Search Results for: ips cells 2013

Dr T Medical Tips – Stem Cells 14 May 2013 – Video

Posted: June 4, 2013 at 9:41 pm


Dr T Medical Tips - Stem Cells 14 May 2013
http://mtv.com.lb/Dr_T_Medical_Tips.

By: MTVAlive

Read more here:
Dr T Medical Tips - Stem Cells 14 May 2013 - Video

Posted in Stem Cell Videos | Comments Off on Dr T Medical Tips – Stem Cells 14 May 2013 – Video

Stem cells: a comprehensive review of origins and emerging clinical …

Posted: March 19, 2024 at 2:38 am

Orthop Rev (Pavia). 2022; 14(3): 37498.

1Department of Anesthesiology, Mount Sinai Medical Center

1Department of Anesthesiology, Mount Sinai Medical Center

2 LSU Health Science Center Shreveport School of Medicine, Shreveport, LA

3 University of Arizona College of Medicine-Phoenix, Phoenix, AZ

3 University of Arizona College of Medicine-Phoenix, Phoenix, AZ

4Department of Emergency Medicine, University of Central Florida

5Department of Anesthesiology, Louisiana State University Health Sciences Center Shreveport

5Department of Anesthesiology, Louisiana State University Health Sciences Center Shreveport

5Department of Anesthesiology, Louisiana State University Health Sciences Center Shreveport

6Department of Anesthesiology, Louisiana State University Health Sciences Center Shreveport, Innovative Pain and Wellness, Creighton University School of Medicine

5Department of Anesthesiology, Louisiana State University Health Sciences Center Shreveport

1Department of Anesthesiology, Mount Sinai Medical Center

2 LSU Health Science Center Shreveport School of Medicine, Shreveport, LA

3 University of Arizona College of Medicine-Phoenix, Phoenix, AZ

4Department of Emergency Medicine, University of Central Florida

5Department of Anesthesiology, Louisiana State University Health Sciences Center Shreveport

6Department of Anesthesiology, Louisiana State University Health Sciences Center Shreveport, Innovative Pain and Wellness, Creighton University School of Medicine

Corresponding author: Salomon Poliwoda MD; Telephone: 7862716678; email: salomon.pb@gmail.com

Stem cells are types of cells that have unique ability to self-renew and to differentiate into more than one cell lineage. They are considered building blocks of tissues and organs. Over recent decades, they have been studied and utilized for repair and regenerative medicine. One way to classify these cells is based on their differentiation capacity. Totipotent stem cells can give rise to any cell of an embryo but also to extra-embryonic tissue as well. Pluripotent stem cells are limited to any of the three embryonic germ layers; however, they cannot differentiate into extra-embryonic tissue. Multipotent stem cells can only differentiate into one germ line tissue. Oligopotent and unipotent stem cells are seen in adult organ tissues that have committed to a cell lineage. Another way to differentiate these cells is based on their origins. Stem cells can be extracted from different sources, including bone marrow, amniotic cells, adipose tissue, umbilical cord, and placental tissue. Stem cells began their role in modern regenerative medicine in the 1950s with the first bone marrow transplantation occurring in 1956. Stem cell therapies are at present indicated for a range of clinical conditions beyond traditional origins to treat genetic blood diseases and have seen substantial success. In this regard, emerging use for stem cells is their potential to treat pain states and neurodegenerative diseases such as Parkinsons and Alzheimers disease. Stem cells offer hope in neurodegeneration to replace neurons damaged during certain disease states. This review compares stem cells arising from these different sources of origin and include clinical roles for stem cells in modern medical practice.

Keywords: Stem cells, regenerative medicine, bone marrow, umbilical cord, placental tissue

Stem cells are a unique population of cells present in all stages of life that possess the ability to self-renew and differentiate into multiple cell lineages. These cells are key mediators in the development of neonates and in restorative processes after injury or disease as they are the source from which specific cell types within differentiated tissues and organs are derived.1 Within the neonate stage of life stem cells serve to differentiate and proliferate into the multitude of cell types and lineages required for continuing development, while in adults their primary role is regenerative and restorative in nature.2 Stem cells have unique properties that set them apart from terminally differentiated cells allowing for their specific physiological roles. The ability of stem cells to differentiate into multiple cell types is termed potency, and stem cells can be classified by their potential for differentiation as well as by their origin. Totipotent or omnipotent stem cells can form embryonic tissues and can differentiate into all cell lineages required for an adult. Pluripotent stem cells can differentiate into all three germ layers while multipotent stem cells may only differentiate into one kind of germ line tissue. Oligopotent and unipotent stem cells are the type seen in adult organ tissues that have committed to a cell lineage and can only diversify into cell types within that lineage.1 Embryonic stem cells are derived from the inner cell mass of a blastocysts and are totipotent. The range of their use is typically restricted due to legal and ethical factors and for this reason mesenchymal stem cells are typically preferred. Mesenchymal stem cells can be isolated from a variety of both neonate and adult tissues but still maintain the ability to differentiate into multiple cell types allowing for their clinical and research utilization without the ethical issues associated with embryonic stem cells.3

Another key feature of stem cells is their ability to self-renew and proliferate providing a continuous supply of progeny to replace aging or damaged cells. During the developmental phase this proliferation allows for the growth necessary to mature into an adult. After the developmental phase has concluded, this continued proliferation allows for healing and restoration on a cellular level after tissue or organ injury has taken place.2 These physiological and developmental characteristics make stem cells an integral part in the field of regenerative medicine due to their ability to generate entire tissues and organs from just a handful of progenitor cells.

Stem cells began their role in modern regenerative medicine in the 1950s with the first bone marrow transplantation occurring in 1956. This breakthrough shed light on the potential treatments possible in the future with further development and refinement of clinical techniques and paved the way for the stem cell therapies that are now available.4,5 Stem cell therapies are now indicated for a range of clinical conditions beyond traditional origins to treat genetic blood diseases and have seen substantial success where other treatments have fallen short. One emerging use for stem cells is their potential to treat paint states and neurodegenerative diseases such as Parkinsons and Alzheimers disease. Stem cells offer the hope in the setting of neurodegeneration to replace the neurons damaged during the pathogenesis of certain diseases, a goal not achievable utilizing current technologies and methods.6

Organ bioengineering is yet another a rapidly developing and exciting new application for stem cells with both clinical and research implications.7 Immunosuppression free organ transplants are now a possibility with the advancement organ manufacturing utilizing the patients own cells.8 This along with the potential for eliminating organ donor waiting lists is an enticing prospect, but many technological developments are necessary before this technology can be implemented in clinical settings on a wide scale. Research has already benefitted greatly from this field because organ like tissues can be grown in lab settings to model disease progression. This offers the potential to develop new treatments while determining their efficacy on a cellular level without risk to patients.9,10

Currently one of the most prolific clinical uses of stem cells in the field of regenerative medicine is to treat inherited blood diseases. Within these diseases a genetic defect or defects prevents the proper function of cells derived from the hematopoietic stem cell lineage. Treatment includes implantation of genetically normal cells from a healthy donor to serve as a lifelong self-renewing source of normally functioning blood cells. However these treatments are limited by the availability of suitable donors.11

Stem cells can be derived from multiple sources including adult tissues or neonatal tissues such as the umbilical cord or placenta. Embryonic stem cells have been utilized in the past for research, but ethical concerns have led to them being replaced largely by stem cells derived from other origins.12 Common tissues from which adult oligopotent and unipotent stem cells are isolated include bone marrow, adipose tissue, and trabecular bone.13 Bone marrow has traditionally been the most common site from which to extract non neonatal derived stem cells but involves an invasive and painful procedure. Peripheral blood progenitor cells have been utilized to avoid harvesting cells from bone marrow. However, this technique has issues and risks of its own and was initially a less potent source of stem cells. It is also now known that stem cells differ in their proliferative and differentiation potential based on their origin. Cells sourced from umbilical Whartons jelly and adipose tissue were found to proliferate significantly more quickly than cells sourced from bone marrow and placental sources.14,15

A rapidly advancing source of stem cells known as induced pluripotent stem cells (iPSCs) are now being utilized clinically as well. These iPSCs are derived from somatic cells that have been reprogrammed back to a pluripotent state utilizing reprogramming factors and require less invasive techniques to harvest in comparison to traditional sources.16,17 Once returned to a pluripotent state, the cells then undergo a process called directed differentiation in which they are converted into desired cell types. Directed differentiation is achieved by mimicking microenvironments and extracellular signals in vitro in a manner that produces predictable cell types.18 In the future, this technique could provide a novel form of personalized gene therapy in which oligopotent or unipotent cells are procured from tissue, reprogrammed back to a less differentiated state, and then reintroduced into a different location within the patient. Work is also being done to combine this technique with modern gene editing methods to provide an entirely new subset of therapies.19 This method of transplantation would greatly reduce the chance for rejection and does not require a suitable donor, as the cells are sourced from the patient being treated.20,21

Stem cells are required by self-renewing tissues to replace damaged and aging cells because of normal biological processes. Both myeloid and lymphoid lineage cells derived from hematopoietic stem cells are relatively short-lived cell types and require a continuous source of newly differentiated replacement cells.22 Hematopoietic stem cells (HSCs) are those that reside within the bone marrow and provide a source for the multiple types of blood cells required for normal physiological and immunological functions. These cells inhabit a physiological niche which allows them to undergo the process of asymmetric division. When stem cells divide asymmetrically the progeny of the division includes one identical daughter cell but also results in the production of a differentiated daughter cell. Differentiation of these daughter cell into specialized cell types is guided by certain microenvironments, extrinsic cues, and growth factors that the cell comes in contact with.23,24 This mechanism allows for bone marrow stem cell numbers to stay relatively constant despite sustained proliferation and differentiation of progeny taking place.22,25,26

HSCs are the most studied class of adult tissue derived stem cells and their clinical potential was recognized early in the history of regenerative medicine. At the beginning of the 1960s, HSCs were isolated from bone marrow and therapeutic models in mice induced with leukemia were developed in order to show the efficacy of bone marrow derived stem cell treatments. Success in these experiments led to further refinement of techniques and by the 1970s and 80s clinical stem cell transplants were a regular occurrence and began to make the impact on blood diseases that we continue to see today.27,28

Bone marrow has historically been the predominant harvesting site for stem cell collection due to its accessibility, early identification as a source, and lengthy research history. Isolating stem cell from bone marrow involves an invasive and painful surgical procedure and does come with a risk hospitalization or other complications. Patients also report increased post procedural pain and pre-procedural anxiety when compared with other harvesting techniques.29,30 Bone marrow however has proved to be a denser source of cells than other harvesting methods yielding 18 times more cells than peripheral blood progenitor cell harvesting techniques initially. As technology and methods improved however, it was found that treating patients with a cytokine treatment prior to peripheral blood progenitor cell harvesting mobilized many of the desired cells into the blood stream and drastically increased the efficacy of this technique, making it clinically viable.3133 In a double blinded randomized study 40 patients underwent bone marrow and peripheral blood progenitor cell collections and the yield of useable harvested cells were compared. It was found that blood progenitor cell collection yielded significantly more useable stem cells and patients were able to undergo the collection procedure more frequently when compared to the bone marrow harvesting method.32 This, coupled with the invasiveness and risks associated with harvesting stem cells from bone marrow have increased peripheral blood progenitor cell collections popularity.

Overall, bone marrow as a reservoir of stem cells continues to be a clinical and research necessity related to its well understood and documented history as a source of viable stem cells and track record of efficacy. According to the European Group for Blood and Marrow Transplantation, only one fatal event was recorded stemming from the first 27,770 hematopoietic stem cell transplants sourced from bone marrow during the period of 1993-2005.34 This undeniable track record of safety coupled with clinicians experience performing bone marrow transplant procedures guarantees the continued use of bone marrow as a source of HSCs for the near future.

Historically, the two most common types of pluripotent stem cells include embryonic stem cells (ESCs) and induced pluripotent stem cells (iPSCs).35 However, despite the many research efforts to improve ESC and iPSC technologies, there are still enormous clinical challenges.35 Two significant issues posed by ESC and iPSC technologies include low survival rate of transplanted cells and tumorigenicity.35 Recently, researchers have isolated pluripotent stem cells from gestational tissues such as amniotic fluid and the placental membrane.35 Human amnion-derived stem cells (hADSCs), including amniotic epithelial cells and amniotic mesenchymal cells, are a relatively new stem cell source that have been found to have several advantageous characteristics.35,36

For background, human amniotic stem cells begin emerging during the second week of gestation when a small cavity forms within the blastocyst and primordial cells lining this cavity are differentiated into amnioblasts.36 Human amniotic epithelial stem cells (hAESCs) are formed when epiblasts differentiate into amnioblasts, whereas human amniotic mesenchymal stem cells (hAMSCs) are formed when hypoblasts differentiate into amnioblasts.35,36 This differentiation occurs prior to gastrulation, so amnioblasts do not belong to one of the 3 germ layers, making them theoretically pluripotent.3537

Previously, pluripotency and immunomodulation are qualities that have been thought to be mutually exclusive, as pluripotency has traditionally been regarded as a characteristic limited to embryonic stem cells whereas immunomodulation has been a recognized property of mesenchymal stem cells.36 However, many recent studies have found that these two qualities coexist in hADSCs.35,36

In recent years, hADSCs, including human amniotic epithelial stem cells (hAESCs) and human amniotic mesenchymal stem cells (hAMSCs) have been attractive cell sources for clinical trials and medical research, and have been shown to have advantages over other stem cells types.35,37 These advantages include low immunogenicity and high histocompatibility, no tumorigenicity, immunomodulatory effects, and significant paracrine effects.35 Also, several studies have evaluated the proangiogenic ability of hADSCs.35 Interestingly, they found that hAMSCs were shown to augment blood perfusion and capillary architecture when transplanted into ischemic limbs of mice, suggesting that hAMSCs stimulate neovascularization.35,38 Additionally, another advantage is that hADSCs are easier to obtain compared to other stem cell sources, such as bone marrow stem cells (BMSCs).35

Regarding the low immunogenicity, hADSCs have been shown to have a low expression of major histocompatibility class I antigen (HLA-ABC), and no expression of major histocompatibility class II antigen (HLA-DR), 2 microglobulin, and HLA-ABC costimulatory molecules, including CD40, CD80 and CD8635. Notably, there have been reports of transplantation of hAMSCs into patients with lysosomal diseases who had no obvious rejection.35 Moreover, a recent study demonstrated no hemolysis, allergic reactions, or tumor formations in mice who received intravenous hAESCs.35,39

Additionally, studies have demonstrated that both hAESCs and hAMSCs have great potential to play an important role in regenerative medicine. They both have demonstrated that they can differentiate into several specialized cells, including adipocytes, bone cells, nerve cells, cardiomyocytes, skeletal muscle cells, hepatocytes, hematopoietic cells, endothelial cells, kidney cells, and retinal cells.35

Multiple preclinical studies have revealed the potential for hADSCs to be used in the treatment of several diseases including premature ovarian failure, diabetes mellitus, inflammatory bowel disease, brain/spine diseases, and more.35,40,41 For example, one preclinical study investigated the effect of hAMSC-therapy on ovarian function in natural aging ovaries within mice.40 They found that after the hAMSCs were transplanted into the mice, the hAMSCs significantly improved follicle proliferation and therefore ovarian function.40 Another study investigated the effect of hAESC-therapy on outcomes after stroke in mice.41 They found that, administration of hAESCs after acute (within 1.5 hours) stroke in mice reduced brain infarct development, inflammation, and functional deficits.41 Additionally, they found that after late administration (1-3 days poststroke) of hAESCs, functional recovery in the mice was still improved.41 Overall, they concluded that administration of hAESCs following a stroke in mice showed a significant neuroprotective effect and facilitated repair and recovery of the brain.41

Although a number of preclinical studies, like the ones previously described, have shown considerable promise regarding the use of ADSC-therapy, more studies are needed. Future studies can continue to work toward determining if hADSCs are capable of being used for cell replacement and better elucidate the mechanisms by which hADSCs work.

Although the use of bone marrow stem cells (BMSCs) is now standard, dilemmas regarding harvesting techniques and the potential for low cell yields has driven researchers to search for other mesenchymal stem cell (MSCs) sources.42 One source that has been investigated is human adipose tissue.42

After enzymatic digestion of adipose tissue, a heterogenous group of adipocyte precursors are generated within a group of cells called the stromal vascular fraction (SVF).42 Adipose-derived stem cells (ADSCs) are found in the SVF.42,43 Studies have demonstrated that ADSCs possess properties typically associated with MSCs, and that they have been found to express several CD markers that MSCs characteristically express.43 ADSCs are multipotent and have been shown to differentiate into other cells of mesodermal origin, including osteoblasts, chondroblasts, myocytes, tendocytes, and more, upon in vitro induction.4245 Additionally, ADSCs have demonstrated in vitro capacity for multi-lineage differentiation into specialized cells, like insulin-secreting cells.43,46

A significant advantage of ADSCs over BMSCs is how easy they are to harvest.43,45 White adipose tissue (WAT) contains an abundance of ADSCs.43 The main stores of WAT in humans are subcutaneous stores in the buttocks, thighs, abdomen and visceral depots.43 Due to this, ADSCs can be harvested relatively easily by liposuction procedures from these areas of the body.43,45 Moreover, ADSCs make up as much as 1-2% of the SVF within WAT, sometimes even nearing 30% in some tissues.43,45 This is a significant difference from the .0001-.0002% stem cells present in bone marrow.43 Given this difference in stem cell concentration between the sources, there will be more ADSCs per sample of WAT compared to stem cells per bone marrow sample, further demonstrating an easier acquisition of stem cells when using adipose tissue.

Another advantage of ADSCs is their immune privilege status due to a lack of major histocompatibility complex II (MHC II) and costimulatory molecules.42,43,45,47 Some studies have even demonstrated a higher immunosuppression capacity in ADSCs compared to BMSCs as ADSCs expressed lower levels of human antigen class I (HLA I) antigen.47 They also have a unique secretome and can produce immunomodulatory, anti-apoptotic, hematopoietic, and angiogenic factors that can help with repair of tissues characteristics that may support successful transplantations without the need for immunosuppression.4245 Moreover, ADSCs have the ability to be reprogrammed to induced pluripotent stem (iPS) cells.43

The number of ADSC clinical trials has risen over the past decade, and some have shown significant promise. They have demonstrated abilities to differentiate into multiple cell lines in a reproducible manner and be safe for both autogenetic and allogeneic transplantations.45 Several recent studies have demonstrated that ADSC-therapy may potentially be useful in the treatment of several conditions, including diabetes mellitus, Crohns disease, multiple sclerosis, fistulas, arthritis, ischemic pathologies, cardiac injury, spinal injury, bone injuries and more.4448

One clinical trial conducted in 2013 investigated the therapeutic effect of co-infusion of autologous adipose-derived differentiated insulin-secreting stem cells and hematopoietic stem cells (HSCs) on patients with insulin-dependent diabetes mellitus.46 Ten patients were followed over an average of about thirty-two months, and they found that all the patients had improvement in C-peptide, HbA1c, blood sugar status, and exogenous insulin requirement.46 Notably, there were no unpleasant side effects of the treatment and all ten patients had rehabilitated to a normal, unrestricted diet and lifestyle.46

In another 4-patient clinical trial in which ADSCs were used to heal fistulas in patients with Crohns disease, full healing occurred in 6 out of the 8 fistulas with partial healing in the remaining two.44 No complications were observed in the patients 12 months following the trial.44 Although these results are promising, the mechanism by which the healing took place remains unclear. When considering the properties of ADSCs, there are a number of factors that could have played a role in the healing, such as the result of paracrine expression of angiogenic and/or anti-apoptotic factors, stem cell differentiation, and/or local immunosuppression.44

Other exciting studies have demonstrated a use of ADSCs in the treatment of osteoarthritis (OA). One meta-analysis compared the use of ADSCs and BMSCs in the treatment of osteoarthritis.47 This meta-analysis included 14 studies comprising 461 original patient records.47 Overall, the comparison between treatment of OA didnt show a significant difference in the disease severity score change rate between patients treated with ADSCs and those treated with BMSCs.47 However, there was significantly more variability in the outcomes of those treated with BMSCs with the highest change rate being 79.65% in one study and the lowest being 22.57% in another study.47 Given this, ADSCs may represent a more stable cell source for the treatment of OA.47 Although this study is specific to OA treatment, it is worth acknowledging the possibility that ADSCs may also represent a more stable cell source for treatment of other diseases as well.

Though recent ADSC research, as described above, has been promising, unfortunately reproducible in vivo studies are still lacking in both quality and quantity.42 Therefore, further studies are necessary prior to progression to routine patient administration.42

Umbilical Cord stem cells can be drawn from a variety of locations including umbilical cord blood, umbilical cord perivascular cells, umbilical vein endothelial cells, umbilical lining, chorion, and amnion. Umbilical cord blood can be drawn with minimal risk to the donor, and it has been used since 1988 as a source for hematopoietic stem cells.49 When compared to stem cells obtained from bone marrow, umbilical cord derived stem cells are much more readily available. With a birth rate of more than a 100 million people per year globally, there is a lot of opportunity to use umbilical cord blood as a source for stem cells.

The process of extracting the blood is very simple and involves a venipuncture followed by drainage into a sterile anti-coagulant-filled blood bag. It is then cryopreserved and stored in liquid nitrogen. There are quite a few benefits to utilizing umbilical cord stem cells rather than stem cells drawn from adults. One of the biggest benefits is that the cells are more immature which means that there is a lower chance of rejection after implantation in a host and would lead to decreased rates of graft-versus-host disease. They also can differentiate into a very wide variety of tissues. For example, when compared with bone marrow stem cells or mobilized peripheral blood, umbilical cord blood stem cells have a greater repopulating ability.50 Cord blood derived CD34+ cells have very potent hematopoietic abilities, and this is attributed to the immaturity of the stem cells relative to adult derived cells. Studies have been done that analyze long term survival of children with hematologic disorders who were transplanted with umbilical cord blood from a sibling donor. These studied revealed the same or better survival in the children that received the umbilical cord blood relative to those that got transplantation from bone marrow cells. Furthermore, rates of relapse were the same for both umbilical cord blood and bone marrow transplant.51

One of the unique features of stem cells taken from umbilical cord blood is the potential to differentiate into a wide variety of cell types. There are three different kinds of stem cells that can be found in the umbilical cord blood which include hematopoietic, mesenchymal, and embryonic-like stem cells. Not only can these cell types all renew themselves, but they can differentiate into many different mature cell types through a complex number of signaling pathways. This means that these cells could give rise to not only hematopoietic cells but bone, neural and endothelial cells. There are studies taking place currently to see if umbilical cord blood derived stem cells can be utilized for cardiomyogenic purposes. Several studies have showed the ability to transform umbilical cord blood mesenchymal stem cells into cells of cardiomyogenic lineage utilizing activations of Wnt signaling pathways.52 Studies are also being conducted on the potential of neurological applications. If successful, this could help diseases such as cerebral palsy, stroke, spinal cord injury and neurodegenerative diseases. Given these cells ability to differentiate into tissues from the mesoderm, endoderm and ectoderm, they could be utilized for neurological issues in place of embryonic stem cells that are currently extremely controversial.53 There are currently studies involving in vitro work, pre-clinical animal studies, and patient clinical trials, all for the application of stem cells in neurological applications. There is big potential for the use of umbilical blood stem cells in the future of regenerative medicine.

Placental tissue contains both stem cells and epithelial cells that can differentiate into a wide variety of tissue types which include adipogenic, myogenic, hepatogenic, osteogenic, cardiac, endothelial, pancreatic, pulmonary, and neurological. Placental cells can differentiate in to all these different kinds of tissues due to lineages originating from different parts of the placenta such as the hematopoietic cells that come from the chorion, allantois, and yolk sac while the mesenchymal lineages come from the chorion and the amnion.54 It can be helpful to think of human fetal placental cells as being divided into four different groups: amniotic epithelial cells, amniotic mesenchymal stromal cells, chorionic mesenchymal stromal cells and chorionic trophoblast cells.54

Human amniotic epithelial cells (hAECs) can be obtained from the amnion membrane where they are then enzymatically digested to be separated from the chorion. When cultured under certain settings hAECs have been found to be able to produce neuronal cells that synthesize acetylcholine, norepinephrine as well as dopamine.55,56 This ability would mean they have potential for regenerative purposes in diseases such as Parkinsons Disease, multiple sclerosis, and spinal cord injury. There is also research being done to utilize hAECs for ophthalmological purposes, lung fibrosis, liver disease, metabolic diseases, and familial hypercholesterolemia. Once cultured, hAECs have been shown to produce both albumin and alpha-fetoprotein as well as showing ability to store glycogen. Furthermore, they have been found to metabolize ammonia and testosterone. In more recent studies conducted in mouse models, these cells have been found to have therapeutic efficacy after transplantation for cirrhosis.57

Mesenchymal stem cells are in many different tissues such as the bone marrow, umbilical cord blood, adipose tissue, Whartons jelly, amniotic fluid, lungs, muscle and the placenta. Placental mesenchymal stromal cells specifically originate from the extraembryonic mesoderm. Human amniotic mesenchymal stromal cells (hAMSCs) and chorionic mesenchymal stromal cells (hCMSCs) have both been found to have very low levels of HLA-A,B,C. This means that they have immune privileged profiles for potential transplantation.58,59 Placental derived mesenchymal stem cells have been shown to have expression of CD29, CD44, CD105 and CD166 which is the same as adipose derived mesenchymal stem cells. These markers have been shown to have osteogenic differentiating abilities.57 An interesting element of placental mesenchymal stem cells is that their properties differ depending on the gestational age of the placenta. When cells are harvested at lower gestational ages, they show faster generation doubling times, better proliferative abilities, wider differentiation potential and more phenotypic stability than cells harvested from placental tissue that is considered to be at term.60 Furthermore, they have great potential to be used clinically. Placental mesenchymal stromal cells have been studied for use in treating acute graft-versus-host disease that was refractory to steroid treatment. Studies have shown that the 1-year survival rates in patients treated with placenta derived stromal cells were 73% while retrospective control only showed 6% survival.61 Placenta derived MSCs have also been found to aid in wound healing and could potentially be used to aid with certain inherited skin conditions such as epidermolysis bullosa.62

Stem cells are diverse in their differentiation capacity as well as their source of origin. As we can see from this review, there are similarities and differences when these cells are extracted from different sources. Research has shown initial promise in neurodegenerative diseases such as Alzheimers and Parkinsons Disease. It has also shown to be beneficial in the areas of musculoskeletal regenerative medicine and other pain states. Organ bioengineering for transplantation is another potential benefit that stem cells may offer. For these reasons, extensive research is still needed in this area of medicine to pave the way for new developing therapy modalities.

none

This review is dedicated to Dr.Justine C. Goldberg MD

Read more here:
Stem cells: a comprehensive review of origins and emerging clinical ...

