Induced stem cells

Welcome to the Wikiversity learning project for Induced stem cells (Guide to publications). This project provides learning resources that help participants learn about Induced stem cells and efforts to produce useful stem cells and obtaining their derivatives for medical therapies. Participants should feel free to ask questions on discuss page and explore related topics.

Induced stem cells (iSC) are stem cells artificially derived from somatic, reproductive, pluripotent or other cell types by deliberate w:epigenetic reprogramming. They are classified as either totipotent (iTC), pluripotent (iPSC) or progenitor (multipotent—iMSC, also called an induced multipotent progenitor cell—iMPC) or unipotent -- (iUSC) according to their developmental potential and degree of dedifferentiation. Progenitors are obtained by so-called direct reprogramming or directed differentiation and are also called induced somatic stem cells.

Three techniques are widely recognized:[1]

Natural processes of induction edit

Back in 1895, Thomas Morgan remove one of the two frog w:blastomeres and found that w:amphibians are able to form whole w:embryo from the remaining part. This meant that the cells can change their differentiation pathway. Later, in 1924, Spemann and Mangold demonstrated the key importance of cell–cell inductions during animal development.[20] The reversible transformation of cells of one differentiated cell type to another is called w:metaplasia.[21] This transition can be a part of the normal maturation process, or caused by an inducing stimulus. For example: transformation of iris cells to lens cells in the process of maturation and transformation of w:retinal pigment epithelium cells into the neural retina during regeneration in adult w:newt eyes. This process allows the body to replace cells not suitable to new conditions with more suitable new cells. In w:Drosophila imaginal discs, cells have to choose from a limited number of standard discrete differentiation states. The fact that transdetermination (change of the path of differentiation) often occurs for a group of cells rather than single cells shows that it is induced rather than part of maturation.[22]

The researchers were able to identify the minimal conditions and factors that would be sufficient for starting the cascade of molecular and cellular processes to instruct pluripotent cells to organize the w:embryo. They show that opposing gradients of w:bone morphogenetic protein (BMP) and Nodal, two w:transforming growth factor family members that act as w:morphogens, are sufficient to induce molecular and cellular mechanisms required to organize, w:in vivo or w:in vitro, uncommitted cells of the w:zebrafish w:blastula animal pole into a well-developed w:embryo.[23]

Some types of mature, specialized adult cells can naturally revert to stem cells. For example, differentiated airway epithelial cells can revert into stable and functional stem cells in vivo after the ablation of airway[24]. Another example, "chief" cells express the stem cell marker Troy. While they normally produce digestive fluids for the stomach, they can revert into stem cells to make temporary repairs to stomach injuries, such as a cut or damage from infection. Moreover, they can make this transition even in the absence of noticeable injuries and are capable of replenishing entire gastric units, in essence serving as quiescent “reserve” stem cells.[25]

After injury, mature terminally differentiated kidney cells dedifferentiate into more primordial versions of themselves, and then differentiate into the cell types needing replacement in the damaged tissue[26] Macrophages can self-renew by local proliferation of mature differentiated cells.[27] In newts, muscle tissue is regenerated from specialized muscle cells that dedifferentiate and forget the type of cell they had been. This capacity to regenerate does not decline with age and may be linked to their ability to make new stem cells from muscle cells on demand.[28]

A variety of nontumorigenic stem cells display the ability to generate multiple cell types. For instance, multilineage-differentiating stress-enduring (Muse) cells are stress-tolerant adult human stem cells that can self-renew. They form characteristic cell clusters in suspension culture that express a set of genes associated with pluripotency and can differentiate into w:endodermal, ectodermal and mesodermal cells both in vitro and in vivo.[29][30][31][32][33][34]

Other well-documented examples of w:transdifferentiation and their significance in development and regeneration were described in detail.[35]

 
Induced totipotent cells usually can be obtained by reprogramming somatic cells by w:somatic-cell nuclear transfer (SCNT).

Induced totipotent cells edit

SCNT-mediated edit

Induced totipotent cells usually can be obtained by reprogramming somatic cells by w:somatic-cell nuclear transfer (SCNT) to the recipient eggs or oocytes.[3][5][36][37]

Using an approach based on the protocol outlined by Tachibana et al.,[3] hESCs can be generated by SCNT using dermal fibroblasts nuclei from both a middle-aged 35-year-old male and an elderly, 75-year-old male, suggesting that age-associated changes are not necessarily an impediment to SCNT-based nuclear reprogramming of human cells.[38] Such reprogramming of somatic cells to a pluripotent state holds huge potentials for w:regenerative medicine. Unfortunately, the cells generated by this technology, potentially are not completely protected from the immune system of the patient (donor of nuclei), because they have the same w:mitochondrial DNA, as a donor of oocytes, instead of the patients mitochondrial DNA. This reduces their value as a source for w:autologous stem cell transplantation therapy, as for the present, it is not clear whether it can induce an immune response of the patient upon treatment.

Induced androgenetic haploid embryonic stem cells can be used instead of sperm for cloning. These cells, synchronized in M phase and injected into the oocyte can produce viable offspring.[39]

These developments, together with data on the possibility of unlimited oocytes from mitotically active reproductive stem cells,[40] offer the possibility of industrial production of transgenic farm animals.

Repeated recloning of viable mice through a SCNT method that includes a w:histone deacetylase inhibitor, trichostatin, added to the cell culture medium,[41] show that it may be possible to reclone animals indefinitely with no visible accumulation of reprogramming or genomic errors[42]

Concerns still exist regarding telomere length resetting in cloned embryos and nuclear transfer ES cells, and possibilities of premature aging of cloned animals achieved by SCNT. It was shown that telomeres of cloned pigs generated by standard SCNT methods are not effectively restored, compared with those of donor cells, however trichostatin A significantly increases telomere lengths in cloned pigs and this could be one of the mechanisms underlying improved development of cloned embryos and animals treated with trichostatin.[43]

However, research into technologies to develop sperm and egg cells from stem cells raises bioethical issues.[44]

Such technologies may also have far-reaching clinical applications for overcoming cytoplasmic defects in human oocytes.[3][45] For example, the technology could prevent inherited w:mitochondrial disease from passing to future generations. Mitochondrial genetic material is passed from mother to child. Mutations can cause diabetes, deafness, eye disorders, gastrointestinal disorders, heart disease, dementia and other neurological diseases. The nucleus from one human egg has been transferred to another, including its mitochondria, creating a cell that could be regarded as having two mothers. The eggs were then fertilised, and the resulting embryonic stem cells carried the swapped mitochondrial DNA.[46] As evidence that the technique is safe author of this method points to the existence of the healthy monkeys that are now more than four years old — and are the product of mitochondrial transplants across different genetic backgrounds.[47]

In late-generation w:telomerase-deficient (Terc−/−) mice, SCNT-mediated reprogramming mitigates telomere dysfunction and mitochondrial defects to a greater extent than iPSC-based reprogramming.[48]

Other cloning and totipotent transformation achievements have been described.[49]

Obtained without SCNT edit

Recently some researchers succeeded to get the totipotent cells without the aid of SCNT. Totipotent cells were obtained using the epigenetic factors such as oocyte germinal isoform of histone.[50] Reprogramming in vivo, by transitory induction of the four factors Oct4, Sox2, Klf4 and c-Myc in mice, confers totipotency features. Intraperitoneal injection of such in vivo iPS cells generates embryo-like structures that express embryonic and extraembryonic (w:trophectodermal) markers.[51]

Rejuvenation to iPSc edit

iPSc were first obtained in the form of transplantable w:teratocarcinoma induced by grafts taken from mouse embryos.[52] Teratocarcinoma formed from somatic cells.[53] Genetically mosaic mice were obtained from malignant teratocarcinoma cells, confirming the cells' pluripotency.[54][55][56] It turned out that teratocarcinoma cells are able to maintain a culture of pluripotent w:embryonic stem cell in an undifferentiated state, by supplying the culture medium with various factors.[57] In the 1980s, it became clear that transplanting pluripotent/embryonic stem cells into the body of adult mammals, usually leads to the formation of w:teratomas, which can then turn into a malignant tumor teratocarcinoma.[58] However, putting teratocarcinoma cells into the embryo at the blastocyst stage, caused them to become incorporated in the w:inner cell mass and often produced a normal chimeric (i.e. composed of cells from different organisms) animal.[59][60][61] This indicated that the cause of the teratoma is a dissonance - mutual miscommunication between young donor cells and surrounding adult cells (the recipient's so-called "niche").

In August 2006, Japanese researchers circumvented the need for an oocyte, as in SCNT. By reprograming mouse embryonic w:fibroblasts into pluripotent stem cells via the ectopic expression of four transcription factors, namely w:Oct4, w:Sox2, w:Klf4 and w:c-Myc, they proved that the overexpression of a small number of factors can push the cell to transition to a new stable state that is associated with changes in the activity of thousands of genes.[7]

 
Human somatic cells are made pluripotent by transducing them with factors that induces pluripotency (OCT 3/4, SOX2, Klf4, c-Myc, NANOG and LIN28). This results in the production of IPS cells, which can differentiate into any cells of the three embryonic germ layers (Mesoderm, Endoderm, Ectoderm).

Later has been found a reprogramming factor BRD3R that increases the efficiency of creating human induced pluripotent stem cells (HiPSCs) from skin fibroblasts in xeno-free media more than 20-fold, speeds the reprogramming time by several days and enhances the quality of reprogramming.[62] Reprogramming mechanisms are thus linked, rather than independent and are centered on a small number of genes.[63] IPSC properties are very similar to ESCs[64]. iPSCs have been shown to support the development of all-iPSC mice using a w:tetraploid (4n) embryo,[65] the most stringent assay for developmental potential. However, some genetically normal iPSCs failed to produce all-iPSC mice because of aberrant epigenetic silencing of the imprinted Dlk1-Dio3 gene cluster.[18]

An important advantage of iPSC over ESC is that they can be derived from adult cells, rather than from embryos. Therefore, it became possible to obtain iPSC from adult and even elderly patients.[9][66][67]

Reprogramming somatic cells to iPSC leads to rejuvenation. It was found that reprogramming leads to telomere lengthening and subsequent shortening after their differentiation back into fibroblast-like derivatives.[68] Thus, reprogramming leads to the restoration of embryonic telomere length,[69] and hence increases the potential number of cell divisions otherwise limited by the w:Hayflick limit.[70]

However, because of the dissonance between rejuvenated cells and the surrounding niche of the recipient's older cells, the injection of his own iPSC usually leads to an w:immune response,[71] which can be used for medical purposes,[72] or the formation of tumors such as teratoma.[73] The reason has been hypothesized to be that some cells differentiated from ESC and iPSC in vivo continue to synthesize embryonic w:protein isoforms.[74] So, the immune system might detect and attack cells that are not cooperating properly.

A small molecule called MitoBloCK-6 can force the pluripotent stem cells to die by triggering apoptosis (via w:cytochrome c release across the w:mitochondrial outer membrane) in human pluripotent stem cells, but not in differentiated cells. Shortly after differentiation, daughter cells became resistant to death. When MitoBloCK-6 was introduced to differentiated cell lines, the cells remained healthy. The key to their survival, was hypothesized to be due to the changes undergone by pluripotent stem cell mitochondria in the process of cell differentiation. This ability of MitoBloCK-6 to separate the pluripotent and differentiated cell lines has the potential to reduce the risk of teratomas and other problems in regenerative medicine.[75]

In 2012 other w:small molecules (selective cytotoxic inhibitors of human pluripotent stem cells—hPSCs) were identified that prevented human pluripotent stem cells from forming teratomas in mice. The most potent and selective compound of them (PluriSIn #1) inhibits stearoyl-coA desaturase (the key enzyme in w:oleic acid biosynthesis), which finally results in apoptosis. With the help of this molecule the undifferentiated cells can be selectively removed from culture.[76] An efficient strategy to selectively eliminate pluripotent cells with teratoma potential is targeting pluripotent stem cell-specific antiapoptotic factor(s) (i.e., w:survivin or Bcl10). A single treatment with chemical survivin inhibitors (e.g., w:quercetin or YM155) can induce selective and complete cell death of undifferentiated hPSCs and is claimed to be sufficient to prevent teratoma formation after transplantation.[77] However, it is unlikely that any kind of preliminary clearance,[78] is able to secure the replanting iPSC or ESC. After the selective removal of pluripotent cells, they re-emerge quickly by reverting differentiated cells into stem cells, which leads to tumors.[79] This may be due to the disorder of let-7 regulation of its target Nr6a1 (also known as w:Germ cell nuclear factor - GCNF), an embryonic transcriptional repressor of pluripotency genes that regulates gene expression in adult fibroblasts following w:micro-RNA miRNA loss.[80]

Yijie Geng et al., identified a small molecule, Displurigen, that potently disrupts pluripotency by targeting heat shock 70-kDa protein 8 (HSPA8), which maintains pluripotency by facilitating the DNA-binding activity of OCT4[81]

Teratoma formation by pluripotent stem cells may be caused by low activity of PTEN enzyme, reported to promote the survival of a small population (0,1-5% of total population) of highly tumorigenic, aggressive, teratoma-initiating embryonic-like carcinoma cells during differentiation. The survival of these teratoma-initiating cells is associated with failed repression of w:Nanog as well as a propensity for increased glucose and cholesterol metabolism.[82] These teratoma-initiating cells also expressed a lower ratio of p53/p21 when compared to non-tumorigenic cells.[83] In connection with the above safety problems, the use iPSC for cell therapy is still limited.[84] However, they can be used for a variety of other purposes - including the modeling of disease,[85] screening (selective selection) of drugs, toxicity testing of various drugs.[86]

 
Small molecule modulators of stem-cell fate.

It is interesting to note that the tissue grown from iPSCs, placed in the "chimeric" embryos in the early stages of mouse development, practically do not cause an immune response (after the embryos have grown into adult mice) and are suitable for autologous transplantation[87] At the same time, full reprogramming of adult cells in vivo within tissues by transitory induction of the four factors Oct4, Sox2, Klf4 and c-Myc in mice results in teratomas emerging from multiple organs.[51] Furthermore, partial reprogramming of cells toward pluripotency in vivo in mice demonstrates that incomplete reprogramming entails epigenetic changes (failed repression of w:Polycomb targets and altered w:DNA methylation) in cells that drive cancer development.[88]

 
Cell culture example of a small molecule as a tool instead of a protein. in w:cell culture to obtain a pancreatic lineage from w:mesodermal stem cells the w:retinoic acid signalling pathway must be activated while the w:sonic hedgehog pathway inhibited, which can be done by adding to the media anti-shh w:antibodies, Hedgehog interacting protein or w:cyclopamine, the first two are protein and the last a small molecule.[89]

Safeguard System edit

Several methods have been reported that may increase the safety and eventually the efficacy of iPSC-based regenerative medicine. The first safety approach eliminates potential oncogenic factors, such as the expression of oncogene c-myc, or integrates the reprogramming transgenes into chromosomes. The latter would be eliminated by using so-called nonintegrating viral vectors. The second safety approach is based on the isolation of desired differentiated cells from other cell types and undifferentiated human pluripotent stem cells (hPSCs), such as the removal of the residual pluripotent cells using fluoresecent activated cell sorting or magnetic beads coated with antibodies against a particular antigen, including SSEA-5 and Claudin-6, and fucose-specific lectin UEA (Ulex europaeus agglutinin)-I. The third safety approach entails the direct targeting and killing of oncogenic cells by using cytotoxic antibody recognizing podocalyxin-like protein-1, a chemical inhibitor of stearoyl-coA desaturase, specific monoclonal antibodies, DNA topoisomerase II inhibitor, and suicide gene therapy under transcriptional control of a pluripotency-related promoter[90][91]. However, these strategies may not suffice to lower risk to acceptable levels, because the tumorigenic risk of iPSC-based cell therapy arises not just from contamination with undifferentiated iPSCs but also from other unexpected events associated with long-term culture for reprogramming and redifferentiation. There is always a chance of unexpected issues associated with first-in-human clinical studies. An efficient and reliable approach to provide safety for future regenerative therapy and first-in-human cell therapy can be a suicide gene engineered from human caspase-9, that is not immunogenic, and can kill transduced cells in a cell-cycle-independent manner[92][93][94][95].

Miki Ando et al.[96] demonstrated the efficacy of suicide gene therapy by introducing inducible caspase-9 (iC9) into iPSCs. Activation of iC9 system in vivo with a specific chemical inducer of dimerization (CID) initiates a caspase cascade that eliminates iPSCs and tumors originated from iPSCs. They introduced this iC9/CID safeguard system into a previously reported iPSC-derived, rejuvenated cytotoxic T lymphocyte (rejCTL) therapy model and confirmed that rejCTLs from iPSCs are expressing high levels of iC9 without disturbing antigen-specific killing activity. iC9-expressing rejCTLs exert antitumor effects in vivo. Upon induction, the iC9 system efficiently leads to apoptosis in rejuvenated CTLs. This safeguard system can eliminate contaminating iPSCs, debulk tumors originated from iPSCs, stop cytokine release syndrome associated with iPSC-derived CTL therapy, and control “on-target, off-tumor toxicities”. It should be applicable to other cell therapies using iPSC-derived cells.

Strategies for good manufacturing practice edit

The potential to develop patient-derived cells into any cell type makes human pluripotent stem cells one of the most promising sources for regenerative treatments. The proper differentiation of autologous iPSCs sometimes results in a loss of immunogenicity and leads to the induction of tolerance.[97] This differentiation of iPSCs to mature cell types—and ultimately to functional tissues and organs—holds great promise for personalized disease modeling, drug screening, and the development of cell-based therapies. [98] However there are some problems that need to be solved previously:

  • Need to develop a standardized methodology for the induction of reprogramming, which would be effective and would not leave traces after delivery and expression of the reprogramming factors which would lead to the genomic instability;[99][100]
  • Hurdles to their clinical applications also include (1) formation of heterogeneously differentiated cultures, (2) the risk of teratoma formation from residual undifferentiated cells, and (3) immune rejection of engrafted cells.[101]

The main steps for the production of human pluripotent stem cell-derived progenitor cells under safe and good manufacturing practice (GMP) conditions must include:

  1. the expansion of a clone of pluripotent stem cells to generate a master cell bank under good manufacturing practice conditions (GMP);
  2. induced specification caused by a growth factor or in some other way;
  3. the purification of committed cells by immunomagnetic sorting to yield a stage-specific embryonic antigen such as for example (SSEA)-1-positive cell population strongly expressing the early cardiac transcription factor Isl-1;
  4. the incorporation of these cells into a fibrin scaffold;
  5. a safety assessment focused on the loss of teratoma-forming cells by in vitro (transcriptomics) and in vivo (cell injections in immunodeficient mice) measurements;
  6. an extensive cytogenetic and viral testing;
  7. the characterization of the final cell product and its release criteria.

The data collected throughout such process already have led to approval for a first-in-man clinical trial of transplantation of SSEA-1+ progenitors in patients with severely impaired cardiac function. [102].

Lonza attempted to develop clinically compliant processes to generate cGMP-compliant human iPSC lines and have described a step-by-step cGMP-compliant process to generate clinically compliant cell lines[103].

Several works have reported evidence of genomic instability in iPSC, raising concerns on their biomedical use. The reasons behind the genomic instability observed in iPSC remain mostly unknown. Sergio Ruiz et al.[104] show that, similar to the phenomenon of oncogene-induced replication stress, the expression of reprogramming factors induces replication stress. Increasing the levels of the checkpoint kinase 1 (CHK1) reduces reprogramming-induced replication stress and increases the efficiency of iPSC generation. Similarly, nucleoside supplementation during reprogramming reduces the load of DNA damage and genomic rearrangements on iPSC. So, lowering replication stress during reprogramming, genetically or chemically, provides a simple strategy to reduce genomic instability on mouse and human iPSC.

Chemically induced pluripotent cells (CiPSCs) edit

By using solely w:small molecules, Deng Hongkui and colleagues demonstrated that endogenous “master genes” are enough for cell fate reprogramming. They induced a pluripotent state in adult cells from mice using seven small-molecule compounds.[17] The effectiveness of the method is quite high: it was able to convert 0.2% of the adult tissue cells into iPSCs, which is comparable to the gene insertion conversion rate. The authors note that the mice generated from CiPSCs were "100% viable and apparently healthy for up to 6 months”.So. This chemical reprogramming strategy has potential use in generating functional desirable cell types for clinical applications.[105]

iPS-like cells (iPSLCs) were also generated from mouse somatic cells in two steps with small molecule compounds. In the first step, stable intermediate cells were generated from mouse astrocytes by Shh activators (oxysterol and purmorphamine) to replace Bmi1 function. These cells called induced epiblast stem cell (EpiSC)-like cells (iEpiSCLCs) are similar to EpiSCs in terms of expression of specific markers, epigenetic state, and ability to differentiate into three germ layers. In the second step, treatment with MEK/ERK and GSK3 pathway inhibitors in the presence of leukemia inhibitory factor resulted in conversion of iEpiSCLCs into iPSLCs that were similar to mESCs, suggesting that Bmi1 is sufficient to reprogram astrocytes to partially reprogrammed pluripotency. So, combinations of small molecules can compensate for reprogramming factors and are sufficient to directly reprogram mouse somatic cells into iPSLCs. The chemically induced pluripotent stem cell-like cells (ciPSLCs) showed similar gene expression profiles, epigenetic status, and differentiation potentials to mESCs.[106]

Differentiation from induced teratoma edit

The fact that human iPSCs capable of forming teratomas not only in humans but also in some animal body, in particular in mice or pigs, allowed to develop a method for differentiation of iPSCs in vivo. For this purpose, iPSCs with an agent for inducing differentiation into target cells are injected to genetically modified pig (such as biallelic RAG2 mutants[107]) or mouse that has suppressed immune system activation on human cells. The formed after a while teratoma is cut out and used for the isolation of the necessary differentiated human cells[108] by means of w:monoclonal antibody to tissue-specific markers on the surface of these cells. This method has been successfully used for the production of functional myeloid, erythroid, and lymphoid human cells suitable for transplantation (yet only to mice).[109] Mice engrafted with human iPSC teratoma-derived hematopoietic cells produced human B and T cells capable of functional immune responses. These results offer hope that in vivo generation of patient customized cells is feasible, providing materials that could be useful for transplantation, human antibody generation, and drug screening applications. Using MitoBloCK-6 [75] and / or PluriSIn # 1 the differentiated progenitor cells can be further purified from teratoma forming pluripotent cells. The fact, that the differentiation takes place even in the teratoma niche, offers hope that the resulting cells are sufficiently stable to stimuli able to cause their transition back to the dedifferentiated (pluripotent) state, and therefore safe. A similar in vivo differentiation system, yielding engraftable hematopoietic stem cells from mouse and human iPSCs in teratoma-bearing animals in combination with a maneuver to facilitate hematopoiesis, was described by Suzuki et al.[110] They noted that neither leukemia nor tumors were observed in recipients after intravenous injection of iPSC-derived hematopoietic stem cells into irradiated recipients. Moreover, this injection resulted in multilineage and long-term reconstitution of the hematolymphopoietic system in serial transfers. Such system provides a useful tool for practical application of iPSCs in the treatment of hematologic and immunologic diseases.[111]

For further development of this method animal in which is grown the human cell graft, for example mouse, must have so modified genome that all its cells express and have on its surface human SIRPα.[112] To prevent rejection after transplantation to the patient of the allogenic organ or tissue, grown from the pluripotent stem cells in vivo in the animal, these cells should express two molecules: CTLA4-Ig, which disrupts T cell costimulatory pathways, and w:PD-L1, which activates T cell inhibitory pathway.[113]

Methods based on the detection of reporter gene-GFP-positive cells in the teratoma derived from iPSCs, will help to identify different types of induced adult stem cells which were previously difficult to pick out and to grow from selected cells tissue cultures.[114]

See also: US 20130058900  patent.

