Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Phenotyping of Human iPSC-derived Neurons: Patient-Driven Research
Phenotyping of Human iPSC-derived Neurons: Patient-Driven Research
Phenotyping of Human iPSC-derived Neurons: Patient-Driven Research
Ebook841 pages5 hours

Phenotyping of Human iPSC-derived Neurons: Patient-Driven Research

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Phenotyping of Human iPSC-derived Neurons: Patient-Driven Research examines the steps in a preclinical pipeline that utilizes iPSC-derived neuronal technology to better understand neurological disorders and identify novel therapeutics, also providing considerations and best practices. By presenting example projects that identify phenotypes and mechanisms relevant to autism spectrum disorder and epilepsy, this book allows readers to understand what considerations are important to assess at the start of project design. Sections address reproducibility issues and advances in technology at each stage of the pipeline and provide suggestions for improvement. From patient sample collection and proper controls to neuronal differentiation, phenotyping, screening, and considerations for moving to the clinic, these detailed descriptions of each stage of the pipeline will help everyone, regardless of stage in the pipeline.

In recent years, drug discovery in the neurosciences has struggled to identify novel therapeutics for patients with varying indications, including epilepsy, chronic pain, and psychosis. Current treatment options for such patients are decades old and offer little relief with many side effects. One explanation for this lull in novel therapeutics is a lack of novel target identification for neurological disorders (and target identification requires exemplar preclinical data). To improve on the preclinical work that often relies on rodent modeling, the field has begun utilizing patient-derived induced pluripotent stem cells (iPSCs) to differentiate neurons in vitro for preclinical characterization of neurological disease and target identification.

  • Discusses techniques and new technology for iPSC culturing and neuronal differentiation to establish best practices in the lab
  • Outlines considerations for phenotypic assay development
  • Provides information about the successes, failures, and implications of phenotyping and screening with iPSC-derived neurons
  • Describes how human iPSC-derived neurons are being used for preclinical discovery research as well as the development of therapeutics utilizing hiPSC-derived neurons
LanguageEnglish
Release dateSep 9, 2022
ISBN9780128222782
Phenotyping of Human iPSC-derived Neurons: Patient-Driven Research

Related to Phenotyping of Human iPSC-derived Neurons

Related ebooks

Medical For You

View More

Related articles

Related categories

Reviews for Phenotyping of Human iPSC-derived Neurons

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Phenotyping of Human iPSC-derived Neurons - Elizabeth D. Buttermore

    Section I

    Best practices and considerations when designing a new project

    Outline

    Chapter 1. iPSC culture: best practices from sample procurement to reprogramming and differentiation

    Chapter 2. Phenotypic assay development with iPSC-derived neurons: technical considerations from plating to analysis

    Chapter 3. Derivation of cortical interneurons from human pluripotent stem cells to model neurodevelopmental disorders

    Chapter 4. Development of transcription factor-based strategies for neuronal differentiation from pluripotent stem cells

    Chapter 5. Differentiation of Purkinje cells from pluripotent stem cells for disease phenotyping in vitro

    Chapter 6. Brain organoids: models of cell type diversity, connectivity, and disease phenotypes

    Chapter 1: iPSC culture

    best practices from sample procurement to reprogramming and differentiation

    Laurence Daheron ¹ , and Ivy Pin-Fang Chen ² , ³       ¹ Harvard Stem Cell Institute, Harvard University, Cambridge, MA, United States      ² Tessera Therapeutics, Cambridge, MA, United States      ³ Human Neuron Core, Boston Children's Hospital, Boston, MA, United States

    Abstract

    In this chapter, we describe best practices and recommendations for patient sample collection, derivation, and differentiation of human induced pluripotent stem cells (iPSCs) in a research laboratory setting. The purpose of this chapter is to set the stage for best practices that should be considered when one is implementing iPSC culture in their lab. Utility of iPSCs in specific disorders or applications, technical difficulties, and challenges are addressed in subsequent chapters, with additional requirements and recommendations for each topic.

    Keywords

    Biology experimental methods; Cell; Experimental design; Human primary sample; Induced pluripotent stem cells; Neuron differentiation; Quality control; Reprogramming; Stem cells research; Tissue culture logistics

    In this chapter, we describe best practices and recommendations for patient sample collection, derivation and differentiation of human induced pluripotent stem cells (iPSCs) in a research laboratory setting. The purpose of this chapter is to set the stage for best practices that should be considered when one is implementing iPSC culture in their lab. Utility of iPSCs in specific disorders or applications, technical difficulties, and challenges are addressed in subsequent chapters, with additional requirements and recommendations for each topic.

