Cell potency

Last updated

Cell potency is a cell's ability to differentiate into other cell types. [1] [2] The more cell types a cell can differentiate into, the greater its potency. Potency is also described as the gene activation potential within a cell, which like a continuum, begins with totipotency to designate a cell with the most differentiation potential, pluripotency, multipotency, oligopotency, and finally unipotency.

Contents

Pluripotent, embryonic stem cells originate as inner mass cells within a blastocyst. These stem cells can become any tissue in the body, excluding a placenta. Only the morula's cells are totipotent, able to become all tissues and a placenta. Stem cells diagram.png
Pluripotent, embryonic stem cells originate as inner mass cells within a blastocyst. These stem cells can become any tissue in the body, excluding a placenta. Only the morula's cells are totipotent, able to become all tissues and a placenta.

Totipotency

Totipotency (Latin: totipotentia, lit.'ability for all [things]') is the ability of a single cell to divide and produce all of the differentiated cells in an organism. Spores and zygotes are examples of totipotent cells. [3] In the spectrum of cell potency, totipotency represents the cell with the greatest differentiation potential, being able to differentiate into any embryonic cell, as well as any extraembryonic tissue cell. In contrast, pluripotent cells can only differentiate into embryonic cells. [4] [5]

A fully differentiated cell can return to a state of totipotency. [6] The conversion to totipotency is complex and not fully understood. In 2011, research revealed that cells may differentiate not into a fully totipotent cell, but instead into a "complex cellular variation" of totipotency. [7] Stem cells resembling totipotent blastomeres from 2-cell stage embryos can arise spontaneously in mouse embryonic stem cell cultures [8] [9] and also can be induced to arise more frequently in vitro through down-regulation of the chromatin assembly activity of CAF-1. [10]

The human development model can be used to describe how totipotent cells arise. [11] Human development begins when a sperm fertilizes an egg and the resulting fertilized egg creates a single totipotent cell, a zygote. [12] In the first hours after fertilization, this zygote divides into identical totipotent cells, which can later develop into any of the three germ layers of a human (endoderm, mesoderm, or ectoderm), or into cells of the placenta (cytotrophoblast or syncytiotrophoblast). After reaching a 16-cell stage, the totipotent cells of the morula differentiate into cells that will eventually become either the blastocyst's Inner cell mass or the outer trophoblasts. Approximately four days after fertilization and after several cycles of cell division, these totipotent cells begin to specialize. The inner cell mass, the source of embryonic stem cells, becomes pluripotent.

Research on Caenorhabditis elegans suggests that multiple mechanisms including RNA regulation may play a role in maintaining totipotency at different stages of development in some species. [13] Work with zebrafish and mammals suggest a further interplay between miRNA and RNA-binding proteins (RBPs) in determining development differences. [14]

Primordial germ cells

In mouse primordial germ cells, genome-wide reprogramming leading to totipotency involves erasure of epigenetic imprints. Reprogramming is facilitated by active DNA demethylation involving the DNA base excision repair enzymatic pathway. [15] This pathway entails erasure of CpG methylation (5mC) in primordial germ cells via the initial conversion of 5mC to 5-hydroxymethylcytosine (5hmC), a reaction driven by high levels of the ten-eleven dioxygenase enzymes TET-1 and TET-2. [16]

Pluripotency

A: Human embryonic stem cells (cell colonies that are not yet differentiated).
B: Nerve cells Human embryonic stem cells.png
A: Human embryonic stem cells (cell colonies that are not yet differentiated).
B: Nerve cells

In cell biology, pluripotency (Latin: pluripotentia, lit.'ability for many [things]') [17] refers to a stem cell that has the potential to differentiate into any of the three germ layers: endoderm (gut, lungs and liver), mesoderm (muscle, skeleton, blood vascular, urogenital, dermis), or ectoderm (nervous, sensory, epidermis), but not into extra-embryonic tissues like the placenta or yolk sac. [18] However, cell pluripotency is a continuum, ranging from the completely pluripotent cell that can form every cell of the embryo proper, e.g., embryonic stem cells and iPSCs, to the incompletely or partially pluripotent cell that can form cells of all three germ layers but that may not exhibit all the characteristics of completely pluripotent cells.

