Startseite Modelling phenotypes, variants and pathomechanisms of syndromic diseases in different systems
Artikel Open Access

Modelling phenotypes, variants and pathomechanisms of syndromic diseases in different systems

  • Anne Gregor

    Dr. rer. nat Anne Gregor

    ORCID logo
    und Christiane Zweier

    Prof. Dr. Dr. med. Christiane Zweier

    ORCID logo EMAIL logo
Veröffentlicht/Copyright: 6. Juni 2024
Veröffentlichen auch Sie bei De Gruyter Brill

Abstract

In this review we describe different model organisms and systems that are commonly used to study syndromic disorders. Different use cases in modeling diseases, underlying pathomechanisms and specific effects of certain variants are elucidated. We also highlight advantages and limitations of different systems. Models discussed include budding yeast, the nematode worm, the fruit fly, the frog, zebrafish, mice and human cell-based systems.

Introduction

With the recent advances of sequencing technologies and therefore growing numbers of identified novel diseases and (candidate) disease genes, also the utilization of suitable models has become more and more important. Reasons and aims for utilizing model systems or organisms in syndromic diseases include the following: a) to obtain a better understanding of both the phenotypic expression and the underlying pathomechanisms as well as to characterize genotype-phenotype correlations, b) to confirm novel disease genes, c) to validate variants of unknown significance, d) to obtain platforms for studying potential therapeutic approaches in a pre-clinical setting. Depending on the disease and the research questions and aims, a variety of model systems and organisms is available. The following overview outlines commonly used models with a description of application areas, advantages and limitations.

Budding yeast (Saccharomyces cerevisiae)

The budding yeast Saccharomyces cerevisiae is a single-cell eukaryotic microorganism known as baker’s yeast. It is a widely used model organism in functional genetics studies as it is relatively cheap, easy to grow and to genetically manipulate due to its haploid status. Its smaller genome and fewer gene duplications allow studying biological and biochemical functions and pathways in a simplified system compared to complex eukaryotes, but limits testing of functions that are specific for higher order organisms [1,2]. 31.5 % of human rare disease genes are conserved in yeast [3].

S. cerevisiae has been successfully utilized to investigate pathomechanisms for various disease groups such as neurodegenerative disorders including Huntington’s and Alzheimer’s disease [1] or mitochondria-related disorders [4]. For example, by observing dysfunction of iron homeostasis and mitochondria upon knockdown of the yeast frataxin, important insights into its function and thus into the pathomechanisms of Friedreich ataxia were obtained [5].

Furthermore, S. cerevisiae can be a powerful tool to confirm pathogenicity of single candidate variants, as recently been demonstrated for a de novo missense variant in KIF21A in an individual with developmental delay, neurodegenerative decline, microcephaly and myelination abnormalities [6]. Loss of a conserved glutamate residue within the switch II motif of KIF21A (and yeast Cin8/Kip3) resulted in kinesin motor activity indicated by impaired yeast proliferation [6].

S. cerevisiae also represents an ideal first tier drug screening platform, particularly for disease groups with highly conserved pathways and mechanisms as for example for mitochondrial disorders [4]. In the so-called drug drop test, mutant yeast cells with defects in oxidative growth are spread on plates and covered with sterile filters containing different compounds. After several days, growth behaviour of mutant cells is compared to a positive control, indicating either rescue, toxic or lacking effects. This allows high throughput testing of >10 000 compounds and identified several rescuing substances in yeast models for different mitochondrial disorders as reviewed by [4].

Nematode worm (Caenorhabditis elegans)

C. elegans is a transparent, free-living soil nematode with a length of 1 mm and a diameter of approximately 80 µm. It is easy to grow on a bacterial diet and has a short life cycle. There are only 959 somatic cells in the adult hermaphrodite, 302 of which are neurons. Despite its rudimentary organization, many cell types with complex functions in mammals such as muscle cells, neurons, gut and excretory cells are present and identifiable in C. elegans [7]. Phenotypes that can be assessed include viability, locomotion, feeding, reproduction, responses to stimuli and learning and adaption [3]. 66 % of human rare disease genes are conserved in the worm [3]. Loss of function or haploinsufficiency of specific genes can be modeled by targeted knockdown via RNA interference approaches by feeding worms with dsRNA expressing bacteria [8]. More recently, CRISPR/Cas9 technology has been utilized for complete knockout or precise genome editing [9]. By gonadal microinjection of DNA, transgenic animals can be generated [10], either overexpressing a worm gene or ectopically a human gene. Such gain-of-function or toxicity models have been particularly used to investigate pathomechanisms of neurodegenerative diseases such as amyotrophic lateral sclerosis and Huntington’s, Alzheimer’s and Parkinson’s disease [11]. Apart from that, C. elegans has been proven to be a valuable model in investigating pathomechanisms of mitochondrial diseases [12], but also of other disorders such as fibrillinopathies [13], spinal muscular atrophy, polycystic kidney diseases and dystrophinopathies [7].

