Category Archives: Induced Pluripotent Stem Cells


Stem Cell Manufacturing Market Share, Size, Trends 2020- Segments worth Observing Aiding Growth Factors | Merck Group, Becton, Dickinson And Company….

Latest Study on Growth of Global Stem Cell Manufacturing Market 2020-2027. A detailed study accumulated to offer Latest insights about acute features of the Stem Cell Manufacturing market. This Report studies the latest industry trends, market development aspects, market gains, and industry scenario during the forecast period. The report provides the details related to fundamental overview, development status, latest advancements, market dominance and market dynamics. While emphasizing the key driving and restraining forces for this market, the report also offers a complete study of the future trends and developments of the market. It also examines the role of the leading market players involved in the industry including their corporate overview, financial summary and SWOT analysis. This Stem Cell Manufacturing Industry report is consist of the worlds crucial region market share, size, trends including the product profit, price, value, production capacity, capability utilization, supply and demand and industry growth rate.

Stem cell manufacturing is forecasted to grow at CAGR of 6.42% to an anticipated value of USD 18.59 billion by 2027 with factors like rising awareness towards diseases like cancer, degenerative disorders and hematopoietic disorders is driving the growth of the market in the forecast period of 2020-2027.

Download Exclusive Sample (350 Pages PDF) Report: To Know the Impact of COVID-19 on this [emailprotected]https://www.databridgemarketresearch.com/request-a-sample/?dbmr=global-stem-cell-manufacturing-market&AB

Stem cell manufacturing has shown an exceptional penetration in North America due to increasing research in stem cell. Increasing research and development activities in biotechnology and pharmaceutical sector is creating opportunity for the stem cell manufacturing market.

The Global Stem Cell Manufacturing Market 2020 research provides a basic overview of the industry including definitions, classifications, applications and industry chain structure. The Global Stem Cell Manufacturing Market Share analysis is provided for the international markets including development trends, competitive landscape analysis, and key regions development status. Development policies and plans are discussed as well as manufacturing processes and cost structures are also analyzed.

Global Stem Cell Manufacturing Market Segematation By Product (Stem Cell Line, Instruments, Culture Media, Consumables), Application (Research Applications, Clinical Applications, Cell and Tissue Banking), End Users (Hospitals and Surgical Centers, Pharmaceutical and Biotechnology Companies, Clinics, Community Healthcare, Others)

List of TOP KEY PLAYERS in Stem Cell Manufacturing Market Report are

Thermo Fisher Scientific Merck KGaA BD JCR Pharmaceuticals Co., Ltd Organogenesis Inc Osiris Vericel Corporation AbbVie Inc AM-Pharma B.V ANTEROGEN.CO.,LTD Astellas Pharma Inc Bristol-Myers Squibb Company FUJIFILM Cellular Dynamics, Inc RHEACELL GmbH & Co. KG Takeda Pharmaceutical Company Limited Teva Pharmaceutical Industries Ltd ViaCyte,Inc VistaGen Therapeutics Inc GlaxoSmithKline plc ..

Complete Report is Available (Including Full TOC, List of Tables & Figures, Graphs, and Chart)@https://www.databridgemarketresearch.com/toc/?dbmr=global-stem-cell-manufacturing-market&AB

The report can help to understand the market and strategize for business expansion accordingly. In the strategy analysis, it gives insights from marketing channel and market positioning to potential growth strategies, providing in-depth analysis for new entrants or exists competitors in the Stem Cell Manufacturing industry. This report also states import/export consumption, supply and demand Figures, cost, price, revenue and gross margins. For each manufacturer covered, this report analyzes their Stem Cell Manufacturing manufacturing sites, capacity, production, ex-factory price, revenue and market share in global market.

The Global Stem Cell Manufacturing Market Trends, development and marketing channels are analysed. Finally, the feasibility of new investment projects is assessed and overall research conclusions offered.

Global Stem Cell Manufacturing Market Scope and Market Size

Stem cell manufacturing market is segmented on the basis of product, application and end users. The growth amongst these segments will help you analyse meagre growth segments in the industries, and provide the users with valuable market overview and market insights to help them in making strategic decisions for identification of core market applications.

Based on product, the stem cell manufacturing market is segmented into stem cell lines, instruments, culture media and consumables. Stem cell lines are further segmented into induced pluripotent stem cells, embryonic stem cells, multipotent adult progenitor stem cells, mesenchymal stem cells, hematopoietic stem cells, neural stem cells. Instrument is further segmented into bioreactors and incubators, cell sorters and other instruments.

On the basis of application, the stem cell manufacturing market is segmented into research applications, clinical applications and cell and tissue banking. Research applications are further segmented into drug discovery and development and life science research. Clinical applications are further segmented into allogenic stem cell and autologous stem cell therapy.

On the basis of end users, the stem cell manufacturing market is segmented into hospitals and surgical centers, pharmaceutical and biotechnology companies, research institutes and academic institutes, community healthcare, cell banks and tissue banks and others.

Healthcare Infrastructure growth Installed base and New Technology Penetration

Stem cell manufacturing market also provides you with detailed market analysis for every country growth in healthcare expenditure for capital equipment, installed base of different kind of products for stem cell manufacturing market, impact of technology using life line curves and changes in healthcare regulatory scenarios and their impact on the stem cell manufacturing market. The data is available for historic period 2010 to 2018.

The Global Stem Cell Manufacturing Market is highly fragmented and the major players have used various strategies such as new product launches, expansions, agreements, joint ventures, partnerships, acquisitions, and others to increase their footprints in this market. The report includes market shares of stem cell manufacturing market for global, Europe, North America, Asia Pacific and South America.

Key Insights in the report:

Historical and current market size and projection up to 2025

Market trends impacting the growth of the global taste modulators market

Analyze and forecast the taste modulators market on the basis of, application and type.

Trends of key regional and country-level markets for processes, derivative, and application Company profiling of key players which includes business operations, product and services, geographic presence, recent developments and key financial analysis

For More Information or Query or Customization Before Buying, Visit @https://databridgemarketresearch.com/inquire-before-buying/?dbmr=global-stem-cell-manufacturing-market&AB

Opportunities in the market

To describe and forecast the market, in terms of value, for various segments, by region North America, Europe, Asia Pacific (APAC), and Rest of the World (RoW)

The key findings and recommendations highlight crucial progressive industry trends in the Stem Cell manufacturing Market, thereby allowing players to develop effective long term strategies

To strategically profile key players and comprehensively analyze their market position in terms of ranking and core competencies, and detail the competitive landscape for market leaders Extensive analysis of the key segments of the industry helps in understanding the trends in types of point of care test across Europe.

To get a comprehensive overview of the Stem Cell manufacturing market.

With tables and figures helping analyses worldwide Global Stem Cell Manufacturing Market Forecast this research provides key statistics on the state of the industry and is a valuable source of guidance and direction for companies and individuals interested in the market. There are 15 Chapters to display the Stem Cell Manufacturing market.

Chapter 1, About Executive Summary to describe Definition, Specifications and Classification of Stem Cell Manufacturing market, By Product Type, by application, by end users and regions.

Chapter 2, objective of the study.

Chapter 3, to display Research methodology and techniques.

Chapter 4 and 5, to show the Stem Cell Manufacturing Market Analysis, segmentation analysis, characteristics;

Chapter 6 and 7, to show Five forces (bargaining Power of buyers/suppliers), Threats to new entrants and market condition;

Chapter 8 and 9, to show analysis by regional segmentation[North America, Europe, Asia-Pacific etc ], comparison, leading countries and opportunities; Regional Marketing Type Analysis, Supply Chain Analysis

Chapter 10, to identify major decision framework accumulated through Industry experts and strategic decision makers;

Chapter 11 and 12, Stem Cell Manufacturing Market Trend Analysis, Drivers, Challenges by consumer behavior, Marketing Channels

Chapter 13 and 14, about vendor landscape (classification and Market Ranking)

Chapter 15, deals with Stem Cell Manufacturing Market sales channel, distributors, Research Findings and Conclusion, appendix and data source.

Thanks for reading this article; you can also get individual chapter wise section or region wise report version like North America, Europe or Asia or Oceania [Australia and New Zealand]

Contact Us:

Data Bridge Market Research

US: +1 888 387 2818

UK: +44 208 089 1725

Hong Kong: +852 8192 7475

Email:[emailprotected]

Our Upcoming Event:Future of Healthcare Robotics|Digital Conference

HEALTHCARE ROBOTICSis the upcoming future of Medical Sciences. The usage of robotics is already prevalent in surgeries across developed and emerging geographies. It has been quite a while where integration of robotics across healthcare supply chain is being discussed. The Covid 19 Pandemic is expected to speed up the timeline for these innovations and concepts to become reality. With the requirement of telemedicine and robotics in handling samples the healthcare industry is expected to see a digital paradigm shift in the next 2 years. The point here to be discussed is where the future is for healthcare robotics Surgical, Diagnostics or Telemedicine.

REGISTER NOW FOR FREE@https://www.databridgemarketresearch.com/digital-conference/future-of-healthcare-robotics?AB

In the pioneer edition of DBMR Healthcare Robotics Conference come join us in this enriching knowledge fest where experts will speak on new technologies and market dynamics perceptions. Together lets move ahead in the future of medical science and technology.

Read this article:
Stem Cell Manufacturing Market Share, Size, Trends 2020- Segments worth Observing Aiding Growth Factors | Merck Group, Becton, Dickinson And Company....

Janus-Faced PCL2? Alzheimer’s Risk Protein Toggles TREM2 and TLR Pathways – Alzforum

12 Jun 2020

Rare variants in TREM2 and PCLG2 influence a persons odds of developing Alzheimers disease, but that is far from all the two genes have in common. According to a study published June 8 in Nature Neuroscience, phospholipase C 2 acts downstream of TREM2 in a signaling pathway that supports critical microglial functions. Using human microglia derived from induced pluripotent stem cells, researchers led by Joseph Lewcock at Denali Therapeutics in South San Francisco reported that knocking out either gene product prevented the immune cells from efficiently processing lipids and neuronal debris. The researchers also found that, independently of TREM2, PLC2 is involved in a pro-inflammatory side hustle dictated by toll-like receptors, which, it so happens, is exacerbated by intracellular lipid build-up. Taken together, the findings strongly implicate faulty microglial lipid handling in the etiology of AD, and support therapeutic strategies that aim to rev up TREM2 signaling.

Using an impressive array of experimental conditions in gene-edited iPSC-microglia, [the authors] demonstrate that PLC2 is a downstream effector of TREM2 and a regulator of lipid metabolism. This exciting discovery directly connects PLC2 to well-established AD pathways involving APOE, TREM2, and microglial activation, commented Rik van der Kant, Vrije University, Amsterdam (full comment below). Florent Ginhoux of the Agency for Science, Technology and Research in Singapore, agreed. The study elegantly links TREM2 and PLC2 signaling pathways, and offers mechanistic insight into how variants in these genes affect the pathophysiology of AD, Ginhoux wrote (full comment below).

Double Dealing. When triggered by TREM2, PLC2 supports lipid metabolism and survival (left). When triggered by TLRs, PLC2 triggers inflammation. In TREM2 KO microglia (right), lipids accumulate and this exacerbates the pro-inflammatory, TLR-driven pathway. [Courtesy of Andreone et al., Nature Neuroscience, 2020.]

Since the discovery, in 2012, that rare variants in the coding region of TREM2 triple the risk of AD, researchers have pegged the receptor as supporting myriad microglial functions, including phagocytosis, walling off A plaques, and promoting an anti-inflammatory, neuroprotective environment (May 2016 news; Apr 2017 conference news;Jul 2018 conference news).

Separately, researchers discovered a rare variant in phospholipase C 2 (PLCG2) that protects against AD (Aug 2017 conference news on Sims et al., 2017). PLCs are a large family of intracellular enzymes that cleave the membrane phospholipid phosphatidylinositol-4,5-bisphosphate (PIP2) to diacylglycerol (DAG) and inositol-1,4,5-trisphosphate (IP3), a process that facilitates calcium signaling. In the brain, the 2 isoform is predominantly expressed by microglia, and initial studies suggest that the protective variant munches phospholipids with more gusto than the common one does (Zhang et al., 2014; May 2019 news).

Might the functions of TREM2 and PLC2 intersect in microglia? To study this question, co-first authors Benjamin Andreone and Laralynne Przybyla derived human microglia. They wove together elements from three recently developed protocols to coax so-called induced microglia (iMGs) from induced pluripotent stem cells (Muffat et al., 2016; Pandya et al., 2017; McQuade et al., 2018). They then used CRISPR to wipe out expression of TREM2 or PLCG2 in these cell-based models.

Under normal conditions, iMGs missing either TREM2 or PLCG2 appeared healthy and viable. When the going got toughi.e., when growth factors were depleted from the culture mediaboth types of knockout suffered a similar fate, dying sooner than their wild-type counterparts. The transcriptomes of each of the two iMG knockouts also differed from those of wild-type cells in similar ways. Specifically, half of the genes differentially expressed in TREM2 KO iMGs were similarly affected in PLCG KO iMGs. These common genes were part of signal transduction pathways downstream of DAP12, the adaptor protein that mediates TREM2 signaling. Using biochemical approaches, the researchers ultimately pieced together a signaling cascade by which lipids activate TREM2, leading to the phosphorylation of Syk2, which directly interacts with PLC2, unleashing its phospholipase activity and downstream signaling events.

Disabling the pathway, either by knocking out TREM2 or PLC2, had a dramatic impact on the processing of lipids, including cholesterol-laden myelin. All microglial lines in this study readily engulfed this type of fluorescently labeled debris; however, while wild-type cells had largely disposed of it after four days, TREM2 or PLCG2 knockouts were still chock-full of it by then. Tellingly, perhaps, the knockout cells failed to ramp up expression of several lipid processing genes in response to the myelin challenge.

Choking on Lipids? Wild-type microglia (left) readily digested lipids after treatment with myelin, while microglia lacking PLCG2 (middle) and TREM2 (right) accumulated the lipids. [Courtesy of Andreone et al., Nature Neuroscience, 2020.]

Lipidomics experiments revealed that the knockouts became burdened with a backlog of several subtypes of unprocessed lipid, including free cholesterol, cholesteryl esters, and myelin-derived ceramides. Similarly, in co-culture experiments with iPSC-derived neurons, both types of microglial knockout were unable to properly digest detritus from injured axons.

How might AD risk variants shift these phenotypes? The researchers generated iMGs that expressed the R47H variant of TREM2, or the protective P522R variant of PLCG2. As might be expected from prior findings on these variants, the R47H-TREM2 iMGs processed lipids more sluggishly than wild-type, whereas the P522R-PLCG2 microglia more deftly disposed of them than wild-type. Together, the findings support the idea that TREM2 and PLCG2 variants influence AD risk via lipid metabolism.

Lest a reader be tempted to tie a neat little bow on this set of results, here comes the twist: PLC2 also takes marching orders from toll-like receptors. This was previously reported in peripheral immune cells. The Denali researchers found the same in iMGs, as PLCG2 knockouts failed to mount a pro-inflammatory response to the TLR2 ligand zymosan.

Interestingly, the same pro-inflammatory cytokines that were down in response to zymosan in PLCG2 knockout iMGs were up in TREM2 knockout iMGs. For example, compared with wild-type iMGs treated with zymosan, PLCG2 knockouts secreted 50 percent less IL-1, while TREM2 knockouts secreted 64 percent more.