    Facility setup

    Tissue culture room design

    General guidelines and best practices for setting up a tissue culture room for iPSC culture are similar to those for any other mammalian cell culture system at Biological Safety Level 2 (BSL2). ¹ , ² However, iPSCs are typically cultured in the absence of antibiotics and additional measures should be taken to maintain sterility. Here we briefly touch on best practices for facility setup as well as equipment that comes in handy for culturing iPSCs. To reduce dust collection, it is ideal for a tissue culture room to have a balanced neutral or slightly positive airflow with respect to neighboring areas, unless biohazardous agents are used. Floors and walls must be durable and liquid tight and, although not required, epoxy coating can help maintain cleanliness. Similarly, smooth and cleanable drop ceiling tiles are ideal. Porous materials, such as fabric and wood, should be avoided wherever possible. A hand washing sink (and eyewash station in some cases) is now a requirement in a tissue culture room due to safety concerns. However, still water and damp environment encourage microorganism growth and extensive measures—including leak proofing, antifungal foam spraying, waterproofing, and regular cleaning—should be taken to prevent the sink from turning into a source of contamination.

    Tissue culture equipment

    Basic equipment required includes storage racks, a refrigerator, a freezer, a vortex mixer, centrifuges, microscopes, a water/bead bath, an incubator, a biosafety cabinet, and a liquid nitrogen storage unit. For centrifuges, a swing bucket centrifuge (e.g., Eppendorf 5702, Cole–Parmer MS-3400) with biocontainment lids is essential and a mini centrifuge is nice to have for quick spins of microcentrifuge tubes. For microscopes, an entry-level inverted light microscope with 4×, 10×, and 20× phase contrast objectives (e.g., Leica DMi1, Zeiss PrimoVert, Nikon ECLIPSE Ts2) is recommended and a compact fluorescence microscope (e.g., ThermoFisher EVOS M5000, Leica DMi8) is helpful if using fluorescence reporters. A standard noncirculating 37°C water bath is commonly used to warm media and reagents, but we recommend replacing water with autoclavable metal beads due to contamination concerns mentioned above. If a water bath were to be used, weekly cleaning and water treatments (e.g., Spectrum Labs Clear Bath, Promo Cell Aquaguard, SP Scienceware Aqua-Clear) would be necessary to reduce growth of microorganisms. Similar practices should be done for the water pan in an incubator to maintain sterility. For incubators, self-sterilizing incubators are recommended (e.g., Panasonic MCO-170AICUVH, ThermoFisher Heracell 150i). Incubator temperature and gas level should be monitored and calibrated monthly, as well as certified annually. Complete sterilization, including replacing HEPA filter (if used), should be done annually at a minimum. Incubators should be placed away from foot traffic, doorways, or air vents to reduce exposure to contaminants. As vibration can cause cells to detach or distribute unevenly, it is not advised for incubators to be near or share one surface with a source of vibration, such as a large centrifuge. It is important to leave enough space around incubators for venting, tubing/cords, as well as maintenance and to ensure that the desired gas inlets (i.e., CO2, N2) are readily available nearby. For BSL2, class II type A2 biosafety cabinets are most commonly used, which should also be placed away from doorways or air vents to allow optimal laminar airflow. Similarly, it is important to leave enough space around biosafety cabinets for venting, tubing/cords, as well as vacuum. Biosafety cabinets should be certified annually for airflow and HEPA filter function. ³ A liquid nitrogen storage unit is required for long-term storage of iPSCs, which should be stored in vapor phase below −135°C. ⁴ Although manually filled liquid nitrogen storage dewars can be used, an automated cryopreservation system is highly recommended to have better control over liquid level and temperature, both of which are critical for iPSC viability post-thaw.