Furthermore, C. elegans can be utilized for rapid variant validation and for drug screening [14]. Approximately 13 % of variants in humans already have a corresponding variant in C. elegans [15]. Also genotype-phenotype correlations can be determined, as for example a missense variant in ion channel NALCN identified in a child with a particularly, severe lethal clinical presentation of CLIFAHDD syndrome (Congenital contractures of the limbs and face, hypotonia, and developmental delay, MIM: 616266) was shown to result in a gain-of-function effect in C. elegans with hypercontraction and uncoordinated movement. Other missense variants in NALCN identified in affected individuals with CLIFAHDD resulted not only in gain-of-function but also in loss-of-function effects [16]. Furthermore, variants in non-conserved amino-acids can be tested in whole-gene humanized animal models [14]. For this, the native gene is replaced by the human orthologue. If the human gene rescues the phenotype associated with loss of the worm gene, most human coding variants can be studies by assessing rescue effects compared to the wildtype. For example, human PTEN and KLC4 have been successfully introduced into C. elegans, respectively, and variant pathogenicity was investigated [17,18].

Fruit fly (Drosophila melanogaster)

Drosophila melanogaster has been one of the most important model organisms in biological and biomedical research within the last 100 years. It has a very short life cycle and is easy and inexpensive to maintain. Moreover, a huge variety of transgenic strains and tools for genetic manipulation are publicly available, and many established protocols for phenotype assessment exist [19]. 73.1 % of human rare disease genes are conserved in the fly [3].

Approaches to manipulate gene expression in a targeted way include the UAS/GAL4 system for ubiquitous or tissue specific knockdown or overexpression [20] as well as CRISPR/Cas9 related approaches for complete knockout or knock-in of specific variants [21]. Apart from investigating the effect of reduced (modeling loss of function and/or haploinsufficiency) or increased (modeling gain of function) dosage of a particular gene, there is a multitude of approaches to investigate the effect of specific missense variants. Patient-derived variants can be introduced in the fly ortholog if the amino acid is conserved, or rescue experiments can be performed with overexpressing the wildtype and mutant fly gene in a fly strain deficient for the endogenous gene. Variants affecting non-conserved amino acids can be assessed in humanized rescue approaches. For this, wildtype and mutant human gene constructs can be expressed in a fly deficient background to perform rescue experiments. Alternatively, these constructs can be overexpressed in an unaltered fly background to assess possible toxic effects. For example, overexpression of a human transgenic construct carrying one of five missense variants identified in PPFIA3 in individuals with a neurodevelopmental disorder showed that variants in the N-terminal coiled-coil domain result in stronger phenotypes than variants in the C-terminal region. Additionally, in contrast to the human wildtype, overexpression of three of the variant transgenic constructs failed to rescue embryonic lethality caused by homozygous loss of its orthologue, suggesting they are dominant negative loss-of-function alleles [22] Therefore, such experiments do not only allow conclusions on the pathogenicity of a particular variant but also indicate loss-of-function, gain-of-function or dominant negative effects [3,23].

Assessable and quantifiable phenotypes in Drosophila include amongst others viability, morphology (e.g. wings, eyes, bristles, brain, neurons, synapses) as well as habituation and complex learning and behavior. Drosophila is therefore an established and very broadly used model to investigate neurodevelopmental and neurodegenerative disorders. Its utilization has contributed to validating and obtaining pathomechanistic insights into numerous neurodevelopmental disorders (NDDs), both on gene and on variant level [24,25]. Using easily quantifiable phenotypes (e.g. wing morphology) as phenologs, Drosophila also allows systematic, large scale screens to characterize common molecular links between NDD genes [26] or to identify functional links between genes implicated in clinically overlapping disorders by genetic interaction experiments [27]. Apart from its prominent role in modelling neurodevelopmental disorders, Drosophila is also utilized as a model for e.g. kidney [28] or cardiac diseases [29] and mitochondrial disorders [30].

Drosophila also can serve as a platform for drug screening or identification. The first hint that mGluR antagonists might pharmacologically rescue Fragile-X syndrome associated symptoms came from a Drosophila study [31].

Frog (Xenopus tropicalis and Xenopus laevis)

The western clawed frog Xenopus tropicalis is a diploid species, while the South African clawed frog Xenopus laevis contains a duplicated set of genes and chromosomes. Xenopus is a rapid, cost effective, high-throughput vertebrate organism to model particularly developmental defects and congenital organ malformations [32]. 91.4 % of human rare disease genes are conserved in X. tropicalis [3]. One of the advantages of Xenopus is the easy accessibility to its eggs and embryos which allows intracellular microinjections and thus precise and organ specific manipulation up to the late tadpole stages. Unique among animal models is the possibility of one-sided injections with the contralateral side serving as an internal control [32]. Gene expression in Xenopus embryos can be manipulated by injecting mRNA for gain of function or morpholino oligonucleotides for loss of function, respectively. Furthermore, CRISPR/Cas9 has been used to induce either a complete knockout of a gene of interest or to induce specific genomic changes [33].

As early tadpoles are transparent, thus allowing assessment of organogenesis by light microscopy, Xenopus has been broadly used to model congenital cardiac defects, heterotaxy, primary ciliary dyskinesis and kidney defects [32]. For example, targeting two ciliary chondrodysplasia loci (ift80 and ift172) by CRISPR/Cas9 resulted in severe limb deformities, polydactyly and cystic kidney in froglets, closely matching the phenotype in humans with skeletal ciliopathies [34]. However, also other specific human phenotypes can be reproduced. For example, knockdown of the Xenopus ortholog of PYCR1 resulted in skin hypoplasia and blistering of the tadpole skin, thus resembling the human phenotype of cutis laxa caused by bi-allelic variants in PYCR1 [35]. Also complex developmental syndromes and neurocristopathies affecting several organs and presenting with craniofacial abnormalities such as CHARGE and Kabuki syndromes [36] or RASopathies [37] have been successfully modeled in Xenopus.