Differentiated cell types edit

Retinal cells edit

In the near-future, clinical trials designed to demonstrate the safety of the use of iPSCs for cell therapy of the people with age-related macular degeneration, a disease causing blindness through retina damaging, will begin. There are several articles describing methods for producing retinal cells from iPSCs[115][116] and how to use them for cell therapy.[117] Reports of iPSC-derived retinal pigmented epithelium transplantation showed enhanced visual-guided behaviors of experimental animals for 6 weeks after transplantation.[118] However, clinical trials have been successful: ten patients suffering from retinitis pigmentosa have had their eyesight restored—including a woman who had only 17 percent of her vision left. [119]

Lung and airway epithelial cells edit

Chronic lung diseases such as idiopathic pulmonary fibrosis and cystic fibrosis or w:chronic obstructive pulmonary disease and w:asthma are leading causes of morbidity and mortality worldwide with a considerable human, societal, and financial burden. So there is an urgent need for effective cell therapy and w:lung w:tissue engineering.[120][121] Several protocols have been developed for generation of the most cell types of the respiratory system, which may be useful for deriving patient-specific therapeutic cells.[122][123]

Reproductive cells edit

Some lines of iPSCs have the potentiality to differentiate into male germ cells and oocyte-like cells in an appropriate niche (by culturing in retinoic acid and porcine follicular fluid differentiation medium or seminiferous tubule transplantation). Moreover, iPSC transplantation make a contribution to repairing the testis of infertile mice, demonstrating the potentiality of gamete derivation from iPSCs in vivo and in vitro.[124]

The region-specific induced stem cells edit

Wu and his colleagues found that a combination of serum-free media plus fibroblast growth factor 2 (FGF2) and Wnt signaling inhibitors resulted in stable line of human rsPSCs (region-specific Induced stem cells)[125][126].

The transcriptomes of these cells resembled those of the posterior cells of the early mouse embryo, and grafting these cells into 7.5-day-old mouse embryos resulted in efficient incorporation in the posterior, but not the other parts of the embryo. After 36 hours of culturing these chimaeric embryos, the rsPSCs proliferated and could differentiate into the developing three germ layers, providing the first demonstration that human pluripotent cells can begin a differentiation program inside mice.

The region-specific cells could provide tremendous advantages -- the cells at this stage of an early embryo undergo dynamic changes to give rise to all cells, tissues and organs of the body. Each germ layer was theoretically capable of giving rise to specific tissues and organs. Whether human rsPSCs can generate more complicated tissue structures within mice or other animals requires further study[127].

These cells also have a lot of favorable characteristics for laboratory manipulation, including high cloning efficiency, stable passage in culture, and ease of genetic engineering.

The ease of culturing and editing the genome of human rsPSCs offers advantages for regenerative medicine applications.


Induced progenitor stem cells edit

Direct transdifferentiation edit

The risk of cancer and tumors creates the need to develop methods for safer cell lines suitable for clinical use. An alternative approach is so-called "direct reprogramming" - transdifferentiation of cells without passing through the pluripotent state.[128][129][130][131][132][133] The basis for this approach was that 5-azacytidine - a DNA demethylation reagent - can cause the formation of w:myogenic, chondrogenic and adipogeni] clones in the immortal cell line of mouse embryonic fibroblasts[134] and that the activation of a single gene, later named MyoD1, is sufficient for such reprogramming.[135] Compared with iPSC whose reprogramming requires at least two weeks, the formation of induced progenitor cells sometimes occurs within a few days and the efficiency of reprogramming is usually many times higher. This reprogramming does not always require cell division.[136] The cells resulting from such reprogramming are more suitable for cell therapy because they do not form teratomas.[133]

Single transcription factor transdifferentiation edit

Originally only early embryonic cells could be coaxed into changing their identity. Mature cells are resistant to changing their identity once they've committed to a specific kind. However, brief expression of a single transcription factor, the ELT-7 GATA factor, can convert the identity of fully differentiated, specialized non-endodermal cells of the w:pharynx into fully differentiated intestinal cells in intact w:larvae and adult roundworm w:Caenorhabditis elegans with no requirement for a dedifferentiated intermediate.[137]

Mogrify algorithm edit

Determining the unique set of cellular factors that is needed to be manipulated for each cell conversion is a long and costly process that involved much trial and error. As a result, this first step of identifying the key set of cellular factors for cell conversion is the major obstacle researchers face in the field of cell reprogramming. An international collaboration of researchers from the Duke-NUS Medical School in Singapore, the University of Bristol in the United Kingdom, Monash University in Australia, and RIKEN in Japan have developed an algorithm, called Mogrify(1), that can predict the optimal set of cellular factors required to convert one human cell type to another. That will drastically reduce the time and effort needed to create induced stem cells When tested, Mogrify was able to accurately predict the set of cellular factors required for previously published cell conversions correctly. To further validate Mogrify's predictive ability, the team conducted two novel cell conversions in the laboratory using human cells, and these were successful in both attempts solely using the predictions of Mogrify.[138][139][140]

The future medical implications of this novel breakthrough in cellular reprogramming are not hard to imagine. A bewildering range of diseases and disorders could be relegated to the dustbin of medical history—from arthritis to macular degeneration, from lost limbs to cancer itself. Mogrify has been made available online for other researchers and scientists.

Phased process modeling regeneration edit

Another way of reprogramming is the simulation of the processes that occur during w:amphibian limb regeneration. In w:urodele amphibians, an early step in limb regeneration is skeletal muscle fiber dedifferentiation into a cellulate that proliferates into limb tissue. However, sequential small molecule treatment of the muscle fiber with myoseverin, w:reversine (the w:aurora B kinase inhibitor) and some other chemicals: BIO (glycogen synthase-3 kinase inhibitor), w:lysophosphatidic acid (pleiotropic activator of G-protein-coupled receptors), w:SB203580 (w:p38 MAP kinase inhibitor), or w:SQ22536 (adenylyl cyclase inhibitor) causes the formation of new muscle cell types as well as other cell types such as precursors to fat, bone and nervous system cells.[141]

Antibody-based transdifferentiation edit

The researchers discovered that GCSF-mimicking w:antibody can activate a growth-stimulating receptor on marrow cells in a way that induces marrow stem cells that normally develop into white blood cells to become neural progenitor cells. The technique[142] enables researchers to search large libraries of antibodies and quickly select the ones with a desired biological effect.[143]

Conditionally reprogrammed cells edit

Schlegel and Liu[144] demonstrated that the combination of feeder cells[145][146][147] and a w:Rho kinase inhibitor (Y-27632) [148][149] induces normal and tumor epithelial cells from many tissues to proliferate indefinitely in vitro. This process occurs without the need for transduction of exogenous viral or cellular genes. These cells have been termed "Conditionally Reprogrammed Cells (CRC)". The induction of CRCs is rapid and results from reprogramming of the entire cell population. CRCs do not express high levels of proteins characteristic of iPSCs or embryonic stem cells (ESCs) (e.g., Sox2, Oct4, Nanog, or Klf4). This induction of CRCs is reversible and removal of Y-27632 and feeders allows the cells to differentiate normally.[144][150][151] CRC technology can generate 2×106 cells in 5 to 6 days from needle biopsies and can generate cultures from cryopreserved tissue and from fewer than four viable cells. CRCs retain a normal w:karyotype and remain nontumorigenic. This technique also efficiently establishes cell cultures from human and rodent tumors.[144][152][153]

The ability to rapidly generate many tumor cells from small biopsy specimens and frozen tissue provides significant opportunities for cell-based diagnostics and therapeutics (including chemosensitivity testing) and greatly expands the value of biobanking.[144][152][153] Using CRC technology, researchers were able to identify an effective therapy for a patient with a rare type of lung tumor.[154] In addition, the CRC method allows for the genetic manipulation of epithelial cells ex vivo and their subsequent evaluation in vivo in the same host. While initial studies revealed that co-culturing epithelial cells with Swiss 3T3 cells J2 was essential for CRC induction, with transwell culture plates, physical contact between feeders and epithelial cells is not required for inducing CRCs, and more importantly that irradiation of the feeder cells is required for this induction. Consistent with the transwell experiments, conditioned medium induces and maintains CRCs, which is accompanied by a concomitant increase of cellular telomerase activity. The activity of the conditioned medium correlates directly with radiation-induced feeder cell apoptosis. Thus, conditional reprogramming of epithelial cells is mediated by a combination of Y-27632 and a soluble factor(s) released by apoptotic feeder cells.[155]

A different approach to CRC is to inhibit w:CD47 - a w:membrane protein that is the w:thrombospondin-1 receptor. Loss of CD47 permits sustained proliferation of primary w:murine endothelial cells, increases asymmetric division, and enables these cells to spontaneously reprogram to form multipotent w:embryoid body-like clusters. CD47 knockdown acutely increases w:mRNA levels of c-Myc and other stem cell transcription factors in cells in vitro and in vivo. Thrombospondin-1 is a key environmental signal that inhibits stem cell self-renewal via CD47. Thus, CD47 antagonists enable cell self-renewal and reprogramming by overcoming negative regulation of c-Myc and other stem cell transcription factors.[156] In vivo blockade of CD47 using an antisense w:morpholino increases survival of mice exposed to lethal total body irradiation due to increased proliferative capacity of bone marrow-derived cells and radioprotection of radiosensitive gastrointestinal tissues.[157]

Indirect lineage conversion edit

Indirect lineage conversion is a reprogramming methodology in which somatic cells transition through a plastic intermediate state of partially reprogrammed cells (pre-iPSC), induced by brief exposure to reprogramming factors, followed by differentiation in a specially developed chemical environment (artificial niche).[158]

This method could be both more efficient and safer, since it does not seem to produce tumors or other undesirable genetic changes, and results in much greater yield than other methods. However, the safety of these cells remains questionable. Since lineage conversion from pre-iPSC relies on the use of iPSC reprogramming conditions, a fraction of the cells could acquire pluripotent properties if they do not stop the de-differentation process in vitro or due to further de-differentiation in vivo.[159]

Outer membrane glycoprotein edit

A common feature of pluripotent stem cells is the specific nature of protein w:glycosylation of their outer membrane. That distinguishes them from most nonpluripotent cells, although not w:white blood cells.[160] The w:glycans on the stem cell surface respond rapidly to alterations in cellular state and signaling and are therefore ideal for identifying even minor changes in cell populations. Many w:stem cell markers are based on cell surface glycan epitopes including the widely used markers SSEA-3, SSEA-4, Tra 1-60, and Tra 1-81.[161] Suila Heli et al.[162] speculate that in human stem cells extracellular O-GlcNAc and extracellular O-LacNAc, play a crucial role in the fine tuning of w:Notch signaling pathway - a highly conserved cell signaling system, that regulates cell fate specification, differentiation, left–right asymmetry, apoptosis, somitogenesis, angiogenesis, and plays a key role in stem cell proliferation (reviewed by Perdigoto and Bardin[163] and Jafar-Nejad et al.[164])

Changes in outer membrane protein glycosylation are markers of cell states connected in some way with pluripotency and differentiation.[165] The glycosylation change is apparently not just the result of the initialization of gene expression, but perform as an important gene regulator involved in the acquisition and maintenance of the undifferentiated state.[166]

For example, activation of w:glycoprotein ACA,[167] linking glycosylphosphatidylinositol on the surface of the progenitor cells in human peripheral blood, induces increased expression of genes Wnt, w:Notch-1, w:BMI1 and w:HOXB4 through a signaling cascade w:PI3K/w:Akt/mTor/PTEN, and promotes the formation of a self-renewing population of hematopoietic stem cells.[168]

Furthermore, dedifferentiation of progenitor cells induced by ACA-dependent signaling pathway leads to ACA-induced pluripotent stem cells, capable of differentiating in vitro into cells of all three w:germ layers.[169] The study of w:lectins' ability to maintain a culture of pluripotent human stem cells has led to the discovery of lectin w:Erythrina crista-galli (ECA), which can serve as a simple and highly effective matrix for the cultivation of human pluripotent stem cells.[170]

Reprogramming through a physical approach edit

w:Cell adhesion protein E-cadherin is indispensable for a robust pluripotent w:phenotype.[171] During reprogramming for iPS cell generation, N-cadherin can replace function of E-cadherin.[172] These functions of cadherins are not directly related to adhesion because sphere morphology helps maintaining the "stemness" of stem cells.[173] Moreover, sphere formation, due to forced growth of cells on a low attachment surface, sometimes induces reprogramming. For example, neural progenitor cells can be generated from fibroblasts directly through a physical approach without introducing exogenous reprogramming factors.

Physical cues, in the form of parallel microgrooves on the surface of cell-adhesive substrates, can replace the effects of small-molecule epigenetic modifiers and significantly improve reprogramming efficiency. The mechanism relies on the mechanomodulation of the cells’ epigenetic state. Specifically, "decreased histone deacetylase activity and upregulation of the expression of WD repeat domain 5 (WDR5)—a subunit of H3 methyltranferase—by microgrooved surfaces lead to increased histone H3 acetylation and methylation". Nanofibrous scaffolds with aligned fibre orientation produce effects similar to those produced by microgrooves, suggesting that changes in cell morphology may be responsible for modulation of the epigenetic state.[174]

 
w:Role of cell adhesions in neural development. Image courtesy of Wikipedia user JWSchmidt under the GNU Free Documentation License

Substrate rigidity is an important biophysical cue influencing neural induction and subtype specification. For example, soft substrates promote neuroepithelial conversion while inhibiting w:neural crest differentiation of hESCs in a BMP4-dependent manner. Mechanistic studies revealed a multi-targeted mechanotransductive process involving mechanosensitive Smad w:phosphorylation and nucleocytoplasmic shuttling, regulated by rigidity-dependent Hippo/YAP activities and w:actomyosin w:cytoskeleton integrity and w:contractility.[175]

An initial sensing event of tissue and extracellular matrix (ECM) stiffness includes a pathway consisting of focal adhesion kinase (FAK), the adaptor protein p130Cas (Cas - Crk-associated substrates), and the guanosine triphosphatase Rac which selectively transduce ECM stiffness into stable intracellular stiffness, to increase the abundance of the cell cycle protein cyclin D1, and to promote S-phase entry. Rac-dependent intracellular stiffening involve its binding partner lamellipodin, a protein that transmits Rac signals to the cytoskeleton during cell migration. Such mechanotransduction by a FAK-Cas-Rac-lamellipodin signaling module converts the external information encoded by ECM stiffness into stable intracellular stiffness and mechanosensitive cell cycling.[176]

Mouse embryonic stem cells (mESCs) undergo self-renewal in the presence of the w:cytokine w:leukemia inhibitory factor (LIF). Following LIF withdrawal, mESCs differentiate, accompanied by an increase in cell–substratum w:adhesion and cell spreading. Restricted cell spreading in the absence of LIF by either culturing mESCs on chemically defined, weakly adhesive biosubstrates, or by manipulating the w:cytoskeleton allowed the cells to remain in an undifferentiated and pluripotent state. The effect of restricted cell spreading on mESC self-renewal is not mediated by increased intercellular adhesion, as inhibition of mESC adhesion using a function blocking anti E-cadherin antibody or w:siRNA does not promote differentiation.[177] Possible mechanisms of stem cell fate predetermination by physical interactions with the extracellular matrix have been described.[178]

Cells involved in the reprogramming process change morphologically as the process proceeds. This results in physical difference in adhesive forces among cells. Substantial differences in 'adhesive signature' between pluripotent stem cells, partially reprogrammed cells, differentiated progeny and somatic cells allowed to develop separation process for isolation of pluripotent stem cells in w:microfluidic devices,[179][180] which is: • fast (separation takes less than 10 minutes); • efficient (separation results in a greater than 95 percent pure iPS cell culture); • innocuous (cell survival rate is greater than 80 percent and the resulting cells retain normal transcriptional profiles, differentiation potential and karyotype).

Discussion on potential future applications of lab-on-a-chips for stem cell research, see inː[181]

A novel method for cell reprogramming and fully automating stem cell cultures entire process is been developed by using smart surfaces that make cell adhesion and de-adhesion possible depending on changes in the environment.[182] This iterative method of cell culture enables to completely automate and remove the need for human involvement in the cell separation and washing stages, without using any additives that increase the toxicity level (such as trypsin).[183]

Stem cells possess mechanical memory (they remember past physical signals)—with the w:Hippo signaling pathway factors:[184] Yes-associated protein (YAP) and transcriptional coactivator with PDZ-binding domain (TAZ) acting as an intracellular mechanical rheostat—that stores information from past physical environments and influences the cells’ fate.[185][186]

Neural stem cells edit

Stroke and many neurodegenerative disorders such as Parkinson's disease, Alzheimer’s disease, amyotrophic lateral sclerosis need cell replacement therapy. The successful use of converted neural cells (cNs) in transplantations open a new avenue to treat such diseases.[187] Nevertheless, induced neurons (iNs), directly converted from fibroblasts are terminally committed and exhibit very limited proliferative ability that may not provide enough autologous donor cells for transplantation.[188] Self-renewing induced neural stem cells (iNSCs) provide additional advantages over iNs for both basic research and clinical applications.[131][132][133][189][190]

For example, under specific growth conditions, mouse fibroblasts can be reprogrammed with a single factor, Sox2, to form iNSCs that self-renew in culture and after transplantation can survive and integrate without forming tumors in mouse brains.[191] INSCs can be derived from adult human fibroblasts by non-viral techniques, thus offering a safe method for autologous transplantation or for the development of cell-based disease models.[190]

Neural chemicaly-induced progenitor cells (ciNPCs) can be generated from mouse tail-tip fibroblasts and human urinary somatic cells without introducing exogenous factors, but - by a chemical cocktail, namely VCR (V, VPA, an inhibitor of HDACs; C, CHIR99021, an inhibitor of GSK-3 kinases and R, RepSox, an inhibitor of w:TGF beta signaling pathways), under a physiological hypoxic condition.[192] Alternative cocktails with inhibitors of histone deacetylation, glycogen synthase kinase, and TGF-β pathways (where: w:sodium butyrate (NaB) or w:Trichostatin A (TSA) could replace VPA, w:Lithium chloride (LiCl) or lithium carbonate (Li2CO3) could substitute CHIR99021, or Repsox may be replaced with w:SB-431542 or w:Tranilast) show similar efficacies for ciNPC induction.[192]

Multiple methods of direct transformation of somatic cells into induced neural stem cells have been described.[193]

Proof of principle experiments demonstrate that it is possible to convert transplanted human fibroblasts and human w:astrocytes directly in the brain that are engineered to express inducible forms of neural reprogramming genes, into neurons, when reprogramming genes (Ascl1, Brn2a and w:Myt1l) are activated after transplantation using a drug.[194]

w:Astrocytes—the most common w:neuroglial brain cells, which contribute to w:scar formation in response to injury—can be directly reprogrammed in vivo to become functional neurons that formed networks in mice without the need of cell transplantation.[195] The researchers followed the mice for nearly a year to look for signs of tumor formation and reported finding none. The same researchers have turned scar-forming astrocytes into progenitor cells called neuroblasts that regenerated into neurons in the injured adult spinal cord.[196]

Oligodendrocyte precursor cells edit

Without w:myelin to insulate neurons, nerve signals quickly lose power. Diseases that attack myelin, such as multiple sclerosis, result in nerve signals that cannot propagate to nerve endings, and as a consequence lead to cognitive, motor and sensory problems. Transplantation of w:oligodendrocyte precursor cells (OPCs), which can successfully create myelin sheaths around nerve cells, is a promising potential therapeutic response. Direct lineage conversion of mouse and rat fibroblasts into oligodendroglial cells provides a potential source of OPCs. Conversion by forced expression of both eight[197] or of the three[198] transcription factors Sox10, Olig2 and Zfp536, may provide such cells.