The same pattern emerged when the researchers used the TLR4 ligand LPS to trigger the microglial NLRP3 inflammasome, which itself has been tied to AD (Nov 2019 news). Loading up the microglia with myelin prior to triggering the inflammasome dramatically enhanced the inflammatory response in TREM2 KO iMGs, the scientists report. This implies that intracellular lipid accumulation may exacerbate damaging inflammatory pathways. The findings dovetail with those of a recent study that tied lipid droplet-accumulating microglia (LAM) in the aging hippocampus to neuroinflammation (Aug 2019 news).

Overall, the findings cast PLC2 as a two-faced player in microglia. When triggered via TREM2, this phospholipase facilitates processing of lipids and microglial survival. When tripped off by TLRs, it ramps up potentially damaging pro-inflammatory responses. And when lipids build up, as might occur in the aging brain, they exacerbate the pro-inflammatory pathway, Andreone told Alzforum. He believes the balance between these two PLC2 signaling pathways could dictate whether microglia help or harm.

The findings lend support to a therapeutic strategy of agonizing TREM2 signaling, Lewcock told Alzforum. That the protective PLC2 variant enhances lipid processing in microglia fits with the idea that even people whose TREM2 functions normally could stand to benefit from a boost in this pathway. Activating PLC2 is also a potential strategy, Lewcock said, although it would come with the risk of rousing its pro-inflammatory side. More work is needed to dissect how the PLC2 protective variant influences signaling downstream of TREM2 versus TLRs.

This is a very important paper, wrote Christian Haass at the German Center for Neurodegenerative Diseases in Munich. Haass noted that its findings fit with fresh data from his and other groups, but also cautioned that the molecular signature of a protective subpopulation of microglia needs to be defined in much greater detail (full comment below).

Denali is collaborating with Haass group to develop an activating antibody for TREM2, which will come with a blood-brain barrier transport vehicle to shuttle it into the brain (May 2019 conference news;May 2020 news).AL002, a TREM2-activating antibody developed by Alector and Abbvie, entered early clinical trials last year (see clinicaltrials.gov).Jessica Shugart

No Available Further Reading

Read more here:
Janus-Faced PCL2? Alzheimer's Risk Protein Toggles TREM2 and TLR Pathways - Alzforum

Evera, A Harvard Consumer Biotech Company, Brings Stem Cell Banking To You – Forbes

Throughout the past decade, consumer biology tests have been all the rage. Companies such as 23andMe and Ancestry DNA have made their test kits accessible to every day Americans. One can screen for anomalies in their genetic code or identify their lineage. With recent advances in stem cell research, a new opportunity within the consumer biology market has appeared. Nabeel Quryshi, Michael Chen and Zeel Patel are three Harvard undergraduates who observed the unmet, rising demand of control over ones stem cells. They worked together to create Evera, the first at-home stem cell banking company. The three Harvard students are joined by the schools world-renowned biology professor, Dr. George Church. The Cambridge, Massachusetts-based company was incubated at the Harvard Innovation Lab, and has former NASA astronaut Scott Kelly as a investor.

Evera cofounders from left to right: Nabeel Quryshi, Michael Chen and Zeel Patel.

Kelly says, "I did a lot of my independent research, consulted with NASA physicians and scientists, and experts in the stem cell for cancer treatment fields. All those discussions and research indicated that this technology has merit."

Frederick Daso: What led you and your team to identify that stem cells could be potentially used to prevent neurodegenerative disease?

Nabeel Quryshi: I wouldn't single out a focus on neurodegenerative diseases. However, over the last decade, there has been a flurry of research around the use of stem cells to treat conditions such as Parkinson's, Dementia, Alzheimer's, etc. People are working on prevention, but there are two main use cases of stem cells currently. One is for treatment (replacement of damaged or lost cells), and the other is disease modeling (being able to model diseases and test the effects of new drugs completely in vitro without having to get a biopsy).

Daso: In the same ways that blood banks function, how did you manage to apply that concept to the storage of stem cells over a long time?

Quryshi: Cord blood banks and academic stem cell banks that use standardized cryopreservation protocols have been around for a while. The main innovation behind Evera was developing technology around the collection and preservation of urine-derived cells.

Daso: Why don't more mothers store their children's cord blood in stem cell banks? Is it mostly due to a price issue, or is there some other factor at play?

Quryshi: From the countless interviews we've done, it seems to be a price issue. Additionally, it's hard to make a sale around the time of birth as families have countless other things to worry about that are more immediate to the birth of a child.

Daso: What would be driving the growth of this market both now and in the future?

Quryshi: The growth of new cutting edge cell therapies is certainly further demonstrating the need for personal cell biobanking. Furthermore, the success of the direct to consumer genetic testing industry (23andMe, Ancestry, etc.) is a significant driver of growth. From the research we've conducted and the customers we have spoken to, individuals who have already taken 23andMe or another genetic test and know what they are at risk for genetically are looking for ways to take tangible action. Evera is that next step. Instead of just understanding what your future genetic risk is, Evera allows you to make a real biological investment in your future health and wellbeing. While knowing you're at risk for saying Parkinson's is excellent, being able to set aside your youngest cells so that one day you may be able to combat the effects of such a disease is terrific.

However, one should note that although the growth and technology coming from the cell therapy and stem cell therapy industry is astonishing, these are still projections. We have yet to see a fully FDA approved therapy that utilizes the specific types of stem cells we use (induced pluripotent stem cells). Nevertheless, by the time such treatments make it to the clinic, your cells will have aged significantly, and thus it makes sense to save them away now.

Daso: Could you walk me through the thought process of figuring out how to extract stem cells from urine? (From what I know, stem cells usually come from other parts of your body!)

Quryshi: Until around 2011/2012, you would have been right. However, utilizing the fantastic technology that comprised Dr. Yamanaka's 2006 Nobel Prize, scientists have been able to convert any cell in the human body to a kind of stem cell called an induced pluripotent stem cell. This cell has the capability of being able to differentiate into any cell type in the human body. We have advanced tech around the conversation of urine-derived cells to these iPSCs.

Daso: How have you designed your D2C service to ensure that a customer's DNA and associated data are not at risk?

Quryshi: We take data and privacy extremely seriously. We are well aware of the concerns people already have to D2C genetics products. To ensure the confidentiality and privacy of your data and sample, we separate your personally identifiable information from sample information and simultaneously use multiple layers of encryption and cryptography. Your sample and associated data cannot be associated with you individually. Furthermore, our facility is monitored 24/7 with top of the line security measures. We believe that your sample is your property.

Daso: What was the turning point during your undergrad to pursue this idea?

Quryshi: Having worked at 23andMe, I was able to get the lucky opportunity to be a part of arguably the world's most successful consumer genetics company. I saw first hand the benefits of providing customers with their genetic risk. Yet, I discovered that merely providing such risk predictions may not be enough led me to found Evera on the notion that tangibly investing in one's future health and wellbeing through cell banking will propel us into the age of personalized medicine.

Daso: How do you leverage your advisory board to navigate regulations and moral hazards in this space?

Quryshi: We have assembled a dream team consisting of experts in stem cell banking and cell therapy. Our co-founders and advisors comprise of professors from Harvard and Stanford, executives from companies such as Verily as well as top grad students and postdocs in stem cell biology from Harvard and Stanford. We work collaboratively to make sure we adhere to all regulations and ensure the secure preservation of our customer's cells.

If you enjoyed this article, feel free to check out my other work onLinkedInand my personal website,frederickdaso.com. Follow me on Twitter@fredsoda, on Medium@fredsoda, and on Instagram@fred_soda.

Originally posted here:
Evera, A Harvard Consumer Biotech Company, Brings Stem Cell Banking To You - Forbes

Here’s Why Fate Therapeutics Rose 18.4% in May – The Motley Fool

What happened

Shares of Fate Therapeutics (NASDAQ:FATE) gained over 18% last month, according to data provided by S&P Global Market Intelligence. Most of the stock's gains in May can be traced to a single announcement by the cell-therapy developer.

On May 20, the development-stage company announced the U.S. Food and Drug Administration (FDA) cleared a new drug candidate to begin clinical trials. Identified as FT538, the drug candidate is the first cell therapy that has been both engineered with CRISPR gene-editing tools and derived from induced pluripotent stem cells (iPSC). The combination could lead to safer, more effective, and significantly lower-cost drug products.

Investors cheered the latest sign of progress for the early stage pipeline -- and the momentum hasn't waned. In fact, a public offering of common stock on June 9 triggered additional gains for the pharma stock. Apparently, investors are content with dilution so long as Fate Therapeutics maintains a well-funded balance sheet.

Image source: Getty Images.

Fate Therapeutics has one of the most ambitious pipelines in cell therapy, spanning 13 unique programs and multiple cell types. Until recently, investors had few tangibles to analyze, but promising (very) early-stage data and a multi-billion-dollar partnership with Johnson & Johnson subsidiary Janssen have de-risked the stock.

It might be a bit silly to get excited about a preclinical asset moving to clinical trials, but FT538 could prove to be an important bellwether for Fate Therapeutics. If researchers prove that gene-editing tools can be used with reproducible results on cells grown from master cell lines, such as iPSCs, then it would be a big step forward for the field of cell therapy. The capabilities would enable the relatively quick engineering of cell therapies, both for efficacy and safety, and allow living drug products to be manufactured at scales and costs simply not possible today.

Including cash on hand at the end of March and the expected proceeds from the stock offering on June 9, Fate Therapeutics should begin the second half of 2020 with at least $350 million in cash. That should be enough to generate results from a handful of ongoing clinical trials, but investors shouldn't forget that the company's ambitious pipeline will require many hundreds of millions of dollars to develop.

See the original post here:
Here's Why Fate Therapeutics Rose 18.4% in May - The Motley Fool

Nanion Technologies and Nexel Partner to Open a New Reference Demonstration Laboratory in South Korea – Labmate Online

Nanion Technologies and Nexel are pleased to announce a partnership, focused on combining Nanions CardioExcyte 96 and FLEXcyte 96 cell monitoring technology with Nexels hiPSC-derived cells for demonstration purposes. Bringing together the two companies infrastructure and expertise serves to meet the growing demand for a reliable, high throughput cell monitoring technology in Asia.

The Nanion- Nexel partnership brings together profound skills in comprehensivein vitroelectrophysiology technology and development of human induced pluripotent stem cells (hiPSCs), with focus on cardiomyocytes. Under the partnership, Nexel opens a reference demonstration laboratory for Nanions systems at Nexels headquarters in Seoul, whereby both companies aim to significantly upscale support of their clients in Asia.

Dr Choong-Seong Han, CEO of Nexel, said: Nexel is proud to start this partnership with Nanion Technologies. We believe it will further build on the excellent relationship we have developed together in the last year. The Cardiosight-S cardiomyocytes have been fully validated on the CardioExcyte 96 and FLEXcyte 96 systems and our expert scientists are dedicated to provide the best demo settings as well as product experience for customers, as part of the collaboration. We hope interest in both Nanions and Nexels offerings will increase with our collaborative efforts.

Frank Henrichsen, Director of Global Sales of Nanion Technologies added: We are very eager to strengthen our position in the Asian market and especially in Korea. In Nexel, we see a valuable partner to help us develop our presence, in this case through opening their laboratories and enabling the use Nanions technology for demo purposes at their premises. Combining Nexels hiPSC-derived cardiomyocytes and cardiacin vitroassays with Nanions CardioExcyte 96 and FLEXcyte 96 systems, we are confident that our customers will get an excellent package solution for use in safety pharmacology and toxicology assays. We are also very happy that Nexel has already implemented the systems into their quality control procedure of Cardiosight-S cardiomyocytes.

Go here to see the original:
Nanion Technologies and Nexel Partner to Open a New Reference Demonstration Laboratory in South Korea - Labmate Online

Induced Pluripotent Stem Cells Market Growth Dynamics …

Induced pluripotent stem cells (iPSCs) hold profound potential in replacing the use of embryonic stem cells (ESCs) as important tool for drug discovery and development, disease modeling, and transplantation medicine. Advent of new approaches in reprogramming of somatic cells to produce iPSCs have considerably advanced stem cell research, and hence the induced pluripotent stem cells market. The iPSC technology has shown potential for disease modeling and gene therapy in various areas of regenerative medicine. Notable candidates are Parkinsons disease, spinal cord trauma, myocardial infarction, diabetes, leukemia, and heart ailments.

Over the past few years, researchers have come out with several clinically important changes in reprogramming process; a case in point is silencing retroviruses in the human genome. Molecular mechanisms that underlie reprogramming have gained better understanding. However, the tools based on this growing understanding are still in nascent stage. Several factors affect the efficiency of reprogramming, most notably chromosomal instability and tumor expression. These have hindered researchers to utilize the full therapeutic potential of iPSCs, reflecting an unmet need, and hence, a vast potential in the induced pluripotent stemcellsmarket.

Get Brochure of the Report @ https://www.tmrresearch.com/sample/sample?flag=B&rep_id=6245

Global Induced Pluripotent Stem Cells Market: Growth Dynamics

The growing application of induced pluripotent stem cells in generating patient-specific stem cells for drug development and human disease models is a key dynamic shaping their demands. Growing focus on personalized regenerative cell therapies among medical researchers and healthcare proponents in various countries have catalyzed their scope of induced pluripotent stem cells market. Advent of new methods to induce safe reprogramming of cells have helped biotechnology companies improve the clinical safety and efficacy of the prevailing stem cells therapies. The relentless pursuit of alternative source of cell types for regenerative therapies has led industry players and the research fraternity to pin hopes on iPSCs to generate potentially a wide range of human cell types with therapeutic potential.

Advances pertaining to better utilizing of retrovirus and lentivirus as reprogramming transcription factors in recent years have expanded the avenue for players in the induced pluripotent stem cells market. Increasing focus on decreasing the clinical difference between ESCs and iPSCs in all its entirety has shaped current research in iPSC technologies, thus unlocking new, exciting potential for biotechnology and pharmaceutical industries.

Global Induced Pluripotent Stem Cells Market: Notable Development

Over the past few years, fast emerging markets in the global induced pluripotent stem cells are seeing the advent of patents that unveil new techniques for reprogramming of adult cells to reach embryonic stage. Particularly, the idea that these pluripotent stem cells can be made to form any cells in the body has galvanized companies to test their potential in human cell lines. Also, a few biotech companies have intensified their research efforts to improve the safety of and reduce the risk of genetic aberrations in their approved human cell lines. Recently, this has seen the form of collaborative efforts among them.

Buy this Premium Report @ https://www.tmrresearch.com/checkout?rep_id=6245&ltype=S

Lineage Cell Therapeutics and AgeX Therapeutics have in December 2019 announced that they have applied for a patent for a new method for generating iPSCs. These are based on NIH-approved human cell lines, and have been undergoing clinical-stage programs in the treatment of dry macular degeneration and spinal cord injuries. The companies claim to include multiple techniques for reprogramming of animal somatic cells.

Such initiatives by biotech companies are expected to impart a solid push to the evolution of the induced pluripotent stem cells.

Global Induced Pluripotent Stem Cells Market: Regional Assessment

North America is one of the regions attracting colossal research funding and industry investments in induced pluripotent stem cells technologies. Continuous efforts of players to generate immune-matched supply of pluripotent cells to be used in disease modelling has been a key accelerator for growth. Meanwhile, Asia Pacific has also been showing a promising potential in the expansion of the prospects of the market. The rising number of programs for expanding stem cell-based therapy is opening new avenues in the market.