Zebrafish (Danio rerio)

Zebrafish are small, 3–4 cm long freshwater fish with a life span of 2 years and a generation time of 3 months. They present a vertebrate model with ex-utero development, making it easily feasible to observe all stages of development, especially given the fact that embryos are transparent, and all cells are visible from the outside until early larval stages. It is cost effective with limited space requirements for maintenance, and can produce large numbers of eggs [38]. Studies on embryos in the first 5 days post fertilization are not regulated, only after that zebrafish larvae will be considered experimental animals under Animal Welfare legislation [39]. Zebrafish has a high degree of genetic, anatomical and physiological similarity to humans with orthologs for 94.5 % of human rare disease genes [3,40]. The existence of an efficient genetic toolbox and the ease of genome manipulation through various techniques (morpholinos, ZF-nucleases, TALENs, CRISPR/Cas9) has also contributed to its success as a model organism [41].

Zebrafish has been widely used to study effects of loss-of-gene-function either through morpholino studies or more recently through knockouts using CRISPR/Cas9 genome editing. Additionally, effects of specific variants can be tested either through overexpression of human wildtype or mutant versions for the gene of interest in wildtype fish or through rescue experiments upon gene knockdown/out using morpholinos or CRISPR/Cas9 and simultaneous co-injection of the wildtype and mutant human gene [42]. Studies can be both performed during early embryonic development and in adult fish. A wide variety of organ systems have been studied making zebrafish a very suitable model to study syndromic disorders. Studied diseases include epilepsy, behavioural anomalies, kidney, heart defects, skeletal malformations and retinal and hearing defects. Zebrafish has a high capability for tissue regeneration, which has been especially studied in the heart. Ease of substance administration through the water has also allowed for drug screenings in zebrafish, at least for water soluble substances [43].

A prototype example of disease modelling in zebrafish are ciliopathies and especially Joubert syndrome, a multisystem disorder with common symptoms, among others being cerebellar and brainstem malformations, retinal degeneration, cystic kidney disease and polydactyly [44–46]. A high level of conservation is present both for specialized cilia and morphologically for the retina [47], the pronephros as a simplified human nephron [48] and bones [49]. Knockouts of many ciliopathy-related genes in zebrafish display cystic kidneys, retinal dystrophy and spinal curvature in larvae [50] and also show functional consequences of knockouts at the organismal level such as impaired visual function assessed in the oculo-kinetic response [51]. Several available tools furthermore have allowed to study mechanisms of how different ciliary genes work and have provided crucial insights into ciliary function and dysfunction [50].

As with any model system, it does come with several limitations. Zebrafish has many duplicated paralogs due to an early genome duplication [40], which can create difficulties in generating full gene knockouts, as it often requires targeting of multiple paralogs in parallel [52]. Generation of genetic knock-ins using CRISPR/Cas9 genome editing still has a rather low efficiency in zebrafish, thereby hindering studies of specific mutations in zebrafish embryos and larvae and requiring at least F2 generations of fish [52]. While zebrafish is a good model for many organ systems, differences in brain anatomy limit its use in neuroscience [53]. However, more recently, functional similarities have been recognized, and zebrafish has been used to model for example behavioural abnormalities and epilepsies [54]. Furthermore, loss of fertility upon inbreeding of fish makes maintenance of mutants challenging, and only a limited number of mutant strains is deeply characterized and maintained.

Mouse (Mus musculus)

Mouse models are the most widely used mammalian model organisms in rare disorder research and are making up around 60 % of research animals to date [55]. They are small, relatively cost-effective to maintain with an efficient reproduction cycle and share many similarities in anatomy and physiology with humans with 97.8 % of rare human disease genes conserved [3,56]. Large scale projects have generated knockout mouse models for many human disease genes (e.g. international knockout mouse consortium) [57], and the international mouse phenotyping consortium has performed systematic phenotyping for a large number of mouse lines [58]. Additionally, genome editing techniques such as CRISPR/Cas9 have also allowed to generate mouse models for specific disease-associated variants [59].

Mouse models have been particularly useful to model limb- and skeletal malformations, often recapitulating the human phenotype well. Here, also models for non-coding variation e.g. affecting chromatin topology have provided striking insights into disease pathology of various human disorders [60]. Mouse models have also been frequently used to study neurodevelopmental syndromic disorders, such as for example Angelman syndrome, where genomic organization is conserved in the mouse and neurological phenotypes recapitulate human phenotypes [61]. Those mouse models have then also been used for pre-clinical studies of gene-replacement and antisense-oligonucleotide (ASO) therapies [62,63].

However, challenges remain regarding high variability between different mouse mutants for the same disorder in different inbred backgrounds [64], which can confound results and impair transfer of knowledge. Furthermore, mouse models do not necessarily mimic human disease pathology accurately [65], which has led cofounding results and contributed to a lack of translatability of pre-clinical results to human clinical trials, e.g. in Fragile-X syndrome [66]. Additionally, societal and regulatory pressures have called for reduction of animal experiments, wherever possible [67], also increasing the search for other, possibly human-related disease and pre-clinical models.

Other mammals

While mice present the most used rodent model, it has some limits regarding specific functions or organs. For some phenotypes that require a better reflection of the human situation, other mammalian models are utilized.