Cardiomyocytes edit

Cell-based in vivo therapies may provide a transformative approach to augment vascular and muscle growth and to prevent non-contractile scar formation by delivering transcription factors[128] or microRNAs[14] to the heart.[199] Cardiac fibroblasts, which represent 50% of the cells in the mammalian heart, can be reprogrammed into cardiomyocyte-like cells in vivo by local delivery of cardiac core transcription factors ( GATA4, MEF2C, TBX5 and for improved reprogramming plus ESRRG, MESP1, Myocardin and ZFPM2) after coronary ligation.[128][200] These results implicated therapies that can directly remuscularize the heart without cell transplantation. However, the efficiency of such reprogramming turned out to be very low and the phenotype of received cardiomyocyte-like cells does not resemble those of a mature normal cardiomyocyte. Furthermore, transplantation of cardiac transcription factors into injured murine hearts resulted in poor cell survival and minimal expression of cardiac genes.[201]

Meanwhile, advances in the methods of obtaining cardiac myocytes in vitro occurred.[202][203] Efficient cardiac differentiation of human iPS cells gave rise to progenitors that were retained within infarcted rat hearts, and reduced remodeling of the heart after ischemic damage.[204]

Furthermore, w:ischemic cardiomyopathy in the murine infarction model was targeted by iPS cell transplantation. It synchronized failing ventricles, offering a regenerative strategy to achieve resynchronization and protection from w:decompensation by dint of improved left ventricular conduction and contractility, reduced scarring and reversal of structural remodelling.[205] One protocol generated populations of up to 98% cardiomyocytes from hPSCs simply by modulating the canonical w:Wnt signaling pathway at defined time points in during differentiation, using readily accessible small molecule compounds.[206]

Discovery of the mechanisms controlling the formation of cardiomyocytes led to the development of the drug ITD-1, which effectively clears the cell surface from TGF-β receptor type II and selectively inhibits intracellular TGF-β signaling. It thus selectively enhances the differentiation of uncommitted w:mesoderm to cardiomyocytes, but not to vascular smooth muscle and endothelial cells.[207]

One project seeded decellularized mouse hearts with human iPSC-derived multipotential cardiovascular progenitor cells. The introduced cells migrated, proliferated and differentiated in situ into cardiomyocytes, smooth muscle cells and endothelial cells to reconstruct the hearts. In addition, the heart's extracellular matrix (the substrate of heart scaffold) signalled the human cells into becoming the specialised cells needed for proper heart function. After 20 days of perfusion with growth factors, the engineered heart tissues started to beat again and were responsive to drugs.[208]

See also: review[209]

Treatment for cardiac arrhythmias edit

w:Tbx18 transduction is a method of turning on genes in heart muscle cells as a treatment for certain w:cardiac arrhythmias. Tbx18 gene therapy is aimed at treating a group of arrhythmias known as sick sinus syndrome. In a healthy heart, w:sinoatrial node (SAN) cells act as the heart’s pacemaker and cause the heart to beat in a regular rhythm. Approximately 10 thousand of the 10 billion cells in the heart are SAN cells.[210] The Tbx18 gene is required for development of pacemaker cells in the heart during fetal development but is normally not functional after birth[211] Tbx18 transduction converts atrial muscle cells into SAN cells that initiate the heartbeat. An engineered virus carrying the Tbx18 gene is injected into animals and infects atrial muscle cells. Inside atrial muscle cells the Tbx18 gene is expressed. Tbx18 turns on genes that drive SA node cell development, simultaneously turning off genes that create atrial muscle cells. Tbx18 gene therapy has been successful in rodent hearts, converting atrial muscle cells into SAN cells by expression of the Tbx18 transcription factor. Tbx18 expression in atrial myocytes was shown to convert them into functional SAN cells in an experiment done in rodents. These converted SAN cells are able to respond to the nervous system, allowing the heart to be regulated as normal. Adenoviral TBX18 gene transfer could create biological pacemaker activity in vivo in a large-animal model of complete heart block. Biological pacemaker activity, originating from the intramyocardial injection site, was evident in TBX18-transduced animals starting at day 2 and persisted for the duration of the study (14 days) with minimal backup electronic pacemaker use. Relative to controls transduced with a reporter gene, TBX18-transduced animals exhibited enhanced autonomic responses and physiologically superior chronotropic support of physical activity. Induced sinoatrial node cells could be identified by their distinctive morphology at the site of injection in TBX18-transduced animals, but not in controls. No local or systemic safety concerns arose. Thus, minimally invasive TBX18 gene transfer creates physiologically relevant pacemaker activity in complete heart block, providing evidence for therapeutic somatic reprogramming in a clinically relevant disease model.[212]

Rejuvenation of the muscle stem cell edit

The elderly often suffer from progressive w:muscle weakness and regenerative failure owing in part to elevated activity of the p38α and p38β mitogen-activated kinase pathway in senescent skeletal muscle stem cells. Subjecting such stem cells to transient inhibition of p38α and p38β in conjunction with culture on soft w:hydrogel substrates rapidly expands and rejuvenates them that result in the return of their strength.[213]

In geriatric mice, resting satellite cells lose reversible quiescence by switching to an irreversible pre-senescence state, caused by derepression of w:p16INK4a (also called Cdkn2a). On injury, these cells fail to activate and expand, even in a youthful environment. p16INK4a silencing in geriatric satellite cells restores quiescence and muscle regenerative functions.[214]

Myogenic progenitors for potential use in disease modeling or cell-based therapies targeting skeletal muscle could also be generated directly from induced pluripotent stem cells using free-floating spherical culture (EZ spheres) in a culture medium supplemented with high concentrations (100 ng/ml) of fibroblast growth factor-2 (w:FGF-2) and w:epidermal growth factor.[215]

Hepatocytes edit

Unlike current protocols for deriving w:hepatocytes from human fibroblasts, Saiyong Zhu et al., (2014)[216] did not generate iPSCs but, using small molecules, cut short reprogramming to pluripotency to generate an induced multipotent progenitor cell (iMPC) state from which endoderm progenitor cells and subsequently hepatocytes (iMPC-Heps) were efficiently differentiated. After transplantation into an immune-deficient mouse model of human liver failure, iMPC-Heps proliferated extensively and acquired levels of hepatocyte function similar to those of human primary adult hepatocytes. iMPC-Heps did not form tumours, most probably because they never entered a pluripotent state. Acute inactivation of Hippo pathway signaling in vivo is sufficient to dedifferentiate, at very high efficiencies, adult hepatocytes into cells bearing progenitor characteristics. These hepatocyte-derived progenitor cells demonstrate self-renewal and engraftment capacity at the single-cell level.[217]

 
An intestinal crypt - an accessible and abundant source of intestinal epithelial cells for conversion into β-like cells.

These results establish the feasibility of significant liver repopulation of mice with human hepatocytes generated in vitro, which removes a long-standing roadblock on the path to autologous liver cell therapy.

Insulin-producing cells edit

Complications of Diabetes mellitus such as w:cardiovascular diseases, retinopathy, neuropathy, nephropathy, and peripheral circulatory diseases depend on sugar dysregulation due to lack of w:insulin from pancreatic w:beta cells and can be lethal if they are not treated. One of the promising approaches to understand and cure diabetes is to use pluripotent stem cells (PSCs), including embryonic stem cells (ESCs) and induced PCSs (iPSCs).[218] Unfortunately, human PSC-derived insulin-expressing cells resemble human fetal β cells rather than adult β cells. In contrast to adult β cells, fetal β cells seem functionally immature, as indicated by increased basal w:glucose secretion and lack of glucose stimulation and confirmed by w:RNA-seq of whose transcripts.[219]

Overexpression of the three transcription factors, w:PDX1 (required for pancreatic bud outgrowth and beta-cell maturation), NGN3 (required for endocrine precursor cell formation) and MAFA (for beta-cell maturation) combination (called PNM) can lead to the transformation of some cell types into a beta cell-like state.[220] An accessible and abundant source of functional insulin-producing cells is intestine. PMN expression in human intestinal “w:organoids” stimulates the conversion of intestinal epithelial cells into β-like cells possibly acceptable for transplantation.[221]

Nephron Progenitors edit

Adult proximal tubule cells were directly transcriptionally reprogrammed to w:nephron progenitors of the embryonic w:kidney, using a pool of six genes of instructive transcription factors (SIX1, SIX2, OSR1, Eyes absent homolog 1(EYA1), Homeobox A11 (HOXA11) and Snail homolog 2 (SNAI2)) that activated genes consistent with a cap mesenchyme/nephron progenitor phenotype in the adult proximal tubule cell line.[222] The generation of such cells may lead to cellular therapies for adult w:renal disease. Embryonic kidney organoids placed into adult rat kidneys can undergo onward development and vascular development.[223]

Blood vessel cells edit

As blood vessels age, they often become abnormal in structure and function, thereby contributing to numerous age-associated diseases including myocardial infarction, ischemic stroke and atherosclerosis of arteries supplying the heart, brain and lower extremities. So, an important goal is to stimulate vascular growth for the w:collateral circulation to prevent the exacerbation of these diseases. Induced Vascular Progenitor Cells (iVPCs) are useful for cell-based therapy designed to stimulate coronary collateral growth. They were generated by partially reprogramming endothelial cells.[158] The vascular commitment of iVPCs is related to the epigenetic memory of endothelial cells, which engenders them as cellular components of growing blood vessels. That is why, when iVPCs were implanted into w:myocardium, they engrafted in blood vessels and increased coronary collateral flow better than iPSCs, mesenchymal stem cells, or native endothelial cells.[224]

Ex vivo genetic modification can be an effective strategy to enhance stem cell function. For example, cellular therapy employing genetic modification with Pim-1 kinase (a downstream effector of w:Akt, which positively regulates neovasculogenesis) of w:bone marrow–derived cells[225] or human cardiac progenitor cells, isolated from failing myocardium[226] results in durability of repair, together with the improvement of functional parameters of myocardial hemodynamic performance.

Stem cells extracted from fat tissue after w:liposuction may be coaxed into becoming progenitor smooth muscle cells (iPVSMCs) found in arteries and veins.[227]

The 2D culture system of human iPS cells[228] in conjunction with triple marker selection (w:CD34 (a surface glycophosphoprotein expressed on developmentally early embryonic fibroblasts), w:NP1 (receptor - neuropilin 1) and KDR (kinase insert domain-containing receptor)) for the isolation of vasculogenic precursor cells from human iPSC, generated endothelial cells that, after transplantation, formed stable, functional mouse blood vessels in vivo, lasting for 280 days.[229]

To treat infarction, it is important to prevent the formation of fibrotic scar tissue. This can be achieved in vivo by transient application of w:paracrine factors that redirect native heart progenitor stem cell contributions from scar tissue to cardiovascular tissue. For example, in a mouse myocardial infarction model, a single intramyocardial injection of human w:vascular endothelial growth factor A mRNA (VEGF-A modRNA), modified to escape the body's normal defense system, results in long-term improvement of heart function due to mobilization and redirection of epicardial progenitor cells toward cardiovascular cell types.[230]

Blood stem cells edit

Definitive w:hematopoiesis emerges during embryogenesis via an endothelial-to-hematopoietic transition. Combination of four transcription factors, Gata2, Gfi1b, cFos, and Etv6, is sufficient to induce in vitro development leading to the formation of endothelial-like precursor cells, with the subsequent appearance of hematopoietic cells.[231] Transient expression of six transcription factors (Run1t1, Hlf, Lmo2, Prdm5, Pbx1, Zfp37) and also Mycn with Meis1, to improve reprogramming efficacy, is sufficient to activate the gene networks governing hematopoietic stem cells functional identity in committed blood cells. This findings mark a significant step toward one of the most sought-after goals of regenerative medicine: the ability to produce hematopoietic stem cells suitable for transplantation, using more mature or differentiated blood cells to make up the shortfall of w:bone marrow transplants[232] It should be noted, however, that the transcription factors used in this study belong to a group of proto-oncogenes and therefore these cells could be dangerous to humans. Still to be answered are the precise contribution of each of the eight factors to the reprogramming process and whether approaches that do not rely on viruses and transcription factors can have similar success. It also is not yet known whether the same results can be achieved using human cells or whether other, non-blood cells can be reprogrammed to iHSCs.[233]

See also OPEN Reviewː [234]

Red blood cells edit

RBC transfusion is necessary for many patients. However, to date the supply of RBCs remains labile. In addition, transfusion risks infectious disease transmission. A large supply of safe RBCs generated in vitro would help to address this issue. Ex vivo erythroid cell generation may provide alternative transfusion products to meet present and future clinical requirements.[235][236] Red blood cells (RBC)s generated in vitro from mobilized w:CD34 positive cells have normal survival when transfused into an autologous recipient.[237] RBC produced in vitro contained exclusively w:fetal hemoglobin (HbF), which rescues the functionality of these RBCs. In vivo the switch of fetal to adult hemoglobin was observed after infusion of nucleated w:erythroid precursors derived from iPSCs.[238] Although RBCs do not have nuclei and therefore can not form a tumor, their immediate erythroblasts precursors have nuclei. The terminal maturation of erythroblasts into functional RBCs requires a complex remodeling process that ends with extrusion of the nucleus and the formation of an enucleated RBC.[239] Cell reprogramming often disrupts enucleation. Transfusion of in vitro-generated RBCs or erythroblasts does not sufficiently protect against tumor formation.

The w:aryl hydrocarbon receptor (AhR) pathway (which has been shown to be involved in the promotion of cancer cell development) plays an important role in normal blood cell development. AhR activation in human hematopoietic progenitor cells (HPs) drives an unprecedented expansion of HPs, megakaryocyte- and erythroid-lineage cells.[240] See also Concise Review:[241][242]

Platelets edit

w:Platelets help prevent hemorrhage in w:thrombocytopenic patients and patients with w:thrombocythemia. A significant problem for multitransfused patients is refractoriness to platelet transfusions. Thus, the ability to generate platelet products ex vivo and platelet products lacking HLA antigens in serum-free media would have clinical value. An w:RNA interference-based mechanism used a lentiviral vector to express short-hairpin RNAi targeting β2-microglobulin transcripts in CD34-positive cells. Generated platelets demonstrated an 85% reduction in class I HLA antigens. These platelets appeared to have normal function in vitro[243]

One clinically-applicable strategy for the derivation of functional platelets from human iPSC involves the establishment of stable immortalized megakaryocyte progenitor cell lines (imMKCLs) through w:doxycycline-dependent overexpression of w:BMI1 and BCL-XL. The resulting imMKCLs can be expanded in culture over extended periods (4–5 months), even after w:cryopreservation. Halting the overexpression (by the removal of doxycycline from the medium) of c-MYC, w:BMI1 and BCL-XL in growing imMKCLs led to the production of CD42b+ platelets with functionality comparable to that of native platelets on the basis of a range of assays in vitro and in vivo.[244]

Immune cells edit

A specialised type of w:white blood cell, known as cytotoxic T w:lymphocytes (CTLs), are produced by the w:immune system and are able to recognise specific markers on the surface of various infectious or tumour cells, causing them to launch an attack to kill the harmful cells. Thence, immunotherapy with functional antigen-specific T cells has potential as a therapeutic strategy for combating many cancers and viral infections.[245] However, cell sources are limited, because they are produced in small numbers naturally and have a short lifespan.

A potentially efficient approach for generating antigen-specific CTLs is to revert mature immune T cells into iPSCs, which possess indefinite proliferative capacity in vitro, and after their multiplication to coax them to redifferentiate back into T cells.[246][247][248]

Another method combines iPSC and w:chimeric antigen receptor (CAR) [249] technologies to generate human T cells targeted to w:CD19, an antigen expressed by malignant w:B cells, in tissue culture.[250] This approach of generating therapeutic human T cells may be useful for cancer immunotherapy and other medical applications.

Invariant natural killer T (iNKT) cells have great clinical potential as w:adjuvants for cancer immunotherapy. iNKT cells act as innate T lymphocytes and serve as a bridge between the innate and w:acquired immune systems. They augment anti-tumor responses by producing w:interferon-gamma (IFN-γ).[251] The approach of collection, reprogramming/dedifferentiation, re-differentiation and injection has been proposed for related tumor treatment.[252]

w:Dendritic cells (DC) are specialized to control T-cell responses. DC with appropriate genetic modifications may survive long enough to stimulate antigen-specific CTL and after that be completely eliminated. DC-like antigen-presenting cells obtained from human induced pluripotent stem cells can serve as a source for w:vaccination therapy.[253]

CCAAT/enhancer binding protein-α (C/EBPα) induces transdifferentiation of w:B cells into w:macrophages at high efficiencies[254] and enhances reprogramming into iPS cells when co-expressed with transcription factors Oct4, Sox2, Klf4 and Myc.[255] with a 100-fold increase in iPS cell reprogramming efficiency, involving 95% of the population.[256] Furthermore, C/EBPa can convert selected human B cell lymphoma and leukemia cell lines into macrophage-like cells at high efficiencies, impairing the cells’ tumor-forming capacity.[257]

Phagocytic cell line edit

Inoue et al. developed and implemented a technique to present a molecule of interest at the cell surface in an inducible manner on a time scale of minutes. They simultaneously induced the cell surface display of the C2 domain of milk fat globule epidermal growth factor factor 8 (MFGE8) and activated the intracellular small guanosine triphosphatase Rac, which stimulates w:actin polymerization at the cell periphery. The C2 domain binds to w:phosphatidylserine, a lipid exposed on the surface of w:apoptotic cells. By integrating the stimulation of these two processes, they converted w:HeLa cells into a phagocytic cell line that bound to and engulfed apoptotic human w:Jurkat cells.[258] This cell surface display technique might be useful as part of a targeted, cell-based therapy in which unwanted cells with characteristic surface molecules could be rapidly consumed by engineered cells.

Thymic epithelial cells rejuvenation edit

The w:thymus is the first organ to deteriorate as people age. This shrinking is one of the main reasons the immune system becomes less effective with age. Diminished expression of the thymic epithelial cell transcription factor w:FOXN1 has been implicated as a component of the mechanism regulating age-related involution.[259][260]

Clare Blackburn and colleagues show that established age-related thymic involution can be reversed by forced upregulation of just one transcription factor - FOXN1 in the thymic epithelial cells in order to promote rejuvenation, proliferation and differentiation of these cells into fully functional thymic epithelium.[261] This rejuvenation and increased proliferation was accompanied by upregulation of genes that promote w:cell cycle progression (w:cyclin D1, ΔNp63, FgfR2IIIb) and that are required in the thymic epithelial cells to promote specific aspects of w:T cell development (Dll4, Kitl, Ccl25, Cxcl12, Cd40, Cd80, Ctsl, Pax1). Furthermore, Clare Blackburn and colleagues show that enforced ​Foxn1 expression is sufficient to reprogramme fibroblasts into functional thymic epithelial cells (TECs), an unrelated cell type across a germ-layer boundary. These ​FOXN1-induced TECs (iTECs) supported efficient development of both ​CD4+ and CD8+ T cells in vitro. On transplantation, iTECs established a complete, fully organized and functional thymus, that contained all of the TEC subtypes required to support T-cell differentiation and populated the recipient immune system with T cells. iTECs thus demonstrate that cellular reprogramming approaches can be used to generate an entire organ, and open the possibility of widespread use of thymus transplantation to boost immune function in patients.[262]

Induction of Mesenchymal stem cells edit

Mesenchymal stem/stromal cells (MSCs) are under investigation for applications in cardiac, renal, neural, joint and bone repair, as well as in inflammatory conditions and hemopoietic cotransplantation.[263] This is because of their immunosuppressive properties and ability to differentiate into a wide range of mesenchymal-lineage tissues. MSCs are typically harvested from adult bone marrow or fat, but these require painful invasive procedures and are low-frequency sources, making up only 0.001%– 0.01% of bone marrow cells and 0.05% in liposuction aspirates.[264] Of concern for autologous use, in particular in the elderly most in need of tissue repair, MSCs decline in quantity and quality with age.[263][265][266]

IPSCs could be obtained by the cells rejuvenation of even centenarians.[9][38] Because iPSCs can be harvested free of ethical constraints and culture can be expanded indefinitely, they are an advantageous source of MSCs.[267][268][269] IPSC treatment with w:SB-431542 leads to rapid and uniform MSC generation from human iPSCs. (SB-431542 is an inhibitor of activin/TGF- pathways by blocking w:phosphorylation of w:ALK4, w:ALK5, and w:ALK7 receptors.) These iPS-MSCs may lack teratoma-forming ability, display a normal stable karyotype in culture and exhibit growth and differentiation characteristics that closely resemble those of primary MSCs. It has potential for in vitro scale-up, enabling MSC-based therapies.[270] MSC derived from iPSC have the capacity to aid periodontal regeneration and are a promising source of readily accessible stem cells for use in the clinical treatment of periodontitis.[271][272][273]

Besides cell therapy in vivo, the culture of human mesenchymal stem cells can be used in vitro for mass-production of exosomes, which are ideal vehicles for drug delivery.[274]

Dedifferentiated adipocytes edit

Adipose tissue, because of its abundance and relatively less invasive harvest methods, represents a source of mesenchymal stem cells (MSCs). Unfortunately, liposuction aspirates are only 0.05% MSCs.[264] However, a large amount of mature adipocytes, which in general have lost their proliferative abilities and therefore are typically discarded, can be easily isolated from the adipose cell suspension and dedifferentiated into lipid-free fibroblast-like cells, named dedifferentiated fat (DFAT) cells. DFAT cells re-establish active proliferation ability and express multipotent capacities.[275] Compared with adult stem cells, DFAT cells show unique advantages in abundance, isolation and homogeneity. Under proper induction culture in vitro or proper environment in vivo, DFAT cells could demonstrate adipogenic, osteogenic, chondrogenic, and myogenic potentials. They could also exhibit perivascular characteristics and elicit neovascularization.[276][277][278]

Chondrogenic cells edit

w:Cartilage is the connective tissue responsible for frictionless joint movement. Its degeneration ultimately results in complete loss of joint function in the late stages of w:osteoarthritis. As an avascular and hypocellular tissue, cartilage has a limited capacity for self-repair. w:Chondrocytes are the only cell type in cartilage, in which they are surrounded by the extracellular matrix that they secrete and assemble.

One method of producing cartilage is to induce it from iPS cells.[279] Alternatively, it is possible to convert fibroblasts directly into induced chondrogenic cells (iChon) without an intermediate iPS cell stage, by inserting three reprogramming factors (c-MYC, KLF4, and SOX9).[280] Human iChon cells expressed marker genes for chondrocytes (type II collagen) but not fibroblasts.

Implanted into defects created in the articular cartilage of rats, human iChon cells survived to form cartilaginous tissue for at least four weeks, with no tumors. The method makes use of c-MYC, which is thought to have a major role in tumorigenesis and employs a w:retrovirus to introduce the reprogramming factors, excluding it from unmodified use in human therapy.[246][248][281]

Sources of cells for reprogramming edit

The most frequently used sources for reprogramming are blood cells[282][283][284][285] and fibroblasts, obtained by biopsy of the skin,[286] but taking cells from urine is less invasive.[287] The latter method does not require a biopsy or blood sampling. As of 2013, urine-derived stem cells had been differentiated into endothelial, osteogenic, chondrogenic, adipogenic, skeletal myogenic and neurogenic lineages, without forming teratomas.[288] Therefore, their epigenetic memory is suited to reprogramming into iPS cells. However, few cells appear in urine, only low conversion efficiencies had been achieved and the risk of bacterial contamination is relatively high.

Another promising source of cells for reprogramming are mesenchymal stem cells derived from human hair follicles.[289]

The origin of somatic cells used for reprogramming may influence the efficiency of reprogramming,[290][291] the functional properties of the resulting induced stem cells[292] and the ability to form tumors.[293]

IPSCs retain an epigenetic memory of their tissue of origin, which impacts their differentiation potential.[281][292][294][295][296][297] This epigenetic memory does not necessarily manifest itself at the pluripotency stage – iPSCs derived from different tissues exhibit proper morphology, express pluripotency markers and are able to differentiate into the three embryonic layers in vitro and in vivo. However, this epigenetic memory may manifest during re-differentiation into specific cell types that require the specific loci bearing residual epigenetic marks.