Induced pluripotency

Induced pluripotent stem cells, commonly abbreviated as iPS cells or iPSCs, are a type of pluripotent stem cell artificially derived from a non-pluripotent cell, typically an adult somatic cell, by inducing a "forced" expression of certain genes and transcription factors. [19] These transcription factors play a key role in determining the state of these cells and also highlights the fact that these somatic cells do preserve the same genetic information as early embryonic cells. [20] The ability to induce cells into a pluripotent state was initially pioneered in 2006 using mouse fibroblasts and four transcription factors, Oct4, Sox2, Klf4 and c-Myc; [21] this technique, called reprogramming, later earned Shinya Yamanaka and John Gurdon the Nobel Prize in Physiology or Medicine. [22] This was then followed in 2007 by the successful induction of human iPSCs derived from human dermal fibroblasts using methods similar to those used for the induction of mouse cells. [23] These induced cells exhibit similar traits to those of embryonic stem cells (ESCs) but do not require the use of embryos. Some of the similarities between ESCs and iPSCs include pluripotency, morphology, self-renewal ability, a trait that implies that they can divide and replicate indefinitely, and gene expression. [24]

Epigenetic factors are also thought to be involved in the actual reprogramming of somatic cells in order to induce pluripotency. It has been theorized that certain epigenetic factors might actually work to clear the original somatic epigenetic marks in order to acquire the new epigenetic marks that are part of achieving a pluripotent state. Chromatin is also reorganized in iPSCs and becomes like that found in ESCs in that it is less condensed and therefore more accessible. Euchromatin modifications are also common which is also consistent with the state of euchromatin found in ESCs. [24]

Due to their great similarity to ESCs, the medical and research communities are interested iPSCs. iPSCs could potentially have the same therapeutic implications and applications as ESCs but without the controversial use of embryos in the process, a topic of great bioethical debate. The induced pluripotency of somatic cells into undifferentiated iPS cells was originally hailed as the end of the controversial use of embryonic stem cells. However, iPSCs were found to be potentially tumorigenic, and, despite advances, [19] were never approved for clinical stage research in the United States until recently. Currently, autologous iPSC-derived dopaminergic progenitor cells are used in trials for treating Parkinson's disease. [25] Setbacks such as low replication rates and early senescence have also been encountered when making iPSCs, [26] hindering their use as ESCs replacements.

Somatic expression of combined transcription factors can directly induce other defined somatic cell fates (transdifferentiation); researchers identified three neural-lineage-specific transcription factors that could directly convert mouse fibroblasts (connective tissue cells) into fully functional neurons. [27] This result challenges the terminal nature of cellular differentiation and the integrity of lineage commitment; and implies that with the proper tools, all cells are totipotent and may form all kinds of tissue.

Some of the possible medical and therapeutic uses for iPSCs derived from patients include their use in cell and tissue transplants without the risk of rejection that is commonly encountered. iPSCs can potentially replace animal models unsuitable as well as in vitro models used for disease research. [28]

Naive human pluripotent stem cell colony here seen growing on feeder cells (mouse). Naive-hPSC.tif
Naive human pluripotent stem cell colony here seen growing on feeder cells (mouse).

Naive vs. primed pluripotency states

Findings with respect to epiblasts before and after implantation have produced proposals for classifying pluripotency into two states: "naive" and "primed", representing pre- and post-implnatation epiblast, respectively. [29] Naive-to-primed continuum is controlled by reduction of Sox2/Oct4 dimerization on SoxOct DNA elements controlling naive pluripotency. [30] Primed pluripotent stem cells from different species could be reset to naive state using a cocktail containing Klf4 and Sox2 or "super-Sox" − a chimeric transcription factor with enhanced capacity to dimerize with Oct4. [30]

The baseline stem cells commonly used in science that are referred as embryonic stem cells (ESCs) are derived from a pre-implantation epiblast; such epiblast is able to generate the entire fetus, and one epiblast cell is able to contribute to all cell lineages if injected into another blastocyst. On the other hand, several marked differences can be observed between the pre- and post-implantation epiblasts, such as their difference in morphology, in which the epiblast after implantation changes its morphology into a cup-like shape called the "egg cylinder" as well as chromosomal alteration in which one of the X-chromosomes under random inactivation in the early stage of the egg cylinder, known as X-inactivation. [31] During this development, the egg cylinder epiblast cells are systematically targeted by Fibroblast growth factors, Wnt signaling, and other inductive factors via the surrounding yolk sac and the trophoblast tissue, [32] such that they become instructively specific according to the spatial organization. [33]

Another major difference is that post-implantation epiblast stem cells are unable to contribute to blastocyst chimeras, [34] which distinguishes them from other known pluripotent stem cells. Cell lines derived from such post-implantation epiblasts are referred to as epiblast-derived stem cells, which were first derived in laboratory in 2007. Both ESCs and EpiSCs are derived from epiblasts but at difference phases of development. Pluripotency is still intact in the post-implantation epiblast, as demonstrated by the conserved expression of Nanog, Fut4, and Oct-4 in EpiSCs, [35] until somitogenesis and can be reversed midway through induced expression of Oct-4. [36]