As many genetic mouse models of Huntington’s disease do not show the typical neurodegeneration known from human individuals, transgenic rats may be used. They either express a human/rat combined fragment of the HTT gene or the full-length HTT genomic sequence with 97 CAG/CAA repeats and all regulatory regions. These rats show an early or later onset progressive neurodegeneration mimicking the human disease course [68].

For modelling cardiac diseases such as long-QT syndrome, short-QT syndrome and hypertrophic cardiomyopathy, transgenic rabbits expressing human pathogenic variants are used. Their electrophysiological, mechanical and structural cardiac characteristics resemble the human situation better than small rodent mouse models [69]. For the generation of transgenic rabbits, pronuclear microinjection, the sleeping beauty transposon system and genome-editing methods such as zinc finger nuclease (ZNF), transcription activator-like effector nuclease (TALEN) and CRISPR/Cas9 are used [69].

Larger animals such as dogs, pigs and primates are only rarely used as models as they are less accessible to genetic manipulation and due to ethical and societal reasons. As an example, a genetically modified porcine model for Duchenne muscular dystrophy carriers presented with hyperCKemia, abnormal dystrophin expression patterns in skeletal and cardiac muscles, histopathological signs of muscle degeneration, myocardial lesions in adulthood and sporadic death [70]. Furthermore, “naturally” occurring variants in larger animals can reflect particular clinical aspects known from human individuals quite well, e.g. for connective tissue disorders. As reviewed in [71], dogs and cats with Ehlers-Danlos syndrome show thin and hyperextensible skin, and cattle with Marfan syndrome have aortic dilatation, ocular abnormalities and skeletal involvement.

Cell-based models

A variety of different cell-based systems has been used for functional analysis aiding variant classification, for understanding pathomechanisms and for preclinical drug design and testing. Especially human cell-based systems can display important advantages over animal models and can overcome challenges in non-translatability of results from animals to humans, especially in drug discovery efforts. The recent advances in the development of 3D cellular systems have made them a very attractive tool to study diseases not only on a cellular level, but also on an organ and even organismal level. Different systems with different levels of complexity are described below.

Immortalized cell lines

Testing of specific effects of (missense) variants via overexpression studies can be done largely independent of the affected organ systems and can therefore be performed in immortalized cell lines in the absence of disease relevant patient samples/tissue. Common cell lines used for such experiments are HEK293, HeLa, U2OS or neuroblastoma cell lines. Advantages of such approaches include that they are relatively quick and cost-effective, that cell lines are easy to manipulate with common transient transfection technologies, and that they often provide specific information regarding mechanistic effects of specific variants that are more difficult to obtain in some animal models. As an example, in epilepsy research, HEK293 cells are frequently used to heterologously express specific ion channels in the absence of endogenous channel activity to test effects of specific patient-related point mutations [72,73]. Using this approach, it could be deciphered how different variants in SCN2A that lead to different clinical phenotypes have different effects on SCN2A function [72]. Limitations of these overexpression studies include that modelling does not occur in the tissue of interest and therefore specific effects may be masked due to the lack of necessary context-specific co-factors or interaction partners. Additionally, careful selection of experimental controls is necessary as many variants may not express well and functional read-outs of assays may not be appropriate [74].

Patient-derived primary cells (fibroblasts)

Patient-derived fibroblasts represent a very suitable disease model and have been used for several disorders to uncover abrogated pathways or biochemical functions in a patient-related context and to aid prioritization and characterization of potential disease-related variants. Amendable disorders include metabolic and mitochondrial disorders, while diseases affecting specific tissues (e.g. neurodevelopmental disorders or neuromuscular disorders), may not be modelled well in fibroblasts. However, it has been shown for example for Rett syndrome, that abrogated biochemical properties are well conserved in fibroblasts [75]. Additionally, the potential to reprogram fibroblasts into hiPSCs opens many opportunities for their use to model disorders tissue-specifically, which is replacing experimental work on fibroblast models for various disorders, such as for example neurodevelopmental disorders (see below). Patient-derived fibroblasts have the advantage that they are relatively easy to obtain through a skin biopsy, and that they can be cultured, so that amount of material to be used in experiments is not strongly limited. Patient fibroblasts have for example been used as a source of RNA to combine RNA-Seq analysis with whole exome/genome sequencing to aid diagnostics. This has successfully been done for mitochondrial disorders, where expression changes due to deep-intronic splice variants improved identification pathogenic variants [76]. Choice of correct tissue to perform RNA-Seq analysis on may limit the use of fibroblasts for many disorders, e.g. neurodevelopmental or neuromuscular disorders, but has been implemented in clinical diagnostics [77]. Another way of aiding variant classification and providing evidence for variant pathogenicity from fibroblasts are rescue experiments. For this, following identification of a measurable phenotype in patient fibroblasts (e.g. enzymatic readout, defect in OXPHOS, biochemical or microscopic assay), rescue experiments can be performed with wildtype or mutant transgenes [2]. Depending on the assay, transgene expression is either performed through transient transfection for assays assessing single cells, or stable transduction using (lenti)-viral particles and subsequent antibiotic selection when high levels of transduction efficiency are required for the assay (e.g. enzymatic readouts). Possible limitations come from expression of the transgene from an exogenous promotor at non-physiological levels, which can either mask rescue effects if expression is too high or too low or can have detrimental effects on the cells in the long-term. One way to circumvent this issue can be the use of inducible promoters (e.g. tetracycline-responsive element containing promoters). This approach has for example been applied successfully multiple time for various mitochondrial or metabolic disorders [2]. Generally, it can be difficult to obtain proper, matched controls, especially when affected individuals are very young. It also has to be considered that mutational load of fibroblast samples can vary depending on the area of skin that biopsy was obtained from, which may impact assay results [78,79].