    In addition to the equipment listed above, several pieces of specialized equipment are recommended for deriving, culturing, and differentiating iPSCs, as well as characterizing iPSCs and their derivatives. A dissection microscope (e.g., Nikon SMZ800, Zeiss Stemi305, Olympus SZX7) is necessary for processing skin biopsies to isolate and culture fibroblasts and for manually picking iPSC colonies during reprogramming or genome editing. To help maintain sterility, the dissection microscope can be put inside a biosafety cabinet or a laminar flow microscope enclosure with HEPA filtration (e.g., AirClean Systems AC632TLFUVMIC). A microscope objective marker (e.g., Nikon MBW10020), a stamp for marking colonies, is useful for clone isolation. For process standardization in tissue culture and treatments, a cell counter (e.g., ThermoFisher Countess, Bio-Rad TC20), electronic multichannel pipettes, and robotic liquid handlers are helpful, especially when working with a large number of small well formats, such as 96- and 384-well plates. In addition, automated imaging stations (e.g., Sartorius IncuCyte for live cell imaging, Molecular Devices IXM-c for high content imaging) are helpful for documenting reprogramming and/or differentiation progress. These platforms can also increase throughput for imaging and screening through the use of automation. Similarly, for increased throughput on functional readouts in iPSC-derived neurons, multi-electrode array recording devices (e.g., Axion BioSystems Maestro) and fluorescence/luminescence plate readers (e.g., Hamamatsu FDSS7000EX) can be used to measure neuronal network activity. Phenotyping equipment can be outfitted with robotic arms for streamlining and humidity control incubators to reduce edge effects in multi-well plates.

    Primary sample collection

    Somatic cells

    The first human iPSC lines were generated from dermal fibroblasts, ⁵ , ⁶ but a large variety of cell types have been reprogrammed since, including different blood cells, ⁷ endothelial cells, ⁸ keratinocytes, ⁹ and adipose-derived stem cells. ¹⁰ The main consideration for the choice of somatic cells is the ease and safety of collection for the donors. Peripheral blood and skin biopsies are the most commonly used sources of cells for reprogramming. Skin biopsies are minimally invasive, but still more intrusive than a simple blood draw. The derivation of fibroblast culture from skin biopsy can be challenging and can take as long as 2–3weeks. Moreover, the fibroblast proliferation rate is crucial for reprogramming success, as partially senescent fibroblasts have limited capacity for reprogramming. Mutational signatures associated with UV-exposure have been reported for iPSC lines derived from dermal fibroblasts, ¹¹ raising concerns about the use of these cells in generating cGMP iPSC lines. However, fibroblasts remain a common source of cells for reprogramming as they are amenable to reprogramming using the three different non-integrative methods (Sendai virus, Episomal vector, and mRNA transfection).

    From peripheral blood samples, different cell types can be isolated for culture. The first step (which should be done within 24h post-draw to increase reprogramming efficiency) involves the isolation of peripheral blood mononuclear cells (PBMCs) by Ficoll gradient or equivalent. These cells can be frozen or cultured immediately for reprogramming. Hematopoietic progenitors (CD34 positive cells) purified from PBMCs show high reprogramming efficiency compared to more differentiated blood cells, but they are rare in the peripheral blood and therefore not often used as a starting material. Erythroblasts have become the most widely used cells for reprogramming. ¹² By culturing in the presence of a combination of recombinant human cytokines (EPO, SCF and IL3) for 9days, erythroblasts can be enriched to 90%–100% from the mixed population of cells in PBMC. T lymphocytes are another blood cell type that have been used for reprogramming. ¹³ These cells are abundant in PBMCs and are easy to activate in the presence of anti-CD3 and anti-CD8 antibodies, in addition to IL2. Within 3days post blood draw, a high number of T lymphocytes can be ready for reprogramming. However, these cells are difficult to reprogram (low efficiency), and they have the TCR receptor rearrangement unique to each cell, which could be a problem for some applications. Erythroblasts and T lymphocytes can be reprogrammed using either the Sendai virus or the episomal vector methods. However, the mRNA transfection reprogramming method has been unsuccessful for any blood cell type so far. One alternative is to isolate endothelial progenitor cells (EPC) from fresh blood samples. EPCs are circulating cells that express a number of markers specific for vascular endothelial cells. They are rare and therefore difficult to expand from a small volume (less than 10mL) of blood. EPCs are amenable to reprogramming with the mRNA method, but it takes approximately 3weeks before they are ready for reprogramming. Urine samples are another convenient source of cells for reprogramming. ¹⁴ Cells are harvested by centrifugation of urine and two cell types can be expanded: epithelial-like cells with limited proliferation capacity and mesenchymal type cells with higher expansion capacity. Although urine is easy to collect, bacterial contamination of the culture is relatively high and the rate of failure for urine cell expansion is around 30%. Since urine sample collection is noninvasive, it is particularly suitable for children. The cGMP iPSC lines generated to date were derived from blood or fibroblast cells. For autologous transplantation, the use of T lymphocytes was proposed to take advantage of the TCR receptor rearrangement as a natural tag for the cells. This unique tag allows for identification of the transplanted cells post transplantation. For allogenic transplantation, obtaining somatic cells with limited mutational load is crucial. EPC or CD34+ cells isolated from cord blood could provide a safe source of cells for reprogramming.