Posted in Stem Cells | Comments Off on Stem cells: a comprehensive review of origins and emerging clinical …

Stem cells: a brief history and outlook – Science in the News

Posted: December 27, 2022 at 12:40 am

Stem cells have been the object of much excitement and controversy amongst both scientists and the general population. Surprisingly, though, not everybody understands the basic properties of stem cells, let alone the fact that there is more than one type of cell that falls within the stem cell category. Here, Ill lay out the basic concepts of stem cell biology as a background for understanding the stem cell research field, where it is headed, and the enormous promise it offers for regenerative medicine.

Fertilization of an egg cell by a sperm cell results in the generation of a zygote, the single cell that, upon a myriad of divisions, gives rise to our whole body. Because of this amazing developmental potential, the zygote is said to be totipotent. Along the way, the zygote develops into the blastocyst, which implants into the mothers uterus. The blastocyst is a structure comprising about 300 cells that contains two main regions: the inner cell mass (ICM) and the trophoblast. The ICM is made of embryonic stem cells (ES cells), which are referred to as pluripotent. They are able to give rise to all the cells in an embryo proper, but not to extra-embryonic tissues, such as the placenta. The latter originate from the trophoblast [].

Even though it is hard to pinpoint exactly when or by whom what we now call stem cells were first discovered, the consensus is that the first scientists to rigorously define the key properties of a stem cell were Ernest McCulloch and James Till. In their pioneering work in mice in the 1960s, they discovered the blood-forming stem cell, the hematopoietic stem cell (HSC) [2, 3]. By definition, a stem cell must be capable of both self-renewal (undergoing cell division to make more stem cells) and differentiation into mature cell types. HSCs are said to be multipotent, as they can still give rise to multiple cell types, but only to other types of blood cells (see Figure 1, left column). They are one of many examples of adult stem cells, which are tissue-specific stem cells that are essential for organ maintenance and repair in the adult body. Muscle, for instance, also possesses a population of adult stem cells. Called satellite cells, these muscle cells are unipotent, as they can give rise to just one cell type, muscle cells.

Therefore, the foundations of stem cell research lie not with the famous (or infamous) human embryonic stem cells, but with HSCs, which have been used in human therapy (such as bone marrow transplants) for decades. Still, what ultimately fueled the enormous impact that the stem cell research field has today is undoubtedly the isolation and generation of pluripotent stem cells, which will be the main focus of the remainder of the text.

Figure 1: Varying degrees of stem cell potency. Left: The fertilized egg (totipotent) develops into a 300-cell structure, the blastocyst, which contains embryonic stem cells (ES cells) at the inner cell mass (ICM). ES cells are pluripotent and can thus give rise to all cell types in our body, including adult stem cells, which range from multipotent to unipotent. Right: An alternative route to obtain pluripotent stem cells is the generation of induced pluripotent stem cells (iPS cells) from patients. Cell types obtained by differentiation of either ES cell (Left) or iPS cells (Right) can then be studied in the dish or used for transplantation into patients. Figure drawn by Hannah Somhegyi.

Martin Evans (Nobel Prize, 2007) and Matt Kauffman were the first to identify, isolate and successfully culture ES cells using mouse blastocysts in 1981 []. This discovery opened the doors to the creation of murine genetic models, which are mice that have had one or several of their genes deleted or otherwise modified to study their function in disease []. This is possible because scientists can modify the genome of a mouse in its ES cells and then inject those modified cells into mouse blastocysts. This means that when the blastocyst develops into an adult mouse, every cell its body will have the modification of interest.

The desire to use stem cells unique properties in medicine was greatly intensified when James Thomson and collaborators first isolated ES cells from human blastocysts []. For the first time, scientists could, in theory, generate all the building blocks of our body in unlimited amounts. It was possible to have cell types for testing new therapeutics and perhaps even new transplantation methods that were previously not possible. Yet, destroying human embryos to isolate cells presented ethical and technical hurdles. How could one circumvent that procedure? Sir John Gurdon showed in the early 1960s that, contrary to the prevalent belief back then, cells are not locked in their differentiation state and can be reverted to a more primitive state with a higher developmental potential. He demonstrated this principle by injecting the nucleus of a differentiated frog cell into an egg cell from which the nucleus had been removed. (This is commonly known as reproductive cloning, which was used to generate Dolly the Sheep.) When allowed to develop, this egg gave rise to a fertile adult frog, proving that differentiated cells retain the information required to give rise to all cell types in the body. More than forty years later, Shinya Yamanaka and colleagues shocked the world when they were able to convert skin cells called fibroblasts into pluripotent stem cells by altering the expression of just four genes []. This represented the birth of induced pluripotent stem cells, or iPS cells (see Figure 1, right column). The enormous importance of these findings is hard to overstate, and is perhaps best illustrated by the fact that, merely six years later, Gurdon and Yamanaka shared the Nobel Prize in Physiology or Medicine 2012 [].

Since the generation of iPS cells was first reported, the stem cell eld has expanded at an unparalleled pace. Today, these cells are the hope of personalized medicine, as they allow one to capture the unique genome of each individual in a cell type that can be used to generate, in principle, all cell types in our body, as illustrated on the right panel of Figure 1. The replacement of diseased tissues or organs without facing the barrier of immune rejection due to donor incompatibility thus becomes approachable in this era of iPS cells and is the object of intense research [].

The first proof-of-principle study showing that iPS cells can potentially be used to correct genetic diseases was carried out in the laboratory of Rudolf Jaenisch. In brief, tail tip cells from mice with a mutation causing sickle cell anemia were harvested and reprogrammed into iPS cells. The mutation was then corrected in these iPS cells, which were then differentiated into blood progenitor cells and transplanted back into the original mice, curing them []. Even though iPS cells have been found not to completely match ES cells in some instances, detailed studies have failed to find consistent differences between iPS and ES cells []. This similarity, together with the constant improvements in the efficiency and robustness of generating iPS cells, provides bright prospects for the future of stem cell research and stem cell-based treatments for degenerative diseases unapproachable with more conventional methods.

Leonardo M. R. Ferreira is a graduate student in Harvard Universitys Department of Molecular and Cellular Biology

[] Stem Cell Basics: http://stemcells.nih.gov/info/basics/Pages/Default.aspx

[] Becker, A. J., McCulloch, E.A., Till, J.E. Cytological demonstration of the clonal nature of spleen colonies derived from transplanted mouse marrow cells. Nature 1963. 197: 452-4

[] Siminovitch, L., McCulloch, E.A., Till, J.E. The distribution of colony-forming cells among spleen colonies. J Cell Comp Physiol 1963, 62(3): 327-336

[] Evans, M. J. and Kaufman, M. Establishment in culture of pluripotential stem cells from mouse embryos. Nature 1981, 292: 151156

[] Simmons, D. The Use of Animal Models in Studying Genetic Disease: Transgenesis and Induced Mutation. Nature Education 2008,1(1):70

[] Thomson, J. A. et al. Embryonic stem cell lines derived from human blastocysts. Science 1998, 282(5391): 1145-1147

[] Takahashi, K. and Yamanaka S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 2006. 126(4): 663-76

[] The Nobel Prize in Physiology or Medicine 2012:

[] Ferreira, L.M.R. and Mostajo-Radji, M.A. How induced pluripotent stem cells are redefining personalized medicine. Gene 2013. 520(1): 1-6 [] Hanna J. et al. Treatment of sickle cell anemia mouse model with iPS cells generated from autologous skin. Science 2007. 318: 1920-1923

[] Yee,J.Turning Somatic Cells into Pluripotent Stem Cells.Nature Education 2010.3(9):25

Excerpt from:
Stem cells: a brief history and outlook - Science in the News

Posted in Stem Cells | Comments Off on Stem cells: a brief history and outlook – Science in the News

Global Induced Pluripotent Stem Cell ((iPSC) Market to Reach $0 Thousand by 2027 – Yahoo Finance

Posted: October 13, 2022 at 1:43 am

ReportLinker

Abstract: Whats New for 2022?? Global competitiveness and key competitor percentage market shares. Market presence across multiple geographies - Strong/Active/Niche/Trivial.

New York, Oct. 10, 2022 (GLOBE NEWSWIRE) -- Reportlinker.com announces the release of the report "Global Induced Pluripotent Stem Cell (iPSC) Industry" - https://www.reportlinker.com/p05798831/?utm_source=GNW

Online interactive peer-to-peer collaborative bespoke updates

Access to our digital archives and MarketGlass Research Platform

Complimentary updates for one yearGlobal Induced Pluripotent Stem Cell ((iPSC) Market to Reach $0 Thousand by 2027- In the changed post COVID-19 business landscape, the global market for Induced Pluripotent Stem Cell ((iPSC) estimated at US$1.4 Billion in the year 2020, is projected to reach a revised size of US$0 Thousand by 2027, growing at a CAGR of -100% over the analysis period 2020-2027. Vascular Cells, one of the segments analyzed in the report, is projected to record a -100% CAGR and reach US$0 Thousand by the end of the analysis period. Taking into account the ongoing post pandemic recovery, growth in the Cardiac Cells segment is readjusted to a revised -100% CAGR for the next 7-year period.- The U.S. Market is Estimated at $629.2 Million, While China is Forecast to Grow at -100% CAGR- The Induced Pluripotent Stem Cell ((iPSC) market in the U.S. is estimated at US$629.2 Million in the year 2020. China, the world`s second largest economy, is forecast to reach a projected market size of US$0 Thousand by the year 2027 trailing a CAGR of -100% over the analysis period 2020 to 2027. Among the other noteworthy geographic markets are Japan and Canada, each forecast to grow at -100% and -100% respectively over the 2020-2027 period. Within Europe, Germany is forecast to grow at approximately -100% CAGR.Neuronal Cells Segment to Record -100% CAGR- In the global Neuronal Cells segment, USA, Canada, Japan, China and Europe will drive the -100% CAGR estimated for this segment. These regional markets accounting for a combined market size of US$188.9 Million in the year 2020 will reach a projected size of US$0 Thousand by the close of the analysis period. China will remain among the fastest growing in this cluster of regional markets.

Select Competitors (Total 51 Featured)Axol Bioscience Ltd.Cynata Therapeutics LimitedEvotec SEFate Therapeutics, Inc.FUJIFILM Cellular Dynamics, Inc.NcardiaPluricell BiotechREPROCELL USA, Inc.Sumitomo Dainippon Pharma Co., Ltd.Takara Bio, Inc.Thermo Fisher Scientific, Inc.ViaCyte, Inc.

Read the full report: https://www.reportlinker.com/p05798831/?utm_source=GNW

I. METHODOLOGY

II. EXECUTIVE SUMMARY

1. MARKET OVERVIEWInfluencer Market InsightsImpact of Covid-19 and a Looming Global RecessionInduced Pluripotent Stem Cells (iPSCs) Market Gains fromIncreasing Use in Research for COVID-19Studies Employing iPSCs in COVID-19 ResearchStem Cells, Application Areas, and the Different Types: A PreludeApplications of Stem CellsTypes of Stem CellsInduced Pluripotent Stem Cell (iPSC): An IntroductionProduction of iPSCsFirst & Second Generation Mouse iPSCsHuman iPSCsKey Properties of iPSCsTranscription Factors Involved in Generation of iPSCsNoteworthy Research & Application Areas for iPSCsInduced Pluripotent Stem Cell ((iPSC) Market: Growth Prospectsand OutlookDrug Development Application to Witness Considerable GrowthTechnical Breakthroughs, Advances & Clinical Trials to SpurGrowth of iPSC MarketNorth America Dominates Global iPSC MarketCompetitionRecent Market ActivitySelect Innovation/AdvancementInduced Pluripotent Stem Cell (iPSC) - Global Key CompetitorsPercentage Market Share in 2022 (E)Competitive Market Presence - Strong/Active/Niche/Trivial forPlayers Worldwide in 2022 (E)

2. FOCUS ON SELECT PLAYERSAxol Bioscience Ltd. (UK)Cynata Therapeutics Limited (Australia)Evotec SE (Germany)Fate Therapeutics, Inc. (USA)FUJIFILM Cellular Dynamics, Inc. (USA)Ncardia (Belgium)Pluricell Biotech (Brazil)REPROCELL USA, Inc. (USA)Sumitomo Dainippon Pharma Co., Ltd. (Japan)Takara Bio, Inc. (Japan)Thermo Fisher Scientific, Inc. (USA)ViaCyte, Inc. (USA)

3. MARKET TRENDS & DRIVERSEffective Research Programs Hold Key in Roll Out of AdvancediPSC TreatmentsInduced Pluripotent Stem Cells: A Giant Leap in the TherapeuticApplicationsResearch Trends in Induced Pluripotent Stem Cell SpaceWorldwide Publication of hESC and hiPSC Research Papers for thePeriod 2008-2010, 2011-2013 and 2014-2016Number of Original Research Papers on hESC and iPSC PublishedWorldwide (2014-2016)Concerns Related to Embryonic Stem Cells Shift the Focus ontoiPSCsRegenerative Medicine: A Promising Application of iPSCsInduced Pluripotent: A Potential Competitor to hESCs?Global Regenerative Medicine Market Size in US$ Billion for2019, 2021, 2023 and 2025Global Stem Cell & Regenerative Medicine Market by Product(in %) for the Year 2019Global Regenerative Medicines Market by Category: Breakdown(in %) for Biomaterials, Stem Cell Therapies and TissueEngineering for 2019Pluripotent Stem Cells Hold Significance for CardiovascularRegenerative MedicineLeading Causes of Mortality Worldwide: Number of Deaths inMillions & % Share of Deaths by Cause for 2017Leading Causes of Mortality for Low-Income and High-IncomeCountriesGrowing Importance of iPSCs in Personalized Drug DiscoveryPersistent Advancements in Genetics Space and Subsequent Growthin Precision Medicine Augur Well for iPSCs MarketGlobal Precision Medicine Market (In US$ Billion) for the Years2018, 2021 & 2024Increasing Prevalence of Chronic Disorders Supports Growth ofiPSCs MarketWorldwide Cancer Incidence: Number of New Cancer CasesDiagnosed for 2012, 2018 & 2040Number of New Cancer Cases Reported (in Thousands) by CancerType: 2018Fatalities by Heart Conditions: Estimated Percentage Breakdownfor Cardiovascular Disease, Ischemic Heart Disease, Stroke,and OthersRising Diabetes Prevalence Presents Opportunity for iPSCsMarket: Number of Adults (20-79) with Diabetes (in Millions)by Region for 2017 and 2045Aging Demographics Add to the Global Burden of ChronicDiseases, Presenting Opportunities for iPSCs MarketExpanding Elderly Population Worldwide: Breakdown of Number ofPeople Aged 65+ Years in Million by Geographic Region for theYears 2019 and 2030Growth in Number of Genomics Projects Propels Market GrowthGenomic Initiatives in Select CountriesNew Gene-Editing Tools Spur Interest and Investments inGenetics, Driving Lucrative Growth Opportunities for iPSCs:Total VC Funding (In US$ Million) in Genetics for the Years2014, 2015, 2016, 2017 and 2018Launch of Numerous iPSCs-Related Clinical Trials Set to BenefitMarket GrowthNumber of Induced Pluripotent Stem Cells based Studies bySelect Condition: As on Oct 31, 2020iPSCs-based Clinical Trial for Heart DiseasesInduced Pluripotent Stem Cells for Stroke Treatment?Off-the-shelf? Stem Cell Treatment for Cancer Enters ClinicalTrialiPSCs for Hematological DisordersMarket Benefits from Growing Funding for iPSCs-Related R&DInitiativesStem Cell Research Funding in the US (in US$ Million) for theYears 2016 through 2021Human iPSC Banks: A Review of Emerging Opportunities and DrawbacksHuman iPSC Banks Worldwide: An OverviewCell Sources and Reprogramming Methods Used by Select iPSC BanksInnovations, Research Studies & Advancements in iPSCsKey iPSC Research Breakthroughs for Regenerative MedicineResearchers Develop Novel Oncogene-Free and Virus-Free iPSCProduction MethodScientists Study Concerns of Genetic Mutations in iPSCsiPSCs Hold Tremendous Potential in Transforming Research EffortsResearchers Highlight Potential Use of iPSCs for DevelopingNovel Cancer VaccinesScientists Use Machine Learning to Improve Reliability of iPSCSelf-OrganizationSTEMCELL Technologies Unveils mTeSR? PlusChallenges and Risks Related to Pluripotent Stem CellsA Glance at Issues Related to Reprogramming of Adult Cells toiPSCsA Note on Legal, Social and Ethical Considerations with iPSCs

4. GLOBAL MARKET PERSPECTIVETable 1: World Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Geographic Region -USA, Canada, Japan, China, Europe, Asia-Pacific and Rest ofWorld Markets - Independent Analysis of Annual Sales in US$Thousand for Years 2020 through 2025 and % CAGR

Table 2: World 5-Year Perspective for Induced Pluripotent StemCell (iPSC) by Geographic Region - Percentage Breakdown ofValue Sales for USA, Canada, Japan, China, Europe, Asia-Pacificand Rest of World Markets for Years 2021 & 2025

Table 3: World Recent Past, Current & Future Analysis forVascular Cells by Geographic Region - USA, Canada, Japan,China, Europe, Asia-Pacific and Rest of World Markets -Independent Analysis of Annual Sales in US$ Thousand for Years2020 through 2025 and % CAGR

Table 4: World 5-Year Perspective for Vascular Cells byGeographic Region - Percentage Breakdown of Value Sales forUSA, Canada, Japan, China, Europe, Asia-Pacific and Rest ofWorld for Years 2021 & 2025

Table 5: World Recent Past, Current & Future Analysis forCardiac Cells by Geographic Region - USA, Canada, Japan, China,Europe, Asia-Pacific and Rest of World Markets - IndependentAnalysis of Annual Sales in US$ Thousand for Years 2020 through2025 and % CAGR

Table 6: World 5-Year Perspective for Cardiac Cells byGeographic Region - Percentage Breakdown of Value Sales forUSA, Canada, Japan, China, Europe, Asia-Pacific and Rest ofWorld for Years 2021 & 2025

Table 7: World Recent Past, Current & Future Analysis forNeuronal Cells by Geographic Region - USA, Canada, Japan,China, Europe, Asia-Pacific and Rest of World Markets -Independent Analysis of Annual Sales in US$ Thousand for Years2020 through 2025 and % CAGR

Table 8: World 5-Year Perspective for Neuronal Cells byGeographic Region - Percentage Breakdown of Value Sales forUSA, Canada, Japan, China, Europe, Asia-Pacific and Rest ofWorld for Years 2021 & 2025

Table 9: World Recent Past, Current & Future Analysis for LiverCells by Geographic Region - USA, Canada, Japan, China, Europe,Asia-Pacific and Rest of World Markets - Independent Analysisof Annual Sales in US$ Thousand for Years 2020 through 2025 and% CAGR

Table 10: World 5-Year Perspective for Liver Cells byGeographic Region - Percentage Breakdown of Value Sales forUSA, Canada, Japan, China, Europe, Asia-Pacific and Rest ofWorld for Years 2021 & 2025

Table 11: World Recent Past, Current & Future Analysis forImmune Cells by Geographic Region - USA, Canada, Japan, China,Europe, Asia-Pacific and Rest of World Markets - IndependentAnalysis of Annual Sales in US$ Thousand for Years 2020 through2025 and % CAGR

Table 12: World 5-Year Perspective for Immune Cells byGeographic Region - Percentage Breakdown of Value Sales forUSA, Canada, Japan, China, Europe, Asia-Pacific and Rest ofWorld for Years 2021 & 2025

Table 13: World Recent Past, Current & Future Analysis forOther Cell Types by Geographic Region - USA, Canada, Japan,China, Europe, Asia-Pacific and Rest of World Markets -Independent Analysis of Annual Sales in US$ Thousand for Years2020 through 2025 and % CAGR

Table 14: World 5-Year Perspective for Other Cell Types byGeographic Region - Percentage Breakdown of Value Sales forUSA, Canada, Japan, China, Europe, Asia-Pacific and Rest ofWorld for Years 2021 & 2025

Table 15: World Recent Past, Current & Future Analysis forCellular Reprogramming by Geographic Region - USA, Canada,Japan, China, Europe, Asia-Pacific and Rest of World Markets -Independent Analysis of Annual Sales in US$ Thousand for Years2020 through 2025 and % CAGR

Table 16: World 5-Year Perspective for Cellular Reprogrammingby Geographic Region - Percentage Breakdown of Value Sales forUSA, Canada, Japan, China, Europe, Asia-Pacific and Rest ofWorld for Years 2021 & 2025

Table 17: World Recent Past, Current & Future Analysis for CellCulture by Geographic Region - USA, Canada, Japan, China,Europe, Asia-Pacific and Rest of World Markets - IndependentAnalysis of Annual Sales in US$ Thousand for Years 2020 through2025 and % CAGR

Table 18: World 5-Year Perspective for Cell Culture byGeographic Region - Percentage Breakdown of Value Sales forUSA, Canada, Japan, China, Europe, Asia-Pacific and Rest ofWorld for Years 2021 & 2025

Table 19: World Recent Past, Current & Future Analysis for CellDifferentiation by Geographic Region - USA, Canada, Japan,China, Europe, Asia-Pacific and Rest of World Markets -Independent Analysis of Annual Sales in US$ Thousand for Years2020 through 2025 and % CAGR

Table 20: World 5-Year Perspective for Cell Differentiation byGeographic Region - Percentage Breakdown of Value Sales forUSA, Canada, Japan, China, Europe, Asia-Pacific and Rest ofWorld for Years 2021 & 2025

Table 21: World Recent Past, Current & Future Analysis for CellAnalysis by Geographic Region - USA, Canada, Japan, China,Europe, Asia-Pacific and Rest of World Markets - IndependentAnalysis of Annual Sales in US$ Thousand for Years 2020 through2025 and % CAGR

Table 22: World 5-Year Perspective for Cell Analysis byGeographic Region - Percentage Breakdown of Value Sales forUSA, Canada, Japan, China, Europe, Asia-Pacific and Rest ofWorld for Years 2021 & 2025

Table 23: World Recent Past, Current & Future Analysis forCellular Engineering by Geographic Region - USA, Canada, Japan,China, Europe, Asia-Pacific and Rest of World Markets -Independent Analysis of Annual Sales in US$ Thousand for Years2020 through 2025 and % CAGR

Table 24: World 5-Year Perspective for Cellular Engineering byGeographic Region - Percentage Breakdown of Value Sales forUSA, Canada, Japan, China, Europe, Asia-Pacific and Rest ofWorld for Years 2021 & 2025

Table 25: World Recent Past, Current & Future Analysis forOther Research Methods by Geographic Region - USA, Canada,Japan, China, Europe, Asia-Pacific and Rest of World Markets -Independent Analysis of Annual Sales in US$ Thousand for Years2020 through 2025 and % CAGR

Table 26: World 5-Year Perspective for Other Research Methodsby Geographic Region - Percentage Breakdown of Value Sales forUSA, Canada, Japan, China, Europe, Asia-Pacific and Rest ofWorld for Years 2021 & 2025

Table 27: World Recent Past, Current & Future Analysis for DrugDevelopment & Toxicology Testing by Geographic Region - USA,Canada, Japan, China, Europe, Asia-Pacific and Rest of WorldMarkets - Independent Analysis of Annual Sales in US$ Thousandfor Years 2020 through 2025 and % CAGR

Table 28: World 5-Year Perspective for Drug Development &Toxicology Testing by Geographic Region - Percentage Breakdownof Value Sales for USA, Canada, Japan, China, Europe,Asia-Pacific and Rest of World for Years 2021 & 2025

Table 29: World Recent Past, Current & Future Analysis forAcademic Research by Geographic Region - USA, Canada, Japan,China, Europe, Asia-Pacific and Rest of World Markets -Independent Analysis of Annual Sales in US$ Thousand for Years2020 through 2025 and % CAGR

Table 30: World 5-Year Perspective for Academic Research byGeographic Region - Percentage Breakdown of Value Sales forUSA, Canada, Japan, China, Europe, Asia-Pacific and Rest ofWorld for Years 2021 & 2025

Table 31: World Recent Past, Current & Future Analysis forRegenerative Medicine by Geographic Region - USA, Canada,Japan, China, Europe, Asia-Pacific and Rest of World Markets -Independent Analysis of Annual Sales in US$ Thousand for Years2020 through 2025 and % CAGR

Table 32: World 5-Year Perspective for Regenerative Medicine byGeographic Region - Percentage Breakdown of Value Sales forUSA, Canada, Japan, China, Europe, Asia-Pacific and Rest ofWorld for Years 2021 & 2025

Table 33: World Recent Past, Current & Future Analysis forOther Applications by Geographic Region - USA, Canada, Japan,China, Europe, Asia-Pacific and Rest of World Markets -Independent Analysis of Annual Sales in US$ Thousand for Years2020 through 2025 and % CAGR

Table 34: World 5-Year Perspective for Other Applications byGeographic Region - Percentage Breakdown of Value Sales forUSA, Canada, Japan, China, Europe, Asia-Pacific and Rest ofWorld for Years 2021 & 2025

III. MARKET ANALYSIS

UNITED STATESInduced Pluripotent Stem Cell (iPSC) Market Presence - Strong/Active/Niche/Trivial - Key Competitors in the United Statesfor 2022 (E)Table 35: USA Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Cell Type - VascularCells, Cardiac Cells, Neuronal Cells, Liver Cells, Immune Cellsand Other Cell Types - Independent Analysis of Annual Sales inUS$ Thousand for the Years 2020 through 2025 and % CAGR