Figure 1:  Schematic overview over different model systems and tissues that are commonly modelled. Created with BioRender.com.
Figure 1:

Schematic overview over different model systems and tissues that are commonly modelled. Created with BioRender.com.

iPSC-based cellular models in 2D

With the development of somatic cell reprogramming technologies, it became feasible to reprogram patient cells into human pluripotent stem cells (hiPSC) [80]. HiPSCs can theoretically be reprogrammed from all actively dividing somatic cells, most commonly this is done using skin fibroblasts. Other starting tissues that have been used successfully include peripheral or umbilical cord blood cells, where CD34+ are isolated, keratinocytes from hair and tubular epithelial cells isolated from urine samples [81]. They can then be differentiated into different tissues of interest from all three germ layers and have been used among others to model different types of neurons, retinal cells, endothelial cells, intestinal cells, hepatocytes, cardiomyocytes, renal cells and lung epithelial cells [82].

Depending on the questions to be answered, it may be more suitable to either use patient-derived cells for reprogramming or to introduce patient-specific variants using CRISPR/Cas9 in control cells. Advantages of the use of patient-derived cells reprogrammed to hiPSCs include that the genetic landscape of the patient including potential disease modifiers is conserved. However, genomic instability and epigenetic memories during reprogramming [83] in addition to a lack of matching controls may increase variability in experiments due to different background, age, sex, ethnicity, or different reprogramming techniques [84]. In contrast, when introducing variants with CRISPR, isogenic not targeted controls are available from the same experiment, therefore less variability due to different backgrounds may be expected. A possible, but cost- and labour-intensive work-around is to use patient-derived hiPSCs and remove disease causing variants using CRISPR/Cas9 as isogenic experimental controls. Especially for modelling not strictly monogenic, oligogenic or complex disorders or disorders with reduced penetrance or variable expressivity, patient-derived models represent the full genetic complexity of the patient and are therefore ideally used.

Table 1:

Overview over different model systems

Model organism

Model for

Modelled disorders

Genetic manipulation

Yeast

– disease pathways and simple functions

– variant pathogenicity

– drug screening

– mitochondrial and neurodegenerative disorders

– Transformation

C. elegans

– disease pathways and simple functions

– variant pathogenicity

– neurodegenerative disorders

– RNA interference (soaking in or feeding dsRNA)

– Microinjection for transgene expression

– CRISPR for knockout (and specific variants)

Drosophila melanogaster

– development and function

– variant pathogenicity and effects

– pathomechanisms

– cardiac, kidney, nervous system diseases

– CRISPR for knockout (and specific variants)

– UAS-GAL4 system for knockdown (RNAi) and overexpression

Xenopus

– organogenesis

– heart, kidney diseases

– Morpholinos and mRNA injection for knockdown and overexpression

– CRISPR for knockout (and specific variants)

Zebrafish

– variant pathogenicity and effects

– Skeletal, eye, heart, kidney diseases

– CRISPR for knockout (and specific variants)

– Morpholinos for knockdown

– TALENs, transgene overexpression

Mouse

– phenotyping

– pathomechanisms

– Skeletal, brain, heart diseases

– CRISPR for knockout and specific variants

Cell-based modeling

– variant pathogenicity and effects

– pathomechanisms

– Eye, brain, heart, lung, intestine, kidney diseases

– CRISPR for knockout and specific variants

– Transient and stable transfection for knockdown and overexpression

For syndromic disorders, disease-relevant hiPSC lines can be differentiated into different lineages and tissues to model different aspects of disorders. Examples include hiPSC-based models for Williams-Beuren syndrome, where differentiations into neural precursor cells and neurons showed electrophysiological dysregulation [85], and differentiation into smooth muscle cells recapitulated the vascular phenotype observed in patients [86]. In hiPSC-derived disease models for Bardet-Biedl syndrome, defects in insulin and leptin signalling in neurons related to obesity observed in patients [87] and abrogated ciliary structures and signalling in retinal sheets as might be expected in a ciliopathy [88] were observed.

HiPSC models have also been used successfully to study effects of combinations of therapeutic compounds. For instance, in cystic fibrosis research various hiPSC models differentiated into lung epithelial cells, pancreatic ductal epithelial cells and intestinal cells reflecting three of the major affected organs exist and were used to test different therapeutic strategies [89–91].

Limitations of hiPSC 2D models include lack of organismal context as cells are usually cultured in a monoculture not representative of the complex interplay of different cell types and extracellular matrix in different organs [92], therefore, non-cell autonomous pathomechanisms cannot be assessed in 2D [93].

iPSC-based cellular models in 3D (organoids & organs on a chip)

To overcome limitations such as a lack of tissue context in 2D iPSC-based models, recent developments have made it possible to generate 3D models for many organs that have also allowed for important novel insights in modelling syndromic disorders, either through the formation of organoids or organs-on-a-chip [94]. Organoids are mostly self-organizing, grown in matrix and recapitulate some level of tissue architecture and function [95]. Organoids have been generated for several organs, including the brain, heart, lung, liver, pancreas, intestine, kidney and retina. For instance, cortical brain organoids were first used to model microcephaly and could show that premature neuronal differentiation was one of the hallmarks in these organoids [96]. In a model of renal phenotypes in Bardet-Biedl syndrome, normal morphology was observed in a 2D model, but spontaneous degeneration of tubular structures was found in an organoid model, highlighting the added value of 3D disease models to reflect pathogenesis [97].