Get Table of Content of the Report @ https://www.tmrresearch.com/sample/sample?flag=T&rep_id=6245

About TMR Research

TMR Research is a premier provider of customized market research and consulting services to business entities keen on succeeding in todays supercharged economic climate. Armed with an experienced, dedicated, and dynamic team of analysts, we are redefining the way our clients conduct business by providing them with authoritative and trusted research studies in tune with the latest methodologies and market trends.

Contact:TMR Research,3739 Balboa St # 1097,San Francisco, CA 94121United StatesTel: +1-415-520-1050Visit Site: https://www.tmrresearch.com/

Originally posted here:
Induced Pluripotent Stem Cells Market Growth Dynamics ...

Generation of self-organized sensory ganglion organoids and retinal ganglion cells from fibroblasts – Science Advances

INTRODUCTION

A ganglion is a cluster or group of nerve cells found in the peripheral nervous system (PNS) or central nervous system (CNS). They often interconnect with each other and with other structures in the PNS and CNS to form a complex nervous network. There are three groups of ganglia in the PNS, which are the dorsal root ganglia (DRG), cranial nerve ganglia, and autonomic ganglia, and two types of ganglia in the CNS, which are the basal ganglia in the brain and retinal ganglion in the retina. Unlike other ganglia, which are essentially cell clusters, retinal ganglia consist of a layer/sheet of dispersive retinal ganglion cells (RGCs). Diverse types of neurons in the somatosensory ganglia such as DRG are specialized for different sensory modalities such as proprioception, mechanoreception, nociception (i.e., pain perception), thermoception, and pruriception (i.e., itch perception) (1, 2). Similarly, there are numerous subtypes of RGCs that are specialized for transmitting from the retina different visual information (e.g., color, contrast, and motion direction) to the central visual system in the brain (3). In the human, a variety of pain, itch, neurological, and degenerative disorders affect sensory ganglia (SGs) and RGCs. Mutations in the FXN (frataxin) and IKBKAP genes, for example, result in debilitating Friedreichs ataxia and familial dysautonomia, respectively (4, 5). Dominant gain-of-function mutations in the sodium channel Nav1.7 gene SCN9A, which is expressed in sensory neurons, are linked to two severe pain syndromesinherited erythromelalgia and paroxysmal extreme pain disorder, while its recessive loss-of-function mutations cause dangerous congenital insensitivity to pain (6). Recently, peripheral SG dysfunction has also been linked to tactile sensitivity and other behavioral deficits associated with the autism spectrum disorders (7). Both genetic and environmental risk factors contribute to glaucoma, which is a leading cause of blindness worldwide and characterized by progressive degeneration of RGCs and the optic nerve (8).

Despite the difference in morphology and embryonic origin, somatosensory and retinal ganglia share extensive overlap of gene expression and we proposed more than two decades back that both might also share genetic regulatory hierarchies (9, 10). This assumption has largely turned out to be the case. During embryogenesis, somatosensory ganglion neurons arise from the multipotent neural crest (NC) cells through a process of cell migration and coalescence (1). RGCs are also derived from multipotent retinal progenitor cells and destined to the ganglion cell layer by migration. It has been shown that the neurogenic bHLH transcription factors (TFs) Ngn1 and/or Ngn2 are involved in the determination of peripheral sensory neurons (11), and that the homeodomain TFs Isl1 and Brn3a or Brn3b are required for the specification and differentiation of different subtypes of neurons in the somatosensory and retinal ganglia (1217). Moreover, there is substantial functional redundancy between Ngn1 and Ngn2 as well as between Brn3a and Brn3b in the development of sensory neurons and RGCs (11, 18, 19).

Somatic cell reprogramming by defined TFs into sensory neurons provides a powerful strategy for studying mechanisms of SG development and sensory disease pathogenesis and for generating cells for patient-specific cell replacement therapy, drug screening, and in vitro disease modeling. It has been shown recently that nociceptor and other subtypes of sensory neurons can be directly induced from murine and human fibroblasts by Brn3a and Ngn1 or Ngn2 or by a combination of five TFs including Ascl1, Ngn1, Isl2, Myt1l, and Klf7 (20, 21). The induced sensory neurons express characteristic marker proteins and are electrically active and selectively responsive to various agonists known to activate pain- and itch-sensing neurons (20, 21). However, networked SG did not appear to be consistently generated in these cases, and it is unclear whether RGCs were induced by these combinations of TFs.

Given the advantages of organoids in studying developmental mechanisms and modeling and treating relevant diseases, we sought to generate ganglion organoids and RGCs from mouse and human fibroblasts using TFs controlling in vivo development of sensory and retinal ganglia. The extensive molecular homology between SG neurons and RGCs creates a dilemma as to how to distinguish these two types of neurons. In the past, several RGC markers including Brn3a, Brn3b, Isl1, Thy1.2, Sncg, Math5, Rbpms, and RPF-1 were used to identify RGCs induced from embryonic stem cells (ESCs), induced pluripotent stem cells (iPSCs), and somatic cells (2225). However, this is a questionable practice because although these markers are sufficient to identify RGCs within the retina, they are inadequate as specific markers for identifying induced RGCs (iRGCs), given their expression in SG and other CNS regions as well (10, 15). We thus carefully screened for RGC-specific markers by comparing expression patterns of numerous known markers in the retina and DRG. This analysis revealed Pax6 expression in RGCs but not in DRG and that RGCs can be identified as Pax6+Brn3a+ or Pax6+Brn3b+ double-positive cells. Equipped with this knowledge, we set to generate induced SG (iSG) organoids and iRGCs from fibroblasts by testing the combination of Ascl1, the pioneer neurogenic TF for somatic cell reprogramming of neurons (26), with a variety of SG and retinal TFs. This screen identified a triple-factor combination ABI (Ascl1-Brn3a/3b-Isl1) as the most efficient way to induce self-organized and networked iSG and iRGCs from fibroblasts.

Previous studies by our group and others have demonstrated that SGs and RGCs share similar transcriptional regulatory mechanism for their development, for instance, both Brn3 TFs (Brn3a and Brn3b) and Isl1 are involved in the specification and differentiation of DRG neurons and RGCs (1214, 16). More recently, Ascl1 has been shown to play a pioneering role in induced neuron (iN) reprogramming from somatic cells (26). As a first step to generate iSG and iRGCs directly from somatic cells, we sought to induce SG neurons and RGCs from mouse embryonic fibroblasts (MEFs) by testing the combination of Ascl1 with each of 22 SGs and retinal TFs (Brn3b, Isl1, Math5, Ebf1, Pax6, Tfap2a, Nr4a2, Nrl, Crx, Ptf1a, Neurod1, Lhx2, Ngn1, Ngn2, Chx10, Sox2, Rx, Meis1, Foxn4, Otx2, Sox9, and Six3). When MEFs were infected with doxycycline (Dox)inducible Ascl1 and Brn3b (AB) or Isl1 (AI) lentiviruses and cultured in the neural differentiation medium containing Dox, they started to change morphology by day 7 and form visible neuronal clusters by day 14 (Fig. 1, C and D). This phenomenon did not occur when Ascl1 acted alone or was combined with each of the rest of 20 TFs (Fig. 1B). Neither did this happen when MEFs were infected with both Brn3b and Isl1 viruses or with only control green fluorescent protein (GFP) viruses (Fig. 1, A and E). When we combined Ascl1 with both Brn3b and Isl1 (ABI), they again induced morphological changes of MEFs but more importantly induced conspicuously more neuronal clusters than either the AB or AI double-factor combinations (Fig. 1, C, D, F, and N, and fig. S1, A and B), suggesting a synergistic effect between Brn3b and Isl1 in reprogramming MEFs into neuronal clusters.

(A to I) Morphological changes of MEFs infected with the indicated lentiviruses (A, Ascl1; B, Brn3b; I, Isl1) and cultured for 14 days. Networked iSGs were induced by combinations of Ascl1 with Brn3b (AB), Isl1 (AI), or both Brn3b and Isl1 (ABI), with the ABI triple-factor combination as the most efficient. Arrows point to the thick fasciculated nerve fibers interconnecting iSG. Scale bars, 160 m (A to F) and 80 m (G to I). (J to M) Scattered iNs and clustered iSG induced by AI, ABI, A, or BAM (Brn2 + Ascl1 + Myt1l) were immunolabeled for Tuj1 and counterstained with nuclear 4,6-diamidino-2-phenylindole (DAPI). Note the morphological differences of Tuj1-immunoreactive neurons between conditions. Scale bars, 40 m. (N) Quantification of iSG induced by single and combinations of TFs. MEFs (6 104) were seeded into each well of 12-well plates and infected with lentiviruses expressing the indicated TFs or GFP, and iSGs in each well were then counted at day 14 following virus infection. Data are means SD (n = 3). Asterisks indicate significance in one-way analysis of variance test: *P < 0.0001. (O) Snapshots of a time-lapse video showing how individual neurons induced by ABI self-organized into an iSG. The arrow, arrowhead, and asterisk indicate the positions of three individual iNs at different time points. Scale bar, 62.5 m. (P) Schematic indicating the outcome (iNs or iSG) of MEFs induced by BAM, AI, AB, or ABI.

The neuronal clusters induced by either double- or triple-factor combinations (AB, AI, and ABI) appeared to be interconnected by thick fasciculated nerve fibers and resemble SG plexus in morphology (Fig. 1, G to I) and thus were designated as iSG organoids. The iSG neurons and associated nerve fibers were highly immunoreactive for the neuronal marker Tuj1 (Fig. 1, J and K, and fig. S2, D to I). Tuj1 immunolabeling also showed that AI- and ABI-induced neurons mostly formed iSG, and only a small number of them were scattered outside the iSG (Fig. 1, J, K, and P). By contrast, Tuj1 immunoreactivity showed that Ascl1 alone induced neurons mostly with an immature morphology and that the BAM (Brn2, Ascl1, and Mytl1) combination induced mature neurons that were scattered instead of clustered (Fig. 1, L, M, and P, and fig. S2, A to C), consistent with previous reports (27). Therefore, we identified the AB, AI, and ABI combinations of TFs capable of inducing MEFs into iSG, with the ABI triple-factor combination as the most efficient.

To investigate how ABI-reprogrammed neurons are organized into iSG, we used long-term time-lapse microscopy to track them over time in culture. For this purpose, MEFs were prepared from the CAG-GFP transgenic mouse embryos (28) and induced by ABI for 10 days before time-lapse recording. Compared to MEFs, reprogrammed individual neurons appeared to be rounder and neurite-bearing and displayed much higher contrast and brighter GFP fluorescence (Fig. 1O and movies S1 and S2). Over a period of tens of hours, they first formed smaller cellular clusters via migration, which then coalesced into bigger and bigger clusters that resembled SG. We did not observe this self-organization phenomenon for neurons induced by Ascl1 (movies S3 and S4).

The induction of iSG by TFs from MEFs could be through direct cell conversion or might be mediated through an intermediate proliferative progenitor. To distinguish these possibilities, we pulse-labeled cells with 5-ethynyl-2-deoxyuridine (EdU) for 24 hours at day 14 of reprogramming with AI or ABI and found that almost no Tuj1-positive cells were labeled by EdU, whereas approximately 15% of Tuj1-negative cells (e.g., MEFs) were labeled (fig. S2, G to J and N to P). We then reprogrammed MEFs with ABI in the presence of EdU for 13 days starting from day 1 of reprogramming. In this case, only 6.1% of Tuj1-positive cells were labeled by EdU, whereas 73.1% of Tuj1-negative cells were labeled (fig. S2, J to M), suggesting that iSGs are most likely induced by direct cell transdifferentiation without undergoing a proliferative intermediate state. In agreement with these results, as determined by quantitative reverse transcription polymerase chain reaction (qRT-PCR) assays, we detected no increase of expression levels of the neural progenitor marker genes Nestin and Olig2 over the entire time course (from day 1 to day 12) of ABI reprogramming (fig. S2Q). Similarly, the expression of pluripotent factor genes Oct4, Klf4, and Nanog was not induced during the time course of ABI reprogramming (fig. S2R). Furthermore, immunostaining showed that from day 1 to day 12 of ABI reprogramming, no protein expression was seen for the neural progenitor marker Nestin, pluripotent progenitor markers Nanog and Oct4, or Sox2, a marker for both neural and pluripotent progenitor cells (fig. S2, S and T). Thus, iSGs are most likely induced by direct cell transdifferentiation without undergoing an intermediate state of neural or pluripotent progenitors.

Given the demonstrated functional redundancy and similar DNA binding and transcriptional properties between Brn3a and Brn3b (10, 18, 19), we investigated whether these two factors are interchangeable in somatic cell reprogramming. We tested whether Brn3a was able to replace Brn3b in reprogramming MEFs into iSG and found that this indeed was the case (Fig. 1N and fig. S3, A to I).

By immunofluorescent staining and qRT-PCR assays, we examined a variety of molecular neuronal markers, both general and cell type specific, to characterize the iSG reprogrammed from MEFs by ABI (Ascl1 + Brn3b + Isl1 or Ascl1 + Brn3a + Isl1). We found that they were highly immunoreactive for Tuj1 and Map2 (Fig. 2, A and O), two general neuronal hallmarks. They also expressed synapsin and Vamp (synaptobrevin) (Fig. 2, B and C), suggesting that the networked iSG neurons were capable of forming synapses and releasing synaptic vesicles. In the normal SG, the heavy neurofilament NF200 and intermediate neurofilament peripherin are expressed in the A-fiber and C-fiber neurons, respectively, and both were seen to be expressed in the iSG (Fig. 2, D, E, and P). Many neurons in the iSG were also immunoreactive for the vesicular glutamate transporters 1 and 2 (vGLUT1 and vGLUT2) (Fig. 2, F and G), consistent with the fact that peripheral sensory neurons are mostly excitatory glutamatergic neurons. As determined by qRT-PCR, these immunolabeling results were confirmed by the marked up-regulation of expression of Tuj1, Map2, NF200, vGlut1, vGlut2, and vGlut3 genes in the ABI-induced iSG compared to MEFs infected by GFP lentiviruses (Fig. 2W).

(A to P) iSGs induced by Ascl1, Brn3b, and Isl1 (A to N) or Ascl1, Brn3a, and Isl1 (O and P) were double-immunostained with the indicated antibodies and counterstained with nuclear DAPI. They were immunoreactive for Tuj1, Map2, synapsin, Vamp, NF200, peripherin, vGLUT1, vGLUT2, TrkA, TrkB, TrkC, c-Ret, TH, p75NTR, and Brn3a. Scale bars, 80 m (A) and 40 m (B to P). (Q to V) Sections from iSG induced by Ascl1, Brn3b, and Isl1 were immunostained with the indicated antibodies and counterstained with nuclear DAPI. Scale bars, 12.7 m. (W) qRT-PCR analysis showing that in MEFs infected with ABI (Ascl1 + Brn3b + Isl1) viruses, compared to those infected with GFP viruses, there was a significant increase in expression of the indicated genes, which represent general and subtype-specific sensory neuron markers. Data are means SD (n = 3 or 4). Asterisks indicate significance in unpaired two-tailed Students t test: *P < 0.05, **P < 0.001, ***P < 0.0001. (X) qRT-PCR analysis showing that in MEFs infected with ABI viruses, compared to those infected with GFP viruses, there was a significant increase in expression of the indicated genes, which represent nociception pathway genes of sensory neurons. Data are means SD (n = 3 or 4). Asterisks indicate significance in unpaired two-tailed Students t test: *P < 0.05, **P < 0.005, ***P < 0.0005. (Y) Quantification of Tuj1-positive neurons that express each of the three Trk receptors (TrkA, TrkB, or TrkC) individually or combined (TrkABC) in MEFs infected with the ABI viruses. Data are means SD (n = 3).