See also edit

̽* WikiJournal_Preprints/Induced_stem_cells

Notes edit

References for further reading edit

References edit

  1. Yamanaka, S.; Blau, H. M. (2010). "Nuclear reprogramming to a pluripotent state by three approaches". Nature 465 (7299): 704–712. doi:10.1038/nature09229. PMID 20535199. PMC 2901154. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2901154/. 
  2. Gurdon J. B. and Ian Wilmut (2011) Nuclear Transfer to Eggs and Oocytes Cold Spring Harb Perspect Biol; 3: a002659
  3. 3.0 3.1 3.2 3.3 Tachibana, Masahito; Amato, Paula; Sparman, Michelle; Gutierrez, Nuria Marti; Tippner-Hedges, Rebecca; Ma, Hong; Kang, Eunju; Fulati, Alimujiang et al. (2013). "Human Embryonic Stem Cells Derived by Somatic Cell Nuclear Transfer". Cell 153 (6): 1228–1238. doi:10.1016/j.cell.2013.05.006. PMID 23683578. PMC 3772789. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3772789/. 
  4. Noggle, Scott; Fung, Ho-Lim; Gore, Athurva; Martinez, Hector; Satriani, Kathleen Crumm; Prosser, Robert; Oum, Kiboong; Paull, Daniel et al. (2011). "Human oocytes reprogram somatic cells to a pluripotent state". Nature 478 (7367): 70–75. doi:10.1038/nature10397. PMID 21979046. 
  5. 5.0 5.1 Pan, Guangjin; Wang, Tao; Yao, Hongjie; Pei, Duanqing (2012). "Somatic cell reprogramming for regenerative medicine: SCNT vs. IPS cells". BioEssays 34 (6): 472–476. doi:10.1002/bies.201100174. PMID 22419173. 
  6. Do, J. T.; Han, D. W.; Gentile, L.; Sobek-Klocke, I.; Stehling, M.; Lee, H. T.; Schöler, H. R. (2007). "Erasure of cellular memory by fusion with pluripotent cells". Stem Cells (Dayton, Ohio) 25 (4): 1013–1020. doi:10.1634/stemcells.2006-0691. PMID 17218392. 
  7. 7.0 7.1 Takahashi, Kazutoshi; Tanabe, Koji; Ohnuki, Mari; Narita, Megumi; Ichisaka, Tomoko; Tomoda, Kiichiro; Yamanaka, Shinya (2007). "Induction of Pluripotent Stem Cells from Adult Human Fibroblasts by Defined Factors". Cell 131 (5): 861–872. doi:10.1016/j.cell.2007.11.019. PMID 18035408. 
  8. Wang, W.; Yang, J.; Liu, H.; Lu, D.; Chen, X.; Zenonos, Z.; Campos, L. S.; Rad, R. et al. (2011). "Rapid and efficient reprogramming of somatic cells to induced pluripotent stem cells by retinoic acid receptor gamma and liver receptor homolog 1". Proceedings of the National Academy of Sciences 108 (45): 18283–18288. doi:10.1073/pnas.1100893108. PMID 21990348. PMC 3215025. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3215025/. 
  9. 9.0 9.1 9.2 Lapasset, Laure; Milhavet, Ollivier; Prieur, Alexandre; Besnard, Emilie; Babled, Amelie; Aït-Hamou, Nafissa; Leschik, Julia; Pellestor, Franck et al. (2011). "Rejuvenating senescent and centenarian human cells by reprogramming through the pluripotent state". Genes & Development 25 (21): 2248–2253. doi:10.1101/gad.173922.111. PMID 22056670. PMC 3219229. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3219229/. 
  10. Zhou, Hongyan; Wu, Shili; Joo, Jin Young; Zhu, Saiyong; Han, Dong Wook; Lin, Tongxiang; Trauger, Sunia; Bien, Geoffery et al. (2009). "Generation of Induced Pluripotent Stem Cells Using Recombinant Proteins". Cell Stem Cell 4 (5): 381–384. doi:10.1016/j.stem.2009.04.005. PMID 19398399. 
  11. Li, Zhonghan; Rana, Tariq M. (2012). "Using MicroRNAs to Enhance the Generation of Induced Pluripotent Stem Cells". Current Protocols in Stem Cell Biology. Chapter 4. pp. Unit 4A.4. doi:10.1002/9780470151808.sc04a04s20. ISBN 978-0470151808. 
  12. Anokye-Danso, Frederick; Trivedi, Chinmay M.; Juhr, Denise; Gupta, Mudit; Cui, Zheng; Tian, Ying; Zhang, Yuzhen; Yang, Wenli et al. (2011). "Highly Efficient miRNA-Mediated Reprogramming of Mouse and Human Somatic Cells to Pluripotency". Cell Stem Cell 8 (4): 376–388. doi:10.1016/j.stem.2011.03.001. PMID 21474102. PMC 3090650. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3090650/. 
  13. Miyoshi, Norikatsu; Ishii, Hideshi; Nagano, Hiroaki; Haraguchi, Naotsugu; Dewi, Dyah Laksmi; Kano, Yoshihiro; Nishikawa, Shinpei; Tanemura, Masahiro et al. (2011). "Reprogramming of Mouse and Human Cells to Pluripotency Using Mature MicroRNAs". Cell Stem Cell 8 (6): 633–638. doi:10.1016/j.stem.2011.05.001. PMID 21620789. 
  14. 14.0 14.1 Jayawardena, Tilanthi M.; Egemnazarov, Bakytbek; Finch, Elizabeth A.; Zhang, Lunan; Payne, J. Alan; Pandya, Kumar; Zhang, Zhiping; Rosenberg, Paul et al. (2012). "MicroRNA-Mediated in Vitro and in Vivo Direct Reprogramming of Cardiac Fibroblasts to Cardiomyocytes". Circulation Research 110 (11): 1465–1473. doi:10.1161/CIRCRESAHA.112.269035. PMID 22539765. PMC 3380624. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3380624/. 
  15. Bao, Xichen; Zhu, Xihua; Liao, Baojian; Benda, Christina; Zhuang, Qiang; Pei, Duanqing; Qin, Baoming; Esteban, Miguel A. (2013). "MicroRNAs in somatic cell reprogramming". Current Opinion in Cell Biology 25 (2): 208–214. doi:10.1016/j.ceb.2012.12.004. PMID 23332905. 
  16. Yoshioka, Naohisa; Gros, Edwige; Li, Hai-Ri; Kumar, Shantanu; Deacon, Dekker C.; Maron, Cornelia; Muotri, Alysson R.; Chi, Neil C. et al. (2013). "Efficient Generation of Human iPSCs by a Synthetic Self-Replicative RNA". Cell Stem Cell 13 (2): 246–254. doi:10.1016/j.stem.2013.06.001. PMID 23910086. PMC 3845961. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3845961/. 
  17. 17.0 17.1 Hou, Pingping; Li, Yanqin; Zhang, Xu; Liu, Chun; Guan, Jingyang; Li, Honggang; Zhao, Ting; Ye, Junqing et al. (2013). "Pluripotent Stem Cells Induced from Mouse Somatic Cells by Small-Molecule Compounds". Science 341 (6146): 651–654. doi:10.1126/science.1239278. PMID 23868920. 
    Efe, Jem A.; Ding, Sheng (2011). "The evolving biology of small molecules: Controlling cell fate and identity". Philosophical Transactions of the Royal Society B: Biological Sciences 366 (1575): 2208–2221. doi:10.1098/rstb.2011.0006. PMID 21727126. PMC 3130415. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3130415/. 
  18. 18.0 18.1 Stadtfeld, Matthias; Apostolou, Effie; Ferrari, Francesco; Choi, Jiho; Walsh, Ryan M.; Chen, Taiping; Ooi, Steen S K.; Kim, Sang Yong et al. (2012). "Ascorbic acid prevents loss of Dlk1-Dio3 imprinting and facilitates generation of all–iPS cell mice from terminally differentiated B cells". Nature Genetics 44 (4): 398–405. doi:10.1038/ng.1110. PMID 22387999. PMC 3538378. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3538378/. 
  19. Pandian, Ganesh N.; Sugiyama, Hiroshi (2012). "Programmable genetic switches to control transcriptional machinery of pluripotency". Biotechnology Journal 7 (6): 798–809. doi:10.1002/biot.201100361. PMID 22588775. 
    Pandian, Ganesh N.; Nakano, Yusuke; Sato, Shinsuke; Morinaga, Hironobu; Bando, Toshikazu; Nagase, Hiroki; Sugiyama, Hiroshi (2012). "A synthetic small molecule for rapid induction of multiple pluripotency genes in mouse embryonic fibroblasts". Scientific Reports 2: 544. doi:10.1038/srep00544. PMID 22848790. PMC 3408130. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3408130/. 
  20. Box 3 FROM THE ARTICLE: Edward M. De Robertis (2006). Nature Reviews Molecular Cell Biology 7, 296-302 doi:10.1038/nrm1855
  21. Slack, JMW (2009). "Metaplasia and somatic cell reprogramming". The Journal of Pathology 217 (2): 161–168. doi:10.1002/path.2442. PMID 18855879. 
  22. Wei, Grace; Schubiger, Gerold; Harder, Friedrich; Müller, Albrecht M. (2000). "Stem Cell Plasticity in Mammals and Transdetermination in Drosophila : Common Themes?". Stem Cells 18 (6): 409–414. doi:10.1634/stemcells.18-6-409. PMID 11072028. 
    Worley, Melanie I.; Setiawan, Linda; Hariharan, Iswar K. (2012). "Regeneration and Transdetermination in Drosophila Imaginal Discs". Annual Review of Genetics 46: 289–310. doi:10.1146/annurev-genet-110711-155637. PMID 22934642. 
  23. Peng-Fei Xu, Nathalie Houssin, Karine F. Ferri-Lagneau, Bernard Thisse and Christine Thisse. (April 2014). Construction of a Vertebrate Embryo from Two Opposing Morphogen Gradients. Science: 344(6179), 87-89 doi:10.1126/science.1248252
  24. Tata, P. R., Mou, H., Pardo-Saganta, A., Zhao, R., Prabhu, M., Law, B. M., ... & Rajagopal, J. (2013). Dedifferentiation of committed epithelial cells into stem cells in vivo. Nature 503, 218–223 doi:10.1038/nature12777 PMID: 24196716
  25. Stange, D. E., Koo, B. K., Huch, M., Sibbel, G., Basak, O., Lyubimova, A., ... & Clevers, H. (2013). Differentiated Troy+ Chief Cells Act as Reserve Stem Cells to Generate All Lineages of the Stomach Epithelium. Cell, 155(2), 357-368. DOI: http://dx.doi.org/10.1016/j.cell.2013.09.008
  26. Kusaba, T.; Lalli, M.; Kramann, R.; Kobayashi, A.; Humphreys, B. D. (2014). "Differentiated kidney epithelial cells repair injured proximal tubule". Proceedings of the National Academy of Sciences 111 (4): 1527–1532. doi:10.1073/pnas.1310653110. PMID 24127583. PMC 3910580. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3910580/. 
  27. Sieweke, Michael H., Judith E. Allen (2013) Beyond Stem Cells: Self-Renewal of Differentiated Macrophages. Science : 342(6161) DOI: 10.1126/science.1242974
  28. Sandoval-Guzmán, Tatiana; Wang, Heng; Khattak, Shahryar; Schuez, Maritta; Roensch, Kathleen; Nacu, Eugeniu; Tazaki, Akira; Joven, Alberto et al. (2014). "Fundamental Differences in Dedifferentiation and Stem Cell Recruitment during Skeletal Muscle Regeneration in Two Salamander Species". Cell Stem Cell 14 (2): 174–187. doi:10.1016/j.stem.2013.11.007. PMID 24268695. 
  29. Kuroda Y, Wakao S, Kitada M, Murakami T, Nojima M, Dezawa M. (2013). Isolation, culture and evaluation of multilineage-differentiating stress-enduring (Muse) cells. Nat Protoc.;8(7):1391-415. doi: 10.1038/nprot.2013.076.
  30. Kuroda, Y., Kitada, M., Wakao, S., et al. & Dezawa, M. (2010) Unique multipotent cells in adult human mesenchymal cell populations. PNAS , 107(19), 8639-8643. doi:10.1073/pnas.0911647107
  31. Ogura F, Wakao S, Kuroda Y, Tsuchiyama K, Bagheri M, Heneidi S, Chazenbalk G, Aiba S, Dezawa M. (2014). Human Adipose Tissue Possesses a Unique Population of Pluripotent Stem Cells with Nontumorigenic and Low Telomerase Activities: Potential Implications in Regenerative Medicine. Stem Cells Dev. Epub ahead of print
  32. Heneidi S, Simerman AA, Keller E, Singh P, Li X, et al. (2013). Awakened by Cellular Stress: Isolation and Characterization of a Novel Population of Pluripotent Stem Cells Derived from Human Adipose Tissue. PLoS ONE 8(6): e64752. doi:10.1371/journal.pone.0064752
  33. Shigemoto T, Kuroda Y, Wakao S, Dezawa M (2013). A Novel Approach to Collecting Satellite Cells From Adult Skeletal Muscles on the Basis of Their Stress Tolerance. Stem Cells Trans Med 2 (7) 488-498 doi:10.5966/sctm.2012-0130
  34. Simerman, Ariel A.; Dumesic, Daniel A.; Chazenbalk, Gregorio D. (2014). "Pluripotent muse cells derived from human adipose tissue: A new perspective on regenerative medicine and cell therapy". Clinical and Translational Medicine 3: 12. doi:10.1186/2001-1326-3-12. PMID 24940477. PMC 4041046. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4041046/. 
  35. Sisakhtnezhad, Sajjad; Matin, Maryam M. (2012). "Transdifferentiation: A cell and molecular reprogramming process". Cell and Tissue Research 348 (3): 379–396. doi:10.1007/s00441-012-1403-y. PMID 22526624. 
  36. Dinnyes, Andras; Tian, Xiuchun Cindy; Oback, Bj¨orn (17 April 2013). Robert A. Meyers. ed. Nuclear Transfer for Cloning Animals.. John Wiley & Sons. pp. 299–344. ISBN 978-3-527-66854-0. https://books.google.com/books?id=Xuo0HI2sGngC&pg=PP4. 
    Jullien, Jerome; Pasque, Vincent; Halley-Stott, Richard P.; Miyamoto, Kei; Gurdon, J. B. (2011). "Mechanisms of nuclear reprogramming by eggs and oocytes: A deterministic process?". Nature Reviews Molecular Cell Biology 12 (7): 453–459. doi:10.1038/nrm3140. PMID 21697902. PMC 3657683. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3657683/. 
    Campbell, Keith H. S. (2002). "A background to nuclear transfer and its applications in agriculture and human therapeutic medicine". Journal of Anatomy 200 (3): 267–275. doi:10.1046/j.1469-7580.2002.00035.x. PMID 12033731. PMC 1570687. //www.ncbi.nlm.nih.gov/pmc/articles/PMC1570687/. 
  37. US 8,647,872  patent
  38. 38.0 38.1 Chung, Young Gie; Eum, Jin Hee; Lee, Jeoung Eun; Shim, Sung Han; Sepilian, Vicken; Hong, Seung Wook; Lee, Yumie; Treff, Nathan R. et al. (2014). "Human Somatic Cell Nuclear Transfer Using Adult Cells". Cell Stem Cell 14 (6): 777–780. doi:10.1016/j.stem.2014.03.015. PMID 24746675.  http://dx.doi.org/10.1016/j.stem.2014.03.015
  39. Yang, Hui; Shi, Linyu; Wang, Bang-An; Liang, Dan; Zhong, Cuiqing; Liu, Wei; Nie, Yongzhan; Liu, Jie et al. (2012). "Generation of Genetically Modified Mice by Oocyte Injection of Androgenetic Haploid Embryonic Stem Cells". Cell 149 (3): 605–617. doi:10.1016/j.cell.2012.04.002. PMID 22541431. 
  40. Hayashi, Katsuhiko; Ogushi, Sugako; Kurimoto, Kazuki; Shimamoto, So; Ohta, Hiroshi; Saitou, Mitinori (2012). "Offspring from Oocytes Derived from in Vitro Primordial Germ Cell–like Cells in Mice". Science 338 (6109): 971–975. doi:10.1126/science.1226889. PMID 23042295. 
  41. Kishigami, Satoshi; Mizutani, Eiji; Ohta, Hiroshi; Hikichi, Takafusa; Thuan, Nguyen Van; Wakayama, Sayaka; Bui, Hong-Thuy; Wakayama, Teruhiko (2006). "Significant improvement of mouse cloning technique by treatment with trichostatin a after somatic nuclear transfer". Biochemical and Biophysical Research Communications 340 (1): 183–189. doi:10.1016/j.bbrc.2005.11.164. PMID 16356478. 
  42. Wakayama, Sayaka; Kohda, Takashi; Obokata, Haruko; Tokoro, Mikiko; Li, Chong; Terashita, Yukari; Mizutani, Eiji; Nguyen, Van Thuan et al. (2013). "Successful Serial Recloning in the Mouse over Multiple Generations". Cell Stem Cell 12 (3): 293–297. doi:10.1016/j.stem.2013.01.005. PMID 23472871. 
  43. Kong, Q., Ji, G., Xie, B.,et al., & Liu, Z. (2014). Telomere Elongation Facilitated by Trichostatin A in Cloned Embryos and Pigs by Somatic Cell Nuclear Transfer. Stem Cell Reviews and Reports, 10(3). 399-407. doi:10.1007/S12015-014-9499-Y
  44. Official website of the Presidential Commission for the Study of Bioethical Issues
  45. Paull, Daniel; Emmanuele, Valentina; Weiss, Keren A.; Treff, Nathan; Stewart, Latoya; Hua, Haiqing; Zimmer, Matthew; Kahler, David J. et al. (2013). "Nuclear genome transfer in human oocytes eliminates mitochondrial DNA variants". Nature 493 (7434): 632–637. doi:10.1038/nature11800. PMID 23254936. PMC 7924261. //www.ncbi.nlm.nih.gov/pmc/articles/PMC7924261/. 
  46. Tachibana, Masahito; Amato, Paula; Sparman, Michelle; Woodward, Joy; Sanchis, Dario Melguizo; Ma, Hong; Gutierrez, Nuria Marti; Tippner-Hedges, Rebecca et al. (2013). "Towards germline gene therapy of inherited mitochondrial diseases". Nature 493 (7434): 627–631. doi:10.1038/nature11647. PMID 23103867. PMC 3561483. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3561483/. 
  47. Check Hayden, Erika (2013). "Regulators weigh benefits of 'three-parent' fertilization". Nature 502 (7471): 284–285. doi:10.1038/502284a. PMID 24132269. 
  48. Le, Rongrong; Kou, Zhaohui; Jiang, Yonghua; Li, Ming; Huang, Bo; Liu, Wenqiang; Li, Hui; Kou, Xiaochen et al. (2014). "Enhanced Telomere Rejuvenation in Pluripotent Cells Reprogrammed via Nuclear Transfer Relative to Induced Pluripotent Stem Cells". Cell Stem Cell 14 (1): 27–39. doi:10.1016/j.stem.2013.11.005. PMID 24268696. 
  49. Cibelli, Jose; Lanza, Robert; Campbell, Keith H.S.; West, Michael D. (14 September 2002). Principles of Cloning. Academic Press. ISBN 978-0-08-049215-5. https://books.google.com/books?id=RY0rXE2HgqsC. 
  50. Shinagawa, Toshie; Takagi, Tsuyoshi; Tsukamoto, Daisuke; Tomaru, Chinatsu; Huynh, Linh My; Sivaraman, Padavattan; Kumarevel, Thirumananseri; Inoue, Kimiko et al. (2014). "Histone Variants Enriched in Oocytes Enhance Reprogramming to Induced Pluripotent Stem Cells". Cell Stem Cell 14 (2): 217–227. doi:10.1016/j.stem.2013.12.015. PMID 24506885. 
  51. 51.0 51.1 Abad, María; Mosteiro, Lluc; Pantoja, Cristina; Cañamero, Marta; Rayon, Teresa; Ors, Inmaculada; Graña, Osvaldo; Megías, Diego et al. (2013). "Reprogramming in vivo produces teratomas and iPS cells with totipotency features". Nature 502 (7471): 340–345. doi:10.1038/nature12586. PMID 24025773. 
    Naik, Gautam (2013-09-11). "New Promise for Stem Cells - WSJ.com". Online.wsj.com. Retrieved 2014-01-30.
  52. Stevens, L. C. (1970). "The development of transplantable teratocarcinomas from intratesticular grafts of pre- and postimplantation mouse embryos". Developmental Biology 21 (3): 364–382. doi:10.1016/0012-1606(70)90130-2. PMID 5436899. 
  53. Mintz, B.; Cronmiller, C.; Custer, R. P. (1978). "Somatic cell origin of teratocarcinomas". Proceedings of the National Academy of Sciences of the United States of America 75 (6): 2834–2838. doi:10.1073/pnas.75.6.2834. PMID 275854. PMC 392659. //www.ncbi.nlm.nih.gov/pmc/articles/PMC392659/. 
  54. Mintz B, Illmensee K. (1975) Normal genetically mosaic mice produced from malignant teratocarcinoma cells. Proc Natl Acad Sci U S A; 72 (9) :3585-3589
  55. MARTIN, G. R. & EVANS, M. J. (1975). Differentiation of clonal lines of teratocarcinoma cells: formation of embryoid bodies in vitro. Proc. Natn. Acad. Sci. U.S.A. 72, 1441-1445
  56. Illmensee, K.; Mintz, B. (1976). "Totipotency and normal differentiation of single teratocarcinoma cells cloned by injection into blastocysts". Proceedings of the National Academy of Sciences of the United States of America 73 (2): 549–553. doi:10.1073/pnas.73.2.549. PMID 1061157. PMC 335947. //www.ncbi.nlm.nih.gov/pmc/articles/PMC335947/. 
  57. Martin, GR (1981) Isolation of a pluripotent cell line from early mouse embryos cultured in medium conditioned by teratocarcinoma stem cells. Proc. Natl. Acad. Sci. USA 78, 7634-7638
  58. Martin, G. R. (1980). "Teratocarcinomas and mammalian embryogenesis". Science 209 (4458): 768–776. doi:10.1126/science.6250214. PMID 6250214. 
  59. Papaioannou, V. E.; Gardner, R. L.; McBurney, M. W.; Babinet, C.; Evans, M. J. (1978). "Participation of cultured teratocarcinoma cells in mouse embryogenesis". Journal of Embryology and Experimental Morphology 44: 93–104. PMID 650144. 
  60. GRAHAM, C. F. (January 1977). Teratocarcinoma cells and normal mouse embryogenesiseditor=Michael I. Sherman. MIT Press. ISBN 978-0-262-19158-6. https://books.google.com/books?id=Uyd0QgAACAAJ. 
  61. ILLMENSEE, K. (14 June 2012). L. B. Russell. ed. Reversion of malignancy and normalized differentiation of teratocarcinoma cells in chimeric mice. Springer London, Limited. pp. 3–24. ISBN 978-1-4684-3392-0. https://books.google.com/books?id=WY-ZmQEACAAJ. 
  62. Zhicheng Shao, Ruowen Zhang, Alireza Khodadadi-Jamayran, et al., & Kejin Hu (2016). The acetyllysine reader BRD3R promotes human nuclear reprogramming and regulates mitosis. Nature Communications 7, Article number: 10869 doi:10.1038/ncomms10869
  63. Stuart, Hannah T.; Van Oosten, Anouk L.; Radzisheuskaya, Aliaksandra; Martello, Graziano; Miller, Anzy; Dietmann, Sabine; Nichols, Jennifer; Silva, José C.R. (2014). "NANOG Amplifies STAT3 Activation and They Synergistically Induce the Naive Pluripotent Program". Current Biology 24 (3): 340–346. doi:10.1016/j.cub.2013.12.040. PMID 24462001. PMC 3982126. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3982126/. 
  64. Jiho Choi, Soohyun Lee, William Mallard et al.,(2015). A comparison of genetically matched cell lines reveals the equivalence of human iPSCs and ESCs. Nature Biotechnology, doi:10.1038/nbt.3388