Organs-on-a-chip are 3D microdevices that can be used to control tissue composition and can incorporate vascular perfusion [98]. Especially for organs like lung and heart, this can provide valuable information on physical forces that occur in living organs [94]. One example is the modelling of Barth syndrome associated cardiomyopathy, where mitochondrial dysfunction could be linked to specific contraction defects in an hiPSC-derived cardiomyocyte organ-on-a-chip and where it was shown that variants in the associated TAZ gene are necessary and sufficient to cause these defects through gene-replacement experiments [99].

Advantages of 3D model systems include that they recapitulate organ-level structure or function to some degree and can even be combined for multiple organs through generation of assembloids, e.g. fusion of cortical organoids and skeletal muscle spheroids to form 3D cortico-motor assembloids [100] or combinations of multiple organs-on-a-chip. Several challenges in the field remain such as high costs and labour-intensiveness, high heterogeneity and variability between different batches of organoids and even within batches, likely due to their self-organizing nature. Furthermore, organoid growth is often limited in size due to the lack of nutrients inside organoids once they reach a certain size, although many protocols are being developed to establish vascularized organoids to overcome this. Experimentally, it remains challenging to set up live, in-vivo experimental readouts from organoid cultures that are not endpoint measurements. For organs-on-a-chip, technical limitations remain in throughput, which remains low, and the need for manufacturing of custom-made chambers and devices [101].

With the use of patient-derived iPSCs, these models also present exciting new opportunities in disease modelling and as preclinical models in drug-testing overcoming some of the limitations in translatability that other non-human model systems present.


Affiliations

1Department of Human Genetics, Inselspital Bern, University of Bern, Bern, Switzerland

2Department for Biomedical Research (DBMR), University of Bern, Bern, Switzerland


About the authors

Dr. rer. nat Anne Gregor

Dr. rer. nat Anne Gregor

PRof. Dr. Dr. med. Christiane Zweier

Prof. Dr. Dr. med. Christiane Zweier

  1. Research funding: None declared.

  2. Author contributions: All authors have accepted responsibility for the entire content of this manuscript and approved its submission.

  3. Competing interests: The authors state no conflicts of interest.

  4. Informed consent: Informed consent was obtained from all individuals included in this study.

  5. Ethical approval: The local Institutional Review Board deemed the study exempt from review.

References

[1] Gastelum, S., Michael, A. F., et al. (2023) Saccharomyces cerevisiae as a research tool for RNA-mediated human disease. Wiley Interdiscip Rev RNA:e1814.10.1002/wrna.1814Suche in Google Scholar PubMed

[2] Rodenburg, R. J. (2018) The functional genomics laboratory: functional validation of genetic variants. J Inherit Metab Dis 41(3):297–307.10.1007/s10545-018-0146-7Suche in Google Scholar PubMed PubMed Central

[3] Yamamoto, S., Kanca, O., et al. (2024) Integrating non-mammalian model organisms in the diagnosis of rare genetic diseases in humans. Nat Rev Genet 25(1):46–60.10.1038/s41576-023-00633-6Suche in Google Scholar PubMed

[4] Magistrati, M., Gilea, A. I., et al. (2023) Drug Drop Test: How to Quickly Identify Potential Therapeutic Compounds for Mitochondrial Diseases Using Yeast Saccharomyces cerevisiae. Int J Mol Sci 24(13).10.3390/ijms241310696Suche in Google Scholar PubMed PubMed Central

[5] Babcock, M., de Silva, D., et al. (1997) Regulation of mitochondrial iron accumulation by Yfh1p, a putative homolog of frataxin. Science 276(5319):1709–1712.10.1126/science.276.5319.1709Suche in Google Scholar PubMed

[6] Bhola, P. T., Mishra, R., et al. (2024) Phenotypic heterogeneity associated with KIF21A: Two new cases and review of the literature. Am J Med Genet A 194(3):e63455.10.1002/ajmg.a.63455Suche in Google Scholar PubMed

[7] Culetto, E., and Sattelle, D. B. (2000) A role for Caenorhabditis elegans in understanding the function and interactions of human disease genes. Hum Mol Genet 9(6):869–877.10.1093/hmg/9.6.869Suche in Google Scholar PubMed

[8] Timmons, L., and Fire, A. (1998) Specific interference by ingested dsRNA. Nature 395(6705):854.10.1038/27579Suche in Google Scholar PubMed

[9] Dickinson, D. J., and Goldstein, B. (2016) CRISPR-Based Methods for Caenorhabditis elegans Genome Engineering. Genetics 202(3):885–901.10.1534/genetics.115.182162Suche in Google Scholar PubMed PubMed Central

[10] Mello, C. C., Kramer, J. M., et al. (1991) Efficient gene transfer in C.elegans: extrachromosomal maintenance and integration of transforming sequences. EMBO J 10(12):3959–3970.10.1002/j.1460-2075.1991.tb04966.xSuche in Google Scholar PubMed PubMed Central

[11] Wu, Y., Chen, Y., et al. (2023) Towards Understanding Neurodegenerative Diseases: Insights from Caenorhabditis elegans. Int J Mol Sci 25(1).10.3390/ijms25010443Suche in Google Scholar PubMed PubMed Central