In the DRG, neurotrophin receptor expression marks subtypes of sensory neurons. For instance, TrkA is expressed by cutaneous nociceptive and thermoceptive neurons, TrkB by a subset of cutaneous mechanoreceptive neurons, and TrkC by proprioceptive neurons (1). In the iSG reprogrammed by ABI, qRT-PCR assays revealed that there was a significant up-regulation of TrkA, TrkB, and TrkC gene expression (Fig. 2W). Moreover, immunolabeling confirmed the presence of TrkA, TrkB, and TrkC proteins in both somas and nerve fibers of the induced ganglion neurons (Fig. 2, H to J). Each of the Trk receptors was found in approximately 30% of the iNs, and 87% of the iNs were labeled by costaining for all three Trk receptors (Fig. 2Y), suggesting that each Trk receptor was expressed in a distinct subpopulation of induced ganglion neurons. c-Ret and TH are expressed in subpopulations of nonpeptidergic nociceptors and C-low threshold mechanoreceptors, respectively (1, 2). Correspondingly, we observed expression of both proteins in the iSG and associated nerve fibers (Fig. 2, K and L). In addition, pan-sensory neuron markers Brn3a (for iSG induced by Ascl1 + Brn3b + Isl1) and the nerve growth factor (NGF) receptor p75NTR were also found in iSG neurons (Fig. 2, M and N). qRT-PCR validated the up-regulation of Brn3a and p75NTR expression in the iSG and additionally revealed up-regulation of CGRP, a marker for a subpopulation of peptidergic nociceptive neurons (Fig. 2W).

The TrkA-positive nociceptive neurons in the iSG were further characterized by qRT-PCR analysis. In the iSG, we found significantly up-regulated expression of receptor ion channel genes Trpv1/2/3 and Trpa1, which detect heat and cold, respectively (Fig. 2X) (29). There was also expression of P2X3, Bdkrb1, and Accn1/2, which are receptor genes responsible for damage sensing (Fig. 2X) (29). In addition, induced expression was observed for other pain perception pathway genes including sodium channel gene Scn11a, potassium channel gene Kcnq2, calcium channel genes Cacna1a and Cacna2d1, and neurotransmitter receptor genes Gria1 and Nk1r (Fig. 2X) (29). To further investigate whether distinct types of sensory neurons were aggregated together within the same iSG, we carried out immunostaining analyses of cryosections of ABI-induced iSG. Besides colabeling between Tuj1 and Brn3a in the same iSG, we found that peripherin+ and HuC/D+ neurons, P2X3+ and vGLUT2+ neurons, and TrkA+ and TrkB+ neurons coexisted in the same iSG (Fig. 2, Q to T). Moreover, we detected coexpression of three markers, such as TrkA, P2X3 and NF200, and TrkA, peripherin, and HuC/D, in the same iSG (Fig. 2, U and V), suggesting that individual iSGs are likely aggregated from distinct sensory neuron types.

Over the time course of ABI reprogramming, qRT-PCR assays showed that the expression of general neuronal marker genes Tuj1 and Map2 was progressively induced starting from day 3, whereas other sensory neuronal marker genes including Trpv2, TrkC, and Brn3a were not induced until day 6 or 9 (fig. S1, C to E). Consistent with this, Brn3a-immunoreactive cells did not emerge until day 6 with a mostly scattered pattern, but by day 9 or 12, they mostly coalesced into iSG (fig. S1G). Therefore, as expected, sensory neuronal markers were induced slightly later than general neuronal markers during ABI reprogramming. Concomitant with neuronal induction, the fibroblast marker genes Col1a1 and Twist2 were gradually down-regulated starting from day 3 (fig. S1F).

Immunostaining of iSG induced by AI or AB suggested that they also contained neurons that expressed typical sensory neuronal markers Tuj1, Map2, Dcx (doublecortin), synapsin, NF200, peripherin, vGLUT1, TrkA, TH, HuC/D, and Brn3a (fig. S3, J to S). Together, these data indicate that certain combinations of TFs (ABI, AI, and AB) are capable of reprogramming MEFs into iSG that contain proprioceptive, mechanoreceptive, nociceptive, and thermoceptive sensory neurons.

To assess the electrophysiological properties of neurons within and outside the iSG reprogrammed from MEFs by ABI or AI, we performed whole-cell patch-clamp recordings of cells with neuronal morphology (Fig. 3A). Following 9 days of induction, the recorded neurons (two of two) generated potassium currents and small sodium currents but no action potentials, suggesting that they were functionally immature. At 2 weeks, the great majority of neurons (34 of 37) had typical sodium and potassium currents and exhibited action potential responses (Fig. 3, B to F). Among them, most (70.3%) are multispiking neurons, and the rest (21.6%) are single-spiking (Fig. 3, B, C, E, and K), similar to those reprogrammed from human fibroblasts by Brn3a and Ngn1 or Ngn2 (21). The inward sodium current could be specifically blocked by tetrodotoxin (TTX) and recovered by its removal (Fig. 3, H to J). Moreover, consistent with the synapsin immunoreactivity (Fig. 2B and fig. S3L), some neurons (2 of 37) exhibited spontaneous postsynaptic currents (Fig. 3G), suggesting the formation of functional synapses between iNs. Therefore, the iSG neurons induced by ABI or AI display membrane and physiological properties of mature neurons.

(A) Micrograph showing a typical iSG neuron chosen for patch-clamp recording. (B to D) Current-clamp recordings revealed multiple action potential responses (multiple-spiking) of a differentiated iSG neuron under current injection (B and C). Voltage-clamp recordings of the same neuron indicated fast activated and inactivated inward sodium currents as well as outward potassium currents (D). (E and F) Current injection revealed a single action potential response (single-spiking) of an iSG neuron (E). Voltage-clamp recordings of the same neuron indicated fast activated and inactivated inward sodium currents as well as outward potassium currents (F). (G) Spontaneous postsynaptic currents recorded from a differentiated iSG neuron. (H to J) The sodium currents of an iSG neuron were completely blocked by TTX and were partially restored by its washout. (K) Observed ratios of iSG neurons that are multiple-spiking and single-spiking, or display no action potential (AP). (L to N) iSG induced by ABI and corresponding fluorescent signals after incubation with Fluo-8 AM. Scale bars, 20 m. (O to Q) Calcium changes indicated by fluorescent intensity in normal Ringers solution (O), 10 M capsaicin (P), and 100 mM KCl (Q). Scale bars, 20 m. (R) Representative calcium responses to 100 M menthol and 100 mM KCl. Calcium responses were calculated as the change in fluorescence (F) over the initial baseline fluorescence (F0). (S) Representative calcium responses to 10 M capsaicin and 100 mM KCl. (T and U) Scatter dot plots showing the positive responses of individual cells to menthol, capsaicin, or KCl. Data are means SEM (n = 19 to 44).

The nociceptive sensory neurons express ion channels Trpv1, Trpm8, and Trpa1, which respond to heat, cold, and noxious chemicals, respectively (29). By calcium imaging, we used specific agonists capsaicin (10 M) and menthol (100 M) for Trpv1 and Trpm8 to confirm the functional expression of these two channels in iSG neurons (20, 21). KCl (100 mM) was transiently perfused to monitor the functional viability of the cells at the beginning and end of recording. Only cells that showed responses to KCl were chosen for analysis. Nearly all the iSG clusters induced by ABI showed green fluorescence following incubation with the calcium indicator Fluo-8 AM (Fig. 3, L to N). We found that among all the recorded cells, 56.8% of them (25 of 44) responded to capsaicin and 70.4% (19 of 27) to menthol (Fig. 3, O to U), suggesting that a large number of iSG neurons express ion channels characteristic of nociceptive sensory neurons.

We investigated the ability of iSG neurons to survive and integrate in the DRG by microinjecting dissociated iSG neurons reprogrammed from CAG-GFP mouse embryos (28), into adult rat DRG explants (fig. S4A). Following 2 weeks of culture of the transplanted explants, we found that the GFP+ iSG neurons survived, spread, and integrated in the DRG and were immunoreactive for the pan-sensory neuron marker HuC/D (fig. S4B). Moreover, a large fraction of them were immunoreactive for TrkA, while a small portion expressed TrkB or TrkC (fig. S4, C to E), indicating that iSG neurons maintain subtype specificity in the DRG.

Consistent with their sensory neuron identity, after a week in culture, iSG neurons reprogrammed from Tau-GFP mouse embryos (30) spontaneously aggregated with rhodamine-labeled sensory neurons dissociated from E13.5 mouse DRGs to form DRG-like organoids interconnected by nerve fibers (fig. S4, N to Q). In contrast, when GFP+ iSG neurons were cocultured with P0 mouse skin cells, they did not co-aggregate with skin cells; instead, they projected to and innervated vimentin-immunoreactive epidermal cells with multiple terminal nerve endings (fig. S4, J to M), in agreement with the fact that DRG neurons normally innervate their peripheral targets in the epidermis.

Previous studies have demonstrated that peripheral SG neurons and RGCs share many common molecular hallmarks, making it difficult to distinguish these two types of sensory neurons in cell culture. During the past decade in stem cell research, a number of supposedly specific molecular markers have been used to identify differentiated or induced SG neurons and RGCs (2225); unfortunately, however, no efforts have been made to confirm the specificity of these markers, casting doubt on some of the previous conclusions. Because Brn3a, Brn3b, and Isl1 are TFs crucial for retinal cell development, in particular, RGC development (13, 14, 16), there is a possibility that they may also be able to reprogram MEFs into RGCs. We thus set out to identify molecular markers that can definitively distinguish RGCs from peripheral sensory neurons. We postulated that such unique identifiers could be single-molecule markers or a combination of multiple-molecule markers that must be present only in RGCs within the retina but not in peripheral sensory neurons or any other tissues.

In the mammalian retina, our early studies have identified Brn3a and Brn3b as the gold standard markers for RGCs, but meanwhile revealed their expression in peripheral SG and other CNS areas (10, 15). In the mouse, immunolabeling of retinal and DRG sections confirmed the specificity of Brn3a and Brn3b in RGCs within the retina as well as their widespread expression in DRG neurons (fig. S5A), indicating that Brn3a and Brn3b alone cannot distinguish RGCs from DRG neurons outside the retina. Similarly, many other commonly used RGC markers including Thy1.2, RPF-1, Rbpms, HuC/D, Six6, Ebf, Isl1, Zeb2, Lmo4, Ldb1, and Sncg all displayed expression in the DRG (fig. S5A). Expressed in both RGCs and DRGs were also a number of sensory neuron markers including CGRP, peripherin, vGLUT2, vGLUT3, GABA, TrkA, TrkB, TrkC, and P2X3 (fig. S5A). Pax6 appeared to be the only exception among all the tested markers, which is expressed in RGCs and inner nuclear layer within the retina but absent from DRG (fig. S5A). Given the expression of Brn3a, Brn3b, Thy1.2, RPF-1, and Rbpms only in RGCs within the retina, a combination of Pax6 with any of these proteins could serve as a potential unique identifier for RGCs.

The uniqueness of double-positive markers was tested by immunolabeling sections of other CNS areas. Double-immunostaining showed that neuronal cells immunoreactive for both Pax6 and RPF-1, Thy1.2, Rbpms, HuC/D, or Tuj1, albeit absent from the DRG, were present not only in the retina but also in the spinal cord (Fig. 4A), precluding their use as specific RGC markers. The Isl1+Pax6+ double-positive cells were absent from the DRG and spinal cord but present within both the ganglion cell layer and inner nuclear layer in the retina (Fig. 4A), precluding also this combination as a specific RGC identifier. By contrast, Brn3a+Pax6+ and Brn3b+Pax6+ double-positive cells were exclusively RGCs in the retina and were not found in the DRG or spinal cord (Fig. 4A). Given the detection of Brn3a/Brn3b expression in the midbrain and cerebellum (10, 15), we investigated whether there were Brn3a+Pax6+ and Brn3b+Pax6+ double-positive cells in these two brain regions and found none at stages E13.5, P4, and P21 (Fig. 4A). Thus, these results together demonstrate that a combination of Pax6 and Brn3a or Brn3b double markers can serve as specific identifiers for RGCs.

(A) Cryosections from the indicated regions and stages of mice were stained by double immunofluorescence with the indicated antibodies and counterstained with nuclear DAPI. Arrows point to representative double-positive cells. GCL, ganglion cell layer; INBL, inner neuroblastic layer; INL, inner nuclear layer; IPL, inner plexiform layer; ONBL, outer neuroblastic layer; ONL, outer nuclear layer; OPL, outer plexiform layer. Scale bars, 40 m. (B to E) MEFs were infected with the ABI lentiviruses, cultured for 14 days, and double-immunostained with the indicated antibodies and counterstained with nuclear DAPI. Arrowheads in (D) indicate colocalized cells in the outlined region located outside the iSG. Scale bars, 40 m (B and C) and 20 m (D and E). (F to I) MEFs infected with the ABI lentiviruses and cultured for 14 days were dissociated and double-immunostained with anti-Brn3a and anti-Pax6 antibodies and counterstained with nuclear DAPI. Arrows indicate colocalized cells. Scale bars, 20 m. (J) qRT-PCR analysis of expression levels of the indicated genes (ex, exogenous; en, endogenous) in MEFs infected with ABI or GFP viruses (means SD, n = 3 or 4). *P < 0.05, **P < 0.001, ***P < 0.0001. (K and L) Quantification of DAPI- or Tuj1-positive cells that express Brn3a or Pax6 in MEFs infected with the ABI viruses (means SD, n = 4). (M) Quantification of Brn3a+Pax6+ iRGCs induced by ABI (means SD, n = 4).

We used the single-cell RNA sequencing (scRNA-seq) technology to further confirm the specificity of Brn3a+Pax6+ and Brn3b+Pax6+ double-positive markers. A Brn3b-GFP knockin mouse line was generated, and RGCs were enriched and sequenced by scRNA-seq. In addition, we isolated adult mouse DRG cells, which were then similarly sequenced. Clustering and expression analyses of the sequenced RGCs revealed that most of them expressed Pax6, Brn3a, or Brn3b and both Pax6 and Brn3a or Brn3b; in particular, the great majority of RGCs were positive for both Pax6 and Brn3a (fig. S5B). By contrast, there was a complete absence of DRG cells expressing both Pax6 and Brn3a or Brn3b, although Brn3a and Brn3b were present in most DRG cells (fig. S5B), consistent with the idea that a combination of Pax6 and Brn3a or Brn3b double markers can be used to distinguish RGCs from DRG cells.

To assess whether ABI and AI are able to induce iRGCs in addition to iSG, we immunostained ABI-reprogrammed MEF cells with antibodies against Tuj1, Brn3a, and/or Pax6. Double-labeling between Tuj1 and Brn3a or Pax6 showed that Brn3a-expressing cells were concentrated in the iSG, whereas the great majority of Pax6-expressing cells were distributed outside of the iSG and only few of them were seen in the iSG (Fig. 4, B and C). Moreover, all Pax6-positive cells coexpressed Brn3a and most of them displayed relatively weak Brn3a expression (Fig. 4, D and F to I), indicating that iRGCs were reprogrammed from MEFs by ABI. Similar to the distribution of endogenous RGCs that are spread throughout the RGC layer, the Brn3a+Pax6+ iRGCs were scattered and did not organize into clustered mini-ganglia (Fig. 4, C and D), unlike the induced peripheral SG neurons. Quantification of immunoreactive cells indicated that approximately 21.1% of all cells were induced by ABI into Brn3a+ neurons, whereas only about 2.6% of them were reprogrammed into Pax6+ cells (Fig. 4K). Furthermore, there were 93.1% of Tuj1+ cells coexpressing Brn3a, 10.5% of Tuj1+ cells coexpressing Pax6, and 12.6% Brn3a+ cells coexpressing Pax6 (Fig. 4, L and M), suggesting that only a small fraction of the ABI-reprogrammed neurons are Brn3a+Pax6+ double-positive iRGCs and that most of them are Brn3a+ iSG neurons. Similarly, a small number of Brn3a+Pax6+ iRGCs were induced by AI (fig. S3, T and U).