  65. Boland, M. J.; Hazen, J. L.; Nazor, K. L.; Rodriguez, A. R.; Gifford, W.; Martin, G.; Kupriyanov, S.; Baldwin, K. K. (2009). "Adult mice generated from induced pluripotent stem cells". Nature 461 (7260): 91–94. doi:10.1038/nature08310. PMID 19672243. 
    Kang, Lan; Wang, Jianle; Zhang, Yu; Kou, Zhaohui; Gao, Shaorong (2009). "IPS Cells Can Support Full-Term Development of Tetraploid Blastocyst-Complemented Embryos". Cell Stem Cell 5 (2): 135–138. doi:10.1016/j.stem.2009.07.001. PMID 19631602. 
  66. Yagi, Takuya; Kosakai, Arifumi; Ito, Daisuke; Okada, Yohei; Akamatsu, Wado; Nihei, Yoshihiro; Nabetani, Akira; Ishikawa, Fuyuki et al. (2012). "Establishment of Induced Pluripotent Stem Cells from Centenarians for Neurodegenerative Disease Research". PLOS ONE 7 (7): e41572. doi:10.1371/journal.pone.0041572. PMID 22848530. PMC 3405135. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3405135/. 
  67. Rohani, Leili; Johnson, Adiv A.; Arnold, Antje; Stolzing, Alexandra (2014). "The aging signature: A hallmark of induced pluripotent stem cells?". Aging Cell 13 (1): 2–7. doi:10.1111/acel.12182. PMID 24256351. PMC 4326871. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4326871/. 
  68. Yehezkel, S.; Rebibo-Sabbah, A.; Segev, Y.; Tzukerman, M.; Shaked, R.; Huber, I.; Gepstein, L.; Skorecki, K. et al. (2011). "Reprogramming of telomeric regions during the generation of human induced pluripotent stem cells and subsequent differentiation into fibroblast-like derivatives". Epigenetics 6 (1): 63–75. doi:10.4161/epi.6.1.13390. PMID 20861676. PMC 3052915. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3052915/. 
  69. West, M. D.; Vaziri, H. (2010). "Back to immortality: The restoration of embryonic telomere length during induced pluripotency". Regenerative Medicine 5 (4): 485–488. doi:10.2217/rme.10.51. PMID 20632849. 
  70. Marión, R. M.; Blasco, M. A. (2010). "Telomere rejuvenation during nuclear reprogramming". Current Opinion in Genetics & Development 20 (2): 190–196. doi:10.1016/j.gde.2010.01.005. PMID 20176474. 
    Gourronc, F. A.; Klingelhutz, A. J. (2012). "Therapeutic opportunities: Telomere maintenance in inducible pluripotent stem cells". Mutation Research 730 (1–2): 98–105. doi:10.1016/j.mrfmmm.2011.05.008. PMID 21605571. PMC 3179558. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3179558/. 
  71. Zhao, Tongbiao; Zhang, Zhen-Ning; Rong, Zhili; Xu, Yang (2011). "Immunogenicity of induced pluripotent stem cells". Nature 474 (7350): 212–215. doi:10.1038/nature10135. PMID 21572395. 
  72. Dhodapkar, M. V.; Dhodapkar, K. M. (2011). "Spontaneous and therapy-induced immunity to pluripotency genes in humans: Clinical implications, opportunities and challenges". Cancer Immunology, Immunotherapy : Cii 60 (3): 413–418. doi:10.1007/s00262-010-0944-8. PMID 21104412. PMC 3574640. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3574640/. 
  73. Gutierrez‐Aranda, Ivan; Ramos‐Mejia, Veronica; Bueno, Clara; Munoz‐Lopez, Martin; Real, Pedro. J.; Mácia, Angela; Sanchez, Laura; Ligero, Gertrudis et al. (2010). "Human Induced Pluripotent Stem Cells Develop Teratoma More Efficiently and Faster Than Human Embryonic Stem Cells Regardless the Site of Injection". Stem Cells 28 (9): 1568–1570. doi:10.1002/stem.471. PMID 20641038. PMC 2996086. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2996086/. 
  74. Chang, Chan-Jung; Mitra, Koyel; Koya, Mariko; Velho, Michelle; Desprat, Romain; Lenz, Jack; Bouhassira, Eric E. (2011). "Production of Embryonic and Fetal-Like Red Blood Cells from Human Induced Pluripotent Stem Cells". PLOS ONE 6 (10): e25761. doi:10.1371/journal.pone.0025761. PMID 22022444. PMC 3192723. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3192723/. 
  75. 75.0 75.1 Dabir, Deepa V.; Hasson, Samuel A.; Setoguchi, Kiyoko; Johnson, Meghan E.; Wongkongkathep, Piriya; Douglas, Colin J.; Zimmerman, Johannes; Damoiseaux, Robert et al. (2013). "A Small Molecule Inhibitor of Redox-Regulated Protein Translocation into Mitochondria". Developmental Cell 25 (1): 81–92. doi:10.1016/j.devcel.2013.03.006. PMID 23597483. PMC 3726224. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3726224/. 
  76. Uri Ben-David, Qing-Fen Gan, Tamar Golan-Lev, et al & Nissim Benvenisty (2013)Selective Elimination of Human Pluripotent Stem Cells by an Oleate Synthesis Inhibitor Discovered in a High-Throughput Screen Cell Stem Cell, 12(2), 167-179 http://dx.doi.org/10.1016/j.stem.2012.11.015
    Lou, Kai-Jye (2013). "Small molecules vs. Teratomas". Science-Business Exchange 6 (7): 158. doi:10.1038/scibx.2013.158. 
  77. Lee, M.-O.; Moon, S. H.; Jeong, H.-C.; Yi, J.-Y.; Lee, T.-H.; Shim, S. H.; Rhee, Y.-H.; Lee, S.-H. et al. (2013). "Inhibition of pluripotent stem cell-derived teratoma formation by small molecules". Proceedings of the National Academy of Sciences 110 (35): E3281–E3290. doi:10.1073/pnas.1303669110. PMID 23918355. PMC 3761568. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3761568/. 
  78. Tang, C.; Weissman, I. L.; Drukker, M. (2012). "The safety of embryonic stem cell therapy relies on teratoma removal". Oncotarget 3 (1): 7–8. doi:10.18632/oncotarget.434. PMID 22294556. PMC 3292887. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3292887/. 
  79. Chaffer, C. L.; Brueckmann, I.; Scheel, C.; Kaestli, A. J.; Wiggins, P. A.; Rodrigues, L. O.; Brooks, M.; Reinhardt, F. et al. (2011). "Normal and neoplastic nonstem cells can spontaneously convert to a stem-like state". Proceedings of the National Academy of Sciences 108 (19): 7950–7955. doi:10.1073/pnas.1102454108. PMID 21498687. PMC 3093533. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3093533/. 
    doi:10.1089/scd.2011.0131
  80. Gurtan, Allan M.; Ravi, Arvind; Rahl, Peter B.; Bosson, Andrew D.; Jnbaptiste, Courtney K.; Bhutkar, Arjun; Whittaker, Charles A.; Young, Richard A. et al. (2013). "Let-7 represses Nr6a1 and a mid-gestation developmental program in adult fibroblasts". Genes & Development 27 (8): 941–954. doi:10.1101/gad.215376.113. PMID 23630078. PMC 3650230. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3650230/. 
    Wang, Hongran; Wang, Xiaohong; Xu, Xueping; Zwaka, Thomas P.; Cooney, Austin J. (2013). "Epigenetic Reprogramming of the Germ Cell Nuclear Factor Gene is Required for Proper Differentiation of Induced Pluripotent Cells". Stem Cells 31 (12): 2659–2666. doi:10.1002/stem.1367. PMID 23495137. PMC 3731425. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3731425/. 
  81. Yijie Geng, Yongfeng Zhao, Lisa Corinna Schuster, et al., (2015). A Chemical Biology Study of Human Pluripotent Stem Cells Unveils HSPA8 as a Key Regulator of Pluripotency. Stem Cell Reports DOI: http://dx.doi.org/10.1016/j.stemcr.2015.09.023
  82. Lindgren, Anne G.; Natsuhara, Kyle; Tian, E.; Vincent, John J.; Li, Xinmin; Jiao, Jing; Wu, Hong; Banerjee, Utpal et al. (2011). "Loss of Pten Causes Tumor Initiation Following Differentiation of Murine Pluripotent Stem Cells Due to Failed Repression of Nanog". PLOS ONE 6 (1): e16478. doi:10.1371/journal.pone.0016478. PMID 21304588. PMC 3029365. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3029365/. 
  83. Grad, I., Hibaoui, Y., Jaconi,. et al. & Feki, A. (2011) NANOG priming before full reprogramming may generate germ cell tumours. Eur. Cell Mater, 22, 258-274 Template:Pmid
  84. Okano, Hideyuki; Nakamura, Masaya; Yoshida, Kenji; Okada, Yohei; Tsuji, Osahiko; Nori, Satoshi; Ikeda, Eiji; Yamanaka, Shinya et al. (2013). "Steps Toward Safe Cell Therapy Using Induced Pluripotent Stem Cells". Circulation Research 112 (3): 523–533. doi:10.1161/CIRCRESAHA.111.256149. PMID 23371901. 
    Cunningham, Justine J.; Ulbright, Thomas M.; Pera, Martin F.; Looijenga, Leendert H J. (2012). "Lessons from human teratomas to guide development of safe stem cell therapies". Nature Biotechnology 30 (9): 849–857. doi:10.1038/nbt.2329. PMID 22965062. 
  85. Bellin, Milena; Marchetto, Maria C.; Gage, Fred H.; Mummery, Christine L. (2012). "Induced pluripotent stem cells: The new patient?". Nature Reviews Molecular Cell Biology 13 (11): 713–726. doi:10.1038/nrm3448. PMID 23034453. 
    Sandoe, Jackson; Eggan, Kevin (2013). "Opportunities and challenges of pluripotent stem cell neurodegenerative disease models". Nature Neuroscience 16 (7): 780–789. doi:10.1038/nn.3425. PMID 23799470. 
  86. Takahashi, Kazutoshi; Yamanaka, Shinya (2013). "Induced pluripotent stem cells in medicine and biology". Development 140 (12): 2457–2461. doi:10.1242/dev.092551. PMID 23715538. 
    Fu, X.; Xu, Y. (2012). "Challenges to the clinical application of pluripotent stem cells: Towards genomic and functional stability". Genome Medicine 4 (6): 55. doi:10.1186/gm354. PMID 22741526. PMC 3698533. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3698533/. 
  87. Araki, Ryoko; Uda, Masahiro; Hoki, Yuko; Sunayama, Misato; Nakamura, Miki; Ando, Shunsuke; Sugiura, Mayumi; Ideno, Hisashi et al. (2013). "Negligible immunogenicity of terminally differentiated cells derived from induced pluripotent or embryonic stem cells". Nature 494 (7435): 100–104. doi:10.1038/nature11807. PMID 23302801. 
    Wahlestedt, Martin; Norddahl, Gudmundur L.; Sten, Gerd; Ugale, Amol; Frisk, Mary-Ann Micha; Mattsson, Ragnar; Deierborg, Tomas; Sigvardsson, Mikael et al. (2013). "An epigenetic component of hematopoietic stem cell aging amenable to reprogramming into a young state". Blood 121 (21): 4257–4264. doi:10.1182/blood-2012-11-469080. PMID 23476050. https://lup.lub.lu.se/record/3628314. 
  88. Ohnishi, K., Semi, K., Yamamoto, T., Shimizu, M., Tanaka, A., Mitsunaga, K., ... & Yamada, Y. (2014). Premature Termination of Reprogramming In Vivo Leads to Cancer Development through Altered Epigenetic Regulation. Cell, 156(4), 663-677. doi:10.1016/j.cell.2014.01.005
  89. Mfopou JK, De Groote V, Xu X, Heimberg H, Bouwens L (May 2007). "Sonic hedgehog and other soluble factors from differentiating embryoid bodies inhibit pancreas development". Stem Cells 25 (5): 1156–65. doi:10.1634/stemcells.2006-0720. PMID 17272496. 
  90. Simonson, Oscar E.; Domogatskaya, Anna; Volchkov, Pavel; Rodin, Sergey (2015). "The safety of human pluripotent stem cells in clinical treatment". Annals of Medicine 47 (5): 370–380. doi:10.3109/07853890.2015.1051579. PMID 26140342. http://www.tandfonline.com/doi/full/10.3109/07853890.2015.1051579#abstract. 
  91. Kenzaburo Tani (2015). Towards the safer clinical translation of human induced pluripotent stem cell–derived cells to regenerative medicine. Molecular Therapy — Methods & Clinical Development 2, Article number: 15032 doiː10.1038/mtm.2015.32
  92. Di Stasi, A., Tey, S. K., Dotti, G., Fujita, Y., Kennedy-Nasser, A., Martinez, C., ... & Brenner, M. K. (2011). Inducible apoptosis as a safety switch for adoptive cell therapy. New England Journal of Medicine, 365(18), 1673-1683. DOIː10.1056/NEJMoa1106152
  93. Yagyu, S., Hoyos, V., Del Bufalo, F., & Brenner, M. K. (2015). An Inducible Caspase-9 Suicide Gene to Improve the Safety of Therapy Using Human Induced Pluripotent Stem Cells. Molecular Therapy. doi:10.1038/mt.2015.100
  94. Wu, C., Hong, S. G., Winkler, T., Spencer, D. M., Jares, A., Ichwan, B., ... & Dunbar, C. E. (2014). Development of an inducible caspase-9 safety switch for pluripotent stem cell–based therapies. Molecular Therapy—Methods & Clinical Development 1, Article number: 14053 doi:10.1038/mtm.2014.53
  95. Ivics, Z. (2015). Self-Destruct Genetic Switch to Safeguard iPS Cells. Molecular Therapy, 23(9), 1417-1420. doi:10.1038/mt.2015.139
  96. Miki Ando, Toshinobu Nishimura, Satoshi Yamazaki,et al., & Hiromitsu Nakauchi (2015). A Safeguard System for Induced Pluripotent Stem Cell-Derived Rejuvenated T Cell Therapy. Stem Cell Reports. DOI: http://dx.doi.org/10.1016/j.stemcr.2015.07.011
  97. de Almeida, P. E., Meyer, E. H., Kooreman, N. G., Diecke, S., Dey, D., Sanchez-Freire, V., ... & Wu, J. C. (2014). Transplanted terminally differentiated induced pluripotent stem cells are accepted by immune mechanisms similar to self-tolerance. Nature Communications, 5.Article number: 3903 doi:10.1038/ncomms4903
  98. Inoue, H.; Nagata, N.; Kurokawa, H.; Yamanaka, S. (2014). "IPS cells: A game changer for future medicine". The EMBO Journal 33 (5): 409–417. doi:10.1002/embj.201387098. PMID 24500035. PMC 3989624. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3989624/. 
  99. Kejin Hu.(March, 2014). Vectorology and factor delivery in induced pluripotent stem cell reprogramming. Stem Cells and Development.doi:10.1089/scd.2013.0621
  100. Kejin Hu . (2014). All Roads Lead to Induced Pluripotent Stem Cells: The Technologies of iPSC Generation. Stem Cells and Development. doi:10.1089/scd.2013.0620.
  101. Tang, C., Weissman, I. L., & Drukker, M. (2013). Immunogenicity of in vitro maintained and matured populations: potential barriers to engraftment of human pluripotent stem cell derivatives. In Embryonic Stem Cell Immunobiology (pp. 17-31). Humana Press. doi:10.1007/978-1-62703-478-4_2
  102. Menasché, P., Vanneaux, V., Fabreguettes, J. R., et al., & Larghero, J. (2014). Towards a clinical use of human embryonic stem cell-derived cardiac progenitors: a translational experience. European heart journal, ehu192. doi:10.1093/eurheartj/ehu192
  103. Baghbaderani, B. A., Tian, X., Neo, B. H., Burkall, A., Dimezzo, T., Sierra, G., ... & Rao, M. S. (2015). cGMP-Manufactured Human Induced Pluripotent Stem Cells Are Available for Pre-clinical and Clinical Applications. Stem cell reports. 5(4), 647–659, DOI: http://dx.doi.org/10.1016/j.stemcr.2015.08.015
  104. Sergio Ruiz et al. & Oscar Fernandez-Capetillo (2015). Limiting replication stress during somatic cell reprogramming reduces genomic instability in induced pluripotent stem cells. Nature Communications 6, Article number: 8036 doi:10.1038/ncomms9036
  105. De Los Angeles, A., & Daley, G. Q. (2013) A chemical logic for reprogramming to pluripotency doi:10.1038/cr.2013.119
    Federation, A. J., Bradner, J. E., & Meissner, A. (2013) The use of small molecules in somatic-cell reprogramming. Trends in cell biology. doi:10.1016/j.tcb.2013.09.011
  106. Kang, P. J., Moon, J. H., Yoon, B. S., Hyeon, S., Jun, E. K., Park, G., ... & You, S. (2014). Reprogramming of mouse somatic cells into pluripotent stem-like cells using a combination of small molecules. Biomaterials. doi:10.1016/j.biomaterials.2014.05.015
  107. Lee, K., Kwon, D. N., Ezashi, T., et al., & Kim, J. H. (2014). Engraftment of human iPS cells and allogeneic porcine cells into pigs with inactivated RAG2 and accompanying severe combined immunodeficiency. PNAS, 111(20), 7260-7265. doi:10.1073/pnas.1406376111
  108. NAKAUCHI Hiromitsu, KAMIYA Akihide, SUZUKI Nao, ITO Keiichi, YAMAZAKI Satoshi (2011) METHOD FOR PRODUCING CELLS INDUCED TO DIFFERENTIATE FROM PLURIPOTENT STEM CELLS PATENT COOPERATION TREATY APPLICATION, patno: WO2011071085 (A1) ― 2011-06-16 (C12N5/07)
  109. Amabile, Giovanni; Welner, Robert S.; Nombela-Arrieta, Cesar; d'Alise, Anna Morena; Di Ruscio, Annalisa; Ebralidze, Alexander K.; Kraytsberg, Yevgenya; Ye, Min et al. (2013). "In vivo generation of transplantable human hematopoietic cells from induced pluripotent stem cells". Blood 121 (8): 1255–1264. doi:10.1182/blood-2012-06-434407. PMID 23212524. PMC 3701251. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3701251/. 
  110. Suzuki, Nao; Yamazaki, Satoshi; Yamaguchi, Tomoyuki; Okabe, Motohito; Masaki, Hideki; Takaki, Satoshi; Otsu, Makoto; Nakauchi, Hiromitsu (2013). "Generation of Engraftable Hematopoietic Stem Cells from Induced Pluripotent Stem Cells by Way of Teratoma Formation". Molecular Therapy 21 (7): 1424–1431. doi:10.1038/mt.2013.71. PMID 23670574. PMC 3705943. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3705943/. 
  111. Chou, Bin-Kuan; Ye, Zhaohui; Cheng, Linzhao (2013). "Generation and Homing of iPSC-Derived Hematopoietic Cells in Vivo". Molecular Therapy 21 (7): 1292–1293. doi:10.1038/mt.2013.129. PMID 23812546. PMC 3702102. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3702102/. 
  112. Yamauchi, T.; Takenaka, K.; Urata, S.; Shima, T.; Kikushige, Y.; Tokuyama, T.; Iwamoto, C.; Nishihara, M. et al. (2013). "Polymorphic Sirpa is the genetic determinant for NOD-based mouse lines to achieve efficient human cell engraftment". Blood 121 (8): 1316–1325. doi:10.1182/blood-2012-06-440354. PMID 23293079. 
  113. Rong, Zhili; Wang, Meiyan; Hu, Zheng; Stradner, Martin; Zhu, Shengyun; Kong, Huijuan; Yi, Huanfa; Goldrath, Ananda et al. (2014). "An Effective Approach to Prevent Immune Rejection of Human ESC-Derived Allografts". Cell Stem Cell 14 (1): 121–130. doi:10.1016/j.stem.2013.11.014. PMID 24388175. PMC 4023958. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4023958/. 
  114. Forster, R., Chiba, K., Schaeffer, L., Regalado, S. G., Lai, C. S., Gao, Q., ... & Hockemeyer, D. (2014). Human Intestinal Tissue with Adult Stem Cell Properties Derived from Pluripotent Stem Cells. Stem Cell Reports, 2(6), 838-852. http://dx.doi.org/10.1016/j.stemcr.2014.05.001
  115. Hirami, Yasuhiko; Osakada, Fumitaka; Takahashi, Kazutoshi; Okita, Keisuke; Yamanaka, Shinya; Ikeda, Hanako; Yoshimura, Nagahisa; Takahashi, Masayo (2009). "Generation of retinal cells from mouse and human induced pluripotent stem cells". Neuroscience Letters 458 (3): 126–131. doi:10.1016/j.neulet.2009.04.035. PMID 19379795. 
  116. Buchholz, David E.; Hikita, Sherry T.; Rowland, Teisha J.; Friedrich, Amy M.; Hinman, Cassidy R.; Johnson, Lincoln V.; Clegg, Dennis O. (2009). "Derivation of Functional Retinal Pigmented Epithelium from Induced Pluripotent Stem Cells". Stem Cells 27 (10): 2427–2434. doi:10.1002/stem.189. PMID 19658190. 
  117. Jin Yang, Eva Nong, Stephen H Tsang (2013) Induced pluripotent stem cells and retinal degeneration treatment. Expert Rev. Ophthalmol. 8(1), 5–8 doi: 10.1586/EOP.12.75
    Fields, Mark A.; Hwang, John; Gong, Jie; Cai, Hui; Del Priore, Lucian (9 December 2012). Stephen Tsang. ed. The Eye as a Target Organ for Stem Cell Therapy. Springer. pp. 1–30. ISBN 978-1-4614-5493-9. https://books.google.com/books?id=2jMCbFX7bngC. 
  118. Li Y, Tsai YT, Hsu CW et al. (2012) Long-term safety and efficacy of human induced pluripotent stem cell (iPS) grafts in a preclinical model of retinitis pigmentosa Mol. Med. 18(1), 1312–1319 doi: 10.2119/molmed.2012.00242
  119. Stem cell therapy for RP is now offered at St. Luke’s Medical Center.
  120. Tzouvelekis, Argyris; Ntolios, Paschalis; Bouros, Demosthenes (2013). "Stem Cell Treatment for Chronic Lung Diseases". Respiration 85 (3): 179–192. doi:10.1159/000346525. PMID 23364286. 
  121. Wagner, Darcy E.; Bonvillain, Ryan W.; Jensen, Todd; Girard, Eric D.; Bunnell, Bruce A.; Finck, Christine M.; Hoffman, Andrew M.; Weiss, Daniel J. (2013). "Can stem cells be used to generate new lungs?Ex vivolung bioengineering with decellularized whole lung scaffolds". Respirology 18 (6): 895–911. doi:10.1111/resp.12102. PMID 23614471. PMC 3729745. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3729745/. 
  122. Firth AL, et al. (2014) Generation of multiciliated cells in functional airway epithelia from human induced pluripotent stem cells. Proc Natl Acad Sci USA 111(17), E1723–E1730. doi:10.1073/pnas.1403470111
  123. Wong, A. P., & Rossant, J. (2013) Generation of Lung Epithelium from Pluripotent Stem Cells. Current pathobiology reports, 1(2), 137-145, DOI: 10.1007/s40139-013-0016-9