Native pluripotency in plants

Ranunculus asiaticus example of totipotency of two individuals MHNT (MHNT) Ranunculus asiaticus - example of Totipotency.jpg
Ranunculus asiaticus example of totipotency of two individuals MHNT

Un-induced pluripotency has been observed in root meristem tissue culture, especially by Kareem et al 2015, Kim et al 2018, and Rosspopoff et al 2017. This pluripotency is regulated by various regulators, including PLETHORA 1 and PLETHORA 2; and PLETHORA 3, PLETHORA 5, and PLETHORA 7, whose expression were found by Kareem to be auxin-provoked. (These are also known as PLT1, PLT2, PLT3, PLT5, PLT7, and expressed by genes of the same names.) As of 2019, this is expected to open up future research into pluripotency in root tissues. [37]

Multipotency

Hematopoietic stem cells are an example of multipotency. When they differentiate into myeloid or lymphoid progenitor cells, they lose potency and become oligopotent cells with the ability to give rise to all cells of its lineage. Hematopoiesis (human) diagram en.svg
Hematopoietic stem cells are an example of multipotency. When they differentiate into myeloid or lymphoid progenitor cells, they lose potency and become oligopotent cells with the ability to give rise to all cells of its lineage.

Multipotency is when progenitor cells have the gene activation potential to differentiate into discrete cell types. For example, a hematopoietic stem cell – and this cell type can differentiate itself into several types of blood cell like lymphocytes, monocytes, neutrophils, etc., but it is still ambiguous whether HSC possess the ability to differentiate into brain cells, bone cells or other non-blood cell types.[ citation needed ]

Research related to multipotent cells suggests that multipotent cells may be capable of conversion into unrelated cell types. In another case, human umbilical cord blood stem cells were converted into human neurons. [38] There is also research on converting multipotent cells into pluripotent cells. [39]

Multipotent cells are found in many, but not all human cell types. Multipotent cells have been found in cord blood, [40] adipose tissue, [41] cardiac cells, [42] bone marrow, and mesenchymal stem cells (MSCs) which are found in the third molar. [43]

MSCs may prove to be a valuable source for stem cells from molars at 8–10 years of age, before adult dental calcification. MSCs can differentiate into osteoblasts, chondrocytes, and adipocytes. [44]

Oligopotency

In biology, oligopotency is the ability of progenitor cells to differentiate into a few cell types. It is a degree of potency. Examples of oligopotent stem cells are the lymphoid or myeloid stem cells. [2] A lymphoid cell specifically, can give rise to various blood cells such as B and T cells, however, not to a different blood cell type like a red blood cell. [45] Examples of progenitor cells are vascular stem cells that have the capacity to become both endothelial or smooth muscle cells.

Unipotency

In cell biology, a unipotent cell is the concept that one stem cell has the capacity to differentiate into only one cell type. [46] It is currently unclear if true unipotent stem cells exist. Hepatoblasts, which differentiate into hepatocytes (which constitute most of the liver) or cholangiocytes (epithelial cells of the bile duct), are bipotent. [47] A close synonym for unipotent cell is precursor cell.

See also

Related Research Articles

<span class="mw-page-title-main">Stem cell</span> Undifferentiated biological cells that can differentiate into specialized cells

In multicellular organisms, stem cells are undifferentiated or partially differentiated cells that can change into various types of cells and proliferate indefinitely to produce more of the same stem cell. They are the earliest type of cell in a cell lineage. They are found in both embryonic and adult organisms, but they have slightly different properties in each. They are usually distinguished from progenitor cells, which cannot divide indefinitely, and precursor or blast cells, which are usually committed to differentiating into one cell type.

Transdifferentiation, also known as lineage reprogramming, is the process in which one mature somatic cell is transformed into another mature somatic cell without undergoing an intermediate pluripotent state or progenitor cell type. It is a type of metaplasia, which includes all cell fate switches, including the interconversion of stem cells. Current uses of transdifferentiation include disease modeling and drug discovery and in the future may include gene therapy and regenerative medicine. The term 'transdifferentiation' was originally coined by Selman and Kafatos in 1974 to describe a change in cell properties as cuticle producing cells became salt-secreting cells in silk moths undergoing metamorphosis.