In agreement with the induction of a small proportion of iRGCs by ABI, immunostaining showed that some cells outside the iSG were positive for Thy1.2 (Fig. 4E). qRT-PCR assays revealed a significant up-regulation of several commonly used RGC marker genes including the endogenous Brn3b, Brn3a, RPF-1, Pax6, Sncg, HuC, and HuD in MEFs infected with ABI lentiviruses compared to those infected with GFP viruses (Fig. 4J), consistent with the induction of iRGCs by ABI from MEFs. Moreover, during the time course of ABI reprogramming, we were able to show by qRT-PCR assay that Pax6 expression was progressively induced starting from day 9 (fig. S1E).

We further characterized the iSG neurons and iRGCs by bulk and single-cell transcriptome profiling. First, we carried out bulk RNA-seq analysis of ABI- and GFP-transduced MEFs after 2 weeks of induction (Fig. 5A). Scatter plot and hierarchical cluster analyses showed that there were numerous genes whose expression was down-regulated or up-regulated in ABI-transduced compared to GFP-transduced MEFs (fig. S6, A to C, and table S1). We performed gene set enrichment analysis (GSEA) of the altered genes followed by network visualization (31), and one major group of clustered networks emerged (fig. S6D). This group encompasses only up-regulated genes that are enriched for GO (gene ontology) terms relevant to neural function and development such as synaptic signaling, synaptic vesicle, synapse organization, neurotransmitter transport, regulation of neurotransmitter levels, exocytosis, calcium ion binding, ligand-gated channel activity, neuron projection, axon, and nervous system development. These results are consistent with the induction of functional SG and retinal ganglion neurons by ABI from MEFs. In agreement with this and qRT-PCR assays (Figs. 2, W and X, and 4J), bulk RNA-seq confirmed up-regulation of many SG and retinal ganglion genes in ABI-transduced MEFs, including NF200, Brn3a, TrkB, vGlut3, Trpv1, P2X3, Gria1, Pax6, Sox11, Sncg, and Thy1 (fig. S6, E and F).

(A) Schematic illustration of the processes for bulk RNA-seq and scRNA-seq analyses. (B) t-distributed Stochastic Neighbor Embedding (t-SNE) plot of the 15 cell clusters generated from the sequenced single iSG neurons. (C to J) t-SNE plots colored by expression of the indicated conventional SG marker genes. (K) Violin plots showing expression patterns of the indicated conventional SG marker genes in single-cell clusters.

We separated ABI-induced iSG from MEFs by mild dissociation and filtering and then carried out scRNA-seq analysis of single iSG cells using the 10 Genomics Chromium platform (Fig. 5A) (32). After processing the sequencing data by the Cell Ranger software pipeline, we clustered the 3231 sequenced single cells into 15 clusters using the Seurat software package (Fig. 5B), which is an R toolkit for single-cell genomics (33). Investigation of gene expression patterns showed high levels of expression of general neuronal marker genes such as Tuj1, Tau, and Map2 in clusters 1, 3, 5, 8, and 11, whereas they are expressed much more weakly in the rest of the clusters (fig. S7, A and C to E). By contrast, many of the previously identified MEF marker genes (34) including Klf4, Mmp2, and Postn have, in general, an opposite expression pattern, displaying little expression in clusters 1, 3, 5, 8, and 11 but obvious expression in the rest of the clusters (fig. S7, A and F to H). Pseudotime trajectory of the sequenced cells constructed using Monocle (35) yielded three presumptive states along which Klf4 expression progressively decreases, while the expression of Tuj1, Tau, and Map2 progressively increases (fig. S7, I and J). Thus, in iSG induced from MEFs by ABI for 2 weeks, there are still some cells that express both neuronal and MEF markers, suggesting that MEFs undergo a transitional intermediate stage that exhibits both MEF and neuronal characteristics before completely reprogrammed into mature iSG neurons (fig. S7B). Consistent with this idea, there were many cells coexpressing both Tuj1 and the fibroblast marker gene vimentin in a number of the clusters (fig. S7K). By days 6 to 12 of ABI reprogramming, we also detected by immunolabeling some cells and nerve bundles that were immunoreactive for both Tuj1 and vimentin proteins (fig. S7L).

Consistent with the induction of iSG neurons, there is expression of NF200, peripherin, p75NTR, TrkB, TrkC, Trpv1, Trpv2, P2X3, Accn2, Kcnq2, Cacna1a, and CGRP in various clusters of sequenced iSG cells (Fig. 5, C to K). In particular, NF200, P2X3, Accn2, Kcnq2, and Cacna1a are primarily expressed in clusters 1, 3, 5, 8, and 11, and peripherin, Trpv1, and Trpv2 are mainly present in a small number of cells in clusters 1, 3, and 5 (Fig. 5, C, D, F to I, and K), indicating their expression in mature iSG neurons and their expression specificity. Many genes that are markers for both SG neurons and RGCs, such as Thy1, Sncg, Rbpms, Gap43, HuC, Sox11, Sox12, Zeb2, Brn3a, Brn3c, and RPF-1, are also expressed in various clusters of sequenced iSG cells (Fig. 6). However, the RGC marker Pax6 is only enriched in small cell clusters 12 and 13 and expressed in few cells in other clusters, consistent with the observation that only a very small number of iSG cells were immunoreactive for Pax6 (Figs. 4C and 6I). The Pax6+ cells in clusters 12 and 13 do not appear to be iRGCs because they lack expression of RGC markers Brn3a, Brn3c, and RPF-1 (Fig. 6I). In agreement with the observation that iRGCs were scattered and rarely present in iSG, there are only a small number of cells coexpressing both Pax6 and Tuj1, Thy1, Gap43, HuC, Sox11, Brn3a, or RPF-1, primarily in clusters 5 and 6 (Fig. 4C and fig. S8, A to H).

(A to H) t-SNE plots colored by expression of the indicated conventional RGC marker genes. (I) Violin plots showing expression patterns of the indicated conventional RGC marker genes in single-cell clusters.

We reprogrammed human skin fibroblasts (HSFs) into iSG with a mixture of the three individual ABI lentiviruses only at a low efficiency. To increase the reprogramming efficiency, we created a Dox-inducible lentiviral construct containing Ascl1, Isl1, and Brn3b in a single open reading frame (ORF) tethered by the P2A and T2A self-cleaving peptide sequences (Fig. 7A). HSFs infected by these single ABI-expressing viruses readily formed well-networked iSG in approximately 45 days in the neural differentiation medium (Fig. 7B). Immunostaining of these iSG showed that they contained typical sensory neurons expressing TUJ1, MAP2, NF200, PERIPHERIN, SYNAPSIN, VGLUT1, TRKA, TRKB, TH, and BRN3A (Fig. 7, C to K). Moreover, similar to MEFs, a small number of iRGCs were induced from HSFs by ABI that were immunoreactive for both PAX6 and BRN3A (Fig. 7, L and M). qRT-PCR assays showed that TUJ1 expression was gradually induced by ABI starting from day 10 but the more mature neuron marker gene MAP2 was not induced until day 20 (Fig. 7N). In contrast, the fibroblast marker genes COL1A1 and TWIST2 were progressively down-regulated starting from day 10 (Fig. 7O), concurrent with TUJ1 induction.

(A) Schematic of the lentiviral construct. (B to M) Networked iSG induced by ABI from HSFs (B) and iSG and iRGCs double-immunostained with the indicated antibodies and counterstained with DAPI (C to M). (N to Q) qRT-PCR assay showing the time course [days 1 (D1) to 20 (D20)] of expression changes of the indicated marker genes in HSFs infected with ABI or GFP viruses (means SD, n = 4). *P < 0.0001 for (N), (P), and (Q) and *P < 0.01, **P < 0.001, ***P < 0.0001 for (O). hES, human embryonic stem cell; hiNSC, human neural stem cell. (R) Schematic of EdU labeling schedule. (S to U) ABI-transduced HSFs were labeled by EdU for 29 days and colabeled for both TUJ1 and EdU before (S) and after dissociation (T). (U) Corresponding quantification (means SD, n = 4). *P < 0.0001. (V to X) ABI-transduced HSFs were labeled by EdU for 24 hours and colabeled for both TUJ1 and EdU before (V) and after dissociation (W). (X) Corresponding quantification (means SD, n = 4). *P < 0.0001. (Y, Z, and A) Current-clamp recordings revealed single action potential responses (single-spiking) of a differentiated iSG neuron (Y). Voltage-clamp recordings of the same neuron indicated fast activated and inactivated inward sodium currents as well as outward potassium currents (Z and A). The sodium currents of the iSG neuron were effectively blocked by TTX and were partially restored by its washout (A). (B and C) Current-clamp recordings revealed an iSG neuron with multiple action potential responses (multiple-spiking). (D) Spontaneous postsynaptic currents recorded from a differentiated iSG neuron. Scale bars, 80 m (B) and 20 m (C to M, S, T, V, and W).

To determine whether iSG induction was mediated by a pluripotent or neural progenitor intermediate, we investigated by qRT-PCR assay expression of pluripotent factor genes and neural progenitor marker genes during HSF reprogramming by ABI. We found no significant change in expression levels of pluripotent factor genes OCT4, KLF4, and NANOG during the reprogramming process (from day 1 to day 20) (Fig. 7P). Similarly, there was no induction of NESTIN and OLIG2 expression in the reprogramming process (Fig. 7Q), suggesting that iSGs were reprogrammed from HSFs by ABI without an intermediate state of pluripotent or neural progenitors. Consistent with this, by day 30 of reprogramming, almost no reprogrammed TUJ1+ neurons were labeled by EdU when EdU was added to the reprogramming cell culture for 29 days or 24 hours (Fig. 7, R to X), confirming that iSG reprogramming occurred in the absence of an intermediate state of proliferative progenitors.

The electrophysiological properties of reprogrammed human iSG neurons were evaluated by whole-cell patch-clamp recording. At day 60, most neurons (15 of 17) exhibited typical sodium and potassium currents and showed action potential responses (Fig. 7, Y, Z, and A). In addition, the inward sodium current could be specifically and completely blocked by TTX and partially recovered by its removal (Fig. 7A). Similar to mouse iSG neurons, some (4 of 17) were multi-spiking, while the others (11 of 17) were single-spiking (Fig. 7, Y, B, and C), although in human iSG single-spiking neurons appeared to be more abundant than those in mouse iSG (Fig. 3K). Among all neurons recorded from day 25 to day 39, a small fraction (4 of 44) displayed spontaneous postsynaptic activities (Fig. 7D), indicating the ability for human iSG neurons to form functional synapses, in agreement with their synapsin labeling (Fig. 7F). Thus, the human iSG neurons induced by ABI from HSFs have the physiological properties characteristic of mature neurons.

We further investigated the ability of human iSG neurons to survive and integrate in the DRG by microinjecting GFP-tagged human iSG neurons into adult rat DRG explants (fig. S4A). Two weeks after transplantation, we found that the GFP+ neurons survived and integrated in the DRG, and were all (99 of 99) immunolabeled by an anti-human nuclei antibody (fig. S4F), indicating that material transfer did not occur between the transplanted and host cells. The transplanted GFP+ cells were immunoreactive for pan-sensory neuron markers, and some of them were immunoreactive for TrkA, TrkB, or TrkC (fig. S4, G to I), suggesting that similar to mouse iSG neurons, transplanted human iSG neurons can also survive in the DRG and maintain sensory neuron subtypes.

Although scattered sensory neurons (iSNs) were previously induced from fibroblasts by TFs (20, 21), to our knowledge, this is the first time to demonstrate that self-organized iSG organoids can be consistently induced directly from somatic cells by defined TFs. The bHLH TF Ascl1 has been shown to be a pioneer neurogenic TF in converting fibroblasts into neurons in in vitro somatic cell reprogramming (26). However, the neurons reprogrammed by Ascl1 alone are mostly slow-maturing and excitatory (36). Addition of Brn2 and Myt1l (BAM) improved the reprogramming efficiency, maturing speed, and varieties of the iNs (27, 36, 37). The iNs induced by BAM were rather generic but motor neurons could be specifically induced when BAM were combined with four other TFs (Lhx3, Hb9, Isl1, and Ngn2) (38). Similarly, when trying BAM with other combinations, Wainger et al. (20) found that the combination of five factors (Ascl1, Myt1l, Ngn1, Isl2, and Klf7) could successfully convert fibroblasts into nociceptor neurons. Notably, all of these reprogramming formulas include Ascl1 as a key component. Alternatively, the bHLH TFs Ngn1 and Ngn2 were combined with Brn3a to reprogram fibroblasts into mature iSNs (21).

Our experiments in this study have demonstrated that the ABI TF combination is most effective in inducing MEFs into self-organized mini-SG, while the AI and AB combinations have a weaker activity (fig. S8J). Thus, Brn3a/3b appears to act synergistically with Isl1 to improve the induction efficiency of iSG organoids. As revealed by time-lapse microscopy, the larger iSG organoids are formed by cell migration and coalescing smaller cell aggregates. The mini-SG induced from both murine and human fibroblasts contain mature and functional sensory neurons. They exhibit typical inward sodium currents, which can be blocked by TTX and recover after TTX removal, and are a mixture of neurons displaying multiple-spiking action potentials or single-spiking action potential. They also show calcium responses to potassium chloride, capsaicin, and menthol. All these features closely resemble their endogenous counterparts.

The iSG neurons reprogrammed by ABI display extensive cell diversities in their expression of characteristic receptors, ligands, ion channels, neuropeptides, neurotransmitters, and so on, similar to the endogenous sensory neurons. In agreement with iSNs induced by Ngn1/2 and Brn3a (21), the iSGs contain roughly equivalent percentages (~30%) of TrkA+, TrkB+, and TrkC+ neurons, supporting the notion that Trk receptors may arise in a stochastic manner such that each donor cell has an approximately equivalent chance to express one of the Trk receptor genes. By bulk RNA-seq, scRNA-seq, and/or qRT-PCR analyses, we investigated the characteristic markers involved in sensory signaling pathways including transduction, conduction, and synaptic transmission of sensory signals. At the transduction level, we found up-regulated expression of genes responsible for perceptions to stimuli such as heat (Trpv1, Trpv2, Trpv3), cold (Trpa1), damage (P2X3, Bdkrb1), and touch (Trpc1, Trpc4, Asic2/Accn1, Accn2). Trpv1, also known as capsaicin receptor that is expressed mainly in the nociceptive neurons (29), has been shown to be present and functional in iSG neurons by capsaicin stimulation. The signaling conduction of sensory neurons is primarily mediated by sodium channels, which propagate the signals, and potassium channels, which usually act to reduce excitability. We found that the expression of many Na+ channels (Scn1a, 2a1, 2b, 3a, 3b, 7a, 11a) and K+ channels (Kcnq2, 4; Kcna2, 3, 4, 5, 6; Kcnb2, c1, d2, e4, f1, h2, j2, k3, s3, t1, etc.) were up-regulated in iSG neurons. For synaptic transmission, neurotransmitter receptors and presynaptic voltage-gated Ca2+ channels are two groups of important regulatory molecules. Correspondingly, the expression of a variety of neurotransmitter receptors (Nk1r, Nr3c2; Gria1, 2, 4; Grid1, k1, k2, k4, k5; Grin1, 2a, etc.) and Ca2+ channels (Cacna1a, 1b, 1d, 2d1, 2d2, 2d3; Cacnb1, g4, etc.) were significantly up-regulated in iSG neurons.