    Mou, Hongmei; Zhao, Rui; Sherwood, Richard; Ahfeldt, Tim; Lapey, Allen; Wain, John; Sicilian, Leonard; Izvolsky, Konstantin et al. (2012). "Generation of Multipotent Lung and Airway Progenitors from Mouse ESCs and Patient-Specific Cystic Fibrosis iPSCs". Cell Stem Cell 10 (4): 385–397. doi:10.1016/j.stem.2012.01.018. PMID 22482504. PMC 3474327. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3474327/. 
    Ghaedi, Mahboobe; Calle, Elizabeth A.; Mendez, Julio J.; Gard, Ashley L.; Balestrini, Jenna; Booth, Adam; Bove, Peter F.; Gui, Liqiong et al. (2013). "Human iPS cellâ€"derived alveolar epithelium repopulates lung extracellular matrix". Journal of Clinical Investigation 123 (11): 4950–4962. doi:10.1172/JCI68793. PMID 24135142. PMC 3809786. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3809786/. 
    Ghaedi, Mahboobe; Mendez, Julio J.; Bove, Peter F.; Sivarapatna, Amogh; Raredon, Micha Sam B.; Niklason, Laura E. (2014). "Alveolar epithelial differentiation of human induced pluripotent stem cells in a rotating bioreactor". Biomaterials 35 (2): 699–710. doi:10.1016/j.biomaterials.2013.10.018. PMID 24144903. PMC 3897000. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3897000/. 
    Huang, Sarah X L.; Islam, Mohammad Naimul; O'Neill, John; Hu, Zheng; Yang, Yong-Guang; Chen, Ya-Wen; Mumau, Melanie; Green, Michael D. et al. (2014). "Efficient generation of lung and airway epithelial cells from human pluripotent stem cells". Nature Biotechnology 32 (1): 84–91. doi:10.1038/nbt.2754. PMID 24291815. PMC 4101921. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4101921/. 
  124. Niu, Zhiwei; Hu, Yue; Chu, Zhili; Yu, Meng; Bai, Yaofu; Wang, Long; Hua, Jinlian (2013). "Germ-like cell differentiation from induced pluripotent stem cells (IPSCs)". Cell Biochemistry and Function 31 (1): 12–19. doi:10.1002/cbf.2924. PMID 23086862. 
    Yang, S.; Bo, J.; Hu, H.; Guo, X.; Tian, R.; Sun, C.; Zhu, Y.; Li, P. et al. (2012). "Derivation of male germ cells from induced pluripotent stem cells in vitro and in reconstituted seminiferous tubules". Cell Proliferation 45 (2): 91–100. doi:10.1111/j.1365-2184.2012.00811.x. PMID 22324506. PMC 6496715. //www.ncbi.nlm.nih.gov/pmc/articles/PMC6496715/. 
    Panula, Sarita; Medrano, Jose V.; Kee, Kehkooi; Bergström, Rosita; Nguyen, Ha Nam; Byers, Blake; Wilson, Kitchener D.; Wu, Joseph C. et al. (2011). "Human germ cell differentiation from fetal- and adult-derived induced pluripotent stem cells". Human Molecular Genetics 20 (4): 752–762. doi:10.1093/hmg/ddq520. PMID 21131292. PMC 3024045. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3024045/. 
  125. New Stem Cell Identified
  126. Jun Wu, Daiji Okamura, Mo Li,et al., & Juan Carlos Izpisua Belmonte (2015). An alternative pluripotent state confers interspecies chimaeric competency. Nature, doi:10.1038/nature14413
  127. New Type of Stem Cell Could Make It Easier to Grow Human Orgachimaeras
  128. 128.0 128.1 128.2 Qian, Li; Huang, Yu; Spencer, C. Ian; Foley, Amy; Vedantham, Vasanth; Liu, Lei; Conway, Simon J.; Fu, Ji-Dong et al. (2012). "In vivo reprogramming of murine cardiac fibroblasts into induced cardiomyocytes". Nature 485 (7400): 593–598. doi:10.1038/nature11044. PMID 22522929. PMC 3369107. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3369107/. 
  129. Szabo, Eva; Rampalli, Shravanti; Risueño, Ruth M.; Schnerch, Angelique; Mitchell, Ryan; Fiebig-Comyn, Aline; Levadoux-Martin, Marilyne; Bhatia, Mickie (2010). "Direct conversion of human fibroblasts to multilineage blood progenitors". Nature 468 (7323): 521–526. doi:10.1038/nature09591. PMID 21057492. 
  130. Efe, Jem A.; Hilcove, Simon; Kim, Janghwan; Zhou, Hongyan; Ouyang, Kunfu; Wang, Gang; Chen, Ju; Ding, Sheng (2011). "Conversion of mouse fibroblasts into cardiomyocytes using a direct reprogramming strategy". Nature Cell Biology 13 (3): 215–222. doi:10.1038/ncb2164. PMID 21278734. 
  131. 131.0 131.1 Lujan, E.; Chanda, S.; Ahlenius, H.; Sudhof, T. C.; Wernig, M. (2012). "Direct conversion of mouse fibroblasts to self-renewing, tripotent neural precursor cells". Proceedings of the National Academy of Sciences 109 (7): 2527–2532. doi:10.1073/pnas.1121003109. PMID 22308465. PMC 3289376. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3289376/. 
  132. 132.0 132.1 Thier, Marc; Wörsdörfer, Philipp; Lakes, Yenal B.; Gorris, Raphaela; Herms, Stefan; Opitz, Thoralf; Seiferling, Dominic; Quandel, Tamara et al. (2012). "Direct Conversion of Fibroblasts into Stably Expandable Neural Stem Cells". Cell Stem Cell 10 (4): 473–479. doi:10.1016/j.stem.2012.03.003. PMID 22445518. 
  133. 133.0 133.1 133.2 Han, Dong Wook; Tapia, Natalia; Hermann, Andreas; Hemmer, Kathrin; Höing, Susanne; Araúzo-Bravo, Marcos J.; Zaehres, Holm; Wu, Guangming et al. (2012). "Direct Reprogramming of Fibroblasts into Neural Stem Cells by Defined Factors". Cell Stem Cell 10 (4): 465–472. doi:10.1016/j.stem.2012.02.021. PMID 22445517. 
  134. Taylor, S. M.; Jones, P. A. (1979). "Multiple new phenotypes induced in 10T1/2 and 3T3 cells treated with 5-azacytidine". Cell 17 (4): 771–779. doi:10.1016/0092-8674(79)90317-9. PMID 90553. 
  135. Lassar, A. B.; Paterson, B. M.; Weintraub, H. (1986). "Transfection of a DNA locus that mediates the conversion of 10T1/2 fibroblasts to myoblasts". Cell 47 (5): 649–656. doi:10.1016/0092-8674(86)90507-6. PMID 2430720. 
    Davis, R. L.; Weintraub, H.; Lassar, A. B. (1987). "Expression of a single transfected cDNA converts fibroblasts to myoblasts". Cell 51 (6): 987–1000. doi:10.1016/0092-8674(87)90585-x. PMID 3690668. 
    Weintraub, H.; Tapscott, S. J.; Davis, R. L.; Thayer, M. J.; Adam, M. A.; Lassar, A. B.; Miller, A. D. (1989). "Activation of muscle-specific genes in pigment, nerve, fat, liver, and fibroblast cell lines by forced expression of MyoD". Proceedings of the National Academy of Sciences of the United States of America 86 (14): 5434–5438. doi:10.1073/pnas.86.14.5434. PMID 2748593. PMC 297637. //www.ncbi.nlm.nih.gov/pmc/articles/PMC297637/. 
  136. Vierbuchen, Thomas; Wernig, Marius (2011). "Direct lineage conversions: Unnatural but useful?". Nature Biotechnology 29 (10): 892–907. doi:10.1038/nbt.1946. PMID 21997635. PMC 3222779. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3222779/. 
  137. Riddle, Misty R.; Weintraub, Abraham; Nguyen, Ken C. Q.; Hall, David H.; Rothman, Joel H. (2013). "Transdifferentiation and remodeling of post-embryonic C. Elegans cells by a single transcription factor". Development 140 (24): 4844–4849. doi:10.1242/dev.103010. PMID 24257624. PMC 3848185. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3848185/. 
  138. Mapping out cell conversion
  139. Owen, Rackham; Gough, Julian (2016). "A predictive computational framework for direct reprogramming between human cell types". Nature Genetics 48 (3): 331–335. doi:10.1038/ng.3487. PMID 26780608. https://research-information.bris.ac.uk/en/publications/e6490a78-f3e8-4253-acc4-7ee181c79168. 
  140. New Algorithm May Someday Enable Scientists to Regrow Limbs and Replace Damaged Organs
  141. Jung, Da-Woon; Williams, Darren R. (2011). "Novel Chemically Defined Approach to Produce Multipotent Cells from Terminally Differentiated Tissue Syncytia". ACS Chemical Biology 6 (6): 553–562. doi:10.1021/cb2000154. PMID 21322636. 
    "In This Issue". ACS Chemical Biology 7 (4): 619. 2012. doi:10.1021/cb300127f. 
  142. Hongkai Zhang, Ian A. Wilson, and Richard A. Lerner (2012) Selection of antibodies that regulate phenotype from intracellular combinatorial antibody libraries. PNAS. 109(39), 15728-15733 doi:10.1073/pnas.1214275109
  143. Antibody that Transforms Bone Marrow Stem Cells Directly into Brain Cells
    Jia Xie, Hongkai Zhang, Kyungmoo Yea, and Richard A. Lerner (2013) Autocrine signaling based selection of combinatorial antibodies that transdifferentiate human stem cells PNAS; doi:10.1073/pnas.1306263110
  144. 144.0 144.1 144.2 144.3 Liu, Xuefeng; Ory, Virginie; Chapman, Sandra; Yuan, Hang; Albanese, Chris; Kallakury, Bhaskar; Timofeeva, Olga A.; Nealon, Caitlin et al. (2012). "ROCK Inhibitor and Feeder Cells Induce the Conditional Reprogramming of Epithelial Cells". The American Journal of Pathology 180 (2): 599–607. doi:10.1016/j.ajpath.2011.10.036. PMID 22189618. PMC 3349876. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3349876/. 
  145. Rheinwald, J. G.; Green, H. (1975). "Serial cultivation of strains of human epidermal keratinocytes: The formation of keratinizing colonies from single cells". Cell 6 (3): 331–343. doi:10.1016/s0092-8674(75)80001-8. PMID 1052771. 
  146. Hiew, Y.-L. (2011) Examining the biological consequences of DNA damage caused by irradiated J2-3T3 fibroblast feeder cells and HPV16: characterisation of the biological functions of Mll. Doctoral thesis, UCL (University College London)
  147. Szumiel, Irena (2012). "Radiation hormesis: Autophagy and other cellular mechanisms". International Journal of Radiation Biology 88 (9): 619–628. doi:10.3109/09553002.2012.699698. PMID 22702489. 
  148. Kurosawa, Hiroshi (2012). "Application of Rho-associated protein kinase (ROCK) inhibitor to human pluripotent stem cells". Journal of Bioscience and Bioengineering 114 (6): 577–581. doi:10.1016/j.jbiosc.2012.07.013. PMID 22898436. 
  149. Terunuma, Atsushi; Limgala, Renuka Pudi; Park, Christine J.; Choudhary, Isha; Vogel, Jonathan C. (2010). "Efficient Procurement of Epithelial Stem Cells from Human Tissue Specimens Using a Rho-Associated Protein Kinase Inhibitor Y-27632". Tissue Engineering Part A 16 (4): 1363–1368. doi:10.1089/ten.tea.2009.0339. PMID 19912046. PMC 2862604. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2862604/. 
  150. Sandra Chapman, Xuefeng Liu, Craig Meyers, Richard Schlegel, and Alison A. McBride. ( 2010) Human keratinocytes are efficiently immortalized by a Rho kinase inhibitor
  151. Suprynowicz, F. A.; Upadhyay, G.; Krawczyk, E.; Kramer, S. C.; Hebert, J. D.; Liu, X.; Yuan, H.; Cheluvaraju, C. et al. (2012). "Conditionally reprogrammed cells represent a stem-like state of adult epithelial cells". Proceedings of the National Academy of Sciences 109 (49): 20035–20040. doi:10.1073/pnas.1213241109. PMID 23169653. PMC 3523865. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3523865/. 
  152. 152.0 152.1 Agarwal, Seema; Rimm, David L. (2012). "Making Every Cell Like He La". The American Journal of Pathology 180 (2): 443–445. doi:10.1016/j.ajpath.2011.12.001. PMID 22192626. 
  153. 153.0 153.1 Lisanti MP, Tanowitz HB. (2012) Translational discoveries, personalized medicine, and living biobanks of the future. The American Journal of Pathology, 2012 Apr;180(4):1334-6
  154. Yuan, Hang; Myers, Scott; Wang, Jingang; Zhou, Dan; Woo, Jennifer A.; Kallakury, Bhaskar; Ju, Andrew; Bazylewicz, Michael et al. (2012). "Use of Reprogrammed Cells to Identify Therapy for Respiratory Papillomatosis". New England Journal of Medicine 367 (13): 1220–1227. doi:10.1056/NEJMoa1203055. PMID 23013073. PMC 4030597. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4030597/. 
  155. Palechor-Ceron, Nancy; Suprynowicz, Frank A.; Upadhyay, Geeta; Dakic, Aleksandra; Minas, Tsion; Simic, Vera; Johnson, Michael; Albanese, Christopher et al. (2013). "Radiation Induces Diffusible Feeder Cell Factor(s) That Cooperate with ROCK Inhibitor to Conditionally Reprogram and Immortalize Epithelial Cells". The American Journal of Pathology 183 (6): 1862–1870. doi:10.1016/j.ajpath.2013.08.009. PMID 24096078. PMC 5745544. //www.ncbi.nlm.nih.gov/pmc/articles/PMC5745544/. 
  156. Sukhbir Kaur, David R. Soto-Pantoja, Erica V. Stein et al. & David D. Roberts.( 2013) Thrombospondin-1 Signaling through CD47 Inhibits Self-renewal by Regulating c-Myc and Other Stem Cell Transcription Factors. Scientific Reports; 3, Article number: 1673 DOI:10.1038/srep01673
  157. Soto-Pantoja, David R.; Ridnour, Lisa A.; Wink, David A.; Roberts, David D. (2013). "Blockade of CD47 increases survival of mice exposed to lethal total body irradiation". Scientific Reports 3: 1038. doi:10.1038/srep01038. PMID 23301159. PMC 3539147. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3539147/. 
  158. 158.0 158.1 Leo Kurian, Ignacio Sancho-Martinez, Emmanuel Nivet, et al. & Juan Carlos Izpisua Belmonte (2012) Conversion of human fibroblasts to angioblast-like progenitor cells. Nature Methods. doi:10.1038/nmeth.2255
  159. Morris, S. A., & Daley, G. Q. (2013). A blueprint for engineering cell fate: current technologies to reprogram cell identity. Cell research, 23(1), 33-48. doi:10.1038/cr.2013.1
  160. Wang, Yu-Chieh; Nakagawa, Masato; Garitaonandia, Ibon; Slavin, Ileana; Altun, Gulsah; Lacharite, Robert M.; Nazor, Kristopher L.; Tran, Ha T. et al. (2011). "Specific lectin biomarkers for isolation of human pluripotent stem cells identified through array-based glycomic analysis". Cell Research 21 (11): 1551–1563. doi:10.1038/cr.2011.148. PMID 21894191. PMC 3364725. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3364725/. 
  161. Thomson, James A.; Itskovitz-Eldor, Joseph; Shapiro, Sander S.; Waknitz, Michelle A.; Swiergiel, Jennifer J.; Marshall, Vivienne S.; Jones, Jeffrey M. (1998). "Embryonic Stem Cell Lines Derived from Human Blastocysts". Science 282 (5391): 1145–1147. doi:10.1126/science.282.5391.1145. PMID 9804556. 
  162. Suila, Heli; Hirvonen, Tia; Ritamo, Ilja; Natunen, Suvi; Tuimala, Jarno; Laitinen, Saara; Anderson, Heidi; Nystedt, Johanna et al. (2014). "Extracellular O-Linked N-Acetylglucosamine is Enriched in Stem Cells Derived from Human Umbilical Cord Blood". Bioresearch Open Access 3 (2): 39–44. doi:10.1089/biores.2013.0050. PMID 24804163. PMC 3995142. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3995142/. 
  163. Perdigoto, C. N., & Bardin, A. J. (2013). Sending the right signal: Notch and stem cells. Biochimica et Biophysica Acta (BBA)-General Subjects, 1830(2), 2307-2322. http://dx.doi.org/10.1016/j.bbagen.2012.08.009
  164. Jafar-Nejad, H., Leonardi, J., & Fernandez-Valdivia, R. (2010). Role of glycans and glycosyltransferases in the regulation of Notch signaling. Glycobiology, 20(8), 931-949. doi:10.1093/glycob/cwq053
  165. Frederico Alisson-Silva, Deivid de Carvalho Rodrigues, Leandro Vairo, et al. and Adriane R Todeschini (2014). Evidences for the involvement of cell surface glycans in stem cell pluripotency and differentiation. Glycobiology 24 (5): 458-468. doi:10.1093/glycob/cwu012
  166. Hasehira, Kayo; Tateno, Hiroaki; Onuma, Yasuko; Ito, Yuzuru; Asashima, Makoto; Hirabayashi, Jun (2012). "Structural and Quantitative Evidence for Dynamic Glycome Shift on Production of Induced Pluripotent Stem Cells". Molecular & Cellular Proteomics 11 (12): 1913–1923. doi:10.1074/mcp.M112.020586. PMID 23023295. PMC 3518133. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3518133/. 
  167. Becker Kojić, Zorica A.; Terness, Peter (2002). "A Novel Human Erythrocyte Glycosylphosphatidylinositol (GPI)-anchored Glycoprotein ACA". Journal of Biological Chemistry 277 (43): 40472–40478. doi:10.1074/jbc.M202416200. PMID 12167612. 
  168. Becker-Kojić, Z. A.; Ureña-Peralta, J. R.; Saffrich, R.; Rodriguez-Jiménez, F. J.; Rubio, M. P.; Rios, P.; Romero, A.; Ho, A. D. et al. (2013). "A novel human glycoprotein ACA is an upstream regulator of human hematopoiesis". Bulletin of Experimental Biology and Medicine 155 (4): 536–551. doi:10.1007/s10517-013-2195-0. PMID 24143385. 
  169. ZABecker-Kojič, JRUreña-Peralta, I.Zipančić, et al. & M.Stojkovič ( 2013 ) Activation of surface glycoprotein ACA induced pluripotent hematopoietic progenitor cells. CELL TECHNOLOGIES IN BIOLOGY AND MEDICINE , 9 (2 ) , 85-101
  170. Mikkola, M. (2013) Human pluripotent stem cells: glycomic approaches for culturing and characterization.978-952-10-8444-7
  171. Redmer, Torben; Diecke, Sebastian; Grigoryan, Tamara; Quiroga‐Negreira, Angel; Birchmeier, Walter; Besser, Daniel (2011). "E‐cadherin is crucial for embryonic stem cell pluripotency and can replace OCT4 during somatic cell reprogramming". EMBO Reports 12 (7): 720–726. doi:10.1038/embor.2011.88. PMID 21617704. PMC 3128971. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3128971/. 
  172. Bedzhov, Ivan; Alotaibi, Hani; Basilicata, M. Felicia; Ahlborn, Kerstin; Liszewska, Ewa; Brabletz, Thomas; Stemmler, Marc P. (2013). "Adhesion, but not a specific cadherin code, is indispensable for ES cell and induced pluripotency". Stem Cell Research 11 (3): 1250–1263. doi:10.1016/j.scr.2013.08.009. PMID 24036274. 
  173. Su, Guannan; Zhao, Yannan; Wei, Jianshu; Xiao, Zhifeng; Chen, Bing; Han, Jin; Chen, Lei; Guan, Jian et al. (2013). "Direct conversion of fibroblasts into neural progenitor-like cells by forced growth into 3D spheres on low attachment surfaces". Biomaterials 34 (24): 5897–5906. doi:10.1016/j.biomaterials.2013.04.040. PMID 23680365. 
  174. Downing, Timothy L.; Soto, Jennifer; Morez, Constant; Houssin, Timothee; Fritz, Ashley; Yuan, Falei; Chu, Julia; Patel, Shyam et al. (2013). "Biophysical regulation of epigenetic state and cell reprogramming". Nature Materials 12 (12): 1154–1162. doi:10.1038/nmat3777. PMID 24141451. 
  175. Yubing Sun, Koh Meng Aw Yong, Luis G. Villa-Diaz, et al. & Jianping Fu(2014). Hippo/YAP-mediated rigidity-dependent motor neuron differentiation of human pluripotent stem cells. Nature Materials doi:10.1038/nmat3945
  176. Bae Y. H., Mui K. L., Hsu B. Y., and Assoian R.K. (2014). A FAK-Cas-Rac-Lamellipodin Signaling Module Transduces Extracellular Matrix Stiffness into Mechanosensitive Cell Cycling. Sci. Signal. 7(330), ra57 doi: 10.1126/scisignal.2004838
  177. Murray, Patricia; Prewitz, Marina; Hopp, Isabel; Wells, Nicola; Zhang, Haifei; Cooper, Andrew; Parry, Kristina L.; Short, Robert et al. (2013). "The self-renewal of mouse embryonic stem cells is regulated by cell–substratum adhesion and cell spreading". The International Journal of Biochemistry & Cell Biology 45 (11): 2698–2705. doi:10.1016/j.biocel.2013.07.001. PMID 23871934. PMC 3898852. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3898852/. 
  178. Guilak, F., Cohen, D. M., Estes, B. T., et al. & Chen, C. S. (2009) Control of stem cell fate by physical interactions with the extracellular matrix. Cell stem cell, 5(1), 17-26. doi: 10.1016/j.stem.2009.06.016
    Worley, Kathryn; Certo, Anthony; Wan, Leo Q. (2013). "Geometry–Force Control of Stem Cell Fate". Bionanoscience 3: 43–51. doi:10.1007/s12668-012-0067-0. 
  179. Qiushui Chen, Jing Wu, Qichen Zhuang, Xuexia Lin, Jie Zhang & Jin-Ming Lin (2013). Microfluidic isolation of highly pure embryonic stem cells using feeder-separated co-culture system. Scientific Reports 3, Article number: 2433 doi:10.1038/srep02433
  180. Singh, Ankur; Suri, Shalu; Lee, Ted; Chilton, Jamie M.; Cooke, Marissa T.; Chen, Weiqiang; Fu, Jianping; Stice, Steven L. et al. (2013). "Adhesion strength–based, label-free isolation of human pluripotent stem cells". Nature Methods 10 (5): 438–444. doi:10.1038/nmeth.2437. PMID 23563795. PMC 3641175. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3641175/. 
  181. Ertl, P., Sticker, D., Charwat, V., Kasper, C., & Lepperdinger, G. (2014). Lab-on-a-chip technologies for stem cell analysis. Trends in biotechnology, 32(5), 245-253.http://dx.doi.org/10.1016/j.tibtech.2014.03.004
  182. Horna, D., Ramírez, J. C., Cifuentes, A., Bernad, A., Borrós, S. and González, M. A. (2012), Efficient Cell Reprogramming Using Bioengineered Surfaces. Advanced Healthcare Materials, 1: 177–182. doi:10.1002/adhm.201200017
  183. A Spanish device produces packaged batches of stem cells for regenerative medicine