Table 36: USA 5-Year Perspective for Induced Pluripotent StemCell (iPSC) by Cell Type - Percentage Breakdown of Value Salesfor Vascular Cells, Cardiac Cells, Neuronal Cells, Liver Cells,Immune Cells and Other Cell Types for the Years 2021 & 2025

Table 37: USA Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Research Method -Cellular Reprogramming, Cell Culture, Cell Differentiation,Cell Analysis, Cellular Engineering and Other Research Methods -Independent Analysis of Annual Sales in US$ Thousand for theYears 2020 through 2025 and % CAGR

Table 38: USA 5-Year Perspective for Induced Pluripotent StemCell (iPSC) by Research Method - Percentage Breakdown of ValueSales for Cellular Reprogramming, Cell Culture, CellDifferentiation, Cell Analysis, Cellular Engineering and OtherResearch Methods for the Years 2021 & 2025

Table 39: USA Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Application - DrugDevelopment & Toxicology Testing, Academic Research,Regenerative Medicine and Other Applications - IndependentAnalysis of Annual Sales in US$ Thousand for the Years 2020through 2025 and % CAGR

Table 40: USA 5-Year Perspective for Induced Pluripotent StemCell (iPSC) by Application - Percentage Breakdown of ValueSales for Drug Development & Toxicology Testing, AcademicResearch, Regenerative Medicine and Other Applications for theYears 2021 & 2025

CANADATable 41: Canada Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Cell Type - VascularCells, Cardiac Cells, Neuronal Cells, Liver Cells, Immune Cellsand Other Cell Types - Independent Analysis of Annual Sales inUS$ Thousand for the Years 2020 through 2025 and % CAGR

Table 42: Canada 5-Year Perspective for Induced PluripotentStem Cell (iPSC) by Cell Type - Percentage Breakdown of ValueSales for Vascular Cells, Cardiac Cells, Neuronal Cells, LiverCells, Immune Cells and Other Cell Types for the Years 2021 &2025

Table 43: Canada Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Research Method -Cellular Reprogramming, Cell Culture, Cell Differentiation,Cell Analysis, Cellular Engineering and Other Research Methods -Independent Analysis of Annual Sales in US$ Thousand for theYears 2020 through 2025 and % CAGR

Table 44: Canada 5-Year Perspective for Induced PluripotentStem Cell (iPSC) by Research Method - Percentage Breakdown ofValue Sales for Cellular Reprogramming, Cell Culture, CellDifferentiation, Cell Analysis, Cellular Engineering and OtherResearch Methods for the Years 2021 & 2025

Table 45: Canada Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Application - DrugDevelopment & Toxicology Testing, Academic Research,Regenerative Medicine and Other Applications - IndependentAnalysis of Annual Sales in US$ Thousand for the Years 2020through 2025 and % CAGR

Table 46: Canada 5-Year Perspective for Induced PluripotentStem Cell (iPSC) by Application - Percentage Breakdown of ValueSales for Drug Development & Toxicology Testing, AcademicResearch, Regenerative Medicine and Other Applications for theYears 2021 & 2025

JAPANInduced Pluripotent Stem Cell (iPSC) Market Presence - Strong/Active/Niche/Trivial - Key Competitors in Japan for 2022 (E)Table 47: Japan Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Cell Type - VascularCells, Cardiac Cells, Neuronal Cells, Liver Cells, Immune Cellsand Other Cell Types - Independent Analysis of Annual Sales inUS$ Thousand for the Years 2020 through 2025 and % CAGR

Table 48: Japan 5-Year Perspective for Induced Pluripotent StemCell (iPSC) by Cell Type - Percentage Breakdown of Value Salesfor Vascular Cells, Cardiac Cells, Neuronal Cells, Liver Cells,Immune Cells and Other Cell Types for the Years 2021 & 2025

Table 49: Japan Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Research Method -Cellular Reprogramming, Cell Culture, Cell Differentiation,Cell Analysis, Cellular Engineering and Other Research Methods -Independent Analysis of Annual Sales in US$ Thousand for theYears 2020 through 2025 and % CAGR

Table 50: Japan 5-Year Perspective for Induced Pluripotent StemCell (iPSC) by Research Method - Percentage Breakdown of ValueSales for Cellular Reprogramming, Cell Culture, CellDifferentiation, Cell Analysis, Cellular Engineering and OtherResearch Methods for the Years 2021 & 2025

Table 51: Japan Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Application - DrugDevelopment & Toxicology Testing, Academic Research,Regenerative Medicine and Other Applications - IndependentAnalysis of Annual Sales in US$ Thousand for the Years 2020through 2025 and % CAGR

Table 52: Japan 5-Year Perspective for Induced Pluripotent StemCell (iPSC) by Application - Percentage Breakdown of ValueSales for Drug Development & Toxicology Testing, AcademicResearch, Regenerative Medicine and Other Applications for theYears 2021 & 2025

CHINAInduced Pluripotent Stem Cell (iPSC) Market Presence - Strong/Active/Niche/Trivial - Key Competitors in China for 2022 (E)Table 53: China Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Cell Type - VascularCells, Cardiac Cells, Neuronal Cells, Liver Cells, Immune Cellsand Other Cell Types - Independent Analysis of Annual Sales inUS$ Thousand for the Years 2020 through 2025 and % CAGR

Table 54: China 5-Year Perspective for Induced Pluripotent StemCell (iPSC) by Cell Type - Percentage Breakdown of Value Salesfor Vascular Cells, Cardiac Cells, Neuronal Cells, Liver Cells,Immune Cells and Other Cell Types for the Years 2021 & 2025

Table 55: China Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Research Method -Cellular Reprogramming, Cell Culture, Cell Differentiation,Cell Analysis, Cellular Engineering and Other Research Methods -Independent Analysis of Annual Sales in US$ Thousand for theYears 2020 through 2025 and % CAGR

Table 56: China 5-Year Perspective for Induced Pluripotent StemCell (iPSC) by Research Method - Percentage Breakdown of ValueSales for Cellular Reprogramming, Cell Culture, CellDifferentiation, Cell Analysis, Cellular Engineering and OtherResearch Methods for the Years 2021 & 2025

Table 57: China Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Application - DrugDevelopment & Toxicology Testing, Academic Research,Regenerative Medicine and Other Applications - IndependentAnalysis of Annual Sales in US$ Thousand for the Years 2020through 2025 and % CAGR

Table 58: China 5-Year Perspective for Induced Pluripotent StemCell (iPSC) by Application - Percentage Breakdown of ValueSales for Drug Development & Toxicology Testing, AcademicResearch, Regenerative Medicine and Other Applications for theYears 2021 & 2025

EUROPEInduced Pluripotent Stem Cell (iPSC) Market Presence - Strong/Active/Niche/Trivial - Key Competitors in Europe for 2022 (E)Table 59: Europe Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Geographic Region -France, Germany, Italy, UK and Rest of Europe Markets -Independent Analysis of Annual Sales in US$ Thousand for Years2020 through 2025 and % CAGR

Table 60: Europe 5-Year Perspective for Induced PluripotentStem Cell (iPSC) by Geographic Region - Percentage Breakdown ofValue Sales for France, Germany, Italy, UK and Rest of EuropeMarkets for Years 2021 & 2025

Table 61: Europe Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Cell Type - VascularCells, Cardiac Cells, Neuronal Cells, Liver Cells, Immune Cellsand Other Cell Types - Independent Analysis of Annual Sales inUS$ Thousand for the Years 2020 through 2025 and % CAGR

Table 62: Europe 5-Year Perspective for Induced PluripotentStem Cell (iPSC) by Cell Type - Percentage Breakdown of ValueSales for Vascular Cells, Cardiac Cells, Neuronal Cells, LiverCells, Immune Cells and Other Cell Types for the Years 2021 &2025

Table 63: Europe Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Research Method -Cellular Reprogramming, Cell Culture, Cell Differentiation,Cell Analysis, Cellular Engineering and Other Research Methods -Independent Analysis of Annual Sales in US$ Thousand for theYears 2020 through 2025 and % CAGR

Table 64: Europe 5-Year Perspective for Induced PluripotentStem Cell (iPSC) by Research Method - Percentage Breakdown ofValue Sales for Cellular Reprogramming, Cell Culture, CellDifferentiation, Cell Analysis, Cellular Engineering and OtherResearch Methods for the Years 2021 & 2025

Table 65: Europe Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Application - DrugDevelopment & Toxicology Testing, Academic Research,Regenerative Medicine and Other Applications - IndependentAnalysis of Annual Sales in US$ Thousand for the Years 2020through 2025 and % CAGR

Table 66: Europe 5-Year Perspective for Induced PluripotentStem Cell (iPSC) by Application - Percentage Breakdown of ValueSales for Drug Development & Toxicology Testing, AcademicResearch, Regenerative Medicine and Other Applications for theYears 2021 & 2025

FRANCEInduced Pluripotent Stem Cell (iPSC) Market Presence - Strong/Active/Niche/Trivial - Key Competitors in France for 2022 (E)Table 67: France Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Cell Type - VascularCells, Cardiac Cells, Neuronal Cells, Liver Cells, Immune Cellsand Other Cell Types - Independent Analysis of Annual Sales inUS$ Thousand for the Years 2020 through 2025 and % CAGR

Table 68: France 5-Year Perspective for Induced PluripotentStem Cell (iPSC) by Cell Type - Percentage Breakdown of ValueSales for Vascular Cells, Cardiac Cells, Neuronal Cells, LiverCells, Immune Cells and Other Cell Types for the Years 2021 &2025

Table 69: France Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Research Method -Cellular Reprogramming, Cell Culture, Cell Differentiation,Cell Analysis, Cellular Engineering and Other Research Methods -Independent Analysis of Annual Sales in US$ Thousand for theYears 2020 through 2025 and % CAGR

Table 70: France 5-Year Perspective for Induced PluripotentStem Cell (iPSC) by Research Method - Percentage Breakdown ofValue Sales for Cellular Reprogramming, Cell Culture, CellDifferentiation, Cell Analysis, Cellular Engineering and OtherResearch Methods for the Years 2021 & 2025

Table 71: France Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Application - DrugDevelopment & Toxicology Testing, Academic Research,Regenerative Medicine and Other Applications - IndependentAnalysis of Annual Sales in US$ Thousand for the Years 2020through 2025 and % CAGR

Table 72: France 5-Year Perspective for Induced PluripotentStem Cell (iPSC) by Application - Percentage Breakdown of ValueSales for Drug Development & Toxicology Testing, AcademicResearch, Regenerative Medicine and Other Applications for theYears 2021 & 2025

GERMANYInduced Pluripotent Stem Cell (iPSC) Market Presence - Strong/Active/Niche/Trivial - Key Competitors in Germany for 2022 (E)Table 73: Germany Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Cell Type - VascularCells, Cardiac Cells, Neuronal Cells, Liver Cells, Immune Cellsand Other Cell Types - Independent Analysis of Annual Sales inUS$ Thousand for the Years 2020 through 2025 and % CAGR

Table 74: Germany 5-Year Perspective for Induced PluripotentStem Cell (iPSC) by Cell Type - Percentage Breakdown of ValueSales for Vascular Cells, Cardiac Cells, Neuronal Cells, LiverCells, Immune Cells and Other Cell Types for the Years 2021 &2025

Table 75: Germany Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Research Method -Cellular Reprogramming, Cell Culture, Cell Differentiation,Cell Analysis, Cellular Engineering and Other Research Methods -Independent Analysis of Annual Sales in US$ Thousand for theYears 2020 through 2025 and % CAGR

Table 76: Germany 5-Year Perspective for Induced PluripotentStem Cell (iPSC) by Research Method - Percentage Breakdown ofValue Sales for Cellular Reprogramming, Cell Culture, CellDifferentiation, Cell Analysis, Cellular Engineering and OtherResearch Methods for the Years 2021 & 2025

Table 77: Germany Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Application - DrugDevelopment & Toxicology Testing, Academic Research,Regenerative Medicine and Other Applications - IndependentAnalysis of Annual Sales in US$ Thousand for the Years 2020through 2025 and % CAGR

Table 78: Germany 5-Year Perspective for Induced PluripotentStem Cell (iPSC) by Application - Percentage Breakdown of ValueSales for Drug Development & Toxicology Testing, AcademicResearch, Regenerative Medicine and Other Applications for theYears 2021 & 2025

ITALYTable 79: Italy Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Cell Type - VascularCells, Cardiac Cells, Neuronal Cells, Liver Cells, Immune Cellsand Other Cell Types - Independent Analysis of Annual Sales inUS$ Thousand for the Years 2020 through 2025 and % CAGR

Table 80: Italy 5-Year Perspective for Induced Pluripotent StemCell (iPSC) by Cell Type - Percentage Breakdown of Value Salesfor Vascular Cells, Cardiac Cells, Neuronal Cells, Liver Cells,Immune Cells and Other Cell Types for the Years 2021 & 2025

Table 81: Italy Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Research Method -Cellular Reprogramming, Cell Culture, Cell Differentiation,Cell Analysis, Cellular Engineering and Other Research Methods -Independent Analysis of Annual Sales in US$ Thousand for theYears 2020 through 2025 and % CAGR

Table 82: Italy 5-Year Perspective for Induced Pluripotent StemCell (iPSC) by Research Method - Percentage Breakdown of ValueSales for Cellular Reprogramming, Cell Culture, CellDifferentiation, Cell Analysis, Cellular Engineering and OtherResearch Methods for the Years 2021 & 2025

Table 83: Italy Recent Past, Current & Future Analysis forInduced Pluripotent Stem Cell (iPSC) by Application - DrugDevelopment & Toxicology Testing, Academic Research,Regenerative Medicine and Other Applications - IndependentAnalysis of Annual Sales in US$ Thousand for the Years 2020through 2025 and % CAGR

Table 84: Italy 5-Year Perspective for Induced Pluripotent StemCell (iPSC) by Application - Percentage Breakdown of ValueSales for Drug Development & Toxicology Testing, AcademicResearch, Regenerative Medicine and Other Applications for theYears 2021 & 2025

UNITED KINGDOMInduced Pluripotent Stem Cell (iPSC) Market Presence - Strong/Active/Niche/Trivial - Key Competitors in the United Kingdomfor 2022 (E)Table 85: UK Recent Past, Current & Future Analysis for InducedPluripotent Stem Cell (iPSC) by Cell Type - Vascular Cells,Cardiac Cells, Neuronal Cells, Liver Cells, Immune Cells andOther Cell Types - Independent Analysis of Annual Sales in US$Thousand for the Years 2020 through 2025 and % CAGR

Table 86: UK 5-Year Perspective for Induced Pluripotent StemCell (iPSC) by Cell Type - Percentage Breakdown of Value Salesfor Vascular Cells, Cardiac Cells, Neuronal Cells, Liver Cells,Immune Cells and Other Cell Types for the Years 2021 & 2025

See the original post here:
Global Induced Pluripotent Stem Cell ((iPSC) Market to Reach $0 Thousand by 2027 - Yahoo Finance

Posted in New York Stem Cells | Comments Off on Global Induced Pluripotent Stem Cell ((iPSC) Market to Reach $0 Thousand by 2027 – Yahoo Finance

Hereditary E200K mutation within the prion protein gene alters human iPSC derived cardiomyocyte function | Scientific Reports – Nature.com

Posted: September 25, 2022 at 2:14 am

Takada, L. T. et al. Genetic prion disease: Experience of a rapidly progressive dementia center in the United States and a review of the literature. Am. J. Med. Genet. B. Neuropsychiatr. Genet. 174(1), 3669 (2017).

PubMed PubMed Central Article Google Scholar

Minikel, E. V. et al. Quantifying prion disease penetrance using large population control cohorts. Sci. Transl. Med. 8(322), 322329 (2016).

Google Scholar

Ashwath, M. L., Dearmond, S. J. & Culclasure, T. Prion-associated dilated cardiomyopathy. Arch. Internal Med. 165(3), 338340 (2005).

Article Google Scholar

Itzhaki-Ben-Zadok, O., Orvin, K., Inbar, E. & Rechavia, E. Cardiomyopathy associated with Ceutzfeld-Jakob disease: A diagnosis of exclusion: A case report. Eur. Heart J. Case Rep. 4(1), 15 (2020).

PubMed PubMed Central Article Google Scholar

Cortelli, P. et al. Cardiovascular dysautonomia in fatal familial insomnia. Clin. Auton. Res. 1(1), 1521 (1991).

ADS CAS PubMed Article Google Scholar

Palma, J. A. et al. Cardiac autonomic impairment during sleep as a marker of human prion diseases: A preliminary report. Clin. Neurophysiol. 125(1), 208210 (2014).

PubMed Article Google Scholar

Garza, M. C., Monzn, M., Marn, B., Badiola, J. J. & Monlen, E. Distribution of peripheral PrP(Sc) in sheep with naturally acquired scrapie. PLoS ONE 9(5), e97768 (2014).

ADS PubMed PubMed Central Article CAS Google Scholar

Jewell, J. E., Brown, J., Kreeger, T. & Williams, E. S. Prion protein in cardiac muscle of elk (Cervus elaphus nelsoni) and white-tailed deer (Odocoileus virginianus) infected with chronic wasting disease. J. Gener. Virol. 87(Pt 11), 34433450 (2006).

CAS Article Google Scholar

Balachandran, A. et al. Experimental oral transmission of chronic wasting disease to red deer (Cervus elaphus elaphus): Early detection and late stage distribution of protease-resistant prion protein. Can. Vet. J. 51(2), 169178 (2010).

CAS PubMed PubMed Central Google Scholar

Bessen, R. A. et al. Transmission of chronic wasting disease identifies a prion strain causing cachexia and heart infection in hamsters. PLoS ONE 6(12), e28026 (2011).

ADS CAS PubMed PubMed Central Article Google Scholar

Mitchell, G. B. et al. Experimental oral transmission of chronic wasting disease to reindeer (Rangifer tarandus tarandus). PLoS ONE 7(6), e39055 (2012).

ADS CAS PubMed PubMed Central Article Google Scholar

Krasemann, S. et al. BSE-associated prion-amyloid cardiomyopathy in primates. Emerg. Infect. Dis. 19(6), 985988 (2013).

CAS PubMed PubMed Central Article Google Scholar

Trifilo, M. J. et al. Prion-induced amyloid heart disease with high blood infectivity in transgenic mice. Science (N.Y.) 313(5783), 9497 (2006).

ADS CAS Article Google Scholar

Nico, P. B. et al. Altered behavioural response to acute stress in mice lacking cellular prion protein. Behav. Brain Res. 162(2), 173181 (2005).

CAS PubMed Article Google Scholar

Nico, P. B. et al. Impaired exercise capacity, but unaltered mitochondrial respiration in skeletal or cardiac muscle of mice lacking cellular prion protein. Neurosci. Lett. 388(1), 2126 (2005).

CAS PubMed Article Google Scholar

Klamt, F. et al. Imbalance of antioxidant defense in mice lacking cellular prion protein. Free Radical Biol. Med. 30(10), 11371144 (2001).

CAS Article Google Scholar

Beraldo, F. H. et al. Mechanisms of neuroprotection against ischemic insult by stress-inducible phosphoprotein-1/prion protein complex. J. Neurochem. 145(1), 6879 (2018).

CAS PubMed PubMed Central Article Google Scholar

Rambold, A. S. et al. Stress-protective signalling of prion protein is corrupted by scrapie prions. EMBO J. 27(14), 19741984 (2008).

CAS PubMed PubMed Central Article Google Scholar

Dupiereux, I. et al. Protective effect of prion protein via the N-terminal region in mediating a protective effect on paraquat-induced oxidative injury in neuronal cells. J. Neurosci. Res. 86(3), 653659 (2008).

CAS PubMed Article Google Scholar

Haigh, C. L. Brown, DR: Prion protein reduces both oxidative and non-oxidative copper toxicity. J. Neurochem. 98(3), 677689 (2006).

CAS PubMed Article Google Scholar

Resenberger, U. K., Winklhofer, K. F. & Tatzelt, J. Neuroprotective and neurotoxic signaling by the prion protein. Top. Curr. Chem. 305, 101119 (2011).

CAS PubMed Article Google Scholar

Turnbaugh, J. A., Westergard, L., Unterberger, U., Biasini, E. & Harris, D. A. The N-terminal, polybasic region is critical for prion protein neuroprotective activity. PLoS ONE 6(9), e25675 (2011).

ADS CAS PubMed PubMed Central Article Google Scholar

Zhang, B. et al. Prion protein protects against renal ischemia/reperfusion injury. PLoS ONE 10(9), e0136923 (2015).

PubMed PubMed Central Article CAS Google Scholar

Zocche-Soprana, H., Canes-Souza, L., Debbas, V. & Martins Laurindo, F. R. Cellular prion protein (PrP(C)) and superoxide dismutase (SOD) in vascular cells under oxidative stress. Exp. Toxicol. Pathol. 63(3), 229236 (2011).

PubMed Article CAS Google Scholar

Zanetti, F. et al. The cellular prion protein counteracts cardiac oxidative stress. Cardiovasc. Res. 104(1), 93102 (2014).

CAS PubMed Article Google Scholar

Keller, G., Binyamin, O., Frid, K., Saada, A. & Gabizon, R. Mitochondrial dysfunction in preclinical genetic prion disease: A target for preventive treatment?. Neurobiol. Dis. 124, 5766 (2019).

CAS PubMed Article Google Scholar

Casini, S., Verkerk, A. O. & Remme, C. A. Human iPSC-derived cardiomyocytes for investigation of disease mechanisms and therapeutic strategies in inherited arrhythmia syndromes: Strengths and limitations. Cardiovasc. Drugs Therapy 31(3), 325344 (2017).

Article Google Scholar

Giacomelli, E. et al. Human-iPSC-derived cardiac stromal cells enhance maturation in 3D cardiac microtissues and reveal non-cardiomyocyte contributions to heart disease. Cell Stem Cell 26(6), 862-879.e811 (2020).

CAS PubMed PubMed Central Article Google Scholar

Koivumki, J. T. et al. Structural immaturity of human iPSC-derived cardiomyocytes: In silico investigation of effects on function and disease modeling. Front. Physiol. 9, 80 (2018).

PubMed PubMed Central Article Google Scholar

Bellin, M. et al. Isogenic human pluripotent stem cell pairs reveal the role of a KCNH2 mutation in long-QT syndrome. EMBO J. 32(24), 31613175 (2013).

CAS PubMed PubMed Central Article Google Scholar

Paci, M., Casini, S., Bellin, M., Hyttinen, J. & Severi, S. Large-scale simulation of the phenotypical variability induced by loss-of-function long QT mutations in human induced pluripotent stem cell cardiomyocytes. Int. J. Mol. Sci. 19, 11 (2018).

Article CAS Google Scholar

Rebs, S., Sedaghat-Hamedani, F., Kayvanpour, E., Meder, B. & Streckfuss-Bmeke, K. Generation of pluripotent stem cell lines and CRISPR/Cas9 modified isogenic controls from a patient with dilated cardiomyopathy harboring a RBM20 p.R634W mutation. Stem Cell Res. 47, 101901 (2020).

CAS PubMed Article Google Scholar

Bosman, A. et al. Perturbations of heart development and function in cardiomyocytes from human embryonic stem cells with trisomy 21. Stem Cells (Dayton, Ohio) 33(5), 14341446 (2015).

CAS Article Google Scholar

Foliaki, S. T. et al. Pathogenic prion protein isoforms are not present in cerebral organoids generated from asymptomatic donors carrying the E200K mutation associated with familial prion disease. Pathogens 9, 6 (2020).

Article Google Scholar

Baiardi, S. et al. Phenotypic diversity of genetic Creutzfeldt-Jakob disease: A histo-molecular-based classification. Acta Neuropathol. 2021, 5 (2021).

Google Scholar

Asante, E. A. et al. Absence of spontaneous disease and comparative prion susceptibility of transgenic mice expressing mutant human prion proteins. J. Gen. Virol. 90(Pt 3), 546558 (2009).

CAS PubMed PubMed Central Article Google Scholar

Foliaki, S. T. et al. Neuronal excitatory-to-inhibitory balance is altered in cerebral organoid models of genetic neurological diseases. Mol. Brain 14(1), 156 (2021).

CAS PubMed PubMed Central Article Google Scholar

Miele, G., Jeffrey, M., Turnbull, D., Manson, J. & Clinton, M. Ablation of cellular prion protein expression affects mitochondrial numbers and morphology. Biochem. Biophys. Res. Commun. 291(2), 372377 (2002).

CAS PubMed Article Google Scholar

Faris, R. et al. Cellular prion protein is present in mitochondria of healthy mice. Sci. Rep. 7, 41556 (2017).

ADS CAS PubMed PubMed Central Article Google Scholar

Flnes, I. H. et al. Mitochondrial respiratory chain deficiency correlates with the severity of neuropathology in sporadic Creutzfeldt-Jakob disease. Acta Neuropathol. Commun. 8(1), 50 (2020).

PubMed PubMed Central Article CAS Google Scholar

Paterson, A. W., Curtis, J. C. & Macleod, N. K. Complex I specific increase in superoxide formation and respiration rate by PrP-null mouse brain mitochondria. J. Neurochem. 105(1), 177191 (2008).

CAS PubMed Article Google Scholar

Forkink, M. et al. Mitochondrial hyperpolarization during chronic complex I inhibition is sustained by low activity of complex II, III, IV and V. Biochem. Biophys. Acta. 1837(8), 12471256 (2014).

CAS PubMed Google Scholar

Li, N. et al. Mitochondrial complex I inhibitor rotenone induces apoptosis through enhancing mitochondrial reactive oxygen species production. J. Biol. Chem. 278(10), 85168525 (2003).

CAS PubMed Article Google Scholar

Poleggi, A. et al. Age at onset of genetic (E200K) and sporadic Creutzfeldt-Jakob diseases is modulated by the CYP4X1 gene. J. Neurol. Neurosurg. Psychiatry 89(12), 12431249 (2018).

PubMed Article Google Scholar

Baylin, S. B. & Jones, P. A. Epigenetic determinants of cancer. Cold Spring Harb. Perspect. Biol. 8, 9 (2016).