Apart from the molecular and electrophysiological properties, ABI-reprogrammed iSG neurons also have salient cellular and innervation characteristics of sensory neurons. For instance, when transplanted, they can survive, integrate, and maintain the nociceptive, mechanoreceptive, and proprioceptive subtypes in the DRG. Moreover, the iSG neurons exhibit strong affinity for endogenous DRG neurons and spontaneously aggregate with them to form interconnected DRG-like organoids in culture. In addition, we have demonstrated by coculture that the iSG neurons have the capacity to innervate the peripheral targets of sensory neurons, i.e., epidermal cells, indicating that the iSGs contain bona fide sensory neurons reprogrammed from fibroblasts by ABI.

Therefore, the combination of ABI TFs is able to reprogram murine and human fibroblasts into self-organized iSG organoids composed of heterogeneous sensory neurons, closely resembling the endogenous SG. Previously, Ascl1 in combination with Brn3a, Brn3b, or Brn3c was shown to induce iNs from MEFs (39). Although the sensory neuron identity of the iNs was not investigated, some of the data suggest the formation of iSG organoids by the Ascl1 and Brn3a combination (39). This is consistent with our work that showed that the AB combination enabled induction of iSG organoids, albeit fewer than those induced by the ABI combination (Fig. 1). Similarly, the data reported in a previous study also suggest the formation of iSG organoids by the nociceptive neurons reprogrammed from MEFs using a 5-TF combination (20). However, unlike the ABI combination, the 5-TF combination did not appear to induce iSG organoids from human fibroblasts (20), suggesting a difference in reprogramming capacity and/or efficiency by different combinations of TFs.

The peripheral ganglia, including cranial ganglia, DRG, trigeminal ganglia, enteric system ganglia, autonomic ganglia, and others, are derived from migrating NC cells. The NC is thought to be a unique cell population found in vertebrates and is initially induced at the neural plate border as a result of neural plate folding and fusion (40). After undergoing an epithelial-to-mesenchymal transition, the NC cells delaminate from the neuroepithelium and become highly migratory. Most NC cells migrate as a chain or group in a so-called collective cell migration, in which cell contact and cooperation allow them to migrate directionally. Guided by local cues and long-range chemoattractants, NC cells reach their destination and differentiate into ganglia and other tissue types.

Mutations in crucial genes controlling the migration and differentiation of NC cells may cause aganglionosis such as Hirschsprungs disease, which may occur by itself or in association with other genetic disorders such as Down syndrome, Waardenburg-Shah syndrome, Mowat-Wilson syndrome, or Bardet-Biedl syndrome (41). This group of genes includes RET, ZEB2, EDNRB, SOX10, and PHOX2B, and mutations of them or their regulatory sequences may increase the risk of Hirschsprungs disease more than 1000-fold (41). In our RNA-seq data, the expression of Ret, Zeb2, Ednrb, and Sox10 was significantly up-regulated in the iSG neurons, in agreement with their importance in the differentiation and formation of SG. Other known risk genesBbs4, Bbs10, Edn3, Gfra1, and Arvcf (41)were also significantly elevated in iSG. The hereditary sensory and autonomic neuropathies (HSANs) consist of several clinically heterogeneous disorders characterized by defective development and maintenance, and progressive degeneration of sensory and autonomic nervous systems. Mutations in the SPTLC1, WNK1, IKBKAP, and TRKA genes have been shown to cause HSAN types I to IV, respectively (42). In addition, loss-of-function mutations in SCN9A and PRDM12 result in congenital insensitivity to pain (6, 43). Indifference to pain appears to be desirable but risks the loss of a vital protective mechanism with dangerous consequences such as unknowingly chewing tongues and lips and damaging digits and joints. On the other hand, pain hypersensitivity reduces the quality of life and may increase susceptibility to chronic pain.

The ability to reprogram somatic cells into iSG organoids by ABI presents new possibilities for modeling sensorineural diseases, studying their pathogenesis, screening for counteractive drugs, and developing cell replacement therapies. For example, patient-derived iSG organoids may be used as an in vitro model for pain to screen and evaluate potential drug treatments. In the future, iSG organoids and neurons may also be used in transplantation as a cell replacement therapy for damaged or degenerated SG. In this respect, we found that transplanted iSG neurons were able to integrate and maintain the nociceptive, mechanoreceptive, and proprioceptive subtypes in the DRG. It has long been recognized that genetic factors are a major contributor to personalized pain perception and the efficacy of analgesic drugs (29). Generation of iSG organoids from autologous somatic cells may thus provide an exciting novel approach to model personalized pain and sensory pathology and help to achieve precision medicine for pain.

In this study, we made efforts to define specific molecular markers to identify RGCs both in vitro and in vivo. This is important because it is impossible to apply commonly used RGC markers to distinguish RGCs from SG neurons in vitro given the high molecular similarity between these two cell types. Since the 1990s, we have established the Brn3 family of TFs, Brn3a, Brn3b, and Brn3c, as the gold standards to identify RGCs in the retina (10, 15). However, Brn3 proteins are not unique to the retina but expressed in other sensory and CNS tissues as well, e.g., trigeminal ganglia, DRG, spiral ganglia, and midbrain (10, 15, 44). Apart from Brn3 proteins, Thy1.2, Sncg, and Rbpms are also commonly used as specific RGC markers. But here again, we show their abundant expression in DRG neurons. Therefore, although because of the spatial separation of the retina from SG in the organism, these so called RGC-specific markers are able to distinguish RGCs from SG neurons in vivo, they are unable to do so in vitro. Unfortunately, however, a number of previous studies used these supposedly RGC-specific markers to identify RGCs induced from ESCs, iPSCs, and somatic cells in vitro (2225), casting doubt on some of the arrived conclusions.

To avoid misidentifying iRGCs and iSG neurons in vitro, we screened for molecular markers that can definitively distinguish RGCs from SG neurons. A rigorous criterion was set that these unique identifiers should be single-molecule markers or a combination of multiple-molecule markers that must be present only in RGCs within the retina but not in SGs or any other tissues. Following a careful examination of a large number of known RGC and SG neuron markers, it became apparent that none of them alone were specific to RGCs. Further double-immunolabeling analysis indicated that a combination of Pax6 and Brn3a or Brn3b double markers satisfied the criterion of specific RGC identifiers. Brn3a+Pax6+ and Brn3b+Pax6+ double-positive cells were found exclusively in RGCs of the retina but not in the DRG, spinal cord, midbrain, or cerebellum, where Brn3a, Brn3b, or Pax6 is normally expressed. Moreover, scRNA-seq analysis confirmed Brn3a+Pax6+ and Brn3b+Pax6+ cells as RGCs and their complete absence in the DRG. Thus, we are able to define the combination of Pax6 with either Brn3a or Brn3b double protein markers as specific identifiers for RGCs. Armed with this knowledge, we found that ABI TFs had the capacity to reprogram MEFs into a small number of Brn3a+Pax6+ iRGCs, representing about 13% of all Brn3a+ neurons. Unlike iSG organoids resembling endogenous SG, iRGCs did not coalesce into clusters but remained scattered, similar to the dispersive distribution pattern of endogenous RGCs in the retina (fig. S8, I and J). Therefore, ABI-induced iSG and iRGCs maintain the morphology characteristic of their endogenous equivalents.

In summary, in a screen of multiple SG and RGC TFs, we have identified a triple-factor combination ABI as the most efficient combination to reprogram self-organized and networked iSG organoids from mouse and human fibroblasts. By immunostaining, qRT-PCR, whole-cell patch-clamp recording, calcium imaging, and bulk and scRNA-seq approaches, we are able to demonstrate that the iSG organoids display molecular and cellular features, subtype diversity, electrophysiological properties, and peripheral innervation patterns characteristic of peripheral SGs. Furthermore, using immunolabeling and scRNA-seq analyses, we have identified bona fide RGC-specific molecular markers to demonstrate that the ABI combination has the additional capacity to induce from fibroblasts a small number of iRGCs. Unlike the ABI-reprogrammed iSG organoids characteristic of endogenous SG, iRGCs maintain a dispersive distribution pattern resembling that of endogenous RGCs in the retina. The iSG organoids and iRGCs may be used to model sensorineural/retinal diseases, to screen for effective drugs and potentially, as cell-based replacement therapy.

All experiments on rodents were performed according to the IACUC (Institutional Animal Care and Use Committee) standards and approved by Sun Yat-sen University and Zhongshan Ophthalmic Center. The C57BL/6 mice were purchased from the Vital River Laboratories (Beijing, China).

The full-length ORFs of Brn3a, Brn3b, Isl1, Math5, Ebf1, Pax6, Tfap2a, Nr4a2, Nrl, Crx, Ptf1a, Neurod1, Lhx2, Ngn1, Ngn2, Chx10, Sox2, Rx, Meis1, Foxn4, Otx2, Sox9, or Six3 were subcloned into the Eco RI site of the FUW-TetO vector (45). In addition, by overlapping PCR subcloning, Ascl1, Isl1, and Brn3b were tethered by P2A and T2A self-cleaving peptide sequences into a single ORF, which was inserted into the same FUW-TetO backbone. Lentiviruses were prepared as previously described (34).

The MEFs were prepared as previously described (34). For isolation of mouse epidermal cells, P0 C57BL/6 mice were anesthetized with ice for 5 min and the brain was removed using a sterilizing razor in a 10-cm culture dish containing Hanks balanced salt solution (HBSS) (Gibco). The epidermis was isolated from the remaining tissue using a pair of fine-tip forceps under a dissection microscope, transferred into a fresh 6-cm culture plate containing 1 ml of 0.25% trypsin, thoroughly minced using a pair of surgical scissors and forceps, and incubated for 15 min at 37C in a CO2 incubator. Six-milliliter MEF medium containing Dulbeccos modified Eagles medium (DMEM)/High Glucose (HyClone) supplemented with 10% fetal bovine serum (Gibco), 1 penicillin/streptomycin (Gibco), 1 MEM nonessential amino acids (NEAA) (Gibco), and 0.008% (v/v) 2-mercaptoethanol (Sigma-Aldrich) was added into the plate to terminate the reaction. After being mixed using a 10-ml pipette, the digested tissue was transferred to a 15-ml fresh tube, centrifuged at 1000 rpm for 5 min, and resuspended in 5-ml fresh MEF medium. The isolated epidermal cells were expanded by culture in the MEF medium at 37C in a CO2 incubator. The HSFs were purchased from the American Type Culture Collection (CRL1502, 12-week gestation). MEFs, mouse epidermal cells, and HSFs were all maintained and expanded in the MEF medium.

To induce iSG and iRGCs from MEFs, 3 104 MEF cells (at passage 3) were cultured in 500-l MEF medium in a well of a 24-well plate containing a glass coverslip precoated with Matrigel (Corning). They were infected the next day with 500-l mixture of lentiviruses and fresh MEF medium in the presence of polybrene (10 g/ml). After 16-hour infection, the virus and medium mixture was removed. The cells were induced for 4 days in the neuron basic medium [(DMEM/F12 (1:1) (Life Technologies) supplemented with 1 B27 (Gibco) and basic fibroblast growth factor (bFGF) (10 ng/ml) (R&D Systems)] in the presence of Dox (2 ng/ml) (Sigma-Aldrich) and then for another 4 days in the neuron maintenance medium containing the neuron basic medium supplemented with insulin-like growth factor 1 (IGF-1) (100 ng/ml), brain-derived neurotrophic factor (BDNF) (10 ng/ml), and glial cell linederived neurotrophic factor (GDNF) (10 ng/ml) in the presence of Dox (2 g/ml). The medium was replaced with the neuron maintenance medium without Dox following the 8-day induction period. By 14 days after infection with Ascl1, Brn3b/3a, and Isl1 (ABI), Ascl1 and Brn3b/3a (AB), or Ascl1 and Isl1 (AI) lentiviruses, many visible neuronal clusters were formed.

With modifications, the HSFs were similarly induced. In brief, after virus infection, the human cells were cultured in the neuron basic medium with Dox for 10 days and then in the neuron maintenance medium without Dox for another 10 days. On day 21, the medium was replaced with the neuron mature medium, which is the maintenance medium supplemented with NGF (20 ng/ml), NT-3 (20 ng/ml), and 10 M forskolin. Thirty days after viral infection, many neuronal clusters were visible, which were usually smaller than those induced from MEFs. To improve the induction efficiency of the HSFs, we created a Dox-inducible lentiviral construct containing Ascl1, Isl1, and Brn3b in a single ORF as described above.

RNA extraction and qRT-PCR analysis were carried out as previously described (34). The qRT-PCR primers used are shown in table S2.

Immunostaining of tissue sections and cells was carried out as previously described (34, 46). The following antibodies (with dilution information) were used: mouse anti-Brn3a (Santa Cruz Biotechnology, sc-390780; 1:1000), mouse anti-Brn3a (Santa Cruz Biotechnology, sc-8429; 1:100), goat anti-Brn3b (Santa Cruz Biotechnology, sc-6026; 1:1000), rat anti-Thy1.2 (BD Biosciences, 550543), goat antiRPF-1 (Santa Cruz Biotechnology, sc-104627; 1:100), rabbit anti-Rbpms (PhosphoSolutions, 1830-RBPMS; 1:500), mouse anti-HuC&D (Life Technologies, A-21271; 1:500), rabbit anti-Pax6 (BioLegend, 901301; 1:2000), mouse anti-Pax6 (Developmental Studies Hybridoma Bank, Pax6; 1:1000), rabbit anti-Six6 (Sigma-Aldrich, HPA001403; 1:500), rabbit anti-Ebf (Santa Cruz Biotechnology, sc-33552; 1:1000), mouse anti-Isl1 (Abcam, ab20670; 1:2000), rabbit anti-Zeb2 (Santa Cruz Biotechnology, sc-48789; 1:1000), rat anti-Lmo4 (1:1000; (47), rabbit anti-Ldb1 (Abcam, ab96799; 1:1000), rabbit anti-Sncg (GeneTex, GTX110483; 1:200), rabbit anti-CGRP (Neuromics, RA24112; 1:200), rabbit anti-peripherin (Millipore, ab1530; 1:1000), rabbit anti-vGLUT1 (Synaptic System,135303; 1:500), mouse anti-vGLUT2 (Abcam, ab79157; 1:500), mouse anti-vGLUT3 (Sigma-Aldrich, SAB5200312; 1:500), rabbit anti-GABA (Sigma-Aldrich, A-2052; 1:1000), goat anti-TrkA (Abcam, ab76291; 1:500), rabbit anti-TrkA (Abcam, ab76291; 1:500), goat anti-TrkB (R&D Systems, AF1494; 1:500), goat anti-TrkC (R&D Systems, AF1404; 1:500), rabbit anti-P2X3 (Millipore, AB5895; 1:100), mouse anti-Tuj1 (Millipore, MAB5564; 1:500), rabbit anti-Tuj1 (Abcam, ab18207; 1:2000), mouse anti-Map2 (Sigma-Aldrich, M1406; 1:2000), rabbit anti-synapsin (Calbiochem, 574778; 1:500), goat anti-Dcx (Santa Cruz Biotechnology, sc-8066; 1:500), mouse anti-NF200 (Millipore, MAB5266; 1:500), rabbit anti-TH (Protos Biotech, CA-101bTHrab; 1:1000), rabbit anti-Vamp (Synaptic System, 104203; 1:500), rabbit anti-p75NTR (Abcam, ab8874; 1:500), mouse anti-c-Ret (Sigma-Aldrich, o4886; 1:1000), goat anti-GFP (Abcam, ab6673; 1:2000), rabbit anti-GFP (MBL, 598; 1:2000), chicken anti-GFP (Abcam, ab13970; 1:2000), rabbit anti-vimentin (Abcam, ab92547; 1:2000), and mouse anti-human nuclei (Millipore, MAB1281; 1:200). The secondary antibodies used included donkey anti-rabbit, donkey anti-goat, and donkey anti-mouse Alexa 488 immunoglobulin G (IgG), Alexa 594 IgG, Alexa 546 IgG, Alexa 647 IgG, or Alexa 594 IgM (1:1000; Invitrogen). 4,6-Diamidino-2-phenylindole (DAPI) (Invitrogen) was used for nuclear counterstaining. Images were captured with a laser scanning confocal microscope (Carl Zeiss, LSM700).