    Quality control of somatic cells

    Mycoplasma testing should be performed on all somatic samples before starting the reprogramming process. For cells that have a greater risk of bacterial contamination (such as skin or urine cells), we recommend thoroughly washing the initial sample (skin biopsies and cell pellets from urine) with PBS containing antibiotics and initially expanding the cells in medium with antibiotics. Once established, it is good practice to culture somatic cells without antibiotics before reprogramming to reduce the chance of creating drug-resistant microorganisms. Viral testing (e.g., HIV, HBV, HCV, HTLV, etc.) is not commonly implemented on samples used for research grade iPSC lines. However, some stem cell banks or FACS sorting facilities might require viral testing before accepting new lines. For cGMP iPSC lines, a panel of relevant communicable disease agents or diseases (RCDADs) are tested in the somatic cells of origin. Karyotyping is rarely performed on somatic samples.

    Reprogramming

    In 2007, Shinya Yamanaka generated the first human iPSC lines by overexpressing Oct4, Sox2, Klf4, and c-Myc in fibroblasts using a retrovirus system. ⁵ This system offers three advantages: it efficiently transduces the fibroblast cells, allows high expression of the four transcription factors for a prolonged period of time, and results in silenced transgene expression in induced pluripotent stem cells, as retroviruses are known to be silenced upon reprogramming. However, since the retrovirus is randomly integrated into the genome, the resulting iPSC lines were not transgene free and therefore had limited applications. An optimal delivery system should provide the following advantages: (1) Efficient delivery: including strong expression of the reprogramming factors, sustained expression over time, and a high percentage of transfected/transduced cells; (2) Reliability: ability to deliver the transgenes in any somatic cell type; (3) Simplicity: one step delivery; (4) Limited toxicity: low cell death; (5) Safety: no integration into the genome and no residual expression of transgenes in iPSC lines. Several methods have been reported since the early study from Shinya Yamanaka. Currently, three non-integrating approaches are commonly used: the episomal vector, the Sendai virus and the mRNA transfection methods.

    Pros and cons of each method

    Episomal vector transfection

    Episomal vectors are plasmids containing the cis-acting origin of replication (oriP) and trans-acting EBNA1 viral elements from the Epstein–Barr Virus. These two elements confer their ability to self-replicate in the cells without integration into the genome. The first iPSC lines generated with these vectors showed very low reprogramming efficiency and required the overexpression of Nanog, Lin28 and SV40 large T genes, in addition to the four original Yamanaka transcription factors. ⁶ Improvement in the efficiency of this method was described a few years later. ¹⁵ In this report, three episomal vectors were used: one with Oct4 and a short hairpin RNA (shRNA) for p53, a second that includes Klf4 and Sox2, and a third that includes L-Myc and Lin28. The p53 pathway acts as a roadblock to reprogramming somatic cells into pluripotency. ¹⁶ The downregulation of p53, which is described in this study, enhances the reprogramming efficiency. This method is simple since it requires only one transfection and can be used to reprogram a variety of cell types. One limiting factor is the transfection efficiency of the large episomal vectors (>10Kb). Nucleofection is recommended over the lipofection system to ensure efficient delivery in the cells. The episomal method has been used to generate many research grade iPSC lines and most of the cGMP lines currently available. However, there are a few caveats to this method. The reprogramming efficiency (calculated based on the number of iPSC colonies generated per input cells) is the lowest of the three methods (around 0.01% for fibroblasts). ¹⁷ The addition of the shRNA against p53, a tumor suppressor, could increase the incidence of genetic abnormalities in the resulting iPSC lines. We reported that the percentage of iPSC lines with abnormal karyotype was the highest (11%) when using the episomal vectors. Last, integration of the episomal vectors into the genome of the iPSC lines is frequent. ¹⁷