Article CAS Google Scholar

Ilango, S., Paital, B., Jayachandran, P., Padma, P. R. & Nirmaladevi, R. Epigenetic alterations in cancer. Front. Biosci. (Landmark Ed.) 25, 10581109 (2020).

CAS Article Google Scholar

Mitrov, E. et al. Genetic Creutzfeldt-Jakob disease affected monozygotic twins: Analysis of survival time, age at death and possible exogenous risk factors. J. Clin. Neurosci. 66, 191195 (2019).

PubMed Article Google Scholar

Simon, E. S. et al. Creutzfeldt-Jakob disease profile in patients homozygous for the PRNP E200K mutation. Ann. Neurol. 47(2), 257260 (2000).

CAS PubMed Article Google Scholar

Groveman, B. R. et al. Sporadic creutzfeldt-jakob disease prion infection of human cerebral organoids. Acta Neuropathol. Commun. 7(1), 90 (2019).

PubMed Article Google Scholar

Henderson, D. M. et al. Rapid antemortem detection of CWD prions in deer saliva. PLoS ONE 8(9), e74377 (2013).

ADS CAS PubMed PubMed Central Article Google Scholar

See the original post here:
Hereditary E200K mutation within the prion protein gene alters human iPSC derived cardiomyocyte function | Scientific Reports - Nature.com

Posted in Stem Cells | Comments Off on Hereditary E200K mutation within the prion protein gene alters human iPSC derived cardiomyocyte function | Scientific Reports – Nature.com

Utilization of Modified Induced Pluripotent Stem Cells as the Advance | OPTH – Dove Medical Press

Posted: August 30, 2022 at 2:14 am

Introduction

Glaucoma is one of the optic neuropathy disorders characterized by the progressive degeneration of retinal ganglion cells (RGC), which eventually lead to cupping of the optic disc and decreased visual field.1 Glaucoma is also closely related to an increase in intraocular pressure caused by the damage of trabecular meshwork (TBM), which results in optic nerve damage, characterized by the loss of retinal ganglion cells.2,3 Globally, in 2020, more than 76 million people are suffering from glaucoma, and it is expected to increase to 111.8 million people by 2040.4,5 Glaucoma is also a severe and complex medical problem because it often causes blindness. According to the World Health Organization (WHO), the most common causes of blindness are cataracts (51%), followed by glaucoma (8%), and age-related macular degeneration (5%).6 This data shows that glaucoma is the worlds second most common cause of blindness after cataracts. Symptoms that are often asymptomatic at an early stage and the low public awareness have contributed to the disorders seriousness.

Handling and treating glaucoma cases is difficult, especially because no therapy can cure glaucoma. Current treatment, both medical and surgical, is focused solely on lowering intraocular pressure. Treatment of glaucoma cases should also be carried out for life to maintain normal intraocular pressure and prevent the progression of intraocular damage due to glaucoma.7 Based on these problems, innovation is needed to handle glaucoma effectively. Besides, solutions are also required to repair the damage to retinal ganglion cells in glaucoma. One of the therapies that researchers are trying to take advantage of is stem cell therapy, a technology where cells can develop into many specific cells desired.8 In cases of glaucoma or optic neuropathies, damaged RGCs can be replaced with new ones grown from stem cells.9 Another option for RGC regeneration is to use retinal stem cells to regenerate RGCs. Indeed, stem cell therapy relies on exogenous stem cell sources due to their limited availability. Currently, many stem cell therapies for eye diseases that are created and studied are limited to treating the damage of photoreceptors and retinal pigment epithelium. iPSC-derived RGCs can serve as an excellent model for formulating approaches to promote de novo-generated RGCs to connect with their targets. Therefore, researchers have been looking into the potential use of modified stem cell therapy to treat the intraocular injury in glaucoma cases.10

This review aims to synthesize and prove the efficacy and further modification of this method so that it can be eligible for treatment and can also give data collection for the scientific community. This systematic review is expected to provide detailed information regarding the possible applications of modified stem cell therapy in treating intraocular damage in glaucoma patients.

In the present literature review, literature regarding the potential utilization of stem cells as an advanced therapy for intraocular glaucomatous damage was searched. The stages of this literature review include five steps: i) identifying the research question, ii) identifying relevant studies, iii) study selection, iv) charting the data, and v) summarizing and reporting the results.

This literature review was conducted to answer the following research questions:

The literature search was carried out from January to February 2021. Keywords and synonyms used to conduct literature searches related to the research question are attached in Table 1. Boolean operators (OR, AND, NOT) combine keywords when searching for literature. The search was conducted on seven online databases, namely PubMed, ScienceDirect, ProQuest, EBSCOhost, SAGE, Clinicalkey, and Scopus.

Table 1 Keywords That Were Used in the Database Search

The inclusion criteria for the literature search consisted of journals published in English and journals published in the last ten years. The exclusion criteria for selected studies consisted of journals that were not fully accessible due to the limited facilities owned as supporting access. We thoroughly screened the titles and abstracts of the studies obtained to suit the purpose of this literature review. Abstracts that were not relevant to the research objectives were excluded. Then a full article screening was carried out from the selected abstracts to identify whether the full article was suitable for the research objectives and whether the full article could be used to answer research questions.

Information obtained from all selected study articles is then displayed in the charting table The information displayed includes the author, year of publication, study objectives, location, study design, inclusion and exclusion criteria, results, and conclusions.

The researcher did not assess the quality of the selected articles because this study was only a literature review. The data from selected studies are reported to produce recommendations for further research regarding the use of stem cell therapy in glaucoma cases.

Based on the literature search that has been conducted, a total of 2262 studies and abstracts were included in the journal screening process at an early stage. From this screening process, 362 duplicate articles were excluded from the selection. The remaining 1900 articles then entered the abstract eligibility screening stage. Only 53 articles were selected, while 1879 other articles were excluded. Of the 53 articles, 18 articles appeared relevant to the study and met the inclusion criteria for review throughout the study. Meanwhile, 35 other studies were excluded because the focus in these studies did not match the objectives of this literature review. After assessing the full articles, six studies met the inclusion criteria in this literature review (Figure 1).

Figure 1 Flow diagram of the literature review process.

In Table 2, a summary of the characteristics of the selected studies is presented. The data used from selected studies include research objectivity, study design, results, outputs, and conclusions from the study. Of all the selected studies, there were six studies that had experimental methods. Almost all studies have the aim of evaluating and proving the potential of using stem cells to replace damaged tissue and restore and restore the function of damaged eye tissue, particularly due to degenerative processes such as disease of the retina or glaucoma.

Table 2 Results Summary of the Characteristics of the Selected Studies

Glaucoma is characterized by the degeneration of retinal ganglion cells. Based on the pathophysiology, glaucoma can be divided into two categories, namely open-angle glaucoma and closed-angle glaucoma. In patients with open-angle glaucoma, there is increased resistance to the aqueous humors outflow through the trabecular meshwork. This increased resistance is often caused by apoptosis and senescence of trabecular meshwork cells with increasing age.15 Degradation and abnormalities of the cytoskeleton arrangement of trabecular meshwork cells resulting in thickening of the drainage pathways and abnormal extracellular matrix deposition also worsen trabecular meshwork function in open-angle glaucoma.16 In closed-angle glaucoma, the aqueous humor cannot reach the trabecular meshwork due to obstruction.17 Examples of obstructions that often cause closed-angle glaucoma are anterior synechiae, the attachment of the iris to the trabecular meshwork, and posterior synechiae, where the iris is attached to the lens. This adhesion causes the aqueous humor to fail to reach the drainage system and the trabecular meshwork.18

Glaucoma is closely related to increased intraocular pressure, which is determined by the balance between the production of aqueous humor by the ciliary body and the drainage of the aqueous humor through the trabecular meshwork. The disturbance of the balance between production and drainage increases the humor Aquos, which at a later stage can increase the intraocular pressure.19 Studies have shown a link between increased intraocular pressure and retinal ganglion cell death. This study has also proven that the longer the intraocular pressure increases, the higher the degree of retinal ganglion cell damage.20 However, data show as many as 3040% of patients with glaucoma have normal intraocular pressure. One of the causes of glaucoma at normal intraocular pressure is a decrease in neurotrophic factors needed in the maintenance of neurons in the optic nerve. Neurotrophic factors are required to maintain retinal ganglion cells, including brain-derived neurotrophic factor (BDNF), ciliary neurotrophic factor (CNTF), and cell line-derived neurotrophic factor.21 Furthermore, microcirculation disorders, changes in immune system conditions, and increased levels of oxidative stress can also cause glaucoma at normal intraocular pressure.21

Stem cells are cells with the ability to differentiate and form all tissues in the human body. They are one of the potential therapies used in cases that require tissue repair and regeneration, one of which is glaucoma. For a cell to be called a stem cell, it must have two essential characteristics. The first one is the stem cell must produce offspring with the exact features the cell originates from, and the second one, the stem cell must be able to differentiate into the specific cell desired.22 There are two types of stem cells found in multicellular organisms, including humans. The first stem cells are embryonic stem cells or multipotent cells found in blastocysts, while the second stem cells are adult stem cells or pluripotent cells that can be found in a wide variety of adult tissues.23

Research has also succeeded in inducing adult cells to return to the pluripotent stage using molecular manipulation. The cells produced by this molecular manipulation are then called induced pluripotent stem cells (iPS).24 Most iPS manufacturing uses viruses such as retroviruses and lentiviruses to carry genes encoding transcription factors to adult cells to be modified. This gene will then undergo transcription and translation into a protein capable of inducing the adult cell nucleus to return to an embryonic state.25

An important concept that needs attention in stem cell therapy is how to induce stem cells to become the desired differentiated cells.26 It is necessary so that the cells can be used to treat various diseases, including glaucoma. We can further achieve differentiation of stem cells into specific desired cells by adding various growth factors and signaling pathways to resemble the conditions of their original development.27

The research conducted successfully isolates cultures and confirms that the trabecular meshwork stem cells around the Schwalbe line are multipotent with the ability to differentiate into a wide variety of cells, including trabecular meshwork cells adipocytes osteocytes, and chondrocytes.28 Other studies have also been able to induce stem cells on the Schwalbe line trabecular meshwork to proliferate and differentiate into photoreceptors under certain conditions.29 Apart from trabecular meshwork stem cells, other stem cells that can differentiate into functional meshwork trabecular cells are adipose-derived stem cells (ADSC), mesenchymal stem cells (MSC), and iPS. iPS cells can also differentiate into trabecular meshwork cells after culturing the extracellular matrix with cell-derived trabecular meshwork. The success of a wide variety of stem cells to differentiate into functional meshwork trabecular cells provides a more effective alternative to cutting-edge therapy in treating glaucoma, especially open-angle glaucoma.3

One of the stem cell therapies successfully applied and able to regenerate damaged retinal ganglion cells is iPS cell therapy. This therapy uses induced adult fibroblasts to return to pluripotent cells using four transcription factors, namely Oct3/4, Sox2, Klf4, and c-Myc. The results of the iPS are pluripotent cell colonies that are morphologically similar to ESCs, which are able to differentiate into the three germ cell layers.30

Because iPS can be programmed from the patients somatic cells, this therapy can maintain the unique genome of each individual. Currently, various modifications to the iPS therapy have been made to increase its acceptability and effectiveness of iPS therapy. One of them is the use of plasmid vectors and miRNA instead of retroviruses to avoid mutagenesis of the adult cells used.31,32

One of the significant challenges in stem cell therapy is to achieve the differentiation of stem cells into the desired cells, in this case, the differentiation of stem cells to retinal ganglion cells. Usually, in vivo, the differentiation of stem cells into retinal ganglion cells is regulated by several transcription factors such as Ath5, Brn3, and Notch. The transcription factors Ath5 and Brn3 play a vital role in the differentiation of retinal ganglion cells, and their levels are increased in the process of eye development.33 Meanwhile, Notch is a negative regulator of retinal ganglion cell differentiation, and its levels are decreased in normal eye development. Therefore, the addition of the transcription factors Ath5 and Brn3 and the Notch antagonist is a strategy to differentiate retinal ganglion cells from stem cells.34 Apart from transcription factors, various neurotrophic pathways and factors have been identified in the differentiation of stem cells into retinal ganglion cells. These pathways consist of fibroblast growth factor (FGF), insulin-like growth factor (IGF), bone morphogenetic protein (BMP), nodal, and Wnt signaling pathways. All of these pathways regulate retinal development, whereas FGF and IGF provide positive regulation. Meanwhile, BMP, nodal, and Wnt signaling pathways provide negative regulation.35

Another major challenge in the clinical application of stem cell therapy in glaucoma sufferers is that not only do the stem cells successfully differentiate into retinal ganglion cells, but they must also be able to reach the central nervous system.36 Modifications must be made so that new retinal ganglion cells can reach the visual cortex of the cerebrum. Recent research has found that a combination of genetic modification and stimulation of the signaling pathway stimulates regeneration of the optic nerve until it reaches the central nervous system. The addition of ephrin molecules, proteoglycans, cell-adhesion molecules, and semaphorin is able to guide the axons of the developing retinal ganglion cells to reach the optic chiasm.13 Meanwhile, the addition of cadherin, ephrin, and the Wnt signaling pathway can guide and stimulate synapse formation in the superior colliculus and the visual cortex.12,37

In addition, because of the adverse intraocular environment in glaucoma, stem cell therapy needs to be combined with neuroprotective compounds. It is also associated with a decrease in neurotrophic factors required to maintain neurons and causes progression of retinal ganglion cell damage in glaucoma sufferers. Therefore, the addition of BDNF and other neurotrophic factors such as glial cell-derived neurotrophic factor (GDNF) and ciliary neurotrophic factor (CNTF) should be considered for combined stem cell therapy.38

The stem cells are used in cases of glaucoma, which require repair and regeneration of trabecular meshwork cells and retinal ganglion cells. iPS has been shown the ability to differentiate to replace damaged trabecular meshwork cells and retinal ganglion cells in glaucoma. Some modifications are required so that stem cells that have differentiated into trabecular meshwork cells and retinal ganglion cells can reach the central nervous system. These modifications include the addition of ephrin molecules, proteoglycans, cell-adhesion molecules, semaphorin, cadherin, and the Wnt signaling pathway. The combination of stem cells with neuroprotective factors such as BDNF, GDNF, and CNTF also needs to be considered to maintain neuronal maintenance and inhibit the progression of cell damage.

The development of new stem cell technologies not only paves the way for us to gain a better understanding of the biology associated with glaucoma and create models for the development of new drugs, but it also opens the door to the prospect of cell-based therapies that can help patients regain their vision. More specifically in relation to the field of glaucoma, there have been recent developments in the process of developing protocols for the differentiation of stem cells into trabecular meshwork and retinal ganglion cells. Further research on the effectiveness of using modified stem cells as a therapy for glaucoma and in vivo research can be carried out immediately so that clinical trials can be carried out, which in turn can be used by the community to control symptoms and reduce blindness due to glaucoma.

The authors report no conflicts of interest in this work.

1. Choudhari NS, Neog A, Fudnawala V, George R. Cupped disc with normal intraocular pressure: the long road to avoid misdiagnosis. Indian J Ophthalmol. 2011;59(6):491. doi:10.4103/0301-4738.86320

2. Braunger BM, Ademoglu B, Koschade SE, et al. Identification of adult stem cells in Schwalbes line region of the primate eye. Invest Ophthalmol Vis Sci. 2014;55(11):7499. doi:10.1167/iovs.14-14872

3. Abu-Hassan DW, Li X, Ryan EI, Acott TS, Kelley MJ. Induced pluripotent stem cells restore function in a human cell loss model of open-angle glaucoma. Stem Cells. 2015;33(3):751761. doi:10.1002/stem.1885

4. Cook C, Foster P. Epidemiology of glaucoma: whats new? Can J Ophthalmol. 2012;47(3):223226. doi:10.1016/j.jcjo.2012.02.003

5. Hapsari DM. The Relationship Between Family Support and Adherence in Care in Glaucoma Clients in the Working Area of the Balung Health Center. Jember Regency. Uni; 2017.

6. World Health Organization. Global data on visual impairments 2010; 2012. https://www.iapb.org/wp-content/uploads/GLOBALDATAFINALforweb.pdf. Accessed August 24, 2022.

7. Harasymowycz P, Birt C, Gooi P, et al. Medical management of glaucoma in the 21st century from a canadian perspective. J Ophthalmol. 2016;2016:122. doi:10.1155/2016/6509809

8. Romito A, Cobellis G. Pluripotent stem cells: current understanding and future directions. Stem Cells Int. 2016;2016:120. doi:10.1155/2016/9451492

9. Cen L-P, Ng TK. Stem cell therapy for retinal ganglion cell degeneration. Neural Regen Res. 2018;13(8):1352. doi:10.4103/1673-5374.235237

10. ner A. Stem cell treatment in retinal diseases: recent developments. Turk J Ophthalmol. 2018;48(1). doi:10.4274/tjo.89972

11. Manuguerra-Gagn R, Boulos PR, Ammar A, et al. Transplantation of mesenchymal stem cells promotes tissue regeneration in a glaucoma model through laser-induced paracrine factor secretion and progenitor cell recruitment. Stem Cells. 2013;31(6):11361148. doi:10.1002/stem.1364

12. Osterhout JA, Josten N, Yamada J, et al. Cadherin-6 mediates axon-target matching in a non-image-forming visual circuit. Neuron. 2011;71(4):632639. doi:10.1016/j.neuron.2011.07.006

13. De lima S, Koriyama Y, Kurimoto T, et al. Full-length axon regeneration in the adult mouse optic nerve and partial recovery of simple visual behaviors. Proc Natl Acad Sci U S A. 2012;109(23):91499154. doi:10.1073/pnas.1119449109

14. Deng F, Chen M, Liu Y, et al. Stage-specific differentiation of iPSCs toward retinal ganglion cell lineage. Mol Vis. 2016;22:536.

15. Pulliero A, Seydel A, Camoirano A, Sacc SC, Sandri M, Izzotti A. Oxidative damage and autophagy in the human trabecular meshwork as related with ageing. PLoS One. 2014;9(6):e98106. doi:10.1371/journal.pone.0098106

16. Porter KM, Epstein DL, Liton PB. Up-regulated expression of extracellular matrix remodeling genes in phagocytically challenged trabecular meshwork cells. PLoS One. 2012;7(4):e34792. doi:10.1371/journal.pone.0034792

17. Lai J, Choy BNK, Shum JWH. Management of primary angle-closure glaucoma. Asia Pac J Ophthalmol. 2016;5(1):5962. doi:10.1097/APO.0000000000000180

18. Sun X, Dai Y, Chen Y, et al. Primary angle closure glaucoma: what we know and what we dont know. Prog Retin Eye Res. 2017;57:2645. doi:10.1016/j.preteyeres.2016.12.003

19. Gemenetzi M, Yang Y, Lotery AJ. Current concepts on primary open-angle glaucoma genetics: a contribution to disease pathophysiology and future treatment. Eye. 2012;26(3):355369. doi:10.1038/eye.2011.309

20. Khan Kareem A, Tse DY, van der Heijden ME, et al. Prolonged elevation of intraocular pressure results in retinal ganglion cell loss and abnormal retinal function in mice. Exp Eye Res. 2015;130. doi:10.1016/j.exer.2014.11.007

21. Almasieh M, Wilson AM, Morquette B, Cueva Vargas JL, Di Polo A. The molecular basis of retinal ganglion cell death in glaucoma. Prog Retin Eye Res. 2012;31(2):152181. doi:10.1016/j.preteyeres.2011.11.002

22. Cao J, Ng ES, Mcnaughton D, et al. The characterisation of pluripotent and multipotent stem cells using Fourier transform infrared microspectroscopy. Int J Mol Sci. 2013;14(9):1745317476. doi:10.3390/ijms140917453

23. Makhani K, Ali SM, Yousuf S, Siddiqui S. Therapeutic potential of totipotent, pluripotent and multipotent stem cells. MOJ Cell Sci Rep. 2015;2(5):00041. doi:10.15406/mojcsr.2015.02.00041

24. Zaveri L, Dhawan J. Cycling to meet fate: connecting pluripotency to the cell cycle. Front Cell Dev Biol. 2018;6. doi:10.3389/fcell.2018.00057

25. Wattanapanitch M. Recent updates on induced pluripotent stem cells in hematological disorders. Stem Cells Int. 2019;2019:115. doi:10.1155/2019/5171032

26. Kolios G, Moodley Y. Introduction to stem cells and regenerative medicine. Respiration. 2012;85(1):310. doi:10.1159/000345615

27. Biehl JK, Russell B. Introduction to stem cell therapy. J Cardiovasc Nurs. 2009;24(2):98103; quiz 1045. doi:10.1097/JCN.0b013e318197a6a5

28. Tay CY, Sathiyanathan P, Chu SWL, Stanton LW, Wong TT. Identification and characterization of mesenchymal stem cells derived from the trabecular meshwork of the human eye. Stem Cells Dev. 2012;21(9):13811390. doi:10.1089/scd.2011.0655

29. Nadri S, Yazdani S, Arefian E, et al. Mesenchymal stem cells from trabecular meshwork become photoreceptor-like cells on amniotic membrane. Neurosci Lett. 2013;541:4348. doi:10.1016/j.neulet.2012.12.055

30. Omole AE, Fakoya AOJ. Ten years of progress and promise of induced pluripotent stem cells: historical origins, characteristics, mechanisms, limitations, and potential applications. PeerJ. 2018;2018. doi:10.7717/peerj.4370

31. Miyoshi N, Ishii H, Nagano H, et al. Reprogramming of mouse and human cells to pluripotency using mature microRNAs. Cell Stem Cell. 2011;8(6):633638. doi:10.1016/j.stem.2011.05.001

32. Ikushima S, Ono R, Fukuda K, Sakayori M, Awano N, Kondo K. Trousseaus syndrome: cancer-associated thrombosis. Jpn J Clin Oncol. 2016;46(3):204208. doi:10.1093/jjco/hyv165

33. Ji SL, Tang SB. Differentiation of retinal ganglion cells from induced pluripotent stem cells: a review. Int J Ophthalmol. 2019;12(1):152160. doi:10.18240/ijo.2019.01.22

34. Li L, Chen LP, Liu QH. Effect of the notch signaling pathway on retinal ganglion cells and its neuroprotection in rats with acute ocular hypertension. Int J Ophthalmol. 2018;11(2). doi:10.18240/ijo.2018.02.05

35. Ritchey ER, Zelinka CP, Tang J, Liu J, Fischer AJ. The combination of IGF1 and FGF2 and the induction of excessive ocular growth and extreme myopia. Exp Eye Res. 2012;99(1):116. doi:10.1016/j.exer.2012.03.019

36. Georgiev D. Photons do collapse in the retina not in the brain cortex: evidence from visual illusions. NeuroQuantology. 2011;9(2). doi:10.14704/nq.2011.9.2.403

37. He CW, Liao CP, Pan CL. Wnt signalling in the development of axon, dendrites and synapses. Open Biol. 2018;8(10):180116. doi:10.1098/rsob.180116

38. Mysona BA, Zhao J, Bollinger KE. Role of BDNF/TrkB pathway in the visual system: therapeutic implications for glaucoma. Expert Rev Ophthalmol. 2017;12(1):6981. doi:10.1080/17469899.2017.1259566

Visit link:
Utilization of Modified Induced Pluripotent Stem Cells as the Advance | OPTH - Dove Medical Press

Posted in Stem Cells | Comments Off on Utilization of Modified Induced Pluripotent Stem Cells as the Advance | OPTH – Dove Medical Press

Neural stem cells: developmental mechanisms and disease modeling

Posted: June 13, 2022 at 2:29 am

Cell Tissue Res. Author manuscript; available in PMC 2018 May 22.

Published in final edited form as:

PMCID: PMC5963504

NIHMSID: NIHMS967727

1Department of Neuroscience, University of Wisconsin-Madison, Madison, WI 53705, USA

2Waisman Center, University of Wisconsin-Madison, Madison, WI 53705, USA

1Department of Neuroscience, University of Wisconsin-Madison, Madison, WI 53705, USA

1Department of Neuroscience, University of Wisconsin-Madison, Madison, WI 53705, USA

2Waisman Center, University of Wisconsin-Madison, Madison, WI 53705, USA

The astonishing progress in the field of stem cell biology during the past 40 years has transformed both science and medicine. Neural stem cells (NSCs) are the stem cells of the nervous system. During development they give rise to the entire nervous system. In adults, a small number of NSCs remain and are mostly quiescent; however, ample evidence supports their important roles in plasticity, aging, disease, and regeneration of the nervous system. Because NSCs are regulated by both intrinsic genetic and epigenetic programs and extrinsic stimuli transduced through the stem cell niche, dysregulation of NSCs due to either genetic causes or environmental impacts may lead to disease. Therefore, extensive investigations in the past decades have been devoted to understanding how NSCs are regulated. On the other hand, ever since their discovery, NSCs have been a focal point for cell-based therapeutic strategies in the brain and spinal cord. The limited number of NSCs residing in the tissue has been a limiting factor for their clinical applications. Although recent advancements in embryonic and induced pluripotent stem cells have provided novel sources for NSCs, several challenges remain. In this special issue, leaders and experts in NSCs summarize our current understanding of NSC molecular regulation and the importance of NSCs for disease modeling and translational applications.

The term stem cells first appeared in the scientific literature in 1868 by the German biologist Ernst Haeckel (Haeckel, 1868). In his writings (Haeckel, 1868), stem cells had two distinct meanings: one is the unicellular evolutionary origin of all multicellular organisms, and the other is the fertilized egg giving rise to all other cell types of the body. The latter definition has evolved into the modern definition of stem cells - cells that can divide to self-renew and to differentiate into other cell types in tissues and organs (Li and Zhao, 2008, Ramalho-Santos and Willenbring, 2007).