One day following infection with ABI lentiviruses, the MEFs were cultured in the presence of 10 M EdU (Life Technologies) for 13 days, or 13 days after infection with AI or ABI viruses, the MEFs were cultured for 24 hours in the presence of 10 M EdU. The cells were then fixed, and EdU staining was carried out according to the manufacturers instruction (Life Technologies). For HSF reprogramming by ABI, EdU was added to the reprogramming cell culture for 29 days starting from day 1 of reprogramming or for 24 hours starting at day 29. Images were captured with a confocal microscope.

For time-lapse recording, we used the JuLI Stage (NanoEntek) with a motorized stage, computer-controlled lens change, and a built-in incubator that supplied humidified 5% CO2 at 37C for live cell recording. MEFs (5 104) derived from the CAG-GFP transgenic mice (28) were induced for 10 days by infection with the ABI lentiviruses or Ascl1 lentiviruses in a well of a 12-well plate precoated with Matrigel. The plate was then placed into the incubator of the JuLI Stage for time-lapse recording for 50 hours. A series of pictures were taken from each well of the 12-well plate in a period of 50 hours under the control of the JuLI EDIT software, which can also edit and replay these pictures in a continuous mode like a movie.

To prepare single iSG cells, MEFs were infected with the ABI (Ascl1, Brn3b, and Isl1) lentiviruses and induced for 2 weeks. Following addition of 500-l Accutase (Millipore) into a well of a 12-well plate, neuronal clusters were suspended by gently pipetting up and down several times using a 1-ml pipette and transferred into a 70-m cell strainer (Falcon) to collect neuronal clusters. Most neuronal clusters attached to the Nylon membrane of the 70-m cell strainer, which was cut from the cell strainer using a pair of scissors and placed into a low-adhesion 6-cm plate containing 4-ml neuron basic medium. To separate the neuronal clusters from the Nylon membrane, the plate was shaken left and right 10 times. The neuronal clusters were then transferred into a 15-ml tube, centrifuged at 1000 rpm for 5 min, resuspended with 1-ml Accutase, and incubated for 5 min at 37C in a CO2 incubator. The neuronal clusters were dissociated into many single cells, which were subsequently used for injection of DRG explants, qRT-PCR, and scRNA-seq analysis.

After euthanization of the rat by the asphyxiation method (CO2 inhalation), the vertebral columns were isolated from the rest of the tissue using a pair of sharp scissors and washed three times with HBSS in a 10-cm culture dish. Both sides of the vertebral columns were mounted onto a surgical mat using needles, and a double cut was made using a pair of surgical scissors to expose the ventral side of the spinal cord. After removal of the spinal cord, DRGs were exposed in the contralateral dorsal spinal roots and pulled out using a pair of fine tweezers. They were collected into a 6-cm culture dish containing HBSS after removal of the attached excessive fibers and connective tissues under a dissection microscope. Four DRGs were transferred onto a Millipore Millicell-CM Low Height Culture Plate Insert using a 3-ml Pasteur pipette, and the rest of HBSS was removed using a 200-l pipette. Then, the insert was placed into a well of a six-well plate containing 1-ml DRG culture medium [BEM (Gibco) supplemented with 20 mM glucose, 1 KIT (Gibco), putrescine (16 ng/ml) (Sigma-Aldrich), 10 mM vitamin C (Sigma-Aldrich), NGF (20 ng/ml) (PeproTech), and 10 mM 5-fluoro-2-deoxyuridine (FDU) (Sigma-Aldrich)]. After being cultured for 1 day, each DRG was injected with 4 103 GFP-labeled single iSG cells. Two weeks following iSG cell injection, the explants were processed and immunostained as described above.

Whole-cell patch-clamp recordings of the iNs were performed with the EPC 10 USB amplifier (HEKA Electronik, Lambrecht, Germany) as previously described (34). Neurons induced from MEFs for 9 or 14 days or from HSFs for 25 to 60 days were used for patch-clamp recordings. In brief, coverslips with adhered cells were transferred into a recording chamber and bathed with Ringers containing 125 mM NaCl, 2.5 mM KCl, 1 mM MgSO4, 2 mM CaCl2, 1.25 mM NaH2PO4, 26 mM NaHCO3, and 20 mM glucose, bubbled with 95% O2 and 5% CO2. Cell responses were recorded with 6- to 9-megohm resistance pipettes that were filled with an internal solution containing 105 mM K-gluconate, 5 mM KCl, 5 mM NaOH, 15 mM KOH, 0.5 mM CaCl2, 2 mM MgCl2, 5 mM EGTA, 2 mM adenosine 5-triphosphate, 0.5 mM guanosine 5-triphosphate, 10 mM Hepes, and 2 mM ascorbate (pH 7.2). The cells and recording pipettes were viewed on a monitor that was coupled to a charge-coupled device camera (Evolve, Photometrics, Tucson, USA) mounted on an upright microscope. Oxygenated external solution was continuously perfused into the recording chamber at a flow rate of 1.5 to 2 ml/min by a peristaltic pump (LEAD-2, Longer Pump, Hebei, China). Capacitive transients were compensated via the Patch Master software (PatchMaster, HEKA), and the series resistance was compensated by ~50%. For current-clamp recording, a small, constant holding current was injected to maintain resting membrane potential (Vrest) at 70 mV and current pulses with a step size of 10 pA were applied to induce action potentials. Voltage-clamp recordings were performed on the same cells directly following current clamp recordings. A simple step protocol from 90 to +30 mV for 200 ms was applied to assess the voltage-gated sodium channels and voltage-gated potassium channels. TTX (Tocris, USA) was added to the bath solution to a final concentration of 0.5 M and perfused into the recording chamber for 5 min. After recording of the currents again, TTX was washed out, followed by the third time of recording.

The fluorescent probe Fluo-8 AM (AAT Bioquest, Sunnyvale, Canada) was used to detect the changes of intracellular calcium. As described above, MEFs were induced for 2 to 3 weeks to form neuronal clusters by infection with the ABI lentiviruses. Under dark environment, glass coverslips with adhered neuronal clusters were loaded with Fluo-8 AM (10 m) for 25 to 30 min at room temperature. After three rinses with Ringers solution, the coverslip was placed into a recording chamber. An upright microscope (Olympus, BX51W1) equipped with a mercury lamp with a 488-nm filter was used to excite Fluo-8. A digital camera (Hamamatsu Photonics, Japan) that was also equipped on the microscope was used to record the fluorescent signal. The software HCImage Live (Hamamatsu Corporation, USA) was used to control the camera and ImageJ for data analysis. Following a 30-s recording of the baseline (F0), 100 mM KCl was puffed to detect the activity of the cells. After a 2-min wash with Ringers, fluorescent signals were decreased to the baseline. Then, 100 M menthol or 10 M capsaicin was puffed to stimulate the iNs. KCl (100 mM KCl) was applied again after menthol/capsaicin to confirm the viability of the tested cells. Only the cells that responded to KCl two times successively were chosen for analysis.

Bulk RNA-seq analysis was performed with modification as previously described (48). Two weeks after infection of MEFs with lentiviruses, total RNA was extracted from GFP-transduced and ABI (Ascl1, Brn3b, and Isl1)transduced MEFs using the TRIzol reagent according to the manufacturers instruction. Ribosomal RNA was depleted before preparation of RNA-seq libraries, which were subsequently sequenced using an Illumina HiSeq 4000 sequencer (Biomarker Technologies, China). The obtained sequence reads were trimmed and mapped to the mouse reference genome (mm10) using HISAT2 (https://daehwankimlab.github.io/hisat2/), and gene expression and changes were analyzed using Cufflinks and Cuffdiff. Hierarchical cluster and scatter plot analyses of gene expression levels were performed using the R software (http://cran.r-project.org). GSEA was carried out as described (31), which was followed by network visualization in Cytoscape using the EnrichmentMap plugin (https://enrichmentmap.readthedocs.io/en/latest/).

Single iSG cells were prepared as described above. Single adult mouse DRG cells were prepared as described previously (49). In brief, DRGs were collected, transferred into a low-adhesion 6-cm pate with 2 ml of DMEM/F12 medium containing collagenase IV (1.25 mg/ml), and incubated at 37C in a 5% CO2 incubator for 50 min. Then, the medium was replaced with 2-ml DMEM/F12 medium containing 0.025% trypsin and incubated for 30 min. Following the addition of 2-ml DMEM/F12 medium containing 33% fetal bovine serum, all the medium was removed using a 10-ml pipette. After being washed three times with 2-ml HBSS, the DRGs were transferred into a 1.5-ml tube containing 1.2-ml DMEM/F12 and triturated by pipetting up and down several times using a 1-ml pipette to obtain single DRG cells. A Brn3b-GFP reporter mouse line was created using the CRISPR-Cas9 gene editing system to label adult RGCs by GFP, which were enriched by fluorescence-activated cell sorting. A more detailed description of this mouse line and RGC enrichment procedure will be published elsewhere.

The number and viability of prepared single cells were quantified using Countess II (Thermo Fisher Scientific, AMQAX1000). Next, single-cell libraries were generated with the Chromium Single Cell 3 V2 Chemistry Library Kit, Gel Bead & Multiplex Kit, and Chip Kit from 10x Genomics. In brief, cell suspension at concentration of 1.2 million/ml was loaded in a Single Cell 3 Chip along with the RT Single Cell 3 Gel Beads and the Partitioning oil, and Single Cell Gel Bead-In-Emulsions were generated in the Chromium Controller. Reverse transcription reaction was run to obtain complementary DNA (cDNA), which was amplified by PCR. To generate the libraries, Enzymatic Fragmentation, End Repair, and A-tailing Double Sided Size Selection were used to incorporate the barcodes and index read sequences. The libraries were qualified by bioanalyzer (Agilent Technologies) and quantified by a Qubit dsDNA High Sensitivity Assay kit (Invitrogen) and then sequenced on Illumina X Ten platform in 150 paired-end configuration.

Raw reads were processed using the 10x Genomics Cell Ranger pipeline (https://support.10xgenomics.com/single-cell-gene-expression/software/downloads/latest) with the mm10 as the reference. Cell Ranger can cluster the single cells, identify the marker genes of each cluster, and export a matrix with unique molecular identifier (UMI) values of each gene in a single cell. The R software package Seurat (https://satijalab.org/seurat, version 2.2) (33) was used for further analysis. Default parameters were used for most of the Seurat analyses. For the FeaturePlot function, max.cutoff was 0.5. The pseudotime trajectory analysis of iSG cells was performed using Monocle 2 (http://cole-trapnell-lab.github.io/monocle-release/) (35).

Statistical analysis was performed using the GraphPad Prism 7 and Microsoft Excel computer programs. The results are expressed as means SD for experiments conducted at least in triplicates. Unpaired two-tailed Students t test or one-way analysis of variance test were used to assess differences between two groups, and a value of P < 0.05 was considered statistically significant.

Acknowledgments: We thank E. Shiang for help with the artwork. Funding: This work was supported, in part, by the National Natural Science Foundation of China (81670862, 81721003, 31871497, 81870682, and 31700900), National Key R&D Program of China (2017YFA0104100, 2018YFA0108300, and 2017YFC1001300), National Basic Research Program (973 Program) of China (2015CB964600), Local Innovative and Research Teams Project of Guangdong Pearl River Talents Program, Science and Technology Planning Projects of Guangzhou City (201904020036 and 201904010358), China Postdoctoral Science Foundation (2019 M650223), and the Fundamental Research Funds of the State Key Laboratory of Ophthalmology, Sun Yat-sen University. Author contributions: D.X., K.J., Y.S., and M.X. conceived and designed the research. D.X., Q.D., Y.G., X.H., M.Z., J.Z., P.R., Z.X., Y.L., and Y.H. performed the experiments and analyzed the data. D.X., K.J., and M.X. interpreted the data and wrote the manuscript. All authors contributed to critical reading of the manuscript. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials. Additional data related to this paper may be requested from the authors. The RNA-seq and scRNA-seq data have been deposited in the NCBI Gene Expression Omnibus database under accession codes PRJNA595403 and PRJNA597624, respectively.

See original here:
Generation of self-organized sensory ganglion organoids and retinal ganglion cells from fibroblasts - Science Advances

Hesperos Human-on-a-Chip System Used to Model Preclinical Stages of Alzheimer’s Disease and Mild Cognitive Impairment – Business Wire

ORLANDO, Fla.--(BUSINESS WIRE)--Hesperos Inc., pioneers of the Human-on-a-Chip in vitro system, today announced a new peer-reviewed publication that describes how the companys functional Human-on-a-Chip system can be used as a drug discovery platform to identify therapeutic interventions targeting the preclinical stages of Alzheimers disease (AD) and mild cognitive impairment (MCI). The manuscript, titled A human induced pluripotent stem cell-derived cortical neuron human-on-a-chip system to study A42 and tau-induced pathophysiological effects on long-term potentiation, was published this week in Alzheimer's & Dementia: Translational Research & Clinical Interventions. The work was conducted in collaboration with the University of Central Florida and with David G. Morgan, Ph.D., Professor of Translational Neuroscience at Michigan State University, and expert in AD pathology.

To date, more than 100 potential therapeutics in development for AD have been abandoned or failed during clinical trials. These therapeutics relied on research conducted in preclinical animal studies, which often are unable to accurately capture the full spectrum of the human disease phenotype, including differences in drug metabolism and excretion between humans and animals. Therefore, there is a need for human models, especially those that accurately recapitulate the functional impairments during the preclinical phases of AD and MCI.

Hesperos offers a breakthrough technology that provides a human cell-based assay based on cognitive function metrics to evaluate drugs designed to restore cognition at early stages of the Alzheimers continuum, said Dr. Morgan. This system can serve as a novel drug discovery platform to identify compounds that rescue or alleviate the initial neuronal deficits caused by A1-42 and/or tau oligomers, which is a main focus of clinical trials.