<span class="mw-page-title-main">Cellular differentiation</span> Developmental biology

Cellular differentiation is the process in which a stem cell changes from one type to a differentiated one. Usually, the cell changes to a more specialized type. Differentiation happens multiple times during the development of a multicellular organism as it changes from a simple zygote to a complex system of tissues and cell types. Differentiation continues in adulthood as adult stem cells divide and create fully differentiated daughter cells during tissue repair and during normal cell turnover. Some differentiation occurs in response to antigen exposure. Differentiation dramatically changes a cell's size, shape, membrane potential, metabolic activity, and responsiveness to signals. These changes are largely due to highly controlled modifications in gene expression and are the study of epigenetics. With a few exceptions, cellular differentiation almost never involves a change in the DNA sequence itself. However, metabolic composition does get altered quite dramatically where stem cells are characterized by abundant metabolites with highly unsaturated structures whose levels decrease upon differentiation. Thus, different cells can have very different physical characteristics despite having the same genome.

<span class="mw-page-title-main">Somatic cell nuclear transfer</span> Method of creating a cloned embryo by replacing the egg nucleus with a body cell nucleus

In genetics and developmental biology, somatic cell nuclear transfer (SCNT) is a laboratory strategy for creating a viable embryo from a body cell and an egg cell. The technique consists of taking an denucleated oocyte and implanting a donor nucleus from a somatic (body) cell. It is used in both therapeutic and reproductive cloning. In 1996, Dolly the sheep became famous for being the first successful case of the reproductive cloning of a mammal. In January 2018, a team of scientists in Shanghai announced the successful cloning of two female crab-eating macaques from foetal nuclei.

<span class="mw-page-title-main">Germ cell</span> Gamete-producing cell

A germ cell is any cell that gives rise to the gametes of an organism that reproduces sexually. In many animals, the germ cells originate in the primitive streak and migrate via the gut of an embryo to the developing gonads. There, they undergo meiosis, followed by cellular differentiation into mature gametes, either eggs or sperm. Unlike animals, plants do not have germ cells designated in early development. Instead, germ cells can arise from somatic cells in the adult, such as the floral meristem of flowering plants.

<span class="mw-page-title-main">Embryonic stem cell</span> Type of pluripotent blastocystic stem cell

Embryonic stem cells (ESCs) are pluripotent stem cells derived from the inner cell mass of a blastocyst, an early-stage pre-implantation embryo. Human embryos reach the blastocyst stage 4–5 days post fertilization, at which time they consist of 50–150 cells. Isolating the inner cell mass (embryoblast) using immunosurgery results in destruction of the blastocyst, a process which raises ethical issues, including whether or not embryos at the pre-implantation stage have the same moral considerations as embryos in the post-implantation stage of development.

<span class="mw-page-title-main">Embryoid body</span> Three-dimensional aggregate of pluripotent stem cells

Embryoid bodies (EBs) are three-dimensional aggregates of pluripotent stem cells.

<span class="mw-page-title-main">Oct-4</span> Mammalian protein found in Homo sapiens

Oct-4, also known as POU5F1, is a protein that in humans is encoded by the POU5F1 gene. Oct-4 is a homeodomain transcription factor of the POU family. It is critically involved in the self-renewal of undifferentiated embryonic stem cells. As such, it is frequently used as a marker for undifferentiated cells. Oct-4 expression must be closely regulated; too much or too little will cause differentiation of the cells.

Gametogonium are stem cells for gametes located within the gonads. They originate from primordial germ cells, which have migrated to the gonads. Male gametogonia which are located within the testes during development and adulthood are called spermatogonium. Female gametogonia, known as oogonium, are found within the ovaries of the developing foetus and were thought to be depleted at or after birth. Spermatogonia and oogonia are classified as sexually differentiated germ cells.

<span class="mw-page-title-main">Homeobox protein NANOG</span> Mammalian protein found in humans

Homeobox protein NANOG(hNanog) is a transcriptional factor that helps embryonic stem cells (ESCs) maintain pluripotency by suppressing cell determination factors. hNanog is encoded in humans by the NANOG gene. Several types of cancer are associated with NANOG.

<span class="mw-page-title-main">Stem-cell line</span> Culture of stem cells that can be propagated indefinitely

A stem cell line is a group of stem cells that is cultured in vitro and can be propagated indefinitely. Stem cell lines are derived from either animal or human tissues and come from one of three sources: embryonic stem cells, adult stem cells, or induced stem cells. They are commonly used in research and regenerative medicine.

In biology, reprogramming refers to erasure and remodeling of epigenetic marks, such as DNA methylation, during mammalian development or in cell culture. Such control is also often associated with alternative covalent modifications of histones.