    Sendai virus transduction

    The sendai virus (SeV) is a rodent virus of the paramyxoviridae family causing respiratory infection. It is a single-stranded negative sense RNA virus. In 2009, it was used for the first time for the generation of iPSC lines from human fibroblasts. ¹⁸ Since this virus does not go through a DNA phase, there is no risk of integration into the host genome. This method offers several benefits. First, it provides an efficient system to deliver the Yamanaka factors. Even at low multiplicity of infection (MOI 1 to 3), Fusaki et al. showed close to 100% transfection efficiency in fibroblasts and high expression of the reprogramming factors. Next, it can be used to transduce a wide range of cells, as it binds to sialic acid found on glycoproteins at the surface of the cell membrane. Moreover, only one transduction is sufficient since the RNA polymerase from the Sendai virus maintains the RNA synthesis for an extended time. The first commercially available reprogramming Sendai virus kit (CytoTune-iPS 1.0) was launched in 2011. This method has been widely used by research labs and core facilities since its introduction. The Cytotune 1.0 consisted of four viral preparations, one for each of the reprogramming factors. An optimized version (CytoTune-iPS 2.0) with less cytotoxicity, higher efficiency and faster elimination of the SeV RNA in the iPSC lines was developed later. Besides the cost, there are two main concerns regarding this method: (1) it relies on a rodent virus which could hamper its use in countries or institutions with restrictive virus use policies, and (2) The SeV RNA is not eliminated rapidly in iPSC lines as the majority of the iPSC lines still express the transgenes at passage 5. A cGMP version of the reprogramming SeV kit is now available (CytoTune-iPS 2.1). It includes three viral preparations: one with Oct4, Sox2, and Klf4, one with L-Myc (instead of c-Myc) and a third with Klf4. Our group has successfully used this kit to generate cGMP iPSC lines from T cells from two patients with diabetes.

    Both reprogramming methods described above may result in iPSC lines with transgene expression, either permanently for the episomal vectors (since it can be integrated into the genome) or transiently for the Sendai virus (until the SeV RNA virus is eliminated).

    mRNA reprogramming method

    A third approach eliminating this risk was reported in 2010 by the Rossi group. ¹⁹ It is based on the delivery of mRNA encoding reprogramming factors to the cells. mRNAs are swiftly degraded, so newly derived iPSC lines are instantly transgene free. However, due to the short-lived nature of mRNA, successive daily transfections are required to efficiently reprogram somatic cells. Moreover, the transfection of mRNA triggers an innate immune response in cells, often leading to cell death. To attenuate the immunogenicity of the mRNA, modified ribonucleotide bases need to be incorporated and B18R, an interferon response inhibitor, is required post transfection. This method enables a robust but transient expression of the reprogramming factors. In the first report, 15–17 successive transfections were necessary to obtain iPSC colonies. By adding microRNAs (including clusters miR-290-295 or miR-302-367) in the reprogramming cocktail, pluripotent reprogramming is promoted and the number of transfections can be reduced to 12. More recently, an optimized version of the mRNA/microRNA reprogramming method was described, allowing the reprogramming of somatic cells with only four transfections for fibroblasts and eight transfections for urine epithelial cells or endothelial progenitor cells. ²⁰ Overall, the mRNA transfection method is the most efficient as it can reach 1% efficiency (1 iPSC colony per 100 input cells). However, the consecutive transfections can be challenging since the toxicity varies between samples. It is recommended to plate the cells at different densities to ensure successful reprogramming. Moreover, as stated before, this method is efficient for fibroblast reprogramming, but has not been demonstrated to reprogram blood cells. A comparison of iPSC lines generated by these three methods showed no difference in their differentiation potential. ¹⁷

    iPSC line characterization

    The characterization and quality control of iPSC lines can be divided into five categories: sterility, pluripotency, transgene elimination, identity, and genetic stability.