The behavior and fate of stem cells are strongly influenced by their specific anatomical locations and surrounding cell types, called the stem cell niche. The niche provides physical support to host or anchor stem cells, and supplies factors to maintain and regulate them (Li and Zhao, 2008). Stem cells are also regulated by intrinsic signaling cascades and transcriptional mechanisms, some of which are common among all stem cells, and others that are unique to specific types. Some of the best known regulators include TGF-, BMP, Smad, Wnt, Notch, EGF fibroblast growth factors (Jobe, et al., 2012, Li and Zhao, 2008). Therefore, stem cells are regulated by complex mechanisms in both temporal- and context-specific manners to maintain their unique characteristics. Understanding stem cell regulation gives us the opportunity to explore mechanisms of development, as well as disorders resulting from their dysfunction.

During development, the central nervous system (CNS) is generated from a small number of neural stem cells (NSCs) lining the neural tube (Kriegstein and Alvarez-Buylla, 2009). A great deal of experimental evidence has demonstrated that radial glia, the NSCs during mammalian CNS development, undergo both symmetric divisions to expand the NSC pool, and asymmetric divisions to give rise to intermediate progenitors (IPCs) and the differentiated cell types. The three major cell types in the CNS arise from NSCs in a temporally defined sequence, with neurons appearing first, followed by astrocytes, and then oligodendrocytes (Okano and Temple, 2009). The technical advancement of live imaging and genomic tools have allowed for the identification of human-specific NSC populations (e.g. outer radial glia, or oRG) located at the outer subventricular zone (SVZ) (Gertz, et al., 2014). These oRG are essential for cortical expansion to achieve the large size of the human cortex. Single-cell genomic technologies have identified specific oRG markers that might be used for further characterization of these cells (Liu, et al., 2016, Pollen, et al., 2014). Investigating the regulatory mechanisms governing the self-renewal and fate specification of NSCs, especially human-specific developmental features, has significantly contributed to our understanding of human brain development and developmental diseases. In addition, this knowledge also has helped scientists refine protocols for pluripotent stem cell differentiation into specific nervous system cell types for both therapeutic goals and disease modeling.

In adult brains, NSCs are reduced and become restricted to specific brain regions. In rodents, both NSCs and ongoing neurogenesis have been widely documented in the SVZ of the lateral ventricles and the subgranular zone (SGZ) of the dentate gyrus (DG) of the hippocampus (Kempermann, et al., 2015). In humans, experimental evidence has supported ongoing neurogenesis in the hippocampus (Eriksson, et al., 1998, Spalding, et al., 2013). The confirmation of mammalian adult neurogenesis in the 1990s was one of the most exciting moments in science in the 21st century. Not only did it overthrow the prevailing dogma suggesting no neurons were made in the adult brain, but also it hinted that these adult NSCs could be utilized for neural repair in disease and following injury. Forty years later, we have learned a lot about NSCs. In the adult rodent SVZ, neurogenesis has been shown to be important for olfactory function and olfactory learning (Alonso, et al., 2006). During development, a subset of slowly-dividing NSCs are set aside to be the NSCs of the SVZ in the postnatal and adult brain (Fuentealba, et al., 2015, Furutachi, et al., 2015). The majority of neurogenic radial glia, however, become astrocytes and ependymal cells at the end of embryonic neurogenesis (Noctor, et al., 2004). A subset of these astrocytes persist as NSCs in specialized niches in the adult brain and continuously generate neurons that functionally integrate into restricted brain regions (Doetsch, 2003). In the hippocampus, radial glia-like stem cells of the SGZ make newborn neurons throughout life (Goritz and Frisen, 2012). These newborn neurons integrate into the circuity of the DG, contributing to behaviors such as pattern separation (Aimone, et al., 2011) and spatial learning (Dupret, et al., 2008), as well as hippocampus-associated learning, memory, and executive functions (Kempermann, Song and Gage, 2015).

Significant effort has been devoted into understanding the regulation of adult neurogenesis. As a result, we now know that many extrinsic stimuli and intrinsic mechanisms can affect this process. Mouse genetic studies have clearly demonstrated the important role of transcriptional regulation of NSCs through intrinsic genetic mechanisms (Hsieh and Zhao, 2016). Some examples include SOXC family proteins [Kavyanifar et al, in this issue (Kavyanifar, et al., 2018)], Bmi-1 (Molofsky, et al., 2003), Sox2 (Ferri, et al., 2004, Graham, et al., 1999), PTEN (Bonaguidi, et al., 2011), and Notch [Zhang et al, in this issue (Zhang, et al., 2018)]. In addition epigenetic regulation by DNA methylation pathways (e.g. Mbd1, Mecp2, Dnmt, Tet) (Noguchi, et al., 2015, Smrt, et al., 2007, Tsujimura, et al., 2009, Zhang, et al., 2013, Zhao, et al., 2003), chromatin remodeling (e.g. BAF, BRG1) (Ninkovic, et al., 2013, Petrik, et al., 2015, Tuoc, et al., 2017), and noncoding RNAs (Liu, et al., 2010)[Anderson and Lim, in this issue (Anderson and Lim, 2018)] play important roles. Many growth factors, signaling molecules, and neurotransmitters have been shown to regulate neurogenesis (Kempermann, Song and Gage, 2015). Catavero et al [in this issue (Catavero, et al., 2018)] review the role of GABA circuits, signaling, and receptors in regulating development of adult born cells, as well as the molecular players that modulate GABA signaling. Because progenitors with multipotent differentiation potentials have been found in brain regions without active neurogenesis (Palmer, et al., 1997), it is hypothesized that these progenitors might be manipulated to become neuron-competent in vivo so that they can contribute to brain generation [Wang et al, in this issue (Wang and Zhang, 2018)].

A great amount of literature has documented how physiological activities and enriched environment influences adult neurogenesis (Kempermann, Song and Gage, 2015). However, as summarized by Eisinger and Zhao [in this issue (Eisinger and Zhao, 2018)], the genes and gene network involved in these changes within NSCs have not been systematically analyzed at genome wide levels. Adult neurogenesis is also influenced by diseases including epilepsy (Parent and Lowenstein, 1997), stroke (Zhang and Chopp, 2016), depression (Dranovsky and Hen, 2006, Kempermann, et al., 2003), and injury [(Morshead and Ruddy, in this issue (Morshead and Ruddy, 2018) in this issue). Thodeson et al [in this issue (Thodeson, et al., 2018)] further summarize the contribution and dysregulation of adult neurogenesis in epilepsy and discuss how we can translate these findings to human therapeutics by using patient-derived neurons to study monogenic epilepsy-in-a-dish.

Aging affects every individual and is a major risk factor for many diseases. One of the strongest negative regulators of adult neurogenesis is aging. Both intrinsic and extrinsic components regulate the limitations of NSC proliferation and function (Moore and Jessberger, 2017, Seib and Martin-Villalba, 2015). In this issue, Mosher and Schaffer (Mosher and Schafer, 2018) and Ruddy and Morshead (Morshead and Ruddy, 2018) examine factors such as secreted signals, cell contact- dependent signals, and extracellular matrix cues that control neurogenesis in an age-dependent manner, and define these signals by the extrinsic mechanism through which they are presented to the NSCs. Smith et al [in this issue (Smith, et al., 2018)] discuss how age-related changes in the blood, such as blood-borne-factors, and peripheral immune cells, contribute to the age-related decline in adult neurogenesis in the mammalian brain.

Despite the extensive knowledge we have gained regarding adult neurogenesis, critical questions remain. For example, the control of the functional integration of new neurons remains a mystery. It has been shown that adult NSC-differentiated newborn neurons exhibit a critical period for sensitivity to external stimuli (Bergami, et al., 2015), and a heightened sensitivity to seizures (Kron, et al., 2010). It remains unclear how new neurons choose their connections. Jahn and Bergami [in this issue (Jahn and Bergami, 2018)] further discuss the critical period and its regulators during adult newborn neuron development.

Understanding the extrinsic and intrinsic regulation of adult NSCs and their newborn progeny, and their response to both positive and negative stimuli will further illuminate their role in disease, injury, stress, and brain function.

Human pluripotent stem cells (PSCs), including human embryonic stem cells (ESCs) and induced PSCs (iPSCs), offer a model system to reveal cellular and molecular events underlying normal and abnormal neural development in humans. ESCs are pluripotent cells derived from the inner cell mass of blastocyst stage preimplantation embryos, which were first isolated from mouse by Evans and Kaufman in 1981 (Evans and Kaufman, 1981) and later from humans by James Thompson in 1998 (Thomson, et al., 1998). Human ESCs are invaluable in the study of early embryonic development, allowing us to identify critical regulators of cell commitment, differentiation, and adult cell reprogramming (Dvash, et al., 2006, Ren, et al., 2009). iPSCs are reprogrammed from somatic cells by forced expression of stem cell genes and have the characteristics of ESCs (Okita, et al., 2007, Yu, et al., 2007). The development of iPSC technology has allowed us access to cells of the human nervous system through reprogramming of patient-derived cells, revolutionizing our ability to study human development and diseases.

To generate neural cells from either ESCs or iPSCs, the first step is neural induction. Through actions of a number of activators and inhibitors of cell signaling pathways, this process yields neural epithelial cell-like NSCs and then intermediate neural progenitors, resembling embryonic development. Despite many advances, a major hurdle of neural differentiation is lineage control. Using a standard dorsal forebrain neural differentiation protocol, most neural progenitors obtained are forebrain excitatory progenitors that produce mostly forebrain glutamatergic excitatory neurons. However, the purity and layer-specific composition of these progenitors, as well as neurons, vary significantly from experiment to experiment, cell line to cell line, and lab to lab. In addition, differentiation into specific types of neurons with high purity has always been a challenging goal. Much effort has been devoted into improving the efficiency of dopaminergic neuron and GABAergic neuron differentiation with great success (Hu, et al., 2010). However, the brain has many other types of neurons. Vadodaria et al [in this issue (Vadodaria, et al., 2018)] discuss how to generate serotonergic neurons, a type of neuron highly relevant to psychiatric disorders. To better understand the molecular control of human PSC and NSC differentiation, where protocols result in a large amount of cellular heterogeneity in identity and response, analysis must be done at the level of single cells. Harbom et al [in this issue (Harbom, et al., 2018)] summarizes how new state-of-the-art single-cell analysis methods may help to define differentiation from pluripotent cells.

The advancement in iPSC and gene editing technology has transformed the field of human disease modeling. As in many human disorders, especially neuropsychiatric disorders, mouse models have been useful. Yet there are several critical reasons why it is necessary to use human cells to define the underlying mechanisms that lead to human patient characteristics, particularly those affecting the nervous system. For example, in fragile X syndrome (FXS), the epigenetic silencing of the Fragile X Mental Retardation Gene 1 (FMR1) gene that causes FXS occurs only in humans. Mice engineered to mimic the human mutation in the FMR1 gene do not show the same methylation and silencing characteristics of the gene as in humans (Brouwer, et al., 2007). These results indicate that some epigenetic mechanisms in human and mice are different and preclude the ability to study epigenetic mechanisms of FMR1 silencing in mouse models of FXS (Bhattacharyya and Zhao, 2016). In this issue, Li and Shi discuss disease modeling using human PSC-differentiated neural progenitors (Li, et al., 2018), and Brito et al specifically focus on modeling autism spectrum disorder (Brito, et al., 2018).

The use of NSCs as a treatment strategy in CNS disease and injury has been tested for decades. Parkinsons disease specifically has gained the most momentum for potential therapeutic benefits (Studer, 2017); however, similar work has been performed in Huntingtons disease, stroke, and following spinal cord injury [for a review on this topic, see (Vishwakarma, et al., 2014)]. In some of these paradigms, NSCs are expected to differentiate into a specific cell type in the local CNS environment; in other cases, they are in a supportive role. In this issue, Kameda et al explores progress in using NSCs as a therapy following spinal cord injury (Kameda, et al., 2018).

While the development of PSC technologies has been a scientific breakthrough for future studies, there are limitations and risks that may be associated with their use. ESCs, iPSCs, and their differentiated NSCs are dividing cells. Either transplantation of NSCs or in vivo reprogramming of endogenous cells into NSCs could lead to tumorigenesis. In addition, reprogramming somatic cells into iPSCs results in a loss of some epigenetic signatures of disease and aging which are critical for disease modeling (Mertens, et al., 2015, Miller, et al., 2013, Ocampo, et al., 2016). In recent years, direct reprogramming of fibroblasts or other cell types into induced neurons (iN) has been developed (for review see (Mertens, et al., 2016)). Remarkably a growing number of studies have demonstrated that such direct reprogramming also can be effective in vivo. Wang et al [(Wang and Zhang, 2018) in this issue] will summarize recent progress of in vivo reprogramming into new neurons and present how this method can be used for spinal cord injury.

In cellular reprogramming, the cells targeted and the genetic factors used vary; however, the biggest difference is that some protocols push cells through a NSC stage, whereas others skip these stages (Gascon, et al., 2017, Guo, et al., 2014, Wang, et al., 2016). Bypassing this developmental stage has both pros and cons, and may lead to a completely novel path towards lineage commitment [discussed by Falk and Karow (Falk and Karow, 2018) in this issue].

NSCs are fascinating and promising cells because of their capability, flexibility, and multiplicity. Understanding how NSCs function provides important knowledge in development, adaptation, disease, regeneration, and rehabilitation of the nervous system. The studies of cortical development and adult neurogenesis using rodent models have contributed significantly to our knowledge about NSCs and will continually yield important new information, taking advantage of novel genetic and imaging technologies. However, using human NSCs provides us with a window to investigate human-specific aspects of development and disease mechanisms, which is potentiated by the fast advancement of stem cell and gene editing technologies. Challenges still remain regarding cell lineage control, in vivo NSC behavior, three dimensional cellular interactions, and preservation of epigenetic and aging marks.

We thank Klaus Unsicker for his encouragement and support and Jutta Jger for her help with invitations, and communications with authors and reviewers. This work was supported by grants from the US National Institutes of Health (R01MH078972, R56MH113146, R21NS098767, and R21NS095632 to X.Z, U54HD090256 to the Waisman Center), University of Wisconsin (UW)-Madison Vilas Trust (Kellett Mid-Career Award to X.Z.) and UW-Madison and Wisconsin Alumni Research Foundation (WARF to X.Z.), Jenni and Kyle Professorship (to X.Zhao), a Sloan Research Fellowship (to D.L.M.), a Junior Faculty Grant from the American Federation for Aging Research (to D.L.M.), and startup funds from UW-Madison School of Medicine and Public Health, WARF, and the Neuroscience Department (to D.L.M.).

Read more:
Neural stem cells: developmental mechanisms and disease modeling

Posted in Wisconsin Stem Cells | Comments Off on Neural stem cells: developmental mechanisms and disease modeling

Cancer stem cells in glioblastoma

Posted: January 20, 2022 at 2:17 am

Genes Dev. 2015 Jun 15; 29(12): 12031217.

1Department of Cellular and Molecular Medicine, Lerner Research Institute, Cleveland Clinic, Cleveland, Ohio 44195, USA;

2Department of Molecular Medicine, Cleveland Clinic, Lerner College of Medicine of Case Western Reserve University, Cleveland, Ohio 44195, USA;

3Department of Stem Cell Biology and Regenerative Medicine, Lerner Research Institute, Cleveland Clinic, Cleveland, Ohio 44195, USA

1Department of Cellular and Molecular Medicine, Lerner Research Institute, Cleveland Clinic, Cleveland, Ohio 44195, USA;

3Department of Stem Cell Biology and Regenerative Medicine, Lerner Research Institute, Cleveland Clinic, Cleveland, Ohio 44195, USA

2Department of Molecular Medicine, Cleveland Clinic, Lerner College of Medicine of Case Western Reserve University, Cleveland, Ohio 44195, USA;

3Department of Stem Cell Biology and Regenerative Medicine, Lerner Research Institute, Cleveland Clinic, Cleveland, Ohio 44195, USA

1Department of Cellular and Molecular Medicine, Lerner Research Institute, Cleveland Clinic, Cleveland, Ohio 44195, USA;

2Department of Molecular Medicine, Cleveland Clinic, Lerner College of Medicine of Case Western Reserve University, Cleveland, Ohio 44195, USA;

3Department of Stem Cell Biology and Regenerative Medicine, Lerner Research Institute, Cleveland Clinic, Cleveland, Ohio 44195, USA

Tissues with defined cellular hierarchies in development and homeostasis give rise to tumors with cellular hierarchies, suggesting that tumors recapitulate specific tissues and mimic their origins. Glioblastoma (GBM) is the most prevalent and malignant primary brain tumor and contains self-renewing, tumorigenic cancer stem cells (CSCs) that contribute to tumor initiation and therapeutic resistance. As normal stem and progenitor cells participate in tissue development and repair, these developmental programs re-emerge in CSCs to support the development and progressive growth of tumors. Elucidation of the molecular mechanisms that govern CSCs has informed the development of novel targeted therapeutics for GBM and other brain cancers. CSCs are not self-autonomous units; rather, they function within an ecological system, both actively remodeling the microenvironment and receiving critical maintenance cues from their niches. To fulfill the future goal of developing novel therapies to collapse CSC dynamics, drawing parallels to other normal and pathological states that are highly interactive with their microenvironments and that use developmental signaling pathways will be beneficial.

Keywords: brain tumor, cancer stem cell, glioblastoma, glioma, stem cell, tumor-initiating cell

Development is a coordinated summation of the individual cellular dynamics that build an organ, and the programs responsible for this construction are generally preserved in stem cells for organ homeostasis and tissue repair. Tumors are complex systems that recapitulate the complexity of organs or tissues with dynamic regulation and constituent cellular populations during tumor initiation, maintenance, and progression (Hanahan and Weinberg 2011). While many scientists have sought to reduce the complexity of cancer to a one-dimensional processfor example, characterizing cancers solely based on geneticsmost advanced cancers unfortunately remain nearly as lethal since the declaration of the War on Cancer in 1971. Targeted therapeutics offer a transient benefit for some cancer types with driving mutations, but even these tumors will develop resistance to overcome initially effective therapies that poison driving oncogenes.

Glioblastoma (GBM; World Health Organization grade IV glioma) is the most prevalent and lethal primary intrinsic brain tumor (Stupp et al. 2009). Unlike other solid tumor cell types, GBM widely invades the surrounding brain but rarely metastasizes to other organs. While halting steps to fight GBM are being made using targeted therapies (e.g., bevacizumab) or immunotherapies, GBM therapy remains focused on achieving maximal surgical resection followed by concurrent radiation therapy with temozolomide (TMZ; an orally available methylation chemotherapy) and subsequent additional adjuvant TMZ therapy. Conventional treatment offers patients with GBM additional survival time with generally acceptable quality of life, but a cure is never achieved. GBM represents one of the most comprehensively genomically characterized cancer types, leading to recognition of groups of tumors defined by transcription profiles (proneural, neural, classical, and mesenchymal), genetics (mutations of isocitrate dehydrogenase 1 [IDH1]), and epigenetics (CpG island methylator phenotype [CIMP]) (Weller et al. 2013). Long-term survivors are often, but not exclusively, patients with tumors harboring mutations in IDH1, which likely represent a different disease than most GBMs. Beyond IDH1 mutations and a few other biomarkers (deletion of chromosomes 1p and 19q in oligodendrogliomas, methyl guanine methyltransferase [MGMT] promoter methylation, etc.), the accumulated genetic characterization of GBMs has failed to impact clinical practice, suggesting that other discovery paradigms should also be considered.

The brain, like other organs with clearly defined cellular hierarchies in development and homeostasis (e.g., blood, breast, skin, and colon), gives rise to tumors with defined cellular hierarchies, suggesting that cancer replicates ontogeny (Reya et al. 2001). Atop the apex of cellular hierarchies are stem cells, which have been assumed to be rare, quiescent, self-renewing cells, but several highly proliferative organs (e.g., the intestine and skin) contain at least two pools of stem cells: one quiescent, and the other highly proliferative (Barker et al. 2010). Stem cells generate transient amplifying cells, which in turn create lineage-restricted progeny that are eventually fated to become the terminally differentiated effector cells.

Neural stem and progenitor cell (NSPC) pools vary in location and control during development, suggesting that different cellular hierarchies may be co-opted by brain tumors (Gibson et al. 2010; Lottaz et al. 2010). Informed by techniques used to enrich and characterize NSPCs, several groups in parallel demonstrated that gliomas and other primary brain tumors contain self-renewing, tumorigenic cells (Ignatova et al. 2002; Hemmati et al. 2003; Singh et al. 2003, 2004; Galli et al. 2004). The nomenclature for these cells has been controversial (as discussed below), with the dominant choice being cancer stem cells (CSCs) or tumor-initiating cells (importantly, these terms are not identical, as a CSC designation is more restrictive but also more informative) (). Unlike normal brain organizationwhere the generation of differentiated progeny is stage-specific (neurons and then glia during development) and derived from rapidly dividing progenitor cells and quiescent, multipotent stem cells that persist into adulthood and become activated upon differentiation (Rowitch and Kriegstein 2010)these populations have yet to be delineated in brain tumors. The ability to distinguish between self-renewing cells with stem and progenitor cell cycle properties and transcriptional signatures is likely to provide clarity with respect to nomenclature and the functional interplay between cells at the apex of the hierarchy. The challenges distinguishing CSCs from their progeny are derived, in part, from the limited recognition of points of relative stability (attractor states) in the landscape of cellular identity that define the stem cell state and transitions into (dedifferentiation) and out of (differentiation) a stable state (Chang et al. 2008). Much like the Heisenberg uncertainty principle in physics, our ability to observe the state of a cell is limited by our act of measurement. The presence of subatomic particles is confirmed in retrospect; similarly, the functional definition of both normal and neoplastic stem cells requires retrospective confirmation. The ability to prospectively distinguish glioma stem cells, which reside at the apex of tumor hierarchies, from their differentiated progeny remains challenging; however, stem cell biology faces a similar difficulty with normal stem cell identification. Of note, the CSC hypothesis does not claim a stem cell as the cell of origin for cancers, suggesting that CSCs do not need to adhere to all of the observed features of normal stem cells.

Definitions and functional characteristics of CSCs, tumor-initiating cells, and tumor-propagating cells

In the following sections, we provide an update on intrinsic and extrinsic regulators of the CSC state in GBM and discuss how the integration of genetics, epigenetics, and metabolism has shaped our understanding of how CSCs function to drive GBM growth. We also highlight future opportunities to further understand the complexity of CSC regulation through interaction with other cells (including immune cells) and how the translation of CSC-based therapies needs to take into account the cellular dynamics of CSCs, which rely on developmental signaling programs.

The heterogeneity of tumor cells has long been appreciated, but two decades ago, seminal work from Dick's laboratory (Bonnet and Dick 1997) described the isolation of a leukemia-initiating cell, the first purification of cancer stem-like cells, a population of cells that had originally been proposed to exist >150 years earlier (Sell 2004). The first prospective isolation of human NSPCs was performed using CD133 (Uchida et al. 2000) and prompted a search for brain tumor cells that shared the characteristics of NSPCs. A burst of studies soon followed describing brain CSCs in anaplastic astrocytoma (Ignatova et al. 2002), medulloblastoma, pilocytic astrocytoma, ependymoma, ganglioglioma (Hemmati et al. 2003; Singh et al. 2003), and GBM (Ignatova et al. 2002; Hemmati et al. 2003; Galli et al. 2004). Brain CSCs have subsequently been shown to be resistant to standard-of-care chemotherapy (Chen et al. 2012) and radiotherapy (Bao et al. 2006a), underscoring their role in disease progression and recurrence.

While cellular heterogeneity within CNS tumors is well recognized (Bonavia et al. 2011; Meacham and Morrison 2013), the nomenclature used to describe the self-renewing population of tumor cells with enhanced tumorigenic properties is far from uniform (see ). To date, many terms have been used to describe this population, including cancer/tumor/glioma/brain tumor stem cell, stem-like tumor cell, cancer-/tumor-/glioma-/brain tumor-initiating cell, and cancer-/tumor-/glioma-/brain tumor-propagating cell. This lack of uniformity has generated confusion and controversy by redirecting the focus away from the biology of these cells and their contribution to tumorigenic processes toward identifying markers that the cells express and whether tumor cells can be propagated as free-floating spheres. In addition, while the term stem cell is used, it does not necessarily mean these cells have been generated from a transformed stem cell, as there is evidence that multiple cell typesranging from stem cell to differentiated progeny, depending on the modelare amenable to oncogenic transformation. Therefore, in the current context, it is essential that the strictest functional assays continue to be performed and a singular term be used for studies using models that fulfill this criteria. As the accepted functional definition of a stem cell is the ability to self-renew and generate differentiated progeny, any claims for a CSC population must also demonstrate this capacity (). For brain tumors, this means the ability to generate a tumor upon intracranial transplantation that recapitulates the cellular heterogeneity present in the parental tumor. Unlike the designation of a tumor-initiating cell, CSCs cannot be investigated in isolation due to the required comparison with differentiated progeny. Prospective enrichment and depletion of tumorigenic and nontumorigenic cells demonstrate the presence of a cellular hierarchy. Cells that meet these criteria (tumorigenic and containing a cellular hierarchy) should be referred to as CSCs (or alternatives such as glioma stem cells, glioma CSCs, or brain tumor stem cells in the context of GBM). While the ability to grow as spheres is also evident in CSCs, it is not by default the defining feature of a self-renewing population of cells. In fact, the majority of spheres derived from both normal and neoplastic brain cells come from progenitor cells with limited self-renewal potential, not stem cells (Pastrana et al. 2011). Furthermore, high-passage cell lines, which do not offer the ability to accurately represent tumor complexity in vivo (Lee et al. 2006), should not be replaced with cells grown in long-term passage as spheres but rather with functionally validated CSC models, as this offers the best opportunity to more deeply model the complexity of brain tumors. Thus, although culture of glioma cells as neurospheres may not be required to maintain stemness (Pollard et al. 2009; Cheng et al. 2012), the microenvironment, including medium composition and culture conditions, does necessarily affect the characteristics of CSCs (Pastrana et al. 2011).

Functional criteria of CSCs. CSCs are defined by functional characteristics that include sustained self-renewal, persistent proliferation, and tumor initiation upon secondary transplantation, which is the definitive functional CSC assay. CSCs also share features with somatic stem cells, including frequency within a tissue (or tumor), stem cell marker expression (examples relevant to GBM and the brain are provided), and the ability to generate progeny of multiple lineages.