<span class="mw-page-title-main">Induced pluripotent stem cell</span> Pluripotent stem cell generated directly from a somatic cell

Induced pluripotent stem cells are a type of pluripotent stem cell that can be generated directly from a somatic cell. The iPSC technology was pioneered by Shinya Yamanaka and Kazutoshi Takahashi in Kyoto, Japan, who together showed in 2006 that the introduction of four specific genes, collectively known as Yamanaka factors, encoding transcription factors could convert somatic cells into pluripotent stem cells. Shinya Yamanaka was awarded the 2012 Nobel Prize along with Sir John Gurdon "for the discovery that mature cells can be reprogrammed to become pluripotent."

<span class="mw-page-title-main">SOX2</span> Transcription factor gene of the SOX family

SRY -box 2, also known as SOX2, is a transcription factor that is essential for maintaining self-renewal, or pluripotency, of undifferentiated embryonic stem cells. Sox2 has a critical role in maintenance of embryonic and neural stem cells.

<span class="mw-page-title-main">Shinya Yamanaka</span> Japanese stem cell researcher

Shinya Yamanaka is a Japanese stem cell researcher and a Nobel Prize laureate. He is a professor and the director emeritus of Center for iPS Cell Research and Application, Kyoto University; as a senior investigator at the UCSF-affiliated Gladstone Institutes in San Francisco, California; and as a professor of anatomy at University of California, San Francisco (UCSF). Yamanaka is also a past president of the International Society for Stem Cell Research (ISSCR).

A mesenchymal–epithelial transition (MET) is a reversible biological process that involves the transition from motile, multipolar or spindle-shaped mesenchymal cells to planar arrays of polarized cells called epithelia. MET is the reverse process of epithelial–mesenchymal transition (EMT) and it has been shown to occur in normal development, induced pluripotent stem cell reprogramming, cancer metastasis and wound healing.

Induced stem cells (iSC) are stem cells derived from somatic, reproductive, pluripotent or other cell types by deliberate epigenetic reprogramming. They are classified as either totipotent (iTC), pluripotent (iPSC) or progenitor 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.

A Muse cell is an endogenous non-cancerous pluripotent stem cell. They reside in the connective tissue of nearly every organ including the umbilical cord, bone marrow and peripheral blood. They are collectable from commercially obtainable mesenchymal cells such as human fibroblasts, bone marrow-mesenchymal stem cells and adipose-derived stem cells as 1~several percent of the total population. Muse cells are able to generate cells representative of all three germ layers from a single cell both spontaneously and under cytokine induction. Expression of pluripotency genes and triploblastic differentiation are self-renewable over generations. Muse cells do not undergo teratoma formation when transplanted into a host environment in vivo. This can be explained in part by their intrinsically low telomerase activity, eradicating the risk of tumorigenesis through unbridled cell proliferation. They were discovered in 2010 by Mari Dezawa and her research group. Clinical trials for acute myocardial infarction, stroke, epidermolysis bullosa, spinal cord injury, amyotrophic lateral sclerosis, acute respiratory distress syndrome (ARDS) related to novel coronavirus (SARS-CoV-2) infection, are conducted. Physician-led clinical trial for neonatal hypoxic-ischemic encephalopathy was also started. The summary results of a randomized double-blind placebo-controlled clinical trial in patients with stroke was announced.

Directed differentiation is a bioengineering methodology at the interface of stem cell biology, developmental biology and tissue engineering. It is essentially harnessing the potential of stem cells by constraining their differentiation in vitro toward a specific cell type or tissue of interest. Stem cells are by definition pluripotent, able to differentiate into several cell types such as neurons, cardiomyocytes, hepatocytes, etc. Efficient directed differentiation requires a detailed understanding of the lineage and cell fate decision, often provided by developmental biology.

After the blastocyst stage, once an embryo implanted in endometrium, the inner cell mass (ICM) of a fertilized embryo segregates into two layers: hypoblast and epiblast. The epiblast cells are the functional progenitors of soma and germ cells which later differentiate into three layers: definitive endoderm, mesoderm and ectoderm. Stem cells derived from epiblast are pluripotent. These cells are called epiblast-derived stem cells (EpiSCs) and have several different cellular and molecular characteristics with Embryonic Stem Cells (ESCs). Pluripotency in EpiSCs is essentially different from that of embryonic stem cells. The pluripotency of EpiSCs is primed pluripotency: primed to differentiate into specific cell lineages. Naïve pluripotent stem cells and primed pluripotent stem cells not only sustain the ability to self-renew but also maintain the capacity to differentiate. Since the cell status is primed to differentiate in EpiSCs, however, one copy of the X chromosome in XX cells in EpiSCs is silenced (XaXi). EpiSCs is unable to colonize and is not available to be used to produce chimeras. Conversely, XX cells in ESCs are both active and can produce chimera when inserted into a blastocyst. Both ESC and EpiSC induce teratoma when injected in the test animals which proves pluripotency. EpiSC display several distinctive characteristics distinct from ESCs. The cellular status of human ESCs (hESCs) is similar to primed state mouse stem cells rather than naïve state.