    Sterility

    It is crucial to maintain sterility of cell culture during the different steps associated with deriving new iPSC lines, including reprogramming, picking iPSC clones, expanding, freezing and thawing newly derived iPSC lines. Good cell culture technique, appropriate set up (dissection microscope placed in a biosafety cabinet), and regular monitoring of the culture are key elements for human pluripotent stem cell (hPSC) culture. Bacterial, yeast, and fungal contaminations can easily be detected by observing the cell culture under the microscope on a regular basis. In contrast, mycoplasma contamination is not visible, but can negatively affect the cells and change their characteristics. It is important to test for mycoplasma contamination regularly or at strategic time-points, such as before freezing a new stock or before starting a new experiment. PCR-based or biochemical-based mycoplasma detection assays are commercially available. For the cGMP iPSC lines our group derived in collaboration with Dr. Ritz at Dana Farber, sterility was assessed by the BacT/alert system (BioMerieux, Marcy l’Etoile, France). Samples were inoculated directly into aerobic and anaerobic culture bottles and incubated for 14days to assess microbial growth.

    Pluripotency

    The pluripotent state of newly derived iPSC lines is typically assessed by testing the expression of certain stem cell markers and, additionally, by evaluating the lines' trilineage differentiation potential. Cells that are fully reprogrammed reactivate stem cell associated markers, such as OCT4, NANOG, SOX2, and LIN28, to a level comparable to human embryonic stem cells. ²¹ This reactivation is evaluated by qRT-PCR using primers specific for the endogenous genes or by looking at the DNA methylation of their promoters through bisulfite sequencing. The somatic cells of origin are used as negative controls and hESCs are used as positive control. The protein expression of these markers can also be confirmed by immunocytochemistry or FACS analysis. Moreover, standardized assays are available to facilitate the characterization of iPSC lines. Pluritest is a bioinformatics assay allowing the comparison of newly derived iPSC lines to a large set of reference samples (more than 450 samples) based on gene expression profiles. ²² This assay provides both a pluripotency and a novelty score. The pluripotency score indicates how strong the pluripotency signature is in the new samples. The novelty score detects technical or biological variations in the data. The pluritest analysis tool is compatible with Illumina HT12V3/HT12V4 arrays, RNA-Seq data, and Affymetrix HG-U219 gene expression arrays. It is also supported by Thermo Fisher PrimeView Global Gene Expression profile assay. The Epi-Pluri score is another assay based on epigenetic biomarkers reported by Lenz et al. ²³ It distinguishes pluripotent stem cells based on the DNA methylation level at three specific CpG sites. This test is offered by the company Cygenia. All of these aforementioned assays show the undifferentiated state of the cells, but the validation of their pluripotency state can only be demonstrated by differentiating the cells into the three germ layers. In vitro spontaneous differentiation is achieved using embryoid body (EB) formation or monolayer culture. Alternatively, iPSCs can be specifically differentiated into mesoderm, endoderm and ectoderm by using recombinant proteins that promote the emergence of specific lineages. This directed differentiation method provides a better yield, is faster, and is more consistent than EB formation. Following differentiation, cells from the three germ layers are detected by immunocytochemistry or by quantitative RT-PCR. The development of standardized quantitative assays, such as the scorecard assay, offers the possibility to compare differentiation potential of multiple iPSC lines. The scorecard was first established by monitoring the expression of 500 selected genes using the nanostring technology. ²⁴ It was further simplified by assessing the expression of less than 100 genes with a TaqMan based method (TaqMan hPSC Scorecard Assay, Thermo Fisher). ²⁵ Teratoma formation assay is the classic method used to evaluate differentiation capacity of iPSCs in vivo. When pluripotent stem cells are injected into immunodeficient mice, they give rise to a teratoma, a benign tumor comprising of cells from all three germ layers. Typically, teratomas are analyzed by performing histology sections and the pluripotency state is confirmed if the teratomas contain tissues from each of the three germ layers. Immuno-histochemistry can also be applied to the teratomas to identify different cell types. In 2015, the Benvenisty group developed an algorithm to quantitatively assess the teratomas: TeratoScore. ²⁶ The gene expression profile of the teratomas is compared with a set of references. This assay can be used to estimate the differentiation potential of each iPSC lines. The teratoma formation assay was considered the gold standard for pluripotency testing for many years. However, it is costly, time-consuming, and can be technically challenging depending on the injection site. In 2018, the ISCI group (International Stem Cell Initiative) reported their data on the comparison of the most commonly used methods to assess human pluripotent stem cells. ²⁷ They showed that the Pluritest can predict, but not determine, the pluripotency state of an iPSC line. In addition, random or directed in vitro differentiation is sufficient to confirm the pluripotent differentiation potential of the cells. However, only the teratoma analysis (by histology or TeratoScore) can provide an evaluation of both pluripotency and malignant potential. Therefore, they concluded that in vitro differentiation is adequate for the characterization of research grade iPSC lines, but teratoma formation is recommended for clinical grade iPSC lines.