  184. Wang, Kainan; Degerny, Cindy; Xu, Minghong; Yang, Xiang-Jiao (2009). YAP, TAZ, and Yorkie: A conserved family of signal-responsive transcriptional coregulators in animal development and human disease. Biochemistry and Cell Biology 87 (1): 77–91. doi:10.1139/O08-114
  185. Yang C., Tibbitt M.W., Basta L. & Anseth K.S.(2014). Mechanical memory and dosing influence stem cell fate. Nature Materials, doi:10.1038/nmat3889
  186. Nampe, D., & Tsutsui, H. (2013). Engineered Micromechanical Cues Affecting Human Pluripotent Stem Cell Regulations and Fate. Journal of laboratory automation, 18(6), 482-493. doi:10.1177/2211068213503156
  187. Zhang, Weiqi; Duan, Shunlei; Li, Ying; Xu, Xiuling; Qu, Jing; Zhang, Weizhou; Liu, Guang-Hui (2012). "Converted neural cells: Induced to a cure?". Protein & Cell 3 (2): 91–97. doi:10.1007/s13238-012-2029-2. PMID 22410787. PMC 4104580. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4104580/. 
  188. Yang, Nan; Ng, Yi Han; Pang, Zhiping P.; Südhof, Thomas C.; Wernig, Marius (2011). "Induced Neuronal Cells: How to Make and Define a Neuron". Cell Stem Cell 9 (6): 517–525. doi:10.1016/j.stem.2011.11.015. PMID 22136927. PMC 4377331. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4377331/. 
  189. Sheng, Chao; Zheng, Qinyuan; Wu, Jianyu; Xu, Zhen; Sang, Lisi; Wang, Libin; Guo, Changlong; Zhu, Wanwan et al. (2012). "Generation of dopaminergic neurons directly from mouse fibroblasts and fibroblast-derived neural progenitors". Cell Research 22 (4): 769–772. doi:10.1038/cr.2012.32. PMID 22370632. PMC 3317566. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3317566/. 
  190. 190.0 190.1 Maucksch, C., E. Firmin, et al. (2012). "Non-viral generation of neural precursor-like cells from adult human fibroblasts" J Stem Cells Regen Med 8(3): 1-9.
  191. Ring, Karen L.; Tong, Leslie M.; Balestra, Maureen E.; Javier, Robyn; Andrews-Zwilling, Yaisa; Li, Gang; Walker, David; Zhang, William R. et al. (2012). "Direct Reprogramming of Mouse and Human Fibroblasts into Multipotent Neural Stem Cells with a Single Factor". Cell Stem Cell 11 (1): 100–109. doi:10.1016/j.stem.2012.05.018. PMID 22683203. PMC 3399516. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3399516/. 
  192. 192.0 192.1 Generation of neural progenitor cells by chemical cocktails and hypoxia.Cell Research, 24:665–679 doi:10.1038/cr.2014.32
  193. Liu, Guang-Hui; Yi, Fei; Suzuki, Keiichiro; Qu, Jing; Belmonte, Juan Carlos Izpisua (2012). "Induced neural stem cells: A new tool for studying neural development and neurological disorders". Cell Research 22 (7): 1087–1091. doi:10.1038/cr.2012.73. PMID 22547025. PMC 3391017. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3391017/. 
  194. Torper, O.; Pfisterer, U.; Wolf, D. A.; Pereira, M.; Lau, S.; Jakobsson, J.; Bjorklund, A.; Grealish, S. et al. (2013). "Generation of induced neurons via direct conversion in vivo". Proceedings of the National Academy of Sciences 110 (17): 7038–7043. doi:10.1073/pnas.1303829110. PMID 23530235. PMC 3637783. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3637783/. 
  195. Niu, Wenze; Zang, Tong; Zou, Yuhua; Fang, Sanhua; Smith, Derek K.; Bachoo, Robert; Zhang, Chun-Li (2013). "In vivo reprogramming of astrocytes to neuroblasts in the adult brain". Nature Cell Biology 15 (10): 1164–1175. doi:10.1038/ncb2843. PMID 24056302. PMC 3867822. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3867822/. 
  196. Su, Zhida; Niu, Wenze; Liu, Meng-Lu; Zou, Yuhua; Zhang, Chun-Li (2014). "In vivo conversion of astrocytes to neurons in the injured adult spinal cord". Nature Communications 5: 3338. doi:10.1038/ncomms4338. PMID 24569435. PMC 3966078. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3966078/. 
  197. Najm, Fadi J.; Lager, Angela M.; Zaremba, Anita; Wyatt, Krysta; Caprariello, Andrew V.; Factor, Daniel C.; Karl, Robert T.; Maeda, Tadao et al. (2013). "Transcription factor–mediated reprogramming of fibroblasts to expandable, myelinogenic oligodendrocyte progenitor cells". Nature Biotechnology 31 (5): 426–433. doi:10.1038/nbt.2561. PMID 23584611. PMC 3678540. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3678540/. 
  198. Yang, Nan; Zuchero, J Bradley; Ahlenius, Henrik; Marro, Samuele; Ng, Yi Han; Vierbuchen, Thomas; Hawkins, John S.; Geissler, Richard et al. (2013). "Generation of oligodendroglial cells by direct lineage conversion". Nature Biotechnology 31 (5): 434–439. doi:10.1038/nbt.2564. PMID 23584610. PMC 3677690. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3677690/. 
  199. Xu, Chunhui (2012). "Turning cardiac fibroblasts into cardiomyocytes in vivo". Trends in Molecular Medicine 18 (10): 575–576. doi:10.1016/j.molmed.2012.06.009. PMID 22770847. 
  200. Fu, Ji-Dong; Stone, Nicole R.; Liu, Lei; Spencer, C. Ian; Qian, Li; Hayashi, Yohei; Delgado-Olguin, Paul; Ding, Sheng et al. (2013). "Direct Reprogramming of Human Fibroblasts toward a Cardiomyocyte-like State". Stem Cell Reports 1 (3): 235–247. doi:10.1016/j.stemcr.2013.07.005. PMID 24319660. PMC 3849259. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3849259/. 
  201. Chen, Jenny X.; Krane, Markus; Deutsch, Marcus-Andre; Wang, Li; Rav-Acha, Moshe; Gregoire, Serge; Engels, Marc C.; Rajarajan, Kuppusamy et al. (2012). "Inefficient Reprogramming of Fibroblasts into Cardiomyocytes Using Gata4, Mef2c, and Tbx5". Circulation Research 111 (1): 50–55. doi:10.1161/CIRCRESAHA.112.270264. PMID 22581928. PMC 3390172. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3390172/. 
  202. Burridge, Paul W.; Keller, Gordon; Gold, Joseph D.; Wu, Joseph C. (2012). "Production of de Novo Cardiomyocytes: Human Pluripotent Stem Cell Differentiation and Direct Reprogramming". Cell Stem Cell 10 (1): 16–28. doi:10.1016/j.stem.2011.12.013. PMID 22226352. PMC 3255078. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3255078/. 
  203. Wang, Haixia; Cao, Nan; Spencer, C. Ian; Nie, Baoming; Ma, Tianhua; Xu, Tao; Zhang, Yu; Wang, Xiaojing et al. (2014). "Small Molecules Enable Cardiac Reprogramming of Mouse Fibroblasts with a Single Factor, Oct4". Cell Reports 6 (5): 951–960. doi:10.1016/j.celrep.2014.01.038. PMID 24561253. PMC 4004339. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4004339/. 
  204. Carpenter, Lee; Carr, Carolyn; Yang, Cheng Tao; Stuckey, Daniel J.; Clarke, Kieran; Watt, Suzanne M. (2012). "Efficient Differentiation of Human Induced Pluripotent Stem Cells Generates Cardiac Cells That Provide Protection Following Myocardial Infarction in the Rat". Stem Cells and Development 21 (6): 977–986. doi:10.1089/scd.2011.0075. PMID 22182484. PMC 3315757. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3315757/. 
  205. Yamada, Satsuki; Nelson, Timothy J.; Kane, Garvan C.; Martinez‐Fernandez, Almudena; Crespo‐Diaz, Ruben J.; Ikeda, Yasuhiro; Perez‐Terzic, Carmen; Terzic, Andre (2013). "Induced pluripotent stem cell intervention rescues ventricular wall motion disparity, achieving biological cardiac resynchronization post‐infarction". The Journal of Physiology 591 (17): 4335–4349. doi:10.1113/jphysiol.2013.252288. PMID 23568891. PMC 3779120. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3779120/. 
  206. Lian, X.; Hsiao, C.; Wilson, G.; Zhu, K.; Hazeltine, L. B.; Azarin, S. M.; Raval, K. K.; Zhang, J. et al. (2012). "Cozzarelli Prize Winner: Robust cardiomyocyte differentiation from human pluripotent stem cells via temporal modulation of canonical WNT signaling". Proceedings of the National Academy of Sciences 109 (27): E1848–E1857. doi:10.1073/pnas.1200250109. PMID 22645348. PMC 3390875. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3390875/. 
  207. Willems, Erik; Cabral-Teixeira, Joaquim; Schade, Dennis; Cai, Wenqing; Reeves, Patrick; Bushway, Paul J.; Lanier, Marion; Walsh, Christopher et al. (2012). "Small Molecule-Mediated TGF-β Type II Receptor Degradation Promotes Cardiomyogenesis in Embryonic Stem Cells". Cell Stem Cell 11 (2): 242–252. doi:10.1016/j.stem.2012.04.025. PMID 22862949. PMC 3419596. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3419596/. 
  208. Lu, Tung-Ying; Lin, Bo; Kim, Jong; Sullivan, Mara; Tobita, Kimimasa; Salama, Guy; Yang, Lei (2013). "Repopulation of decellularized mouse heart with human induced pluripotent stem cell-derived cardiovascular progenitor cells". Nature Communications 4: 2307. doi:10.1038/ncomms3307. PMID 23942048. 
  209. Budniatzky, I., & Gepstein, L. (2014). Concise Review: Reprogramming Strategies for Cardiovascular Regenerative Medicine: From Induced Pluripotent Stem Cells to Direct Reprogramming. Stem cells translational medicine, 3(4), 448-457. doi:10.5966/sctm.2013-0163
  210. Kapoor, N., Liang, W., Marbán, E., and Cheol Cho, H. (2013). Direct conversion of quiescent cardiomyocytes to pacemaker cells by expression of Tbx18. Nature Biotechnology. 31: 54-62. doi:10.1038/nbt.2465
  211. Wiese, C., Grieskamp, T., Airik, R., Mommersteeg, M., Gardiwal, A., deVries, C., Gossler, K., Moorman, A., Kispert, A., and Christoffels, V. (2009). Formation of the Sinus Node Head and Differentiation of Sinus Node Myocardium Are Independently Regulated by Tbx18 and Tbx3. Circulation Research. 104: 388-397.
  212. Y-F. Hu, J. F. Dawkins, H. C. Cho, E. Marbán, E. Cingolani,(2014). Biological pacemaker created by minimally invasive somatic reprogramming in pigs with complete heart block. Sci. Transl. Med. 6, 245ra94 doi:10.1126/scitranslmed.3008681
  213. Cosgrove, B. D., Gilbert, P. M., Porpiglia, E., Mourkioti, F., Lee, S. P., Corbel, S. Y., ... & Blau, H. M. (2014). Rejuvenation of the muscle stem cell population restores strength to injured aged muscles. Nature medicine, 20(3), 255-264.doi:10.1038/nm.3464
  214. Sousa-Victor, Pedro; Gutarra, Susana; García-Prat, Laura; Rodriguez-Ubreva, Javier; Ortet, Laura; Ruiz-Bonilla, Vanessa; Jardí, Mercè; Ballestar, Esteban et al. (2014). "Geriatric muscle stem cells switch reversible quiescence into senescence". Nature 506 (7488): 316–321. doi:10.1038/nature13013. PMID 24522534. 
  215. Hosoyama, et al. and Masatoshi Suzuki (March, 2014). Derivation of Myogenic Progenitors Directly From Human Pluripotent Stem Cells Using a Sphere-Based Culture. Stem Cells Trans Med. doi:10.5966/sctm.2013-0143
  216. Zhu, Saiyong; Rezvani, Milad; Harbell, Jack; Mattis, Aras N.; Wolfe, Alan R.; Benet, Leslie Z.; Willenbring, Holger; Ding, Sheng (2014). "Mouse liver repopulation with hepatocytes generated from human fibroblasts". Nature 508 (7494): 93–97. doi:10.1038/nature13020. PMID 24572354. PMC 4161230. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4161230/. 
  217. Dean Yimlamai, Constantina Christodoulou, Giorgio G. Galli, et al., & Fernando D. Camargoemai (2014). Hippo Pathway Activity Influences Liver Cell Fate. Cell, 157(6), 1324–1338 doi:10.1016/j.cell.2014.03.060
  218. Abdelalim, E. M., Bonnefond, A., Bennaceur-Griscelli, A., & Froguel, P. (2014). Pluripotent Stem Cells as a Potential Tool for Disease Modelling and Cell Therapy in Diabetes. Stem Cell Reviews and Reports, 1-11. doi:10.1007/s12015-014-9503-6
  219. Hrvatin, S., O’Donnell, C. W., Deng, F., et al. & Melton, D. A. (2014). Differentiated human stem cells resemble fetal, not adult, β cells. Proceedings of the National Academy of Sciences, 111(8), 3038-3043. doi:10.1073/pnas.1400709111
  220. Akinci E, Banga A, Tungatt K, et al. and Slack, J.M. (2013). Reprogramming of Various Cell Types to a Beta-Like State by Pdx1, Ngn3 and MafA. PLoS ONE 8(11): e82424. doi:10.1371/journal.pone.0082424
  221. Chen, Y. J., Finkbeiner, S. R., Weinblatt, D., et al. & Stanger, B. Z. (2014). De Novo Formation of Insulin-Producing “Neo-β Cell Islets” from Intestinal Crypts. Cell Reports., doi:10.1016/j.celrep.2014.02.013
  222. Hendry, Caroline E.; Vanslambrouck, Jessica M.; Ineson, Jessica; Suhaimi, Norseha; Takasato, Minoru; Rae, Fiona; Little, Melissa H. (2013). "Direct Transcriptional Reprogramming of Adult Cells to Embryonic Nephron Progenitors". Journal of the American Society of Nephrology 24 (9): 1424–1434. doi:10.1681/ASN.2012121143. PMID 23766537. PMC 3752949. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3752949/. 
  223. Xinaris C, Benedetti V, Rizzo P, et al. and Giuseppe Remuzzi (2012) In vivo maturation of functional renal organoids formed from embryonic cell suspensions. J Am Soc Nephrol 23: 1857–1868, doi: 10.1681/ASN.2012050505
  224. Yin, Liya; Ohanyan, Vahagn; Fen Pung, Yuh; Delucia, Angelo; Bailey, Erin; Enrick, Molly; Stevanov, Kelly; Kolz, Christopher L. et al. (2012). "Induction of Vascular Progenitor Cells from Endothelial Cells Stimulates Coronary Collateral Growth". Circulation Research 110 (2): 241–252. doi:10.1161/CIRCRESAHA.111.250126. PMID 22095729. PMC 3974272. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3974272/. 
  225. Quijada, Pearl; Toko, Haruhiro; Fischer, Kimberlee M.; Bailey, Brandi; Reilly, Patrick; Hunt, Kristin D.; Gude, Natalie A.; Avitabile, Daniele et al. (2012). "Preservation of Myocardial Structure is Enhanced by Pim-1 Engineering of Bone Marrow Cells". Circulation Research 111 (1): 77–86. doi:10.1161/CIRCRESAHA.112.265207. PMID 22619278. PMC 3398618. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3398618/. 
  226. Mohsin, Sadia; Khan, Mohsin; Toko, Haruhiro; Bailey, Brandi; Cottage, Christopher T.; Wallach, Kathleen; Nag, Divya; Lee, Andrew et al. (2012). "Human Cardiac Progenitor Cells Engineered with Pim-I Kinase Enhance Myocardial Repair". Journal of the American College of Cardiology 60 (14): 1278–1287. doi:10.1016/j.jacc.2012.04.047. PMID 22841153. PMC 3461098. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3461098/. 
  227. American Heart Association (2012, July 25). Adult stem cells from liposuction used to create blood vessels in the lab. ScienceDaily.
  228. Wang, Zack Z.; Au, Patrick; Chen, Tong; Shao, Ying; Daheron, Laurence M.; Bai, Hao; Arzigian, Melanie; Fukumura, Dai et al. (2007). "Endothelial cells derived from human embryonic stem cells form durable blood vessels in vivo". Nature Biotechnology 25 (3): 317–318. doi:10.1038/nbt1287. PMID 17322871. 
  229. Samuel, R.; Daheron, L.; Liao, S.; Vardam, T.; Kamoun, W. S.; Batista, A.; Buecker, C.; Schafer, R. et al. (2013). "Generation of functionally competent and durable engineered blood vessels from human induced pluripotent stem cells". Proceedings of the National Academy of Sciences 110 (31): 12774–12779. doi:10.1073/pnas.1310675110. PMID 23861493. PMC 3732948. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3732948/. 
  230. Zangi, Lior; Lui, Kathy O.; von Gise, Alexander; Ma, Qing; Ebina, Wataru; Ptaszek, Leon M.; Später, Daniela; Xu, Huansheng et al. (2013). "Modified mRNA directs the fate of heart progenitor cells and induces vascular regeneration after myocardial infarction". Nature Biotechnology 31 (10): 898–907. doi:10.1038/nbt.2682. PMID 24013197. PMC 4058317. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4058317/. 
  231. Pereira, C. F., Chang, B., Qiu, J., Niu, X., Papatsenko, D., Hendry, C. E., ... & Moore, K. (2013). Induction of a hemogenic program in mouse fibroblasts. Cell stem cell, 13(2), 205-218. DOI: http://dx.doi.org/10.1016/j.stem.2013.05.024
  232. Jonah Riddell, Roi Gazit, Brian S. Garrison, et al., & Derrick J. Rossi (2014). Reprogramming Committed Murine Blood Cells to Induced Hematopoietic Stem Cells with Defined Factors. Cell, 157(30, 549–564, DOI: http://dx.doi.org/10.1016/j.cell.2014.04.006
  233. Boston Children's Hospital."Blood cells reprogrammed into blood stem cells in mice". ScienceDaily, 24 April 2014
  234. Focosi, D., Amabile, G., Di Ruscio, A., Quaranta, P., Tenen, D. G., & Pistello, M. (2014).Induced pluripotent stem cells in hematology: current and future applications. Blood Cancer Journal (2014) 4, e211; doi:10.1038/bcj.2014.30
  235. Zeuner, Ann; Martelli, Fabrizio; Vaglio, Stefania; Federici, Giulia; Whitsett, Carolyn; Migliaccio, Anna Rita (2012). "Concise Review: Stem Cell‐Derived Erythrocytes as Upcoming Players in Blood Transfusion". Stem Cells 30 (8): 1587–1596. doi:10.1002/stem.1136. PMID 22644674. PMC 3697769. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3697769/. 
  236. Hirose, Sho-Ichi; Takayama, Naoya; Nakamura, Sou; Nagasawa, Kazumichi; Ochi, Kiyosumi; Hirata, Shinji; Yamazaki, Satoshi; Yamaguchi, Tomoyuki et al. (2013). "Immortalization of Erythroblasts by c-MYC and BCL-XL Enables Large-Scale Erythrocyte Production from Human Pluripotent Stem Cells". Stem Cell Reports 1 (6): 499–508. doi:10.1016/j.stemcr.2013.10.010. PMID 24371805. PMC 3871399. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3871399/. 
  237. Giarratana, Marie-Catherine; Rouard, Hélène; Dumont, Agnès; Kiger, Laurent; Safeukui, Innocent; Le Pennec, Pierre-Yves; François, Sabine; Trugnan, Germain et al. (2011). "Proof of principle for transfusion of in vitro–generated red blood cells". Blood 118 (19): 5071–5079. doi:10.1182/blood-2011-06-362038. PMID 21885599. PMC 3217398. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3217398/. 
  238. Kobari, L.; Yates, F.; Oudrhiri, N.; Francina, A.; Kiger, L.; Mazurier, C.; Rouzbeh, S.; El-Nemer, W. et al. (2012). "Human induced pluripotent stem cells can reach complete terminal maturation: In vivo and in vitro evidence in the erythropoietic differentiation model". Haematologica 97 (12): 1795–1803. doi:10.3324/haematol.2011.055566. PMID 22733021. PMC 3590085. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3590085/. 
  239. Keerthivasan, Ganesan; Wickrema, Amittha; Crispino, John D. (2011). "Erythroblast Enucleation". Stem Cells International 2011: 1–9. doi:10.4061/2011/139851. PMID 22007239. PMC 3189604. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3189604/. 
  240. Smith, Brenden W.; Rozelle, Sarah S.; Leung, Amy; Ubellacker, Jessalyn; Parks, Ashley; Nah, Shirley K.; French, Deborah; Gadue, Paul et al. (2013). "The aryl hydrocarbon receptor directs hematopoietic progenitor cell expansion and differentiation". Blood 122 (3): 376–385. doi:10.1182/blood-2012-11-466722. PMID 23723449. PMC 3716202. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3716202/. 
  241. Siddharth Shah, Xiaosong Huang, Linzhao Cheng (2014). Stem Cell-Based Approaches to Red Blood Cell Production for Transfusion. Stem Cells Trans Med; 3:346-355; doi:10.5966/sctm.2013-0054
  242. Scientific Breakthrough as Artificial Blood is Created from Stem Cells
  243. Figueiredo, Constança; Goudeva, Lilia; Horn, Peter A.; Eiz-Vesper, Britta; Blasczyk, Rainer; Seltsam, Axel (2010). "Generation of HLA-deficient platelets from hematopoietic progenitor cells". Transfusion 50 (8): 1690–1701. doi:10.1111/j.1537-2995.2010.02644.x. PMID 20412529. https://zenodo.org/record/896100. 
  244. Nakamura, Sou; Takayama, Naoya; Hirata, Shinji; Seo, Hideya; Endo, Hiroshi; Ochi, Kiyosumi; Fujita, Ken-Ichi; Koike, Tomo et al. (2014). "Expandable Megakaryocyte Cell Lines Enable Clinically Applicable Generation of Platelets from Human Induced Pluripotent Stem Cells". Cell Stem Cell 14 (4): 535–548. doi:10.1016/j.stem.2014.01.011. PMID 24529595. 
  245. Riddell, Stanley R.; Greenberg, Philip D. (1995). "Principles for Adoptive T Cell Therapy of Human Viral Diseases". Annual Review of Immunology 13: 545–586. doi:10.1146/annurev.iy.13.040195.002553. PMID 7612234. 