References

  1. Cell totipotency was discovered by Habertland and the term was coined by Thomas Hund Morgan. Mahla RS (2016). "Stem Cells Applications in Regenerative Medicine and Disease Therapeutics". International Journal of Cell Biology. 2016 (7): 6940283. doi: 10.1155/2016/6940283 . PMC   4969512 . PMID   27516776.
  2. 1 2 Schöler HR (2007). "The Potential of Stem Cells: An Inventory". In Knoepffler M, Schipanski D, Sorgner SL (eds.). Human biotechnology as Social Challenge. Ashgate Publishing, Ltd. p. 28. ISBN   978-0-7546-5755-2.
  3. Mitalipov S, Wolf D (2009). "Totipotency, pluripotency and nuclear reprogramming". Engineering of Stem Cells. Advances in Biochemical Engineering/Biotechnology. Vol. 114. pp. 185–199. Bibcode:2009esc..book..185M. doi:10.1007/10_2008_45. ISBN   978-3-540-88805-5. PMC   2752493 . PMID   19343304.
  4. Lodish H (2016). Molecular Cell Biology (8th ed.). W. H. Freeman. pp. 975–977. ISBN   978-1319067748.
  5. "What is the difference between totipotent, pluripotent, and multipotent?".
  6. Western P (2009). "Foetal germ cells: striking the balance between pluripotency and differentiation". The International Journal of Developmental Biology. 53 (2–3): 393–409. doi: 10.1387/ijdb.082671pw . PMID   19412894.
  7. Sugimoto K, Gordon SP, Meyerowitz EM (April 2011). "Regeneration in plants and animals: dedifferentiation, transdifferentiation, or just differentiation?". Trends in Cell Biology. 21 (4): 212–218. doi:10.1016/j.tcb.2010.12.004. PMID   21236679.
  8. Macfarlan TS, Gifford WD, Driscoll S, Lettieri K, Rowe HM, Bonanomi D, et al. (July 2012). "Embryonic stem cell potency fluctuates with endogenous retrovirus activity". Nature. 487 (7405): 57–63. Bibcode:2012Natur.487...57M. doi:10.1038/nature11244. PMC   3395470 . PMID   22722858.
  9. Morgani SM, Canham MA, Nichols J, Sharov AA, Migueles RP, Ko MS, Brickman JM (June 2013). "Totipotent embryonic stem cells arise in ground-state culture conditions". Cell Reports. 3 (6): 1945–1957. doi:10.1016/j.celrep.2013.04.034. PMC   3701323 . PMID   23746443.
  10. Ishiuchi T, Enriquez-Gasca R, Mizutani E, Bošković A, Ziegler-Birling C, Rodriguez-Terrones D, et al. (September 2015). "Early embryonic-like cells are induced by downregulating replication-dependent chromatin assembly". Nature Structural & Molecular Biology. 22 (9): 662–671. doi:10.1038/nsmb.3066. PMID   26237512. S2CID   837230.
  11. Seydoux G, Braun RE (December 2006). "Pathway to totipotency: lessons from germ cells". Cell. 127 (5): 891–904. doi: 10.1016/j.cell.2006.11.016 . PMID   17129777. S2CID   16988032.
  12. Asch R, Simerly C, Ord T, Ord VA, Schatten G (July 1995). "The stages at which human fertilization arrests: microtubule and chromosome configurations in inseminated oocytes which failed to complete fertilization and development in humans". Human Reproduction. 10 (7): 1897–1906. doi:10.1093/oxfordjournals.humrep.a136204. PMID   8583008.
  13. Ciosk R, DePalma M, Priess JR (February 2006). "Translational regulators maintain totipotency in the Caenorhabditis elegans germline". Science. 311 (5762): 851–853. Bibcode:2006Sci...311..851C. doi:10.1126/science.1122491. PMID   16469927. S2CID   130017.
  14. Kedde M, Agami R (April 2008). "Interplay between microRNAs and RNA-binding proteins determines developmental processes". Cell Cycle. 7 (7): 899–903. doi: 10.4161/cc.7.7.5644 . PMID   18414021.
  15. Hajkova P, Jeffries SJ, Lee C, Miller N, Jackson SP, Surani MA (July 2010). "Genome-wide reprogramming in the mouse germ line entails the base excision repair pathway". Science. 329 (5987): 78–82. Bibcode:2010Sci...329...78H. doi:10.1126/science.1187945. PMC   3863715 . PMID   20595612.