    Transgene elimination

    Depending on the reprogramming method, the iPSC lines may need to be tested for the residual expression of the Yamanaka factors. The first human iPSC lines were generated using either a retroviral or a lentiviral system and the transgenes were integrated into the genome. Despite silencing of the provirus during the reprogramming process, it was not uncommon to detect residual expression of the transgenes in these iPSC lines. Moreover, the transgene expression could be reactivated during the differentiation of the iPSC lines, leading to tumor formation after transplantation into mice. ²⁸ However, with the non-integrative reprogramming methods, it is possible to obtain lines that are completely free of transgene expression. If the episomal vector method is used for reprogramming, it is important to test any potential integration of parts of the plasmids into the genome. Although the episomal vector replicates in the cytoplasm, we showed that 30% of the iPSC lines tested retained part of the plasmids over time. ¹⁷ A quantitative PCR for the EBNA sequence (found in all the episomal vectors) offers a simple screening method for determining if iPSC lines are free of the vectors. If the iPSC lines are generated using the Sendai virus method, the RNA virus will remain in the newly derived iPSC lines for several passages. The virus will be eliminated over time as the viral RNA polymerase protein is progressively diluted. The RNA virus can be monitored by quantitative RT-PCR using a probe specific for the Sendai virus sequence. We showed that the majority of the iPSC lines are free of transgene expression by passage 10. To accelerate the process, picking individual colonies during the first few passages and screening for SeV negative clones could be performed, but it would be a labor intensive process. Overall, the slow clearance of the SeV RNA is the main shortcoming of the Sendai virus method. The mRNA transfection reprogramming method generates iPSC lines that are instantly transgene free, eliminating the need for additional screening of the iPSC lines. However, as mentioned before, this method is labor intensive and it is not amenable to the reprogramming of blood cells.

    Identity

    To ensure that there was no swapping or mixing of iPSC lines during the reprogramming process or expansion of cell lines, a simple DNA fingerprinting test on genomic DNA extracted from both the iPSC lines and the somatic cells of origin should be performed. Both the UK stem cell bank and Wicell reported that some of the lines that were deposited in their banks were not the expected lines or were mixed with other cells. A short tandem repeat (STR) analysis, looking at the length of tandem repeats at 15 loci, is sufficient to differentiate lines from each donor. It is easy, inexpensive, and highly recommended for laboratories and facilities that are keeping multiple cell lines in culture at one time to perform STR analysis routinely or at strategic time-points, such as immediately after iPSC derivation, before freezing a new stock, or before starting a new experiment.