Great energy and passion have been devoted to the discovery, validation, and use of CSC enrichment methods. Demonstration of a cellular hierarchy demands methods to separate populations that can be functionally studied. Ideally, an enrichment method would be based on a property that defines an essential CSC feature (self-renewal, tumor initiation, etc.) that is immediately lost upon differentiation (i.e., a digital readout) and is usable with live cells. Currently, no such system exists for any cell type (normal or neoplastic) because biologic systems rarely exhibit all or none phenomena. Critics of the CSC hypothesis have held this limitation up as proof against CSCs; while the same limitations exist for even the best-characterized normal stem cell (hematopoietic stem cell), no scientists deny the existence of hematopoietic stem cells. Leukemia stem cells are considered a definitive tumor population, yet no marker signature for these cells is definitive (Eppert et al. 2011). A more sophisticated and nuanced use of enrichment systems that is informed by recognition of the diversity of GBMs can lead to context-specific methods to produce matched tumorigenic and nontumorigenic populations.

Most glioma CSC markers have been appropriated from normal stem cells, but the linkage between glioma CSCs and normal stem cells remains controversial. Many of the transcription factors or structural proteins essential for normal NSPC function also mark glioma CSCs, including SOX2 (Hemmati et al. 2003), NANOG (Ben-Porath et al. 2008; Suva et al. 2014), OLIG2 (Ligon et al. 2007), MYC (Kim et al. 2010), MUSASHI1 (Hemmati et al. 2003), BMI1 (Hemmati et al. 2003), NESTIN (Tunici et al. 2004), and inhibitor of differentiation protein 1 (ID1) (Anido et al. 2010). However, because of the limited utility of intracellular proteins for enriching CSCs from nonstem tumor cells (NSTCs) using traditional methods such as flow cytometry, a multitude of potential cell surface markers have been suggested, including CD133 (Hemmati et al. 2003), CD15 (also called Lewis x and SSEA-1 [stage-specific embryonic antigen 1]) (Son et al. 2009), integrin 6 (Lathia et al. 2010), CD44 (Liu et al. 2006), L1CAM (Bao et al. 2008), and A2B5 (Ogden et al. 2008). These types of cell surface markers mediate interactions between cells and the microenvironment, but dissociation of cells from their surroundings rapidly degrades the informational content of markers, requiring rapid utilization.

The first proposed marker, CD133 (Prominin-1), a cell surface glycoprotein expressed on neural stem cells, enriches for cells with higher rates of self-renewal and proliferation and increased differentiation ability (Singh et al. 2003). However, CD133 expression, rather than the AC133 surface epitope, should be used with care to enrich for any cells: Surface CD133 marks stem cells and decreases with differentiation, but the expression of Prominin-1 mRNA is not regulated with stemness (Kemper et al. 2010), suggesting that only the glycosylated surface protein CD133 is CSC-specific. The AC133 antigen marks the glycosylated molecule localized in lipid rafts that signals through PI3K and other key pathways to mediate interactions between a cell and its microenvironment (Wei et al. 2013). Most studies fail to recognize this role and use CD133 as a marker in cells that have been extensively cultured out of their microenvironment. Furthermore, the information contained in CD133 is context-dependent. CD133 mRNA, protein lysates, immunofluorescence, and FACS analysis for the AC133 glycoprotein have very different relationships to cell biology. Unfortunately, the complexity of these biomarkers has led to a reductionist view that has challenged the field due to the lack of consistency in methodology and models. It is nearly certain that CD133 is not universally informative in all tumors and has a false-negative rate for identifying CSCs (CD133-negative cells can be tumor-propagating in some tumors) (Beier et al. 2007). Additionally, the use of CD133 as a stem marker is complicated by the observation that expression of CD133 can be regulated at the level of the cell cycle, with potentially slow-cycling NSPCs lacking CD133 expression during G0/G1 cell cycle phase but still maintaining multipotency (Sun et al. 2009).

Although CD133 continues to be the most commonly used cell surface marker, other markers, such as integrin 6, have been proposed to segregate CSCs and NSTCs (Lathia et al. 2010). CD15/SSEA-1 and CD44 have also been proposed as possible markers, potentially with an association with specific subgroups of GBM (Bhat et al. 2013). These markers have utility but must be approached with caution. Each can mark a large percentage of cells, consistent with a high false-positive rate. Due to the current limitations in the functional assays defining CSCs, false-positive markers are sometimes claimed to be superior to functional identification, but markers lack significant utility in discovery studies, which benefit from greater specificity. Additionally, it is likely that no marker will ever be uniformly informative for CSCs because most tissue types contain multiple populations of stem cells expressing different markers and due to the inherent adaptability of cancer cells.

Several methods other than marker expression have been used to enrich for glioma CSCs, such as the abilities to grow as neurospheres in serum-free medium or efflux fluorescent dyes (Goodell et al. 1996; Kondo et al. 2004). Many investigators have used neurosphere culture to select for progenitor cells in the normal and neoplastic brain cells, but there are challenges with this approach. Neurosphere culture selects for a small fraction of the original tumor cells with bias toward progenitor features and expression of epidermal growth factor receptor (EGFR) and FGFR based on growth factors (EGF and FGF) added to the medium (Pastrana et al. 2011). This selection process eliminates the ability to prospectively enrich and deplete stem-like cells, preventing the delineation of a cellular hierarchy required to prove the presence of CSCs. Neurosphere culture selects for cells that can grow in stem cell medium; however, the selection of CSCs simply based on culture methods fails to recapitulate the heterogeneity of the original tumor in vivo as assessed by histological morphology, differentiated cell lineage, and gene expression (Lee et al. 2006; Lathia et al. 2011; Venere et al. 2011), a characteristic that CSCs acutely isolated using marker expression maintain (Singh et al. 2004). An alternative approach to CSC enrichment is the use of flow cytometry to isolate a side population containing CSCs, which is based on the hypothesis that stem cells contain drug efflux transporters (Yu et al. 2008). While this approach has identified a population of self-renewing cells in a mouse glioma model (Bleau et al. 2009), it has not been used successfully to enrich for self-renewing cells in human GBM (Broadley et al. 2011; Golebiewska et al. 2013), highlighting the model- and species-specific challenges of enrichment methods.

CSC markers, although useful to enrich populations of stem cells from nonstem cells, are not sufficient to define either population due to the lack of definitive markers. Functional validationthe observation of differences in stem cell characteristics of CSCs and NSTCsis essential to ensure that the enriched cells truly exhibit the functional characteristics of stem cells (). Various methods, both in vitro and in vivo, are employed to assess stem cell characteristics (self-renewal, proliferation, and ability to reproduce the complexity of the original tumor) of enriched cells. In vitro neurosphere formation assays test for both proliferation and self-renewal but fail to address cellular hierarchy and do not recapitulate the tumor microenvironment. Sphere formation is a surrogate of self-renewal capability andwhen performed in a limiting dilution formatstem cell frequency, but in vivo tumor formation assays are essential to claim the presence of CSCs.

The gold standard for CSC determination remains the ability of a limiting dilution of cells to recapitulate the complexity of the original patient tumor when transplanted orthotopically. The ability to derive heterogeneity is essential because populations of transit-amplifying cells may form a tumor but will only give rise to cells from their specific lineage. Heterotypic transplantation of cellsfor example, into the flank of the animalmay also be informative, but this technique lacks the proper microenvironmental cues of orthotopic implantation.

Glioma CSCs are regulated by six main mechanisms, which include intrinsic factors such as genetics, epigenetics, and metabolism as well as extrinsic qualities of niche factors, cellular microenvironment, and the host immune system (). The following sections describe the key features of each of these factors and highlight new advances in the topics of epigenetics mapping, single-cell heterogeneity, metabolism, and immunotherapy.

Regulation of CSCs. Cell-autonomous (intrinsic) and external (extrinsic) forces regulate the CSC state. Key intrinsic regulators include genetic, epigenetic, and metabolic regulation, while extrinsic regulators include interaction with the microenvironment, including niche factors and the immune system.

Through advances in genomic technologies, we now have a comprehensive picture of the genetic mutations and structural variants present in GBM (Atlas 2008; Brennan et al. 2013). Some of the most recurrent alterations include EGFR, IDH1, PDGFRA, HDM2, PIK3CA, and TERT promoter and PI3KR1 gain-of-function mutations or amplifications and mutations or deletions of the tumor suppressors PTEN, TP53, CDKN2A, NF1, ATRX, and RB1. While many of these mutations are prevalent in several other cancer genomes, several mutations are highly enriched in GBM, such as IDH1 mutations, which lead to a CIMP (G-CIMP) (Noushmehr et al. 2010). These studies highlight the significant degree of intertumoral heterogeneity present in GBM, which is further captured at both the transcriptional and epigenetic levels (Phillips et al. 2006; Verhaak et al. 2010), and also underscore the complexity of the clonal evolution and clonal diversity that occur during the genesis of GBM and their bearing on the shape and structure of the CSC hierarchy. While both genetic and epigenetic landscapes define functionally distinct clones during tumor evolution, epigenetic differences likely account for the functional differences between cells within the hierarchy.

Epigenetic maintenance of the CSC state is regulated largely at the level of transcriptional and chromatin regulation. CSC regulation converges on MYC, which occurs in the presence of MYC-mediated cancer cell survival and proliferation programs (Wang et al. 2008; Zheng et al. 2008; Wurdak et al. 2010; Chan et al. 2012; Fang et al. 2014). Additional transcription factors have been identified as important for CSC identity, including STAT3 (Sherry et al. 2009), SOX2 (Gangemi et al. 2009), FOXM1 (Joshi et al. 2013), FOXG1 (Verginelli et al. 2013), GLI1 (Clement et al. 2007), ASCL1 (Rheinbay et al. 2013), ZFX (Fang et al. 2014), NANOG (Zbinden et al. 2010), and ZFHX4 (Chudnovsky et al. 2014), which recruit necessary chromatin remodeling factors to promote maintenance of the glioma CSC state. By using epigenome-wide mapping of cellular chromatin state, Suva et al. (2014) identified a core set of four transcription factors in proneural GBM able to reprogram differentiated tumor cells into glioma CSCs. These investigators showed that POU3F2, SOX2, SALL2, and OLIG2 are master transcription factors required to maintain the tumor-forming capability of these cells, suggesting that mediators of stem cell programs could capture the oncogenic capacity of CSCs. In addition to transcription factors, regulators of nucleosome structure have also been reported to maintain the CSC state. The mixed lineage leukemia 1 (MLL1) protein has been shown to maintain the CSC phenotype through activation of HOXA10, which subsequently regulates a network of homeobox genes that is required for tumor maintenance (Heddleston et al. 2012; Gallo et al. 2013). Similarly, the H3K27 methylase EZH2 has been shown to be important for CSC maintenance through its function as a regulator of both Polycomb-repressive domains and STAT3 signaling (Kim et al. 2013). The BMI1 Polycomb ring finger oncogene regulates both normal neural stem cells and GBM cells (Bruggeman et al. 2007).

These studies highlight the importance of understanding the dynamics of core transcription factors in maintaining stem cell state and the effect that these factors have on shaping the chromatin landscape of cells within the tumor hierarchy.

Single-cell RNA sequencing (RNA-seq) interrogation of cellular heterogeneity within GBMs identified novel genes predominantly present in GBM CSCs compared with differentiated cells and provocatively detected cells of multiple GBM subtypes within single tumors, drawing into question the utility of subtyping tumors and targeting specific subtypes (Patel et al. 2014). Furthermore, these investigators described an inverse correlation between stem signature and cell cycle gene expression, suggesting that the cells that form neurospheres in culture cycle more slowly compared with differentiated and differentiating tumor cells. A parallel, single-cell functional analysis of GBMs confirmed a strong variation of genomics and response to therapy (Meyer et al. 2015). Additional detailed analysis of heterogeneity of this type will greatly expand our understanding of the differences between tumor cells both within and among GBM patients and improve the characterization of glioma CSCs.

GBM CSCs reside in varied tumor microenvironments that limit nutrients, such as glucose and oxygen. Under such conditions, cancer cells, including glioma CSCs, exhibit the Warburg effect, a metabolic shift toward aerobic glycolysis and the accumulation of lactate in exchange for sustained ATP production and metabolite generation for macromolecule synthesis. Glioma CSCs demonstrate plasticity in the metabolic pathways used in response to metabolic restrictions and may shift toward the use of the pentose phosphate shunt (Vlashi et al. 2011; Kathagen et al. 2013). This inherent persistence of CSCs under hypoxic and acidic conditions as well as the preferential utilization of HIF-2 signaling compared with NSTCs and normal progenitors promote the maintenance of self-renewal, proliferation, and survival (Li et al. 2009b). Similarly, in conditions of nutrient deprivation such as low glucose, glioma CSCs outcompete neighboring NSTCs for glucose uptake through preferential up-regulation of the high-affinity GLUT3 transporter (Flavahan et al. 2013). A consequence of altered metabolic state is the production of reactive oxygen species. Glioma CSCs not only are dependent on NOS2 activity for promoting tumor growth but also synthesize nitric oxide through the specific up-regulation of NOS2 protein (Eyler et al. 2011). Importantly, in GBM, cellular metabolic characteristics are often genetically hardwired, such as recurrent IDH1 mutations, which are commonly observed in proneural GBM. Mutant IDH1 leads to a gain-of-function enzymatic activity, causing accumulation of 2-hydroxyglutarate, an oncometabolite that inhibits the TET1 and TET2 demethylases to cause aberrant hypermethylation of DNA and histones. While the function of IDH1 mutations in the context of CSCs is not directly defined, IDH1 mutations induce a loss of differentiation, preventing the terminal differentiation of lineage-specific progenitors (Lu et al. 2012). Moving forward, integrated metabolomic and epigenomic profiling may reveal other examples of intricate relationships between metabolism and epigenetic programs and their influence on the glioma CSC state.

Brain development is orchestrated by a series of regulatory pathways with spatially and temporally controlled activity. Notch and NF-B (nuclear factor B) signaling instructs the fate of NSPCs, with the guidance and lineage commitment of progeny dictated by pathways that include the ephrins and bone morphogenetic proteins (BMPs). In a manner that mimics aberrant differentiation, CSCs co-opt developmental programs to maintain an undifferentiated state, increasing their survival and maintenance. Common pathways activated in CSCs include Notch, BMP, NF-B, and Wnt signaling (Li et al. 2009a; Day et al. 2013; Rheinbay et al. 2013; Lubanska et al. 2014; Yan et al. 2014). Collectively, niche factors represent an overriding theme in CSC biology, where stem and progenitor cell features provide selective advantages to maintain tumor growth (). These pathways may be activated through a combination of genetic and epigenetic alterations in addition to microenvironmental and metabolic factors.

The Notch pathway plays a role during neural development, functioning to inhibit neuronal differentiation and sustain NSPC populations. This pathway is co-opted in GBM, where aberrant NOTCH activation stimulates astrocytes to assume a stem-like state accompanied by increased proliferation (Jeon et al. 2008). The importance of Notch signaling in glioma CSC biology is highlighted by the convergence on this pathway from other pathways and exogenous factors, such as hypoxia, eNOS signaling, and response to radiation (Charles et al. 2010; Wang et al. 2010; Qiang et al. 2012). The dependence of glioma CSCs on Notch signaling is further supported by experiments demonstrating depletion of CSCs by treatment with -secretase inhibitors (Fan et al. 2006, 2010).

As BMPs direct NSPC fate toward an astroglial lineage, these signals have been proposed as a possible differentiation therapy for GBM (Piccirillo et al. 2006). Despite the presence of BMP expression in primary GBM tissue, glioma CSCs are highly resistant to the differentiation effects of BMPs in a process that occurs through at least two distinct cell-autonomous mechanisms: the shift to a fetal BMP receptor expression in glioma CSCs through recruitment of the transcriptional repressor EZH2 (Lee et al. 2008) and the secretion of BMP antagonists, specifically Gremlin1, by CSCs to protect against endogenous BMP-mediated differentiation (Yan et al. 2014). In this manner, CSCs generate differentiated progeny that provide supportive cues to the parental cells (e.g., Notch ligands, interleukin-6 [IL-6], and extracellular matrix) while resisting differentiation signals.

The NF-B pathway has emerged as an important regulator of GBM cell survival and identity through an endogenous cell stress response transcriptional program (Bhat et al. 2013). The A20 protein (TNFAIP3), a mediator of cell survival and the NF-B pathway, is overexpressed in CSCs compared with NSTCs (Hjelmeland et al. 2010). Supporting these findings, Sema3C and its receptors, PlexinA2 and PlexinD1, are also coordinately expressed in CSCs and activate Rac1 and NF-B in an autocrine/paracrine loop to promote CSC survival (Man et al. 2014).

GBM CSCs have also been shown to be highly dependent on Ephrin receptor signaling for survival and the maintenance of stem cell properties. Specifically, Ephrin A molecules and the EPHA2 and EPHA3 receptors are highly expressed in glioma CSCs and potentially function through the negative regulation of mitogen-activated protein kinase (MAPK) signaling (Binda et al. 2012; Day et al. 2013).

Wnt signaling is highly active in CSCs and is critical for the maintenance of the stem cell phenotype. An integrated genomic and biological analysis identified PLAGL2 as highly amplified in gliomas with functional suppression of CSC differentiation through modulation of Wnt/-catenin signaling (Zheng et al. 2010). Comprehensive mapping of chromatin modifications in CSCs and their NSTC counterparts revealed widespread activation of Wnt pathway genes through loss of Polycomb-mediated repression. The CSC chromatin landscape is thought to be dependent on achaete scute family basic helixloophelix (bHLH) transcription factor 1 (ASCL1), which activates Wnt signaling through negative regulation of dickkopf WNT signaling pathway inhibitor 1 (DKK1) (Rheinbay et al. 2013). Hedgehog signaling in the CNS is mediated in part by NSPC communication with the cerebrovascular fluid through primary cilia. Gliomas contain primary cilia, and the resulting CSCs are dependent on hedgehog signaling (Bar et al. 2007; Clement et al. 2007; Ehtesham et al. 2007).

Given the role of growth factors in normal brain development, it is not unexpected that numerous canonical growth factor signaling pathways have been shown to contribute to GBM maintenance and function. PDGFR signaling promotes CSC survival, self-renewal, and invasion and tumor growth through downstream STAT3 activation (Kim et al. 2012). Similarly, glioma CSCs preferentially express the IL-6 receptor, which also promotes convergent signaling upon STAT3 activation (Wang et al. 2009).

EGFR signaling has also been reported to contribute to CSC maintenance through the activation of AKT, the recruitment of SMAD5, and the induction of ID3, IL-6, and IL-8. This suggests a potential hypothesis in which the EGFR and PDGFR pathways are linked by IL-6 signaling. A potential alternate hypothesis is the presence of distinct CSC populations dependent on different growth factor receptor signaling pathways. Supporting this latter notion, EGFR inhibition promotes expansion of a cMET growth factor receptor-positive population of CSCs (Jun et al. 2014). Furthermore, elevated cMET expression is important for CSC maintenance, tumorigenicity, and resistance to radiation (Joo et al. 2012).

Aligned with its role in stress responses, transforming growth factor (TGF-) stimulates CSC self-renewal. Autocrine TGF- signaling permits retention of stemness through positive regulation of SOX2 and SOX4 expression (Ikushima et al. 2009). A distinct subset of TGF--dependent CSCs expresses CD44 and ID1 (Anido et al. 2010), which are markers of functionally distinct CSCs. A crucial mediator of the TGF- response in CSCs is the BMI1 protein, which connects stem cell programs and ER stress pathways through the transcriptional repressor ATF3 (Gargiulo et al. 2013).

Immune suppression is a hallmark of cancer (Hanahan and Weinberg 2011); while the brain possesses a unique series of immune surveillance mechanisms that become active during pathogenic states (Ransohoff and Engelhardt 2012), brain tumors have been characterized as immunosuppressive (Platten et al. 2001; Fecci et al. 2006). There is increasing enthusiasm for immunotherapy strategies based on the limited success of signaling pathway inhibitors and anti-angiogenic agents in brain tumors and the success of immunotherapy in melanoma. Immunotherapies for brain tumors include cellular (adoptive T-cell transfer and chimeric antigen receptor engineered T cells), vaccination, and immunomodulatory therapies targeting immune checkpoints (including anti-programmed death 1 [PD1], PD ligand 1 [PD-L1], and cytotoxic T lymphocyte-associated protein 4 [CTLA4] antibodies) (Reardon et al. 2014). Reversing tumor-induced immune suppression by increasing cytotoxic cell function and reducing suppressor cell function may unleash the endogenous immune response. Immunologic therapies may offer an additional benefit, as most strategies do not require intracranial delivery, a major restriction point for many oncologic treatments. While CSCs are key drivers of tumor growth, CSC interactions with the immune system and potential exploitation in immunotherapy are under active investigation (). These studies will require innovative approaches, as the majority of CSC studies involve xenograft models that lack major immune cell components, and many mouse models have reduced cellular heterogeneity. However, the information obtained from mouse model approaches is likely to be informative for the human immune response, as genetically engineered mouse models can recapitulate key aspects of brain tumor immunosuppression (Kong et al. 2010).

Proposed features of CSCs. Non-cell-autonomous aspects of CSCs may drive tumor growth but also serve as points of fragility. These include the increased ability to invade through the brain parenchyma, immune evasion, relationship with a niche, and promotion of angiogenesis.

Despite these challenges, there is building evidence that CSCs directly modulate the immune system. In coculture studies, CSCs induced regulatory T cells while inhibiting proliferation and cytotoxic T-cell activation with a concomitant induction of cytotoxic T-cell apoptosis, mediated via PD1 and soluble galectin-3 (Di Tomaso et al. 2010; Wei et al. 2010). Other CSC-secreted factors include IL-10 and TGF-, which also suppresses tumor-associated microglia/macrophage function and generates a more immunosuppressive (M2) phenotype (Wu et al. 2010). Another immunotherapy approach that may benefit from CSC targeting is the development of anti-tumor vaccines. Current vaccine efforts have focused on tumor-specific antigens (such as EGFRvIII) or whole tumor cell lysates, and there is evidence from preclinical models that CSC lysates are more effective in generating dendritic cell (DC) vaccines than differentiated cells (Pellegatta et al. 2006; Xu et al. 2009). CSCs modulate T-cell and tumor-associated microglia/macrophage function through secreted factors (Zhou et al. 2015), which may be exploited in the development of vaccine strategies or in combination with other drugs (Sarkar et al. 2014). These data provide a rationale for future studies investigating how the interaction between CSCs and other immune cell populations may drive immune suppression and in vivo interrogations into how CSC targeting may alter the immune activation status. Evaluating changes in CSC populations as a result of immunotherapy will also be essential, as will be evaluating combinatorial targeting strategies using immunotherapies and anti-CSC approaches.

Most conventional anti-neoplastic therapies target proliferating cells, but the malignancy of advanced cancers also derives from effects on the immune system, vasculature, and invasion/metastasis (). GBMs infiltrate the surrounding brain, precluding curative surgical resection. Infiltrative tumors must adapt to new environments, including the formation of new vessels to obtain nutrients. GBMs express proangiogenic growth factors (Batchelor et al. 2007), with CSCs driving neoangiogenesis with high levels of VEGF (Bao et al. 2006b). The humanized monoclonal antibody bevacizumab was developed to target VEGF to inhibit angiogenesis and has been used to treat recurrent GBM (Cohen et al. 2009). Bevacizumab attenuates tumor size, but the surviving tumor may display increased invasion in human and mouse models (de Groot et al. 2010), potentially due to a release of c-MET inhibition (Lu et al. 2012). Cancer cells often activate redundant angiogenic pathways in response to VEGF pathway inhibition (Atlas 2008). CSCs located at the perivascular niche are in close contact with the endothelial cells (Calabrese et al. 2007), permitting engagement of endothelial cell Notch ligands with glioma CSC Notch receptors to activate Notch signaling, which supports self-renewal of glioma CSCs (Zhu et al. 2011). CSCs also contribute to vascular structure through transdifferentiation into pericytes to promote tumor growth (Cheng et al. 2013). Inhibition of CSC-derived pericytes disrupts angiogenesis and inhibits tumor growth, directing attention toward nonendothelial cell targeting strategies. Anti-angiogenic drugs in current use have failed to provide a significant survival benefit to GBM patients (Gilbert et al. 2014), suggesting that a benefit may exist to investigating the mechanisms by which tumor cells regulate angiogenesis and that contribute to tumor growth and maintenance to efficiently target the GBM vasculature.

The mainstay treatment of GBM involves surgery, concurrent radiation with chemotherapy, and adjuvant chemotherapy with TMZ (Stupp et al. 2009). Despite advances in the field, the overall survival rate remains only 1519 mo (Stupp et al. 2009). The high degree of tumor heterogeneity in GBM contributes to treatment failure, to which functional and molecular heterogeneity and aberrant receptor tyrosine kinase (RTK) activity all contribute. CSCs located at the top of the hierarchy initiate and maintain the tumor after treatment (Chen et al. 2012). Glioma CSCs have also been shown to contribute to radiation resistance by increasing the DNA damage response machinery (Bao et al. 2006a). In terms of molecular heterogeneity, different subtypes of GBM with distinct molecular profiles coexist within the same tumor and likely exhibit differential therapeutic responses (Sottoriva et al. 2013). For example, several RTKs, including PDGFR in the proneural and EGFR in the classical subtype, are altered in GBM (Verhaak et al. 2010). The abnormal activation of RTKs involves many pathways that are redundant and can initiate and maintain downstream signaling, making tumors refractory to treatment (Stommel et al. 2007). A recent single-cell analysis of primary GBM patients showed that cells from the same tumor have differential expression of genes involved in oncogenic signaling, proliferation, immune response, and hypoxia (Patel et al. 2014). Furthermore, an increase in tumor heterogeneity was associated with a decrease in patient survival. The addition of TMZ to radiation has increased median survival by several months (Stupp et al. 2009), but lineage tracing studies in mouse models demonstrate that CSCs repopulate brain tumors after TMZ treatment (Chen et al. 2012). A number of molecular mechanisms have been identified that mediate the therapeutic resistance of CSCs to cytotoxic therapies, including the DNA damage checkpoint, Notch, NF-B, EZH2, and PARP (Bao et al. 2006a; Wang et al. 2010; Bhat et al. 2013; Venere et al. 2014; Kim et al. 2015), which suggests that CSCs develop multiple mechanisms of resistance that may require combinations of targeted agents. Moving forward, these studies demonstrate the importance of understanding the molecular alterations that are present in recurrent tumors and how these influence the structure of cells within the tumor hierarchy. In addition, it is necessary to consider that therapeutic resistance mechanisms may not be solely innate but may evolve from exposure to microenvironmental factors such as hypoxia and acidic and metabolic stress (Heddleston et al. 2009; Li et al. 2009b; Hjelmeland et al. 2011; Flavahan et al. 2013; Xie et al. 2015).