  16. Hackett JA, Sengupta R, Zylicz JJ, Murakami K, Lee C, Down TA, Surani MA (January 2013). "Germline DNA demethylation dynamics and imprint erasure through 5-hydroxymethylcytosine". Science. 339 (6118): 448–452. Bibcode:2013Sci...339..448H. doi:10.1126/science.1229277. PMC   3847602 . PMID   23223451.
  17. "Biology Online". Biology-Online.org. Retrieved 25 April 2013.
  18. Binder MD, Hirokawa N, Nobutaka, Windhorst U, eds. (2009). Encyclopedia of neuroscience. Berlin: Springer. ISBN   978-3540237358.
  19. 1 2 Baker M (2007-12-06). "Adult cells reprogrammed to pluripotency, without tumors". Nature Reports Stem Cells: 1. doi: 10.1038/stemcells.2007.124 .
  20. Stadtfeld M, Hochedlinger K (October 2010). "Induced pluripotency: history, mechanisms, and applications". Genes & Development. 24 (20): 2239–2263. doi:10.1101/gad.1963910. PMC   2956203 . PMID   20952534.
  21. Takahashi K, Yamanaka S (August 2006). "Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors". Cell. 126 (4): 663–676. doi:10.1016/j.cell.2006.07.024. hdl: 2433/159777 . PMID   16904174. S2CID   1565219.
  22. "The Nobel Prize in Physiology or Medicine 2012". Nobelprize.org. Nobel Media AB 2013. Web. 28 Nov 2013.
  23. Takahashi K, Tanabe K, Ohnuki M, Narita M, Ichisaka T, Tomoda K, Yamanaka S (November 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. hdl: 2433/49782 . PMID   18035408. S2CID   8531539.
  24. 1 2 Liang G, Zhang Y (January 2013). "Embryonic stem cell and induced pluripotent stem cell: an epigenetic perspective". Cell Research. 23 (1): 49–69. doi:10.1038/cr.2012.175. PMC   3541668 . PMID   23247625.
  25. Schweitzer JS, Song B, Herrington TM, Park TY, Lee N, Ko S, et al. (May 2020). "Personalized iPSC-Derived Dopamine Progenitor Cells for Parkinson's Disease". The New England Journal of Medicine. 382 (20): 1926–1932. doi:10.1056/NEJMoa1915872. PMC   7288982 . PMID   32402162.
  26. Choi, Charles. "Cell-Off: Induced Pluripotent Stem Cells Fall Short of Potential Found in Embryonic Version". Scientific American. Retrieved 25 April 2013.
  27. Vierbuchen T, Ostermeier A, Pang ZP, Kokubu Y, Südhof TC, Wernig M (February 2010). "Direct conversion of fibroblasts to functional neurons by defined factors". Nature. 463 (7284): 1035–1041. Bibcode:2010Natur.463.1035V. doi:10.1038/nature08797. PMC   2829121 . PMID   20107439.
  28. Park IH, Lerou PH, Zhao R, Huo H, Daley GQ (2008). "Generation of human-induced pluripotent stem cells". Nature Protocols. 3 (7): 1180–1186. doi:10.1038/nprot.2008.92. PMID   18600223. S2CID   13321484.
  29. Nichols J, Smith A (June 2009). "Naive and primed pluripotent states". Cell Stem Cell. 4 (6): 487–492. doi: 10.1016/j.stem.2009.05.015 . PMID   19497275.
  30. 1 2 MacCarthy, Caitlin M.; Wu, Guangming; Malik, Vikas; Menuchin-Lasowski, Yotam; Velychko, Taras; Keshet, Gal; Fan, Rui; Bedzhov, Ivan; Church, George M.; Jauch, Ralf; Cojocaru, Vlad; Schöler, Hans R.; Velychko, Sergiy (December 2023). "Highly cooperative chimeric super-SOX induces naive pluripotency across species". Cell Stem Cell. doi: 10.1016/j.stem.2023.11.010 .
  31. Heard E (June 2004). "Recent advances in X-chromosome inactivation". Current Opinion in Cell Biology. 16 (3): 247–255. doi:10.1016/j.ceb.2004.03.005. PMID   15145348.
  32. Beddington RS, Robertson EJ (January 1999). "Axis development and early asymmetry in mammals". Cell. 96 (2): 195–209. doi: 10.1016/s0092-8674(00)80560-7 . PMID   9988215. S2CID   16264083.