    Genetic stability

    The evidence that human pluripotent stem cells acquire genetic alterations during the reprogramming process, or the subsequent expansion of the lines, is well documented. ²⁹ , ³⁰ The most common chromosomal abnormalities are gains of whole chromosomes X and 8, gains of the long q arm of chromosomes 1, 17, and 20, and gains of the short p arm of chromosome 12.³¹ These recurrent abnormalities provide a selective advantage by increasing the proliferation or survival of the cells. Interestingly, the relative frequencies of these alterations have changed over the years as culture conditions of the cells were modified (ISSCR poster presentation from Erik McIntire). Gains of the short arm of chromosome 12 were the most prevalent around 2010, while gains of chromosome 1 and 20 have increased recently, becoming the most predominant chromosomal alterations detected in human iPSCs now. Although chromosomal gains are more common, partial loss of chromosomes 7, 10, 18, and 22 are also recurrent in human iPSC cultures. Additionally, translocations involving chromosome 1 or chromosome 20 have been reported in multiple ESC and iPSC lines. G-Banding karyotyping is the most commonly used method to detect genetic instability in iPSC lines. However, the resolution of this method is only around 5Mb. Higher resolution can be achieved using other methods, such as high-resolution array comparative genomic hybridization (aCGH) or single nucleotide polymorphism (SNP) microarrays. These methods can detect gain or loss of 0.05–0.4Mb and the SNP microarray technology also allows for the detection of loss of heterozygosity (LOH). It was reported that small duplications of chromosome 20, undetectable by G-Banding, lead to additional copies of the gene BCL2L1 and increased expression of BCL-xL (the anti-apoptotic isoform of BCL2L1). ³¹ This genetic abnormality promotes cell survival upon loss of cell-to-cell contact. It is found at high frequency in genetically edited clones which have been through strong selective pressure during transfection and single cell cloning. Our group established a digital droplet PCR (ddPCR) assay to evaluate the copy number variation (CNV) of the gene BCL2L1 (manuscript in preparation) to facilitate the screening of genetically edited clones. Unlike G-banding karyotype or aCGH/SNP microarrays, ddPCR is fast and inexpensive. Similarly, other qPCR- or ddPCR-based methods have been developed to identify the most common gain or loss of chromosome in human iPSCs. The hPSC Genetic Analysis kit (StemCell Technologies) can detect eight of the most common genomic alterations in hPSCs by qPCR, including gains of chromosomes 12, 17, 20, and X. A thorough analysis of all the reported genetic alterations in hPSCs led to the development of the iCS-digital PSC test (StemGenomics). ³² It is ddPCR-based and can detect more than 90% of recurrent genomic abnormalities using 24 probes or 76% using 12 probes. Besides the recurrent gain and loss of genetic material, single point mutations in the p53 gene were demonstrated to confer a selective advantage in hPSCs. ³³ This was discovered by performing exome sequencing on hundreds of iPSC lines. This finding highlights the importance of carefully analyzing the genetic stability of hPSCs or their derivatives before their use in clinical trials. For cGMP iPSC lines, it is evident that G-banding karyotype, with its limited resolution, is not sufficient. In contrast, any mutation found through full genome sequencing or exome sequencing might be difficult to interpret, as a single mutation itself may not be tumorigenic. One approach to create safe cGMP PSC lines is to focus on oncogenes and tumor suppressor genes. Our group used the Oncopanel assay to detect CNV or SNV in 447 oncogenes. Based on this assay, we can eliminate iPSC lines with mutations that have a high risk of tumorigenic potential. Finally, since genetic alterations can occur at any point during in vitro culture, it is critical to verify the genome integrity of newly derived iPSC lines and to continuously monitor during expansion every 5 to 10 passages. ³⁴

    Best practices prior to differentiation

    The quality of starting iPSC culture is key for producing high quality neural culture. Differences in genetic, epigenetic, and pluripotent state can affect cell fate decision during differentiation and lead to variability between line-to-line and/or batch-to-batch. Therefore, when choosing an iPSC line for differentiation purposes, it is important to select a cell line that has as many documented characterizations (listed in the section above) as possible for quality control (QC). If the selected iPSC line comes from an external source, whether from a biorepository or a collaborator, it is important to keep cells in quarantine upon thawing and carry out mycoplasma test initially. Once proven negative for mycoplasma and other bacterial/fungal contaminants, the very first step is to expand iPSCs at low passage number to create a good number of cryovial stocks for cell banking, followed by iPSC line characterization discussed above. An example workflow is depicted in Fig. 1.1.

    Cell banking

    For good cell banking practices, a tiered banking system is recommended, including the initial token stocks, a master cell bank, and a working cell bank. ³⁵ In general, the number of vials for each tier depends on the anticipated level of use. Token stocks are cryovials created upon in-house iPSC derivation or external purchasing and it is recommended to create 3–5 cryovials per iPSC line at this level. For iPSCs derived in-house, all iPSC QC, except reprogramming factor clearance, should be done at this stage. For externally sourced iPSCs that may or may not come with certificate of analysis, it is essential to assess parameters not characterized previously, but it is also a good idea to confirm reported data to ensure cell quality. It is worth noting that iPSC quality can be variable even when purchased from prestigious biobanks. Once the cells pass QC, one of the token stocks can be thawed and expanded to create a master cell bank, usually around 10–20 vials. Similarly, one vial from the master cell bank is thawed and expanded to create a working cell bank, typically around 20–50 vials for academic research purposes and 100–400 vials for therapeutic development. At a minimum, sterility, cell identity, and genomic stability should be tested after each expansion. The clearance of exogenous reprogramming agents should be confirmed around passage 10, ideally before the expansion for working cell bank. It is also important to record post-thawing cell morphology, survival rate, and spontaneous differentiation level. For protection from accidental loss, token stocks and/or several vials of the master cell bank should be stored apart from the working cell bank, ideally in a separate liquid nitrogen storage unit off-site. ³⁶ ,

    Enjoying the preview?
    Page 1 of 1