[12] Onraet, T., and Zuryn, S. (2024) C. elegans as a model to study mitochondrial biology and disease. Semin Cell Dev Biol 154(Pt A):48–58.10.1016/j.semcdb.2023.04.006Suche in Google Scholar PubMed

[13] Summers, K. M. (2024) Genetic models of fibrillinopathies. Genetics 226(1).10.1093/genetics/iyad189Suche in Google Scholar PubMed PubMed Central

[14] Hopkins, C. E., Brock, T., et al. (2023) Phenotypic screening models for rapid diagnosis of genetic variants and discovery of personalized therapeutics. Mol Aspects Med 91:101153.10.1016/j.mam.2022.101153Suche in Google Scholar PubMed PubMed Central

[15] Pir, M. S., Bilgin, H. I., et al. (2022) ConVarT: a search engine for matching human genetic variants with variants from non-human species. Nucleic Acids Res 50(D1):D1172-D1178.10.1093/nar/gkab939Suche in Google Scholar PubMed PubMed Central

[16] Bend, E. G., Si, Y., et al. (2016) NALCN channelopathies: Distinguishing gain-of-function and loss-of-function mutations. Neurology 87(11):1131–1139.10.1212/WNL.0000000000003095Suche in Google Scholar PubMed PubMed Central

[17] Gumusderelioglu, S., Resch, L., et al. (2023) A humanized Caenorhabditis elegans model of hereditary spastic paraplegia-associated variants in KLC4. Dis Model Mech 16(8).10.1242/dmm.050076Suche in Google Scholar PubMed PubMed Central

[18] McDiarmid, T. A., Au, V., et al. (2018) CRISPR-Cas9 human gene replacement and phenomic characterization in Caenorhabditis elegans to understand the functional conservation of human genes and decipher variants of uncertain significance. Dis Model Mech 11(12).10.1101/369249Suche in Google Scholar

[19] Ugur, B., Chen, K., et al. (2016) Drosophila tools and assays for the study of human diseases. Dis Model Mech 9(3):235–244.10.1242/dmm.023762Suche in Google Scholar PubMed PubMed Central

[20] Brand, A. H., and Perrimon, N. (1993) Targeted gene expression as a means of altering cell fates and generating dominant phenotypes. Development 118(2):401–415.10.1242/dev.118.2.401Suche in Google Scholar PubMed

[21] Housden, B. E., and Perrimon, N. (2016) Cas9-Mediated Genome Engineering in Drosophila melanogaster. Cold Spring Harb Protoc 2016(9).10.1101/pdb.top086843Suche in Google Scholar PubMed

[22] Paul, M. S., Michener, S. L., et al. (2024) A syndromic neurodevelopmental disorder caused by rare variants in PPFIA3. Am J Hum Genet 111(1):96–118.10.1016/j.ajhg.2023.12.004Suche in Google Scholar PubMed PubMed Central

[23] Bellen, H. J., Wangler, M. F., et al. (2019) The fruit fly at the interface of diagnosis and pathogenic mechanisms of rare and common human diseases. Hum Mol Genet 28(R2):R207-R214.10.1093/hmg/ddz135Suche in Google Scholar PubMed PubMed Central

[24] Coll-Tane, M., Krebbers, A., et al. (2019) Intellectual disability and autism spectrum disorders ‘on the fly’: insights from Drosophila. Dis Model Mech 12(5).10.1242/dmm.039180Suche in Google Scholar PubMed PubMed Central

[25] Link, N., and Bellen, H. J. (2020) Using Drosophila to drive the diagnosis and understand the mechanisms of rare human diseases. Development 147(21).10.1242/dev.191411Suche in Google Scholar PubMed PubMed Central

[26] Kochinke, K., Zweier, C., et al. (2016) Systematic Phenomics Analysis Deconvolutes Genes Mutated in Intellectual Disability into Biologically Coherent Modules. Am J Hum Genet 98(1):149–164.10.1016/j.ajhg.2015.11.024Suche in Google Scholar PubMed PubMed Central

[27] Straub, J., Gregor, A., et al. (2020) Genetic interaction screen for severe neurodevelopmental disorders reveals a functional link between Ube3a and Mef2 in Drosophila melanogaster. Sci Rep 10(1):1204.10.1038/s41598-020-58182-5Suche in Google Scholar PubMed PubMed Central

[28] Dow, J. A. T., Simons, M., et al. (2022) Drosophila melanogaster: a simple genetic model of kidney structure, function and disease. Nat Rev Nephrol 18(7):417–434.10.1038/s41581-022-00561-4Suche in Google Scholar PubMed

[29] Souidi, A., and Jagla, K. (2021) Drosophila Heart as a Model for Cardiac Development and Diseases. Cells 10(11).10.3390/cells10113078Suche in Google Scholar PubMed PubMed Central

[30] Foriel, S., Willems, P., et al. (2015) Mitochondrial diseases: Drosophila melanogaster as a model to evaluate potential therapeutics. Int J Biochem Cell Biol 63:60–65.10.1016/j.biocel.2015.01.024Suche in Google Scholar PubMed

[31] McBride, S. M., Choi, C. H., et al. (2005) Pharmacological rescue of synaptic plasticity, courtship behavior, and mushroom body defects in a Drosophila model of fragile X syndrome. Neuron 45(5):753–764.10.1016/j.neuron.2005.01.038Suche in Google Scholar PubMed