  33. Lawson KA, Meneses JJ, Pedersen RA (November 1991). "Clonal analysis of epiblast fate during germ layer formation in the mouse embryo". Development. 113 (3): 891–911. doi:10.1242/dev.113.3.891. PMID   1821858. S2CID   17685207.
  34. Rossant J (February 2008). "Stem cells and early lineage development". Cell. 132 (4): 527–531. doi: 10.1016/j.cell.2008.01.039 . PMID   18295568. S2CID   14128314.
  35. Brons IG, Smithers LE, Trotter MW, Rugg-Gunn P, Sun B, Chuva de Sousa Lopes SM, et al. (July 2007). "Derivation of pluripotent epiblast stem cells from mammalian embryos". Nature. 448 (7150): 191–195. Bibcode:2007Natur.448..191B. doi:10.1038/nature05950. PMID   17597762. S2CID   4365390.
  36. Osorno R, Tsakiridis A, Wong F, Cambray N, Economou C, Wilkie R, et al. (July 2012). "The developmental dismantling of pluripotency is reversed by ectopic Oct4 expression". Development. 139 (13): 2288–2298. doi:10.1242/dev.078071. PMC   3367440 . PMID   22669820.
  37. Ikeuchi M, Favero DS, Sakamoto Y, Iwase A, Coleman D, Rymen B, Sugimoto K (April 2019). "Molecular Mechanisms of Plant Regeneration". Annual Review of Plant Biology. Annual Reviews. 70 (1): 377–406. doi: 10.1146/annurev-arplant-050718-100434 . PMID   30786238. S2CID   73498853.
  38. Giorgetti A, Marchetto MC, Li M, Yu D, Fazzina R, Mu Y, et al. (July 2012). "Cord blood-derived neuronal cells by ectopic expression of Sox2 and c-Myc". Proceedings of the National Academy of Sciences of the United States of America. 109 (31): 12556–12561. Bibcode:2012PNAS..10912556G. doi: 10.1073/pnas.1209523109 . PMC   3412010 . PMID   22814375.
  39. Guan K, Nayernia K, Maier LS, Wagner S, Dressel R, Lee JH, et al. (April 2006). "Pluripotency of spermatogonial stem cells from adult mouse testis". Nature. 440 (7088): 1199–1203. Bibcode:2006Natur.440.1199G. doi:10.1038/nature04697. PMID   16565704. S2CID   4350560.
  40. Zhao Y, Mazzone T (December 2010). "Human cord blood stem cells and the journey to a cure for type 1 diabetes". Autoimmunity Reviews. 10 (2): 103–107. doi:10.1016/j.autrev.2010.08.011. PMID   20728583.
  41. Tallone T, Realini C, Böhmler A, Kornfeld C, Vassalli G, Moccetti T, et al. (April 2011). "Adult human adipose tissue contains several types of multipotent cells". Journal of Cardiovascular Translational Research. 4 (2): 200–210. doi:10.1007/s12265-011-9257-3. PMID   21327755. S2CID   36604144.
  42. Beltrami AP, Barlucchi L, Torella D, Baker M, Limana F, Chimenti S, et al. (September 2003). "Adult cardiac stem cells are multipotent and support myocardial regeneration". Cell. 114 (6): 763–776. doi: 10.1016/S0092-8674(03)00687-1 . PMID   14505575. S2CID   15588806.
  43. Ohgushi H, Arima N, Taketani T (December 2011). "[Regenerative therapy using allogeneic mesenchymal stem cells]". Nihon Rinsho. Japanese Journal of Clinical Medicine (in Japanese). 69 (12): 2121–2127. PMID   22242308.
  44. Uccelli A, Moretta L, Pistoia V (September 2008). "Mesenchymal stem cells in health and disease". Nature Reviews. Immunology. 8 (9): 726–736. doi:10.1038/nri2395. PMID   19172693. S2CID   3347616.
  45. Ibelgaufts, Horst. "Cytokines & Cells Online Pathfinder Encyclopedia" . Retrieved 25 April 2013.
  46. Betts, J Gordon; Desaix, Peter; Johnson, Eddie; Johnson, Jody E; Korol, Oksana; Kruse, Dean; Poe, Brandon; Wise, James; Womble, Mark D; Young, Kelly A (June 8, 2023). Anatomy & Physiology. Houston: OpenStax CNX. 3.5 Cell Growth and Division. ISBN   978-1-947172-04-3.
  47. "hepatoblast differentiation". GONUTS. Archived from the original on 2016-03-03. Retrieved 2012-08-31.