  246. 246.0 246.1 Nishimura, Toshinobu; Kaneko, Shin; Kawana-Tachikawa, Ai; Tajima, Yoko; Goto, Haruo; Zhu, Dayong; Nakayama-Hosoya, Kaori; Iriguchi, Shoichi et al. (2013). "Generation of Rejuvenated Antigen-Specific T Cells by Reprogramming to Pluripotency and Redifferentiation". Cell Stem Cell 12 (1): 114–126. doi:10.1016/j.stem.2012.11.002. PMID 23290140. 
  247. Vizcardo, Raul; Masuda, Kyoko; Yamada, Daisuke; Ikawa, Tomokatsu; Shimizu, Kanako; Fujii, Shin-Ichiro; Koseki, Haruhiko; Kawamoto, Hiroshi (2013). "Regeneration of Human Tumor Antigen-Specific T Cells from iPSCs Derived from Mature CD8 + T Cells". Cell Stem Cell 12 (1): 31–36. doi:10.1016/j.stem.2012.12.006. PMID 23290135. 
  248. 248.0 248.1 Lei, Fengyang; Haque, Rizwanul; Xiong, Xiaofang; Song, Jianxun (2012). "Directed Differentiation of Induced Pluripotent Stem Cells towards T Lymphocytes". Journal of Visualized Experiments (63): e3986. doi:10.3791/3986. PMID 22617911. PMC 3389997. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3389997/. 
  249. Sadelain, Michel; Brentjens, Renier; Rivière, Isabelle (2013). "The Basic Principles of Chimeric Antigen Receptor Design". Cancer Discovery 3 (4): 388–398. doi:10.1158/2159-8290.CD-12-0548. PMID 23550147. PMC 3667586. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3667586/. 
  250. Themeli, Maria; Kloss, Christopher C.; Ciriello, Giovanni; Fedorov, Victor D.; Perna, Fabiana; Gonen, Mithat; Sadelain, Michel (2013). "Generation of tumor-targeted human T lymphocytes from induced pluripotent stem cells for cancer therapy". Nature Biotechnology 31 (10): 928–933. doi:10.1038/nbt.2678. PMID 23934177. PMC 5722218. //www.ncbi.nlm.nih.gov/pmc/articles/PMC5722218/. 
  251. Pilones, Karsten A.; Aryankalayil, Joseph; Demaria, Sandra (2012). "Invariant NKT Cells as Novel Targets for Immunotherapy in Solid Tumors". Clinical and Developmental Immunology 2012: 1–11. doi:10.1155/2012/720803. PMID 23118781. PMC 3483734. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3483734/. 
  252. Watarai, Hiroshi; Yamada, Daisuke; Fujii, Shin-Ichiro; Taniguchi, Masaru; Koseki, Haruhiko (2012). "Induced pluripotency as a potential path towards iNKT cell-mediated cancer immunotherapy". International Journal of Hematology 95 (6): 624–631. doi:10.1007/s12185-012-1091-0. PMID 22592322. 
  253. Haruta, M.; Tomita, Y.; Yuno, A.; Matsumura, K.; Ikeda, T.; Takamatsu, K.; Haga, E.; Koba, C. et al. (2013). "TAP-deficient human iPS cell-derived myeloid cell lines as unlimited cell source for dendritic cell-like antigen-presenting cells". Gene Therapy 20 (5): 504–513. doi:10.1038/gt.2012.59. PMID 22875043. 
  254. Xie, Huafeng; Ye, Min; Feng, Ru; Graf, Thomas (2004). "Stepwise Reprogramming of B Cells into Macrophages". Cell 117 (5): 663–676. doi:10.1016/S0092-8674(04)00419-2. PMID 15163413. 
    Bussmann, Lars H.; Schubert, Alexis; Vu Manh, Thien Phong; De Andres, Luisa; Desbordes, Sabrina C.; Parra, Maribel; Zimmermann, Timo; Rapino, Francesca et al. (2009). "A Robust and Highly Efficient Immune Cell Reprogramming System". Cell Stem Cell 5 (5): 554–566. doi:10.1016/j.stem.2009.10.004. PMID 19896445. 
  255. Hanna, Jacob; Markoulaki, Styliani; Schorderet, Patrick; Carey, Bryce W.; Beard, Caroline; Wernig, Marius; Creyghton, Menno P.; Steine, Eveline J. et al. (2008). "Direct Reprogramming of Terminally Differentiated Mature B Lymphocytes to Pluripotency". Cell 133 (2): 250–264. doi:10.1016/j.cell.2008.03.028. PMID 18423197. PMC 2615249. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2615249/. 
  256. Di Stefano, Bruno; Sardina, Jose Luis; Van Oevelen, Chris; Collombet, Samuel; Kallin, Eric M.; Vicent, Guillermo P.; Lu, Jun; Thieffry, Denis et al. (2014). "C/EBPα poises B cells for rapid reprogramming into induced pluripotent stem cells". Nature 506 (7487): 235–239. doi:10.1038/nature12885. PMID 24336202. 
  257. Rapino, Francesca; Robles, Eloy F.; Richter-Larrea, Jose A.; Kallin, Eric M.; Martinez-Climent, Jose A.; Graf, Thomas (2013). "C/EBPα Induces Highly Efficient Macrophage Transdifferentiation of B Lymphoma and Leukemia Cell Lines and Impairs Their Tumorigenicity". Cell Reports 3 (4): 1153–1163. doi:10.1016/j.celrep.2013.03.003. PMID 23545498. 
  258. H. Onuma, T. Komatsu, M. Arita, K. Hanaoka, T. Ueno, T. Terai, T. Nagano, and T. Inoue (2014), Rapidly rendering cells phagocytic through a cell surface display technique and concurrent Rac activation. Sci. Signal. 7, rs4 doi:10.1126/scisignal.2005123
  259. Guo, J., Feng, Y., Barnes, P., Huang, F. F., Idell, S., Su, D. M., & Shams, H. (2012). Deletion of FoxN1 in the thymic medullary epithelium reduces peripheral T cell responses to infection and mimics changes of aging. PloS one, 7(4), e34681. doi:10.1371/journal.pone.0034681
  260. Sun, Liguang; Guo, Jianfei; Brown, Robert; Amagai, Takashi; Zhao, Yong; Su, Dong-Ming (2010). "Declining expression of a single epithelial cell-autonomous gene accelerates age-related thymic involution". Aging Cell 9 (3): 347–357. doi:10.1111/j.1474-9726.2010.00559.x. PMID 20156205. PMC 2894280. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2894280/. 
  261. Nicholas Bredenkamp, Craig S. Nowell and C. Clare Blackburn (April 2014). Regeneration of the aged thymus by a single transcription factor. Development, 141, 1627-1637 doi:10.1242/dev.103614
  262. Bredenkamp N., Ulyanchenko S., O’Neill K. E., Manley N. R., Vaidya H. J. & Blackburn C. C. (2014). An organized and functional thymus generated from ​FOXN1-reprogrammed fibroblasts. Nature Cell Biology, doi:10.1038/ncb3023
  263. 263.0 263.1 Peng, Yan; Huang, Sha; Cheng, Biao; Nie, Xiaohu; Enhe, Jirigala; Feng, Changjiang; Fu, Xiaobing (2013). "Mesenchymal stem cells: A revolution in therapeutic strategies of age-related diseases". Ageing Research Reviews 12: 103–115. doi:10.1016/j.arr.2012.04.005. PMID 22569401. 
  264. 264.0 264.1 Bieback, K.; Kern, S.; Kocaömer, A.; Ferlik, K.; Bugert, P. (2008). "Comparing mesenchymal stromal cells from different human tissues: Bone marrow, adipose tissue and umbilical cord blood". Bio-Medical Materials and Engineering 18 (1 Suppl): S71-6. PMID 18334717. 
  265. Efimenko, Anastasia; Dzhoyashvili, Nina; Kalinina, Natalia; Kochegura, Tatiana; Akchurin, Renat; Tkachuk, Vsevolod; Parfyonova, Yelena (2014). "Adipose-Derived Mesenchymal Stromal Cells from Aged Patients with Coronary Artery Disease Keep Mesenchymal Stromal Cell Properties but Exhibit Characteristics of Aging and Have Impaired Angiogenic Potential". Stem Cells Translational Medicine 3 (1): 32–41. doi:10.5966/sctm.2013-0014. PMID 24353175. PMC 3902283. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3902283/. 
  266. Stolzing, A.; Jones, E.; McGonagle, D.; Scutt, A. (2008). "Age-related changes in human bone marrow-derived mesenchymal stem cells: Consequences for cell therapies". Mechanisms of Ageing and Development 129 (3): 163–173. doi:10.1016/j.mad.2007.12.002. PMID 18241911. 
  267. Joana Frobe, Hatim Hemeda, Michael Lenz, et al., & Wolfgang Wagneremai (Sept. 2014). Epigenetic Rejuvenation of Mesenchymal Stromal Cells Derived from Induced Pluripotent Stem Cells. Stem Cell Reports, 3(3), 414–422, doi: http://dx.doi.org/10.1016/j.stemcr.2014.07.003
  268. Diederichs Solvig and TuanRocky S. (April, 2014). Functional Comparison of Human-Induced Pluripotent Stem Cell-Derived Mesenchymal Cells and Bone Marrow-Derived Mesenchymal Stromal Cells from the Same Donor Stem Cells and Development. doi:10.1089/scd.2013.0477
  269. Irina Eberle, Mohsen Moslem, Reinhard Henschler, Tobias Cantz (2012) Engineered MSCs from Patient-Specific iPS Cells. Advances in Biochemical Engineering Biotechnology
  270. Chen, Yen Shun; Pelekanos, Rebecca A.; Ellis, Rebecca L.; Horne, Rachel; Wolvetang, Ernst J.; Fisk, Nicholas M. (2012). "Small Molecule Mesengenic Induction of Human Induced Pluripotent Stem Cells to Generate Mesenchymal Stem/Stromal Cells". Stem Cells Translational Medicine 1 (2): 83–95. doi:10.5966/sctm.2011-0022. PMID 23197756. PMC 3659681. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3659681/. 
  271. Hynes, K.; Menicanin, D.; Han, J.; Marino, V.; Mrozik, K.; Gronthos, S.; Bartold, P. M. (2013). "Mesenchymal stem cells from iPS cells facilitate periodontal regeneration". Journal of Dental Research 92 (9): 833–839. doi:10.1177/0022034513498258. PMID 23884555. 
  272. iPSC for Dental Tissue Regeneration
  273. Hynes, Kim; Menicanin, Danijela; Mrozik, Krzysztof; Gronthos, Stan; Bartold, P. Mark (2014). "Generation of Functional Mesenchymal Stem Cells from Different Induced Pluripotent Stem Cell Lines". Stem Cells and Development 23 (10): 1084–1096. doi:10.1089/scd.2013.0111. PMID 24367908. PMC 4015475. http://online.liebertpub.com/doi/abs/10.1089/scd.2013.0111. 
  274. Lai, Ruenn Chai; Yeo, Ronne Wee Yeh; Tan, Soon Sim; Zhang, Bin; Yin, Yijun; Sze, Newman Siu Kwan; Choo, Andre; Lim, Sai Kiang (2013). "Mesenchymal Stem Cell Exosomes: The Future MSC-Based Therapy?". Mesenchymal Stem Cell Therapy. pp. 39–61. doi:10.1007/978-1-62703-200-1_3. ISBN 978-1-62703-199-8. 
    Lai, Ruenn Chai; Yeo, Ronne Wee Yeh; Tan, Kok Hian; Lim, Sai Kiang (2013). "Exosomes for drug delivery — a novel application for the mesenchymal stem cell". Biotechnology Advances 31 (5): 543–551. doi:10.1016/j.biotechadv.2012.08.008. PMID 22959595. 
    Kosaka, Nobuyoshi; Takeshita, Fumitaka; Yoshioka, Yusuke; Hagiwara, Keitaro; Katsuda, Takeshi; Ono, Makiko; Ochiya, Takahiro (2013). "Exosomal tumor-suppressive microRNAs as novel cancer therapy". Advanced Drug Delivery Reviews 65 (3): 376–382. doi:10.1016/j.addr.2012.07.011. PMID 22841506. 
  275. Jumabay, Medet; Abdmaulen, Raushan; Ly, Albert; Cubberly, Mark R.; Shahmirian, Laurine J.; Heydarkhan-Hagvall, Sepideh; Dumesic, Daniel A.; Yao, Yucheng et al. (2014). "Pluripotent Stem Cells Derived from Mouse and Human White Mature Adipocytes". Stem Cells Translational Medicine 3 (2): 161–171. doi:10.5966/sctm.2013-0107. 
  276. Poloni, Antonella; Maurizi, Giulia; Leoni, Pietro; Serrani, Federica; Mancini, Stefania; Frontini, Andrea; Zingaretti, M. Cristina; Siquini, Walter et al. (2012). "Human Dedifferentiated Adipocytes Show Similar Properties to Bone Marrow‐Derived Mesenchymal Stem Cells". Stem Cells 30 (5): 965–974. doi:10.1002/stem.1067. PMID 22367678. 
  277. Shen, J. F.; Sugawara, A.; Yamashita, J.; Ogura, H.; Sato, S. (2011). "Dedifferentiated fat cells: An alternative source of adult multipotent cells from the adipose tissues". International Journal of Oral Science 3 (3): 117–124. doi:10.4248/IJOS11044. PMID 21789960. PMC 3470092. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3470092/. 
  278. Melief, Sara M.; Zwaginga, Jaap Jan; Fibbe, Willem E.; Roelofs, Helene (2013). "Adipose Tissue-Derived Multipotent Stromal Cells Have a Higher Immunomodulatory Capacity Than Their Bone Marrow-Derived Counterparts". Stem Cells Translational Medicine 2 (6): 455–463. doi:10.5966/sctm.2012-0184. PMID 23694810. PMC 3673757. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3673757/. 
  279. Cheng, Aixin; Hardingham, Timothy E.; Kimber, Susan J. (2014). "Generating Cartilage Repair from Pluripotent Stem Cells". Tissue Engineering Part B: Reviews 20 (4): 257–266. doi:10.1089/ten.teb.2012.0757. PMID 23957872. PMC 4123466. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4123466/. 
  280. Outani, Hidetatsu; Okada, Minoru; Yamashita, Akihiro; Nakagawa, Kanako; Yoshikawa, Hideki; Tsumaki, Noriyuki (2013). "Direct Induction of Chondrogenic Cells from Human Dermal Fibroblast Culture by Defined Factors". PLOS ONE 8 (10): e77365. doi:10.1371/journal.pone.0077365. PMID 24146984. PMC 3797820. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3797820/. 
  281. 281.0 281.1 Crompton, Joseph G.; Rao, Mahendra; Restifo, Nicholas P. (2013). "Memoirs of a Reincarnated T Cell". Cell Stem Cell 12 (1): 6–8. doi:10.1016/j.stem.2012.12.009. PMID 23290132. PMC 6352969. //www.ncbi.nlm.nih.gov/pmc/articles/PMC6352969/. 
  282. Tan, Hong-Kee; Toh, Cheng-Xu Delon; Ma, Dongrui; Yang, Binxia; Liu, Tong Ming; Lu, Jun; Wong, Chee-Wai; Tan, Tze-Kai et al. (2014). "Human Finger-Prick Induced Pluripotent Stem Cells Facilitate the Development of Stem Cell Banking". Stem Cells Translational Medicine 3 (5): 586–598. doi:10.5966/sctm.2013-0195. PMID 24646489. PMC 4006490. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4006490/. 
  283. Okita, Keisuke; Yamakawa, Tatsuya; Matsumura, Yasuko; Sato, Yoshiko; Amano, Naoki; Watanabe, Akira; Goshima, Naoki; Yamanaka, Shinya (2013). "An Efficient Nonviral Method to Generate Integration‐Free Human‐Induced Pluripotent Stem Cells from Cord Blood and Peripheral Blood Cells". Stem Cells 31 (3): 458–466. doi:10.1002/stem.1293. PMID 23193063. 
  284. Geti, Imbisaat; Ormiston, Mark L.; Rouhani, Foad; Toshner, Mark; Movassagh, Mehregan; Nichols, Jennifer; Mansfield, William; Southwood, Mark et al. (2012). "A Practical and Efficient Cellular Substrate for the Generation of Induced Pluripotent Stem Cells from Adults: Blood-Derived Endothelial Progenitor Cells". Stem Cells Translational Medicine 1 (12): 855–865. doi:10.5966/sctm.2012-0093. PMID 23283547. PMC 3659672. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3659672/. 
  285. Staerk, Judith; Dawlaty, Meelad M.; Gao, Qing; Maetzel, Dorothea; Hanna, Jacob; Sommer, Cesar A.; Mostoslavsky, Gustavo; Jaenisch, Rudolf (2010). "Reprogramming of Human Peripheral Blood Cells to Induced Pluripotent Stem Cells". Cell Stem Cell 7 (1): 20–24. doi:10.1016/j.stem.2010.06.002. PMID 20621045. PMC 2917234. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2917234/. 
    Park, Tea Soon; Huo, Jeffrey S.; Peters, Ann; Talbot, C. Conover; Verma, Karan; Zimmerlin, Ludovic; Kaplan, Ian M.; Zambidis, Elias T. (2012). "Growth Factor-Activated Stem Cell Circuits and Stromal Signals Cooperatively Accelerate Non-Integrated iPSC Reprogramming of Human Myeloid Progenitors". PLOS ONE 7 (8): e42838. doi:10.1371/journal.pone.0042838. PMID 22905176. PMC 3414503. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3414503/. 
  286. Yoshikawa, Katsuhiro; Naitoh, Motoko; Kubota, Hiroshi; Ishiko, Toshihiro; Aya, Rino; Yamawaki, Satoko; Suzuki, Shigehiko (2013). "Multipotent stem cells are effectively collected from adult human cheek skin". Biochemical and Biophysical Research Communications 431 (1): 104–110. doi:10.1016/j.bbrc.2012.12.069. PMID 23268344. 
  287. Zhou, Ting; Benda, Christina; Duzinger, Sarah; Huang, Yinghua; Li, Xingyan; Li, Yanhua; Guo, Xiangpeng; Cao, Guokun et al. (2011). "Generation of Induced Pluripotent Stem Cells from Urine". Journal of the American Society of Nephrology 22 (7): 1221–1228. doi:10.1681/ASN.2011010106. PMID 21636641. PMC 3137570. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3137570/. 
    Zhou, Ting; Benda, Christina; Dunzinger, Sarah; Huang, Yinghua; Ho, Jenny Cy; Yang, Jiayin; Wang, Yu; Zhang, Ya et al. (2012). "Generation of human induced pluripotent stem cells from urine samples". Nature Protocols 7 (12): 2080–2089. doi:10.1038/nprot.2012.115. PMID 23138349. 
    Wang, Lihui; Wang, Linli; Huang, Wenhao; Su, Huanxing; Xue, Yanting; Su, Zhenghui; Liao, Baojian; Wang, Haitao et al. (2013). "Generation of integration-free neural progenitor cells from cells in human urine". Nature Methods 10 (1): 84–89. doi:10.1038/nmeth.2283. PMID 23223155. 
    Cai J, Zhang Y, Liu P, Chen S, Wu X, Sun Y, Li A, Huang K et al (2013)Generation of tooth-like structures from integration-free human urine induced pluripotent stem cells. Cell Regeneration , 2:6 doi:10.1186/2045-9769-2-6
  288. Bharadwaj, Shantaram; Liu, Guihua; Shi, Yingai; Wu, Rongpei; Yang, Bin; He, Tongchuan; Fan, Yuxin; Lu, Xinyan et al. (2013). "Multipotential differentiation of human urine‐derived stem cells: Potential for therapeutic applications in urology". Stem Cells 31 (9): 1840–1856. doi:10.1002/stem.1424. PMID 23666768. 
  289. Wang, Yimei; Liu, Jinyu; Tan, Xiaohua; Li, Gaofeng; Gao, Yunhe; Liu, Xuejuan; Zhang, Lihong; Li, Yulin (2013). "Induced Pluripotent Stem Cells from Human Hair Follicle Mesenchymal Stem Cells". Stem Cell Reviews and Reports 9 (4): 451–460. doi:10.1007/s12015-012-9420-5. PMID 23242965. PMC 3742959. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3742959/. 
  290. Schnabel, Lauren V.; Abratte, Christian M.; Schimenti, John C.; Southard, Teresa L.; Fortier, Lisa A. (2012). "Genetic background affects induced pluripotent stem cell generation". Stem Cell Research & Therapy 3 (4): 30. doi:10.1186/scrt121. PMID 22862934. PMC 3580468. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3580468/. 
  291. Panopoulos, Athanasia D.; Ruiz, Sergio; Yi, Fei; Herrerías, Aída; Batchelder, Erika M.; Belmonte, Juan Carlos Izpisua (2011). "Rapid and Highly Efficient Generation of Induced Pluripotent Stem Cells from Human Umbilical Vein Endothelial Cells". PLOS ONE 6 (5): e19743. doi:10.1371/journal.pone.0019743. PMID 21603572. PMC 3095638. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3095638/. 
  292. 292.0 292.1 Polo, Jose M.; Liu, Susanna; Figueroa, Maria Eugenia; Kulalert, Warakorn; Eminli, Sarah; Tan, Kah Yong; Apostolou, Effie; Stadtfeld, Matthias et al. (2010). "Cell type of origin influences the molecular and functional properties of mouse induced pluripotent stem cells". Nature Biotechnology 28 (8): 848–855. doi:10.1038/nbt.1667. PMID 20644536. PMC 3148605. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3148605/. 
  293. Miura, Kyoko; Okada, Yohei; Aoi, Takashi; Okada, Aki; Takahashi, Kazutoshi; Okita, Keisuke; Nakagawa, Masato; Koyanagi, Michiyo et al. (2009). "Variation in the safety of induced pluripotent stem cell lines". Nature Biotechnology 27 (8): 743–745. doi:10.1038/nbt.1554. PMID 19590502. 
    Liang, Yi; Zhang, Hui; Feng, Qi-Sheng; Cai, Man-Bo; Deng, Wen; Qin, Dajiang; Yun, Jing-Ping; Tsao, George Sai Wah et al. (2013). "The propensity for tumorigenesis in human induced pluripotent stem cells is related with genomic instability". Chinese Journal of Cancer 32 (4): 205–212. doi:10.5732/cjc.012.10065. PMID 22704487. PMC 3845575. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3845575/. 
  294. Kim, K.; Doi, A.; Wen, B.; Ng, K.; Zhao, R.; Cahan, P.; Kim, J.; Aryee, M. J. et al. (2010). "Epigenetic memory in induced pluripotent stem cells". Nature 467 (7313): 285–290. doi:10.1038/nature09342. PMID 20644535. PMC 3150836. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3150836/. 
  295. Kim, Kitai; Zhao, Rui; Doi, Akiko; Ng, Kitwa; Unternaehrer, Juli; Cahan, Patrick; Hongguang, Huo; Loh, Yuin-Han et al. (2011). "Donor cell type can influence the epigenome and differentiation potential of human induced pluripotent stem cells". Nature Biotechnology 29 (12): 1117–1119. doi:10.1038/nbt.2052. PMID 22119740. PMC 3357310. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3357310/. 
  296. Bar-Nur, Ori; Russ, Holger A.; Efrat, Shimon; Benvenisty, Nissim (2011). "Epigenetic Memory and Preferential Lineage-Specific Differentiation in Induced Pluripotent Stem Cells Derived from Human Pancreatic Islet Beta Cells". Cell Stem Cell 9 (1): 17–23. doi:10.1016/j.stem.2011.06.007. PMID 21726830. 
  297. Denker, Hans-Werner (2012). "Time to Reconsider Stem Cell Induction Strategies". Cells 1 (4): 1293–1312. doi:10.3390/cells1041293. PMID 24710555. PMC 3901125. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3901125/. 

Notes edit

  Type classification: this is a notes resource.

External links edit

ESC & iPSC News

Induced Stem cells page on facebook

Efficient cell specific differentiation systems for iPSC. Ask the Expert discussion.