Conventional treatment for GBM promotes a transient elimination of the tumor and is almost always followed by tumor recurrence, possibly with an increase in the percentage of CSCs (Auffinger et al. 2014), as CSCs are involved in tumor recurrence and therapeutic resistance (Bao et al. 2006a; Chen et al. 2012). To effectively eliminate CSCs, it is critical to target their essential functions and their interactions with the microenvironment. Treatment with TMZ may kill CSCs that contain higher expression of the DNA repair protein MGMT; however, TMZ cannot prevent self-renewal of CSCs that contain MGMT (Beier et al. 2008). Another feature of CSCs is their ability to evade apoptosis. A potential therapeutic strategy would be the use of PARP inhibitors to enhance apoptosis under genotoxic damage. When the PARP inhibitor ABT-888 was used in combination with TMZ and radiation in GBM cell lines, apoptosis increased, and cells were sensitized to therapy (Barazzuol et al. 2013). GBMs thrive in harsh microenvironments characterized by hypoxia and limited nutrient availability. The HIF family of transcription factors is involved in promoting angiogenesis and cell migration in hypoxic regions (Kaur et al. 2005), and several drugs have been developed to target this gene family, with a few undergoing clinical trials. For example, as described previously, glioma CSCs reprogram their metabolic machinery and preferentially take up glucose to survive in environments with limited nutrients by expressing the high-affinity glucose transporter GLUT3 (Flavahan et al. 2013). GLUT3 therefore represents a promising therapeutic target for potential selective inhibition of CSCs. Epigenetic modifications are manifest in tumor recurrence (Nagarajan and Costello 2009). Histone acetylation and methylation are reversible and can be targeted by drugs; the histone deacetylase (HDAC) inhibitor vorinostat is currently in clinical trials (Bezecny 2014). Immunotherapy is an additional emerging therapeutic approach for GBM. The development of vaccines based on heat-shock proteins, EGFRvIII (Del Vecchio et al. 2012), and DCs (Terasaki et al. 2011) has shown promising results in clinical trials. ICT-107, a patient-derived DC vaccine developed against six antigens highly expressed in glioma CSCs (Phuphanich et al. 2013), is currently under clinical evaluation for use in patients.

Some of the challenges of developing therapeutic targeting agents are derived from the lack of universally informative markers to identify CSCs and the common molecular pathways shared by CSCs and NSPCs. The understanding of the biology of the CSCs and how these cells interact with their microenvironment in combination with the genetic and epigenetic landscape in GBM will be essential to develop more effective therapies.

As biological observations have revealed increasing levels of complexity, mathematical modeling approaches have provided a framework to understand the dynamic complexity of stem cell self-renewal and differentiation. By use of proliferation data and lineage tracing analysis, quantitative models have been generated for tissue-specific stem cells that have provided insight into the kinetics of cell fate choice and tissue development (Blanpain and Simons 2013). Similar approaches have been taken to reduce the complexity of CSCs. A network-based model has suggested that CSCs can transition between plastic (proliferative, symmetrically dividing, and less invasive) and rigid (quiescent, asymmetrically dividing, and more invasive) networks that can be modulated by extrinsic stressors, such as hypoxia, inflammation, and therapies (Csermely et al. 2015). Testing this model with biological data is likely to provide additional insights into the complexity of CSCs and identify points of fragility for additional therapeutic development. Mathematical approaches have also been used to evaluate the dynamics of GBM growth. Proliferation and invasion are phenotypes that have been modeled (Harpold et al. 2007). By use of a model that takes into account rates of proliferation and invasion in combination with imaging data, it has been proposed that IDH1 mutant tumors are actually less proliferative and more invasive (Baldock et al. 2014). Clinically relevant parameters, such as identifying optimal radiation schedules, have also been modeled using genetically engineered mice (Leder et al. 2014). Additionally, quantitative approaches have been developed to model the events leading to intertumoral and intratumoral heterogeneity in both human patient specimens (Sottoriva et al. 2013) and mouse models (Cheng et al. 2012). Integrating mathematical approaches into future CSC studies will provide an opportunity to identify key pathways essential for self-renewal and will predict responses to therapeutic perturbations.

GBM provides an excellent system to investigate the importance of CSCs. While there is a standard set of assays used to enrich for and identify CSCs, it remains unclear whether multiple CSC populations exist in different niches (perivascular and hypoxic) and possess different characteristics (slow vs. rapid cycling) as well as how key developmental signaling pathways are used by each of these populations. In addition, while a hierarchy is in place for GBM, the current view of CSCs and NSTCs is mutually exclusive and lacks a progenitor cell population that serves as an intermediate for differentiated progeny generation from somatic stem cells. Mouse studies have revealed that multiple stem and progenitor cell populations have the capacity to give rise to tumors upon oncogenic transformation, but it remains unclear whether there is a single cell of origin for the human disease or, more likely, whether multiple cells of origin exist and how this may be linked to genetic diversity. Making inroads into these unresolved questions will refine the experimental foundation upon which translational studies aiming to develop novel anti-CSC therapies are built and provide key signaling pathways responsible for CSC maintenance that are amenable for targeting.

The extensive molecular characterization of gliomas of all grades has permitted the recognition that the continuum of tumor grade has hidden a set of genetically distinct diseases. IDH1 mutations produce an oncometabolite, 2-hydroxyglutarate, that reprograms cellular chromatin to assume a stem-like state (Lu et al. 2012). Thus, IDH1 mutant gliomas may have a relatively flat hierarchy, with most tumor cells acquiring stem-like features early in tumor initiation. In contrast, primary GBMs accumulate a greater diversity of genetic and epigenetic alterations, which is associated with a more vertical cellular hierarchy. This duality of tumor biology resembles that of the two forms of head and neck cancers. Human papilloma virus-induced head and neck cancers are morphologically uniform and, like IDH1 mutant gliomas, are more responsive to therapies. Alcohol- and tobacco-associated head and neck cancers harbor more mutations and display a worse outcome with a reliable cellular hierarchy. Large-scale genomic sequencing has informed commonalities among cancer types based on driving genetic lesions. It is possible that similar patterns will be appreciated with cancer types based on epigenetic and cellular hierarchies, creating broader opportunities to improve diagnostics and therapeutics. In fact, expanding the organizational structures is likely to be a useful approach to increase our understanding of complex disease states. Many diseases display heterogeneous aspects that are governed by both cell-autonomous and microenvironmental forces. With the success of immunotherapy approaches to activate the immune system via immune checkpoint inhibition in cancers such as melanoma, understanding how GBM and, in particular, CSCs interface with the immune system is an immediate priority. An alternative view of heterogeneity and therapeutic response may also be informative for future studies. For example, bacterial infections contain distinct populations of cells that have different proliferative potential and responses to therapy. Viable but nonculturable bacteria and latent infections, including tuberculosis, may be found in particular niches associated with inflammation, hypoxia, acidic and nitrosative stress, and nutrient restriction (Oliver 2010). Most antibiotics, like anti-neoplastic agents, are directed against the proliferative population, leaving a resistant population behind. Novel methods are being used to screen for new agents that target resistant bacteria, such as latent tuberculosis (Bryk et al. 2008). Nathan (2004) suggested that essentiality is conditional, and the conditions defining essentiality are multiple in the context of latent infections. An identical view can instruct CSC targeting efforts as we grow in our understanding of the governing stimuli both internal and external to CSCs.

One infrequently discussed point is a re-equilibration of a cellular hierarchy in tumors generated from CSCs. If cell-autonomous advantages were the sole determinant of the differentiation state of tumor cells, CSCs would represent the majority of tumor cells, as the evolutionary drive toward increased fitness would provide a selective advantage to CSCs. At steady state (in distinction from homeostasis), tissues balance competing requirements through multiple levels of interaction among stem cells, progenitor cells, and differentiated progeny. Collectively, the individual cellular dynamics in cancer permits tumors to respond to exogenous insults (cytotoxic therapies, immunologic attack, etc.) to maintain the aberrant organ system. These dynamics are also at play within the cellular hierarchy in which CSCs give rise to NSTCs, and, when necessary, NSTCs give rise to CSCs to maintain the cellular equilibrium necessary for optimal tumor growth. CSCs should not be considered a model to simplify the modeling of GBMs and other cancers, but rather the CSC hypothesis constitutes an additional level of complexity that contributes to the malignancy of cancers. As CSCs reside in multiple niches governed by different molecular programs, there will not be single anti-CSC targeted therapeutics with broad activity; instead, CSCs will demand multitargeted approaches. Patients with GBMs are in desperate need of improved therapies. The real validation of CSCs will come with better treatments due to the integration of CSCs into drug development.

We sincerely apologize to those investigators whose work we were unable to cite due to space limitations. We thank Amanda Mendelsohn (Center of Medical Art and Photography, Cleveland Clinic) for assistance with figure preparation. We also thank our funding sources: The National Institutes of Health (grants CA154130, CA171652, CA169117, NS087913, and NS089272 to J.N.R., and CA157948, CA191263, and NS083629 to J.D.L.); Sontag Foundation (J.D.L.); Research Programs Committees of Cleveland Clinic (J.N.R); and James S. McDonnell Foundation (J.N.R). S.C.M. is supported by a Canadian Institutes of Health Research Banting Fellowship. Work in the Lathia laboratory is also supported by the Lerner Research Institute, Case Comprehensive Cancer Center, Voices Against Brain Cancer, Blast GBM, the Ohio Cancer Research Associates, Research Scholar Award from the American Cancer Society, V Scholar Award from the V Foundation for Cancer Research, and grant IRG-91-022-18 to the Case Comprehensive Cancer Center from the American Cancer Society.

Original post:
Cancer stem cells in glioblastoma

Posted in Ohio Stem Cells | Comments Off on Cancer stem cells in glioblastoma

Induced Pluripotent Stem Cells and Their Potential for …

Posted: November 22, 2021 at 2:46 am

Curr Cardiol Rev. 2013 Feb; 9(1): 6372.

1Department of Medicine, University of Minnesota Medical School, Minneapolis, Minnesota, USA

2Stem Cell Institute, University of Minnesota Medical School, Minneapolis, Minnesota, USA

1Department of Medicine, University of Minnesota Medical School, Minneapolis, Minnesota, USA

1Department of Medicine, University of Minnesota Medical School, Minneapolis, Minnesota, USA

2Stem Cell Institute, University of Minnesota Medical School, Minneapolis, Minnesota, USA

3Department of Biomedical Engineering, University of Minnesota, Minneapolis, Minnesota, USA

1Department of Medicine, University of Minnesota Medical School, Minneapolis, Minnesota, USA

2Stem Cell Institute, University of Minnesota Medical School, Minneapolis, Minnesota, USA

3Department of Biomedical Engineering, University of Minnesota, Minneapolis, Minnesota, USA

Received 2012 Jun 11; Revised 2012 Jul 31; Accepted 2012 Aug 27.

Induced pluripotent stem (iPS) cells, are a type of pluripotent stem cell derived from adult somatic cells. They have been reprogrammed through inducing genes and factors to be pluripotent. iPS cells are similar to embryonic stem (ES) cells in many aspects. This review summarizes the recent progresses in iPS cell reprogramming and iPS cell based therapy, and describe patient specific iPS cells as a disease model at length in the light of the literature. This review also analyzes and discusses the problems and considerations of iPS cell therapy in the clinical perspective for the treatment of disease.

Keywords: Cellular therapy, disease model, embryonic stem cells, induced pluripotent stem cells, reprogramm.

Induced pluripotent stem (iPS) cells, are a type of pluripotent stem cell derived from adult somatic cells that have been genetically reprogrammed to an embryonic stem (ES) cell-like state through the forced expression of genes and factors important for maintaining the defining properties of ES cells.

Mouse iPS cells from mouse fibroblasts were first reported in 2006 by the Yamanaka lab at Kyoto University [1]. Human iPS cells were first independently produced by Yamanakas and Thomsons groups from human fibroblasts in late 2007 [2, 3]. iPS cells are similar to ES cells in many aspects, including the expression of ES cell markers, chromatin methylation patterns, embryoid body formation, teratoma formation, viable chimera formation, pluripotency and the ability to contribute to many different tissues in vitro.

The breakthrough discovery of iPS cells allow researchers to obtain pluripotent stem cells without the controversial use of embryos, providing a novel and powerful method to "de-differentiate" cells whose developmental fates had been traditionally assumed to be determined. Furthermore, tissues derived from iPS cells will be a nearly identical match to the cell donor, which is an important factor in research of disease modeling and drug screening. It is expected that iPS cells will help researchers learn how to reprogram cells to repair damaged tissues in the human body.

The purpose of this paper is to summarize the recent progresses in iPS cell development and iPS cell-based therapy, and describe patient specific iPS cells as a disease model, analyze the problems and considerations of iPS therapy in the clinical treatment of disease.

The methods of reprogramming somatic cells into iPS cells are summarized in Table . It was first demonstrated that genomic integration and high expression of four factors, Oct4/Sox2/Klf4/c-Myc or Oct4/Sox2/Nanog/LIN28 by virus, can reprogram fibroblast cells into iPS cells [1-3]. Later, it was shown that iPS cells can be generated from fibroblasts by viral integration of Oct4/Sox2/Klf4 without c-Myc [4]. Although these iPS cells showed reduced tumorigenicity in chimeras and progeny mice, the reprogramming process is much slower, and efficiency is substantially reduced. These studies suggest that the ectopic expression of these three transcription factors (Oct4/Klf4/Sox2) is required for reprogramming of somatic cells in iPS cells.

Various growth factors and chemical compounds have recently been found to improve the induction efficiency of iPS cells. Shi et al., [5] demonstrated that small molecules, able to compensate for Sox2, could successfully reprogram mouse embryonic fibroblasts (MEF) into iPS cells. They combined Oct4/Klf4 transduction with BIX-01294 and BayK8644s and derived MEF into iPS cells. Huangfu et al., [6, 7] reported that 5-azacytidine, DNA methyltransferase inhibitor, and valproic acid, a histone deacetylase inhibitor, improved reprogramming of MEF by more than 100 folds. Valproic acid enables efficient reprogramming of primary human fibroblasts with only Oct4 and Sox2.

Kim et al. showed that mouse neural stem cells, expressing high endogenous levels of Sox2, can be reprogrammed into iPS cells by transduction Oct4 together with either Klf4 or c-Myc [19]. This suggests that endogenous expression of transcription factors, that maintaining stemness, have a role in the reprogramming process of pluripotency. More recently, Tsai et al., [20] demonstrated that mouse iPS cells could be generated from the skin hair follicle papilla (DP) cell with Oct4 alone since the skin hair follicle papilla cells expressed endogenously three of the four reprogramming factors: Sox2, c-Myc, and Klf4. They showed that reprogramming could be achieved after 3 weeks with efficiency similar to other cell types reprogrammed with four factors, comparable to ES cells.

Retroviruses are being extensively used to reprogram somatic cells into iPS cells. They are effective for integrating exogenous genes into the genome of somatic cells to produce both mouse and human iPS cells. However, retroviral vectors may have significant risks that could limit their use in patients. Permanent genetic alterations, due to multiple retroviral insertions, may cause retrovirus-mediated gene therapy as seen in treatment of severe combined immunodeficiency [25]. Second, although retroviral vectors are silenced during reprogramming [26], this silencing may not be permanent, and reactivation of transgenes may occur upon the differentiation of iPS cells. Third, expression of exogenous reprogramming factors could occur. This may trigger the expression of oncogenes that stimulate cancer growth and alter the properties of the cells. Fourth, the c-Myc over-expression may cause tumor development after transplantation of iPS derived cells. Okita et al. [10] reported that the chimeras and progeny derived from iPS cells frequently showed tumor formation. They found that the retroviral expression of c-Myc was reactivated in these tumors. Therefore, it would be desirable to produce iPS cells with minimal, or free of, genomic integration. Several new strategies have been recently developed to address this issue (Table ).

Stadtfeld et al. [16] used an adenoviral vector to transduce mouse fibroblasts and hepatocytes, and generated mouse iPS cells at an efficiency of about 0.0005%. Fusaki et al. [22] used Sendai virus to efficiently generate iPS cells from human skin fibroblasts without genome integration. Okita et al. [27] repeatedly transfected MEF with two plasmids, one carrying the complementary DNAs (cDNAs) of Oct3/4, Sox2, and Klf4 and the other carrying the c-Myc cDNA. This generated iPS cells without evidence of plasmid integration. Using a polycistronic plasmid co-expressing Oct4, Sox2, Klf4, and c-Myc, Gonzalez et al., [28] reprogrammed MEF into iPS cells without genomic integration. Yu et al. [29] demonstrated that oriP/EBNA1 (EpsteinBarr nuclear antigen-1)-based episomal vectors could be used to generate human iPS cells free of exogenous gene integration. The reprogramming efficiency was about 36 colonies/1 million somatic cells. Narsinh et al., [21] derived human iPS cells via transfection of human adipocyte stromal cells with a nonviral minicircle DNA by repeated transfection. This produced hiPS cells colonies from an adipose tissue sample in about 4 weeks.

When iPS cells generated from either plasmid transfection or episomes were carefully analyzed to identify random vector integration, it was possible to have vector fragments integrated somewhere. Thus, reprogramming strategies entirely free of DNA-based vectors are being sought. In April 2009, it was shown that iPS cells could be generated using recombinant cell-penetrating reprogramming proteins [30]. Zhou et al. [30] purified Oct4, Sox2, Klf4 and c-Myc proteins, and incorporated poly-arginine peptide tags. It allows the penetration of the recombinant reprogramming proteins through the plasma membrane of MEF. Three iPS cell clones were successfully generated from 5x 104 MEFs after four rounds of protein supplementation and subsequent culture of 2328 days in the presence of valproic acid.

A similar approach has also been demonstrated to be able to generate human iPS cells from neonatal fibroblasts [31]. Kim et al. over-expressed reprogramming factor proteins in HEK293 cells. Whole cell proteins of the transduced HEK293 were extracted and used to culture fibroblast six times within the first week. After eight weeks, five cell lines had been established at a yield of 0.001%, which is one-tenth of viral reprogramming efficiency. Strikingly, Warren et al., [24] demonstrated that human iPS cells can be derived using synthetic mRNA expressing Oct3/4, Klf4, Sox2 and c-Myc. This method efficiently reprogrammed fibroblast into iPS cells without genome integration.

Strenuous efforts are being made to improve the reprogramming efficiency and to establish iPS cells with either substantially fewer or no genetic alterations. Besides reprogramming vectors and factors, the reprogramming efficiency is also affected by the origin of iPS cells.

A number of somatic cells have been successfully reprogrammed into iPS cells (Table ). Besides mouse and human somatic cells, iPS cells from other species have been successfully generated (Table ).

The origin of iPS cells has an impact on choice of reprogramming factors, reprogramming and differentiation efficiencies. The endogenous expression of transcription factors may facilitate the reprogramming procedure [19]. Mouse neural stem cells express higher endogenous levels of Sox2 and c-Myc than ES cells. Thus, two transcription factors, exogenous Oct4 together with either Klf4 or c-Myc, are sufficient to generate iPS cells from neural stem cells [19]. Ahmed et al. [14] demonstrated that mouse skeletal myoblasts endogenously expressed Sox2, Klf4, and c-Myc and can be easily reprogrammed to iPS cells.

It is possible that iPS cells may demonstrate memory of parental source and therefore have low differentiation efficiency into other tissue cells. Kim et al. [32] showed that iPS cells reprogrammed from peripheral blood cells could efficiently differentiate into the hematopoietic lineage cells. It was found, however, that these cells showed very low differentiation efficiency into neural cells. Similarly, Bar-Nur et al. found that human cell-derived iPS cells have the epigenetic memory and may differentiate more readily into insulin producing cells [33]. iPS cells from different origins show similar gene expression patterns in the undifferentiated state. Therefore, the memory could be epigenetic and are not directly related to the pluripotent status.

The cell source of iPS cells can also affect the safety of the established iPS cells. Miura et al. [54] compared the safety of neural differentiation of mouse iPS cells derived from various tissues including MEFs, tail-tip fibroblasts, hepatocyte and stomach. Tumorigenicity was examined. iPS cells that reprogrammed from tail-tip fibroblasts showed many undifferentiated pluripotent cells after three weeks of in vitro differentiation into the neural sphere. These cells developed teratoma after transplantation into an immune-deficient mouse brain. The possible mechanism of this phenomenon may be attributable to epigenetic memory and/or genomic stability. Pre-evaluated, non-tumorigenic and safe mouse iPS cells have been reported by Tsuji et al. [55]. Safe iPS cells were transplanted into non-obese diabetic/severe combined immunodeficiency mouse brain, and found to produce electrophysiologically functional neurons, astrocytes, and oligodendrocytes in vitro.

The cell source of iPS cells is important for patients as well. It is important to carefully evaluate clinically available sources. Human iPS cells have been successfully generated from adipocyte derived stem cells [35], amniocytes [36], peripheral blood [38], cord blood [39], dental pulp cells [40], oral mucosa [41], and skin fibroblasts (Table ). The properties and safety of these iPS cells should be carefully examined before they can be used for treatment.

Shimada et al. [17] demonstrated that combination of chemical inhibitors including A83-01, CHIR99021, PD0325901, sodium butyrate, and Y-27632 under conditions of physiological hypoxia human iPS cells can be rapidly generated from adipocyte stem cells via retroviral transduction of Oct4, Sox2, Klf4, and L-Myc. Miyoshi et al., [42] generated human iPS cells from cells isolated from oral mucosa via the retroviral gene transfer of Oct4, Sox2, c-Myc, and Klf4. Reprogrammed cells showed ES-like morphology and expressed undifferentiated markers. Yan et al., [40] demonstrated that dental tissue-derived mesenchymal-like stem cells can easily be reprogrammed into iPS cells at relatively higher rates as compared to human fibroblasts. Human peripheral blood cells have also been successfully reprogrammed into iPS cells [38]. Anchan et al. [36] described a system that can efficiently derive iPS cells from human amniocytes, while maintaining the pluripotency of these iPS cells on mitotically inactivated feeder layers prepared from the same amniocytes. Both cellular components of this system are autologous to a single donor. Takenaka et al. [39] derived human iPS cells from cord blood. They demonstrated that repression of p53 expression increased the reprogramming efficiency by 100-fold.

All of the human iPS cells described here are indistinguishable from human ES cells with respect to morphology, expression of cell surface antigens and pluripotency-associated transcription factors, DNA methylation status at pluripotent cell-specific genes and the capacity to differentiate in vitro and in teratomas. The ability to reprogram cells from human somatic cells or blood will allow investigating the mechanisms of the specific human diseases.

The iPS cell technology provides an opportunity to generate cells with characteristics of ES cells, including pluripotency and potentially unlimited self-renewal. Studies have reported a directed differentiation of iPS cells into a variety of functional cell types in vitro, and cell therapy effects of implanted iPS cells have been demonstrated in several animal models of disease.

A few studies have demonstrated the regenerative potential of iPS cells for three cardiac cells: cardiomyocytes, endothelial cells, and smooth muscle cells in vitro and in vivo. Mauritz [56] and Zhang [57] independently demonstrated the ability of mouse and human iPS cells to differentiate into functional cardiomyocytes in vitro through embryonic body formation. Rufaihah [58], et al. derived endothelial cells from human iPS cells, and showed that transplantation of these endothelial cells resulted in increased capillary density in a mouse model of peripheral arterial disease. Nelson et al. [59] demonstrated for the first time the efficacy of iPS cells to treat acute myocardial infarction. They showed that iPS cells derived from MEF could restore post-ischemic contractile performance, ventricular wall thickness, and electrical stability while achieving in situ regeneration of cardiac, smooth muscle, and endothelial tissue. Ahmed et al. [14] demonstrated that beating cardiomyocyte-like cells can be differentiated from iPS cells in vitro. The beating cells expressed early and late cardiac-specific markers. In vivo studies showed extensive survival of iPS and iPS-derived cardiomyocytes in mouse hearts after transplantation in a mouse experimental model of acute myocardial infarction. The iPs derived cardiomyocyte transplantation attenuated infarct size and improved cardiac function without tumorgenesis, while tumors were observed in the direct iPS cell transplantation animals.

Strategies to enhance the purity of iPS derived cardiomyocytes and to exclude the presence of undifferentiated iPS are required. Implantation of pre-differentiation or guided differentiation of iPS would be a safer and more effective approach for transplantation. Selection of cardiomyocytes from iPS cells, based on signal-regulatory protein alpha (SIRPA) or combined with vascular cell adhesion protein-1 (VCAM-1), has been reported. Dubois et al. [60] first demonstrated that SIRPA was a marker specifically expressed on cardiomyocytes derived from human ES cells and human iPS cells. Cell sorting with an antibody against SIRPA could enrich cardiac precursors and cardiomyocytes up to 98% troponin T+ cells from human ESC or iPS cell differentiation cultures. Elliott et al. [61] adopted a cardiac-specific reporter gene system (NKX2-5eGFP/w) and identified that VCAM-1 and SIRPA were cell-surface markers of cardiac lineage during differentiation of human ES cells.

Regeneration of functional cells from human stem cells represents the most promising approach for treatment of type 1 diabetes mellitus (T1DM). This may also benefit the patients with type 2 diabetes mellitus (T2DM) who need exogenous insulin. At present, technology for reprogramming human somatic cell into iPS cells brings a remarkable breakthrough in the generation of insulin-producing cells.

Human ES cells can be directed to become fully developed cells and it is expected that iPS cells could also be similarly differentiated. Stem cell based approaches could also be used for modulation of the immune system in T1DM, or to address the problems of obesity and insulin resistance in T2DM.

