Jepsen K.J., Nerlich M., Johnstone B. Cyclic hydrostatic pressure enhances the chondrogenic phenotype of human mesenchymal progenitor cells differentiated in vitro. J. Orthop. Res. 2003; 21:451–457.

Written By

Yvonne Reinwald, Jessica Bratt and Alicia El Haj

Submitted: 06 November 2015 Reviewed: 23 February 2016 Published: 20 July 2016