In 2018, Hesperos received a Phase I Small Business Innovation Research (SBIR) grant from the National Institute on Aging (NIA) division within the US National Institutes of Health (NIH) to help create a new multi-organ human-on-a-chip model for testing AD drugs. Research conducted under this grant included a study to assess therapeutic interventions based on functional changes in neurons, not neuronal death.

In the recent Alzheimer's & Dementia publication, Hesperos describes its in vitro human induced pluripotent stem cell (iPSC)-derived cortical neuron human-on-a-chip system for the evaluation of neuron morphology and function after exposure to toxic A and tau oligomers as well as brain extracts from AD transgenic mouse models.

Researchers are now focusing on biomarker development and therapeutic intervention before symptoms arise in AD and MCI, said James Hickman, Ph.D., Chief Scientist at Hesperos and Professor at the University of Central Florida. By studying functional disruption without extensive cell loss, we now have a screening methodology for drugs that could potentially evaluate therapeutic efficacy even before the neurodegeneration in MCI and AD occurs.

The researchers found that compared to controls, treatment with toxic A and tau oligomers or brain extracts on the iPSC cortical neurons significantly impaired information processing as demonstrated by reduction in high-frequency stimulation-induced long-term potentiation (LTP), a process that is thought to underlie memory formation and learning. Additionally, oligomer and brain extract exposure led to dysfunction in iPSC cortical neuron electrophysiological activity, including decreases in ion current and action potential firing.

While exposure to the toxic oligomers and brain extracts caused morphological defects in the iPSC cortical neurons, there was no significant loss in cell viability.

Clinical success for AD therapies has been challenging since preclinical animal studies often do not translate to humans, said Michael L. Shuler, Ph.D., Chief Executive Officer of Hesperos. With our recent study, we are now one step closer in developing an AD multi-organ model to better evaluate drug metabolism in the liver, penetration through the blood-brain barrier and the effects on neuronal cells.

About Alzheimers Disease/Preclinical Stage AD

AD is a progressive disease that is characterized by memory loss and deterioration of cognitive function. Preclinical AD is the first stage of the disease, and it begins long before any symptoms become apparent. It is thought that symptoms do not manifest until there is a significant death of neuronal cells, which is caused by the aggregation of toxic amyloid beta (A) and tau oligomers, typically during the earliest stages of the disease. Unfortunately, treatment after the diagnosis of MCI may be too late to reverse or modify disease progression.

To read the full manuscript, please visit https://alz-journals.onlinelibrary.wiley.com/doi/full/10.1002/trc2.12029.

About Hesperos

Hesperos, Inc. is a leading provider of Human-on-a-Chip microfluidic systems to characterize an individuals biology. Founders Michael L. Shuler and James J. Hickman have been at the forefront of every major scientific discovery in this realm, from individual organ-on-a-chip constructs to fully functional, interconnected multi-organ systems. With a mission to revolutionize toxicology testing as well as efficacy evaluation for drug discovery, the company has created pumpless platforms with serum-free cellular mediums that allow multi-organ system communication and integrated computational PKPD modeling of live physiological responses utilizing functional readouts from neurons, cardiac, muscle, barrier tissues and neuromuscular junctions as well as responses from liver, pancreas and barrier tissues. Created from human stem cells, the fully human systems are the first in vitro solutions to accurately predict in vivo functions without the use of animal models. More information is available at http://www.hesperosinc.com.

Hesperos and Human-on-a-Chip are trademarks of Hesperos Inc. All other brands may be trademarks of their respective holders.

See more here:
Hesperos Human-on-a-Chip System Used to Model Preclinical Stages of Alzheimer's Disease and Mild Cognitive Impairment - Business Wire

AgeX Therapeutics and Sernova to Collaborate to Engineer Universal Locally Immune Protected Cell Therapies for Type I Diabetes and Hemophilia A -…

ALAMEDA, Calif. & LONDON, Ontario--(BUSINESS WIRE)--AgeX Therapeutics, Inc. (AgeX: NYSE American: AGE), a biotechnology company developing therapeutics for human aging and regeneration, and Sernova Corp. (TSX-V:SVA)(OTCQB:SEOVF)(FSE:PSH), a clinical-stage regenerative medicine therapeutics company, announced today a research collaboration where Sernova will utilize AgeXs UniverCyteTM gene technology to generate immune-protected universal therapeutic cells for use in combination with Sernovas Cell PouchTM for the treatment of type I diabetes and hemophilia A. The goal is to eliminate the need for immunosuppressive medications following Cell Pouch cell transplantation.

The research collaboration will evaluate whether Sernovas pluripotent stem cell-derived pancreatic islet beta cells engineered with AgeXs UniverCyte technology can evade human immune detection. The complementary combination of technologies could enable the transplantation of therapeutic cells in patients with type I diabetes in an off-the-shelf manner using Sernovas Cell Pouch, without human leukocyte antigen (HLA) tissue matching or concurrent administration of immunosuppressive medications. With a similar intent, pluripotent stem cell-derived or adult donor-derived human Factor VIII-releasing cells modified with AgeXs UniverCyte will be evaluated in Sernovas hemophilia A program.

Under the terms of the agreement, Sernova has been granted a time-limited, non-exclusive research license by AgeX. A commercial license for Sernova to utilize UniverCyte to engineer cellular products for therapeutic and commercial purposes may be negotiated between the companies pending successful study outcomes.

The UniverCyte technology aims to mask therapeutic cells derived from pluripotent stem cells or adult donors from human immune detection to allow for off-the-shelf cellular products without the need for immunosuppressant medications which may have potent side effects, or HLA-matching between donor and patient. UniverCyte uses a novel, modified form of HLA-G, a potent immunomodulatory molecule, which in nature protects an unborn child from their mothers immune system. In almost all human cells, native HLA-G expression is silenced after birth. AgeXs modified HLA-G shows evidence of being resistant to this silencing, thereby potentially allowing for long-term, stable and high expression of the immunomodulatory effect.

Sernova plans to utilize the universal therapeutic cells generated through this research collaboration with its Cell Pouch System, a proprietary, scalable, implantable macro-encapsulation device, which, upon implantation, incorporates with tissue and forms highly vascularized chambers. These chambers become a natural environment in the body to house and favor long-term survival and function of therapeutic cells. The Cell Pouch System has shown initial safety and efficacy indicators in an ongoing Phase I/II clinical study at the University of Chicago and in a preclinical model of hemophilia A when assessed with human cells corrected to produce Factor VIII.

We are thrilled with our collaboration with Sernova, which is at the forefront of cellular therapies for diabetes and hemophilia and is already in the clinic for the former. The combination of AgeXs UniverCyte to cloak cells from a patients immune system and Sernovas Cell Pouch technologies to permit cells to function long-term upon transplantation would be a landmark for regenerative medicine. This deal marks another important step in AgeXs collaboration and licensing strategy to work with the very best people, companies and institutions in the world of regenerative medicine, said Dr. Nafees Malik, Chief Operating Officer of AgeX.

We look forward to working with AgeX and its outstanding team as we continue to identify and evaluate technologies complementary to Sernovas therapeutic platform and expand our immune protection offerings. AgeXs UniverCyte technology is a significant advancement in the field of cell therapy and a perfect fit with Sernovas Cell Pouch technologies and therapeutic pipeline with its potential benefit over current immunosuppressive strategies for regenerative medicine therapeutics, said Dr. Philip Toleikis, President and CEO of Sernova Corp.

About AgeX Therapeutics

AgeX Therapeutics, Inc. (NYSE American: AGE) is focused on developing and commercializing innovative therapeutics for human aging. Its PureStem and UniverCyte manufacturing and immunotolerance technologies are designed to work together to generate highly defined, universal, allogeneic, off-the-shelf pluripotent stem cell-derived young cells of any type for application in a variety of diseases with a high unmet medical need. AgeX has two preclinical cell therapy programs: AGEX-VASC1 (vascular progenitor cells) for tissue ischemia and AGEX-BAT1 (brown fat cells) for Type II diabetes. AgeXs revolutionary longevity platform induced Tissue Regeneration (iTR) aims to unlock cellular immortality and regenerative capacity to reverse age-related changes within tissues. AGEX-iTR1547 is an iTR-based formulation in preclinical development. HyStem is AgeXs delivery technology to stably engraft PureStem cell therapies in the body. AgeXs core product pipeline is intended to extend human healthspan. AgeX is seeking opportunities to establish licensing and collaboration arrangements around its broad IP estate and proprietary technology platforms and therapy product candidates.

For more information, please visit http://www.agexinc.com or connect with the company on Twitter, LinkedIn, Facebook, and YouTube.

About Sernova Corp.

Sernova Corp is developing regenerative medicine therapeutic technologies using the Cell Pouch System, a medical device and immune protected therapeutic cells (i.e., human donor cells, corrected human cells and stem-cell-derived cells) to improve the treatment and quality of life of people with chronic metabolic diseases such as insulin-dependent diabetes, blood disorders including hemophilia, and other diseases treated through replacement of proteins or hormones missing or in short supply within the body. For more information, please visit http://www.sernova.com.

Forward-Looking Statements for AgeX

Certain statements contained in this release are forward-looking statements within the meaning of the Private Securities Litigation Reform Act of 1995. Any statements that are not historical fact including, but not limited to statements that contain words such as will, believes, plans, anticipates, expects, estimates should also be considered forward-looking statements. Forward-looking statements involve risks and uncertainties. Actual results may differ materially from the results anticipated in these forward-looking statements and as such should be evaluated together with the many uncertainties that affect the business of AgeX Therapeutics, Inc. and its subsidiaries, particularly those mentioned in the cautionary statements found in more detail in the Risk Factors section of AgeXs most recent Annual Report on Form 10-K and Quarterly Report on Form 10-Q filed with the Securities and Exchange Commissions (copies of which may be obtained at http://www.sec.gov). Subsequent events and developments may cause these forward-looking statements to change. In addition, there can be no assurance that Sernovas planned use of AgeXs UniverCyteTM gene technology will successfully generate immune-protected universal therapeutic cells for use in combination with Sernovas Cell PouchTM for the treatment of type I diabetes and hemophilia A or any other disease, and there can be no assurance that AgeX and Sernova will enter into a commercial license for the use of UniverCyteTM in a therapeutic or other product. AgeX specifically disclaims any obligation or intention to update or revise these forward-looking statements as a result of changed events or circumstances that occur after the date of this release, except as required by applicable law.

Forward-Looking Statements for Sernova

This release may contain forward-looking statements. Forward-looking statements are statements that are not historical facts and are generally, but not always, identified by the words expects, plans, anticipates, believes, intends, estimates, projects, potential and similar expressions, or that events or conditions will, would, may, could or should occur. Although Sernova believes the expectations expressed in such forward-looking statements are based on reasonable assumptions, such statements including those related to the potential of Univercyte combined with Sernovas technologies are not guarantees of future performance, and actual results may differ materially from those in forward-looking statements. Forward-looking statements are based on the beliefs, estimates, and opinions of Sernovas management on the date such statements were made, which include our beliefs about the effect on company operations of the COVID-19 virus and conduct and outcome of discussions, clinical programs, and our clinical trials. Sernova expressly disclaims any intention or obligation to update or revise any forward-looking statements, whether as a result of new information, future events or otherwise.

See the original post here:
AgeX Therapeutics and Sernova to Collaborate to Engineer Universal Locally Immune Protected Cell Therapies for Type I Diabetes and Hemophilia A -...

In Vitro Toxicology Testing Market to Grow at Robust CAGR in the COVID-19 Lockdown Scenario – Cole of Duty

[112 Report Pages] This market research report identifies Laboratory Corporation of America Holdings, Charles River Laboratories, Inc, Thermo Fisher Scientific, Eurofins Scientific, Agilent Technologies, Inc., as the major vendors operating in the global in vitro toxicology testing market. This report also provides a detailed analysis of the market by toxicology end points (systemic toxicity, cytotoxicity testing, genotoxicity testing, ocular toxicity, organ toxicity, dermal toxicity, neurotoxicity, and others), industry type (pharmaceutical and biopharmaceutical, cosmetics, chemical, diagnostics, and food industry), and region (North America, Europe, Asia Pacific, and Rest of the World).

Request For Report [emailprotected]https://www.trendsmarketresearch.com/report/sample/9863

Infoholicsmarket research report predicts that the globalin vitro toxicology testingmarket will grow at a CAGR of8.2%during the forecast period 20182024.The market for in vitro toxicology testing is driven by high opposition to animal testing, increased cost related to animal-based toxicity testing, and increasing R&D expenditure for early stage toxicity testing. Whereas, the lack of in vitro models and decreased adoption rate are limiting the growth of the in vitro toxicology testingmarket to an extent.

According to the in vitro toxicology testingmarket analysis, Europe accounted for the largest share of the global in vitro toxicology testingmarket followed by North America in 2017. The reason is the upsurge in the investments by the European Commission in R&D to develop substitute methods to in vitro testing is driving the demand in this region. Asia Pacific is expected to grow at a high CAGR during the forecast period due to increasing number of contract research organizations offering testing services, advancements in healthcare infrastructure, increasing investments in the biopharmaceutical sector, and upward economic conditions in the region.

Competitive Analysis and Key Vendors:

There is an increase in collaborations between companies on in vitro testing of compounds. For instance, in December 2016, Evotec and Celgene entered into a drug discovery collaboration for neurodegenerative diseases. According to agreement terms, Celgene will use Evotecs unique induced pluripotent stem cell (iPSC) platform that enables systematic drug screening in patient-derived disease models. In June 2017, Censo Biotechnologies Ltd. collaborated with Evotec AG to source and provide patient-derived induced pluripotent stem cells to support Evotecs drug discovery iPSC platform. In addition, the companies are also coming up with new products for in vitro testing. For instance, in January 2018, STEMCELL Technologies Inc. released two product lines for organoid research that will enable scientists to create powerful models for studying human disease in the laboratory.

Some of the In Vitro Toxicology Testing Market key vendorsare:

Other prominent vendors in the global in vitro toxicology testing market are Bio-Rad Laboratories, GE Healthcare, SGS SA, BioIVT, Abbott Laboratories, Gentronix Limited, Promega Corporation, MB Research Laboratories, Evotec AG (Cyprotex plc), Catalent, Inc., Qiagen N.V., and niche players.

In Vitro Toxicology Testing Market by Toxicology End Points:

In 2017, the systemic toxicity accounted for the highest market share due to the availability of a wide range of sub-studies, which ensure total analysis of toxicity and safety margin of the testing compounds.

<<< Get COVID-19 Report Analysis >>>https://www.trendsmarketresearch.com/report/covid-19-analysis/9863

In Vitro Toxicology Testing Market by Industry type:

In 2017, the pharmaceutical and biopharmaceutical industry occupied significant market share and the cosmetics industry is expected to hold a high percentage during the forecast period. Increased support of regulatory authorities to use in vitro and in silico methods instead of animal testing to check toxicology is driving the growth of the cosmetic industry.

In Vitro Toxicology Testing Market Benefits:

The report provides detailed information about the services offered by in vitro toxicology testingin various therapeutic verticals and regions. With that, key stakeholders can find out the major trends, drivers, investments, and vertical players initiatives. Moreover, the report provides details about the major challenges that are going to have an impact on market growth. Additionally, the report gives complete details about the business opportunities to key stakeholders to expand their business and capture revenues in the specific verticals. The report will help companies interested or established in this market to analyze the various aspects of this domain before investing or expanding their business in the in vitro toxicology testingmarket.

Continued here:
In Vitro Toxicology Testing Market to Grow at Robust CAGR in the COVID-19 Lockdown Scenario - Cole of Duty