Tateishi et al., [62] demonstrated that insulin-producing islet-like clusters (ILCs) can be generated from the human iPS cells under feeder-free conditions. The iPS cell derived ILCs not only contain C-peptide positive and glucagon-positive cells but also release C-peptide upon glucose stimulation. Similarly, Zhang et al., [63] reported a highly efficient approach to induce human ES and iPS cells to differentiate into mature insulin-producing cells in a chemical-defined culture system. These cells produce insulin/C-peptide in response to glucose stimuli in a manner comparable to that of adult human islets. Most of these cells co-expressed mature cell-specific markers such as NKX6-1 and PDX1, indicating a similar gene expression pattern to adult islet beta cells in vivo.

Alipo et al. [64] used mouse skin derived iPS cells for differentiation into -like cells that were similar to the endogenous insulin-secreting cells in mice. These -like cells were able to secrete insulin in response to glucose and to correct a hyperglycemic phenotype in mouse models of both T1DM and T2DM after iPS cell transplant. A long-term correction of hyperglycemia could be achieved as determined by hemoglobin A1c levels. These results are encouraging and suggest that induced pluripotency is a viable alternative to directing iPS cell differentiation into insulin secreting cells, which has great potential clinical applications in the treatment of T1DM and T2 DM.

Although significant progress has been made in differentiating pluripotent stem cells to -cells, several hurdles remain to be overcome. It is noted in several studies that the general efficiency of in vitro iPS cell differentiation into functional insulin-producing -like cells is low. Thus, it is highly essential to develop a safe, efficient, and easily scalable differentiation protocol before its clinical application. In addition, it is also important that insulin-producing b-like cells generated from the differentiation of iPS cells have an identical phenotype resembling that of adult human pancreatic cells in vivo.

Currently, the methodology of neural differentiation has been well established in human ES cells and shown that these methods can also be applied to iPS cells. Chambers et al. [65] demonstrated that the synergistic action of Noggin and SB431542 is sufficient to induce rapid and complete neural conversion of human ES and iPS cells under adherent culture conditions. Swistowsk et al. [66] used a completely defined (xenofree) system, that has efficiently differentiated human ES cells into dopaminergic neurons, to differentiate iPS cells. They showed that the process of differentiation into committed neural stem cells (NSCs) and subsequently into dopaminergic neurons was similar to human ES cells. Importantly, iPS cell derived dopaminergic neurons were functional as they survived and improved behavioral deficits in 6-hydroxydopamine-leasioned rats after transplantation. Lee et al. [67] provided detailed protocols for the step-wise differentiation of human iPS and human ES into neuroectodermal and neural crest cells using either the MS5 co-culture system or a defined culture system (Noggin with a small-molecule SB431542), NSB system. The average time required for generating purified human NSC precursors will be 25 weeks. The success of deriving neurons from human iPS cells provides a study model of normal development and impact of genetic disease during neural crest development.

Wernig et al., [68] showed that iPS cells can give rise to neuronal and glial cell types in culture. Upon transplantation into the fetal mouse brain, the cells differentiate into glia and neurons, including glutamatergic, GABAergic, and catecholaminergic subtypes. Furthermore, iPS cells were induced to differentiate into dopamine neurons of midbrain character and were able to improve behavior in a rat model of Parkinson's disease (PD) upon transplantation into the adult brain. This study highlights the therapeutic potential of directly reprogrammed fibroblasts for neural cell replacement in the animal model of Parkinsons disease.

Tsuji et al., [55] used pre-evaluated iPS cells derived for treatment of spinal cord injury. These cells differentiated into all three neural lineages, participated in remyelination and induced the axonal regrowth of host 5HT+ serotonergic fibers, promoting locomotor function recovery without forming teratomas or other tumors. This study suggests that iPS derived neural stem/progenitor cells may be a promising cell source for treatment of spinal cord injury.

Hargus et al., [69] demonstrated proof of principle of survival and functional effects of neurons derived from iPS cells reprogrammed from patients with PD. iPS cells from patients with Parkinsons disease were differentiated into dopaminergic neurons that could be transplanted without signs of neuro-degeneration into the adult rodent striatum. These cells survived and showed arborization, and mediated functional effects in an animal model of Parkinsons disease. This study suggests that disease specific iPS cells can be generated from patients with PD, which be used to study the PD development and in vitro drug screen for treatment of PD.

Reprogramming technology is being applied to derive patient specific iPS cell lines, which carry the identical genetic information as their patient donor cells. This is particularly interesting to understand the underlying disease mechanism and provide a cellular and molecular platform for developing novel treatment strategy.

Human iPS cells derived from somatic cells, containing the genotype responsible for the human disease, hold promise to develop novel patient-specific cell therapies and research models for inherited and acquired diseases. The differentiated cells from reprogrammed patient specific human iPS cells retain disease-related phenotypes to be an in vitro model of pathogenesis (Table ). This provides an innovative way to explore the molecular mechanisms of diseases.

Disease Modeling Using Human iPS Cells

Recent studies have reported the derivation and differentiation of disease-specific human iPS cells, including autosomal recessive disease (spinal muscular atrophy) [70], cardiac disease [71-75], blood disorders [13, 76], diabetes [77], neurodegenerative diseases (amyotrophic lateral sclerosis [78], Huntingtons disease [79]), and autonomic nervous system disorder (Familial Dysautonomia) [80]. Patient-specific cells make patient-specific disease modeling possible wherein the initiation and progression of this poorly understood disease can be studied.

Human iPS cells have been reprogrammed from spinal muscular atrophy, an autosomal recessive disease. Ebert et al., [70] generated iPS cells from skin fibroblast taken from a patient with spinal muscular atrophy. These cells expanded robustly in culture, maintained the disease genotype and generated motor neurons that showed selective deficits compared to those derived from the patients' unaffected relative. This is the first study to show that human iPS cells can be used to model the specific pathology seen in a genetically inherited disease. Thus, it represents a promising resource to study disease mechanisms, screen new drug compounds and develop new therapies.

Similarly, three other groups reported their findings on the use of iPS cells derived cardiomyocytes (iPSCMs) as disease models for LQTS type-2 (LQTS2). Itzhaki et al., [72] obtained dermal fibroblasts from a patient with LQTS2 harboring the KCNH2 gene mutation and showed that action potential duration was prolonged and repolarization velocity reduced in LQTS2 iPS-CMs compared with normal cardiomyocytes. They showed that Ikr was significantly reduced in iPS-CMs derived from LQTS2. They also tested the potential therapeutic effects of nifedipine and the KATP channel opener pinacidil (which augments the outward potassium current) and demonstrated that they shortened the action potential duration and abolished early after depolarization. Similarly, Lahti et al., [73] demonstrated a more pronounced inverse correlation between the beating rate and repolarization time of LQTS2 disease derived iPS-CMs compared with normal control cells. Prolonged action potential is present in LQT2-specific cardiomyocytes derived from a mutation. Matsa et al., [74] also successfully generated iPS-CMs from a patient with LQTS2 with a known KCNH2 mutation. iPS-CMs with LQTS2 displayed prolonged action potential durations on patch clamp analysis and prolonged corrected field potential durations on microelectrode array mapping. Furthermore, they demonstrated that the KATP channel opener nicorandil and PD-118057, a type 2 IKr channel enhancer attenuate channel closing.

LQTS3 has been recapitulated in mouse iPS cells [75]. Malan et al. [75] generated disease-specific iPS cells from a mouse model of a human LQTS3. Patch-clamp measurements of LQTS 3-specific cardiomyocytes showed the biophysical effects of the mutation on the Na+ current, withfaster recovery from inactivation and larger late currents than observed in normal control cells. Moreover, LQTS3-specific cardiomyocytes had prolonged action potential durations and early after depolarizations at low pacing rates, both of which are classic features of the LQTS3 mutation.

Human iPS cells have been used to recapitulate diseases of blood disorder. Ye et al. [13] demonstrated that human iPS cells derived from periphery blood CD34+ cells of patients with myeloproliferative disorders, have the JAK2-V617F mutation in blood cells. Though the derived iPS cells contained the mutation, they appeared normal in phenotypes, karyotype, and pluripotency. After hematopoietic differentiation, the iPS cell-derived hematopoietic progenitor (CD34+/CD45+) cells showed the increased erythropoiesis and expression of specific genes, recapitulating features of the primary CD34+ cells of the corresponding patient from whom the iPS cells were derived. This study highlights that iPS cells reprogrammed from somatic cells from patients with blood disease provide a prospective hematopoiesis model for investigating myeloproliferative disorders.

Raya et al., [76] reported that somatic cells from Fanconi anaemia patients can be reprogrammed to pluripotency after correction of the genetic defect. They demonstrated that corrected Fanconi-anaemia specific iPS cells can give rise to haematopoietic progenitors of the myeloid and erythroid lineages that are phenotypically normal. This study offers proof-of-concept that iPS cell technology can be used for the generation of disease-corrected, patient-specific cells with potential value for cell therapy applications.

Maehr et al., [77] demonstrated that human iPS cells can be generated from patients with T1DM by reprogramming their adult fibroblasts. These cells are pluripotent and differentiate into three lineage cells, including insulin-producing cells. These cells provide a platform to assess the interaction between cells and immunocytes in vitro, which mimic the pathological phenotype of T1DM. This will lead to better understanding of the mechanism of T1DM and developing effective cell replacement therapeutic strategy.

Lee et al., [80] reported the derivation of human iPS cells from patient with Familial Dysautonomia, an inherited disorder that affects the development and function of nerves throughout the body. They demonstrated that these iPS cells can differentiate into all three germ layers cells. However gene expression analysis demonstrated tissue-specific mis-splicing of IKBKAP in vitro, while neural crest precursors showed low levels of normal IKBKAP transcript. Transcriptome analysis and cell-based assays revealed marked defects in neurogenic differentiation and migration behavior. All these recaptured familial Dysautonomia pathogenesis, suggesting disease specificity of the with familial Dysautonomia human iPS cells. Furthermore, they validated candidate drugs in reversing and ameliorating neuronal differentiation and migration. This study illustrates the promise of disease specific iPS cells for gaining new insights into human disease pathogenesis and treatment.

Human iPS cells derived reprogrammed from patients with inherited neurodegenerative diseases, amyotrophic lateral sclerosis [78] and Huntingtons disease 79, have also been reported. Dimos et al., [78] showed that they generated iPS cells from a patient with a familial form of amyotrophic lateral sclerosis. These patient-specific iPS cells possess the properties of ES cells and were reprogrammed successfully to differentiate into motor neurons. Zhang et al., [79] derived iPS cells from fibroblasts of patient with Huntingtons disease. They demonstrated that striatal neurons and neuronal precursors derived from these iPS cells contained the same CAG repeat expansion as the mutation in the patient from whom the iPS cell line was established. This suggests that neuronal progenitor cells derived from Huntingtons disease cell model have endogenous CAG repeat expansion that is suitable for mechanistic studies and drug screenings.

Disease specific somatic cells derived from patient-specific human iPS cells will generate a wealth of information and data that can be used for genetically analyzing the disease. The genetic information from disease specific-iPS cells will allow early and more accurate prediction and diagnosis of disease and disease progression. Further, disease specific iPS cells can be used for drug screening, which in turn correct the genetic defects of disease specific iPS cells.

iPS cells appear to have the greatest promise without ethical and immunologic concerns incurred by the use of human ES cells. They are pluripotent and have high replicative capability. Furthermore, human iPS cells have the potential to generate all tissues of the human body and provide researchers with patient and disease specific cells, which can recapitulate the disease in vitro. However, much remains to be done to use these cells for clinical therapy. A better understanding of epigenetic alterations and transcriptional activity associated with the induction of pluripotency and following differentiation is required for efficient generation of therapeutic cells. Long-term safety data must be obtained to use human iPS cell based cell therapy for treatment of disease.

These works were supported by NIH grants HL95077, HL67828, and UO1-100407.

The authors confirm that this article content has no conflicts of interest.

Follow this link:
Induced Pluripotent Stem Cells and Their Potential for ...

Posted in Minnesota Stem Cells | Comments Off on Induced Pluripotent Stem Cells and Their Potential for …

Stem cells: past, present, and future | Stem Cell Research …

Posted: November 22, 2021 at 2:09 am

Sukoyan MA, Vatolin SY, et al. Embryonic stem cells derived from morulae, inner cell mass, and blastocysts of mink: comparisons of their pluripotencies. Embryo Dev. 1993;36(2):14858

Larijani B, Esfahani EN, Amini P, Nikbin B, Alimoghaddam K, Amiri S, Malekzadeh R, Yazdi NM, Ghodsi M, Dowlati Y, Sahraian MA, Ghavamzadeh A. Stem cell therapy in treatment of different diseases. Acta Medica Iranica. 2012:7996 https://www.ncbi.nlm.nih.gov/pubmed/22359076.

Sullivan S, Stacey GN, Akazawa C, et al. Quality guidelines for clinical-grade human induced pluripotent stem cell lines. Regenerative Med. 2018; https://doi.org/10.2217/rme-2018-0095.

Amps K, Andrews PW, et al. Screening ethnically diverse human embryonic stem cells identifies a chromosome 20 minimal amplicon conferring growth advantage. Nat. Biotechnol. 2011;29(12):112144.

Google Scholar

Amit M, Itskovitz-Eldor J. Atlas of human pluripotent stem cells: derivation and culturing. New York: Humana Press; 2012.

Google Scholar

Ludwig TE, Bergendahl V, Levenstein ME, Yu J, Probasco MD, Thomson JA. Feeder-independent culture of human embryonic stem cells. Nat Methods. 2006;3:63746.

CAS PubMed Google Scholar

Kang MI. Transitional CpG methylation between promoters and retroelements of tissue-specific genes during human mesenchymal cell differentiation. J. Cell Biochem. 2007;102:22439.

CAS PubMed Google Scholar

Vaes B, Craeye D, Pinxteren J. Quality control during manufacture of a stem cell therapeutic. BioProcess Int. 2012;10:505.

Bloushtain-Qimron N. Epigenetic patterns of embryonic and adult stem cells. Cell Cycle. 2009;8:80917.

CAS PubMed Google Scholar

Brindley DA. Peak serum: implications of serum supply for cell therapy manufacturing. Regenerative Medicine. 2012;7:80917.

Google Scholar

Solter D, Knowles BB. Immunosurgery of mouse blastocyst. Proc Natl Acad Sci U S A. 1975;72:5099102.

CAS PubMed PubMed Central Google Scholar

Hoepfl G, Gassmann M, Desbaillets I. Differentiating embryonic stem cells into embryoid bodies. Methods Mole Biol. 2004;254:7998 https://doi.org/10.1385/1-59259-741-6:079.

Google Scholar

Lim WF, Inoue-Yokoo T, Tan KS, Lai MI, Sugiyama D. Hematopoietic cell differentiation from embryonic and induced pluripotent stem cells. Stem Cell Res Ther. 2013;4(3):71. https://doi.org/10.1186/scrt222.

CAS Article PubMed PubMed Central Google Scholar

Mohr JC, de Pablo JJ, Palecek SP. 3-D microwell culture of human embryonic stem cells. Biomaterials. 2006;27(36):603242. https://doi.org/10.1016/j.biomaterials.2006.07.012.

CAS Article PubMed Google Scholar

Doetschman TC, Eistetter H, Katz M, Schmidt W, Kemler R. The in vitro development of blastocyst-derived embryonic stem cell lines: formation of the visceral yolk sac, blood islands, and myocardium. J Embryol Exp Morphol. 1985;87:2745.

CAS PubMed Google Scholar

Kurosawa HY. Methods for inducing embryoid body formation: in vitro differentiation system of embryonic stem cells. J Biosci Bioeng. 2007;103:38998.

Heins N, Englund MC, Sjoblom C, Dahl U, Tonning A, Bergh C, Lindahl A, Hanson C, Semb H. Derivation, characterization, and differentiation of human embryonic stem cells. Stem Cells. 2004;22:36776.

Rosowski KA, Mertz AF, Norcross S, Dufresne ER, Horsley V. Edges of human embryonic stem cell colonies display distinct mechanical properties and differentiation potential. Sci Rep. 2015;5:Article number:14218.

PubMed Google Scholar

Chung Y, Klimanskaya I, Becker S, Li T, Maserati M, Lu SJ, Zdravkovic T, Ilic D, Genbacev O, Fisher S, Krtolica A, Lanza R. Human embryonic stem cell lines generated without embryo destruction. Cell Stem Cell. 2008;2:1137.

CAS PubMed Google Scholar

Zhang X, Stojkovic P, Przyborski S, Cooke M, Armstrong L, Lako M, Stojkovic M. Derivation of human embryonic stem cells from developing and arrested embryos. Stem Cells. 2006;24:266976.

CAS PubMed Google Scholar

Beers J, Gulbranson DR, George N, Siniscalchi LI, Jones J, Thomson JA, Chen G. Passaging and colony expansion of human pluripotent stem cells by enzyme-free dissociation in chemically defined culture conditions. Nat Protoc. 2012;7:202940.

CAS PubMed PubMed Central Google Scholar

Ellerstrm C, Hyllner J, Strehl R. single cell enzymatic dissociation of human embryonic stem cells: a straightforward, robust, and standardized culture method. In: Turksen K, editor. Human embryonic stem cell protocols. Methods in molecular biology: Humana Press; 2009. p. 584.

Heng BC, Liu H, Ge Z, Cao T. Mechanical dissociation of human embryonic stem cell colonies by manual scraping after collagenase treatment is much more detrimental to cellular viability than is trypsinization with gentle pipetting. Biotechnol Appl Biochem. 2010;47(1):337.

Google Scholar

Ellerstrom C, Strehl R, Noaksson K, Hyllner J, Semb H. Facilitated expansion of human embryonic stem cells by single-cell enzymatic dissociation. Stem Cells. 2007;25:16906.

PubMed Google Scholar

Brimble SN, Zeng X, Weiler DA, Luo Y, Liu Y, Lyons IG, Freed WJ, Robins AJ, Rao MS, Schulz TC. Karyotypic stability, genotyping, differentiation, feeder-free maintenance, and gene expression sampling in three human embryonic stem cell lines deri. Stem Cells Dev. 2004;13:58597.

CAS PubMed Google Scholar

Watanabe K, Ueno M, Kamiya D, Nishiyama A, Matsumura M, Wataya T, Takahashi JB, Nishikawa S, Nishikawa S, Muguruma K, Sasai Y. A ROCK inhibitor permits survival of dissociated human embryonic stem cells. Nat Biotechnol. 2007;25:6816.

CAS PubMed Google Scholar

Nie Y, Walsh P, Clarke DL, Rowley JA, Fellner T. Scalable passaging of adherent human pluripotent stem cells. 2014. https://doi.org/10.1371/journal.pone.0088012.

Thomson JA, Itskovitz-Eldor J, Shapiro SS, Waknitz MA, Swiergiel JJ, Marshall VS, Jones JM. Embryonic stem cell lines derived from human blastocysts. Science. 1998;282:11457.

CAS PubMed Google Scholar

Martin MJ, Muotri A, Gage F, Varki A. Human embryonic stem cellsexpress an immunogenic nonhuman sialic acid. Nat. Med. 2005;11:22832.

CAS PubMed Google Scholar

Smith AG, Heath JK, Donaldson DD, Wong GG, Moreau J, Stahl M, Rogers D. Inhibition of pluripotential embryonic stem cell differentiation by purified polypeptides. Nature. 1988;336(6200):68890. https://doi.org/10.1038/336688a0.

CAS Article PubMed Google Scholar

Xu C, Inokuma MS, Denham J, Golds K, Kundu P, Gold JD, Carpenter MK. Feeder-free growth of undifferentiated human embryonic stem cells. Nature Biotechnol. 2001;19:9714. https://doi.org/10.1038/nbt1001-971.

CAS Article Google Scholar

Weathersbee PS, Pool TB, Ord T. Synthetic serum substitute (SSS): a globulin-enriched protein supplement for human embryo culture. J. Assist Reprod Genet. 1995;12:35460.

CAS PubMed Google Scholar

Chen G, Gulbranson DR, Hou Z, Bolin JM, Ruotti V, Probasco MD, Smuga-Otto K, Howden SE, Diol NR, Propson NE, Wagner R, Lee GO, Antosiewicz-Bourget J, Teng JM, Thomson JA. Chemically defined conditions for human iPSC derivation and culture. Nat. Methods. 2011;8:4249.

CAS PubMed PubMed Central Google Scholar

Sommer CA, Mostoslavsky G. Experimental approaches for the generation of induced pluripotent stem cells. Stem Cell Res Ther. 2010;1:26.

PubMed PubMed Central Google Scholar

Takahashi K, Yamanaka S. Induced pluripotent stem cells in medicine and biology. Development. 2013;140(12):245761 https://doi.org/10.1242/dev.092551.

CAS PubMed Google Scholar

Shi D, Lu F, Wei Y, et al. Buffalos (Bubalus bubalis) cloned by nuclear transfer of somatic cells. Biol. Reprod. 2007;77:28591. https://doi.org/10.1095/biolreprod.107.060210.

CAS Article PubMed Google Scholar

Gurdon JB. The developmental capacity of nuclei taken from intestinal epithelium cells of feeding tadpoles. Development. 1962;10:62240 http://dev.biologists.org/content/10/4/622.

CAS Google Scholar

Kain K. The birth of cloning: an interview with John Gurdon. Dis Model Mech. 2009;2(12):910. https://doi.org/10.1242/dmm.002014.

Article PubMed Central Google Scholar

Davis RL, Weintraub H, Lassar AB. Expression of a single transfected cDNA converts fibroblasts to myoblasts. Cell. 1987;24(51(6)):9871000.

Google Scholar

Quinlan AR, Boland MJ, Leibowitz ML, et al. Genome sequencing of mouse induced pluripotent stem cells reveals retroelement stability and infrequent DNA rearrangement during reprogramming. Cell Stem Cell. 2011;9(4):36673.

CAS PubMed PubMed Central Google Scholar

Maherali N, Sridharan R, Xie W, Utika LJ, Eminli S, Arnold K, Stadtfeld M, Yachechko R, Tchieu J, Jaenisch R, Plath K, Hochedlinger K. Directly reprogrammed fibroblasts show global epigenetic remodeling and widespread tissue contribution. Cell Stem Cell. 2007;1:5570.

CAS PubMed Google Scholar

Ohi Y, Qin H, Hong C, Blouin L, Polo JM, Guo T, Qi Z, Downey SL, Manos PD, Rossi DJ, Yu J, Hebrok M, Hochedlinger K, Costello JF, Song JS, Ramalho-Santos M. Incomplete DNA methylation underlines a transcriptional memory of somatic cells in human IPS cells. Nat Cell Biol. 2011;13:5419.

CAS PubMed PubMed Central Google Scholar

Zhou Q, Brown J, Kanarek A, Rajagopal J, Melton DA. In vivo reprogramming of adult pancreatic exocrine cells to beta-cells. Nature. 2008;455:62732 https://doi.org/10.1038/nature07314.

CAS PubMed Google Scholar

Hilfiker A, Kasper C, Hass R, Haverich A. Mesenchymal stem cells and progenitor cells in connective tissue engineering and regenerative medicine: is there a future for transplantation? Langenbecks Arch Surg. 2011;396:48997.

PubMed Google Scholar

Zhang Wendy, Y., de Almeida Patricia, E., and Wu Joseph, C. Teratoma formation: a tool for monitoring pluripotency in stem cell research. StemBook, ed. The Stem Cell Research Community. June 12, 2012. https://doi.org/10.3824/stembook.1.53.1.

Narsinh KH, Sun N, Sanchez-Freire V, et al. Single cell transcriptional profiling reveals heterogeneity of human induced pluripotent stem cells. J Clin Invest. 2011;121(3):121721.

CAS PubMed PubMed Central Google Scholar

Gertow K, Przyborski S, Loring JF, Auerbach JM, Epifano O, Otonkoski T, Damjanov I, AhrlundRichter L. Isolation of human embryonic stem cell-derived teratomas for the assessment of pluripotency. Curr Protoc Stem Cell Biol. 2007, Chapter 1, Unit 1B 4. 3: 1B.4.1-1B.4.29.

Cooke MJ, Stojkovic M, Przyborski SA. Growth of teratomas derived from human pluripotent stem cells is influenced by the graft site. Stem Cells Dev. 2006;15(2):2549.

CAS PubMed Google Scholar

Przyborski SA. Differentiation of human embryonic stem cells after transplantation in immune-deficient mice. Stem Cells. 2005;23:124250.

PubMed Google Scholar

Tannenbaum SE, Turetsky TT, Singer O, Aizenman E, Kirshberg S, Ilouz N, Gil Y, Berman-Zaken Y, Perlman TS, Geva N, Levy O, Arbell D, Simon A, Ben-Meir A, Shufaro Y, Laufer N, Reubinoff BE. Derivation of xeno-free and GMP-grade human embryonic stem cells- platforms for future clinical applications. PLoS One. 2012;7:e35325.

CAS PubMed PubMed Central Google Scholar

Cohen DE, Melton D. Turning straw into gold: directing cell fate for regenerative medicine. Nat Rev Genet. 2011;12:24352.

CAS PubMed Google Scholar

Hwang NS, Varghese S, Elisseeff J. Controlled differentiation of stem cells. Adv Drug Deliv Rev. 2007;60(2):199214. https://doi.org/10.1016/j.addr.2007.08.036.

CAS Article PubMed PubMed Central Google Scholar

Turner N, Grose R. Fibroblast growth factor signalling: from development to cancer. Nat Rev Cancer. 2010;10:11629.

CAS PubMed Google Scholar

Rao TP, Kuhl M. An updated overview on Wnt signaling pathways: a prelude for more. Circ Res. 2010;106:1798806.

CAS PubMed Google Scholar

Moustakas A, Heldin CH. The regulation of TGFbeta signal transduction. Development. 2009;136:3699714.

CAS PubMed Google Scholar

Excerpt from:
Stem cells: past, present, and future | Stem Cell Research ...

Posted in Stem Cell Therapy | Comments Off on Stem cells: past, present, and future | Stem Cell Research …