[32] Blum, M., and Ott, T. (2018) Xenopus: An Undervalued Model Organism to Study and Model Human Genetic Disease. Cells Tissues Organs 205(5–6):303–313.10.1159/000490898Suche in Google Scholar PubMed

[33] Kostiuk, V., and Khokha, M. K. (2021) Xenopus as a platform for discovery of genes relevant to human disease. Curr Top Dev Biol 145:277–312.10.1016/bs.ctdb.2021.03.005Suche in Google Scholar PubMed PubMed Central

[34] Getwan, M., Hoppmann, A., et al. (2021) Ttc30a affects tubulin modifications in a model for ciliary chondrodysplasia with polycystic kidney disease. Proc Natl Acad Sci U S A 118(39).10.1073/pnas.2106770118Suche in Google Scholar PubMed PubMed Central

[35] Reversade, B., Escande-Beillard, N., et al. (2009) Mutations in PYCR1 cause cutis laxa with progeroid features. Nat Genet 41(9):1016–1021.10.1038/ng.413Suche in Google Scholar PubMed

[36] Schwenty-Lara, J., Pauli, S., et al. (2021) Using Xenopus to analyze neurocristopathies like Kabuki syndrome. Genesis 59(1–2):e23404.10.1002/dvg.23404Suche in Google Scholar PubMed

[37] Patterson, V. L., and Burdine, R. D. (2020) Swimming toward solutions: Using fish and frogs as models for understanding RASopathies. Birth Defects Res 112(10):749–765.10.1002/bdr2.1707Suche in Google Scholar PubMed PubMed Central

[38] Santoriello, C., and Zon, L. I. (2012) Hooked! Modeling human disease in zebrafish. J Clin Invest 122(7):2337–2343.10.1172/JCI60434Suche in Google Scholar PubMed PubMed Central

[39] Crouzier, L., Richard, E. M., et al. (2021) Use of Zebrafish Models to Boost Research in Rare Genetic Diseases. Int J Mol Sci 22(24).10.3390/ijms222413356Suche in Google Scholar PubMed PubMed Central

[40] Howe, K., Clark, M. D., et al. (2013) The zebrafish reference genome sequence and its relationship to the human genome. Nature 496(7446):498–503.10.1038/nature12111Suche in Google Scholar PubMed PubMed Central

[41] Sertori, R., Trengove, M., et al. (2016) Genome editing in zebrafish: a practical overview. Brief Funct Genomics 15(4):322–330.10.1093/bfgp/elv051Suche in Google Scholar PubMed

[42] Davis, E. E., Frangakis, S., et al. (2014) Interpreting human genetic variation with in vivo zebrafish assays. Biochim Biophys Acta 1842(10):1960–1970.10.1016/j.bbadis.2014.05.024Suche in Google Scholar PubMed PubMed Central

[43] Patton, E. E., Zon, L. I., et al. (2021) Zebrafish disease models in drug discovery: from preclinical modelling to clinical trials. Nat Rev Drug Discov 20(8):611–628.10.1038/s41573-021-00210-8Suche in Google Scholar PubMed PubMed Central

[44] Bachmann-Gagescu, R., Dempsey, J. C., et al. (2015) Joubert syndrome: a model for untangling recessive disorders with extreme genetic heterogeneity. J Med Genet 52(8):514–522.10.1136/jmedgenet-2015-103087Suche in Google Scholar PubMed PubMed Central

[45] Mitchison, H. M., and Valente, E. M. (2017) Motile and non-motile cilia in human pathology: from function to phenotypes. J Pathol 241(2):294–309.10.1002/path.4843Suche in Google Scholar PubMed

[46] Romani, M., Micalizzi, A., et al. (2013) Joubert syndrome: congenital cerebellar ataxia with the molar tooth. Lancet Neurol 12(9):894–905.10.1016/S1474-4422(13)70136-4Suche in Google Scholar PubMed PubMed Central

[47] Bachmann-Gagescu, R., and Neuhauss, S. C. (2019) The photoreceptor cilium and its diseases. Curr Opin Genet Dev 56:22–33.10.1016/j.gde.2019.05.004Suche in Google Scholar PubMed

[48] Morales, E. E., and Wingert, R. A. (2017) Zebrafish as a Model of Kidney Disease. Results Probl Cell Differ 60:55–75.10.1007/978-3-319-51436-9_3Suche in Google Scholar PubMed

[49] Boswell, C. W., and Ciruna, B. (2017) Understanding Idiopathic Scoliosis: A New Zebrafish School of Thought. Trends Genet 33(3):183–196.10.1016/j.tig.2017.01.001Suche in Google Scholar PubMed

[50] Rusterholz, T. D. S., Hofmann, C., et al. (2022) Insights Gained From Zebrafish Models for the Ciliopathy Joubert Syndrome. Front Genet 13:939527.10.3389/fgene.2022.939527Suche in Google Scholar PubMed PubMed Central

[51] Fleisch, V. C., and Neuhauss, S. C. (2006) Visual behavior in zebrafish. Zebrafish 3(2):191–201.10.1089/zeb.2006.3.191Suche in Google Scholar PubMed

[52] Li, M., Zhao, L., et al. (2016) Zebrafish Genome Engineering Using the CRISPR-Cas9 System. Trends Genet 32(12):815–827.10.1016/j.tig.2016.10.005Suche in Google Scholar PubMed PubMed Central

[53] Meshalkina, D. A., Kysil, E. V., et al. (2017) Adult zebrafish in CNS disease modeling: a tank that’s half-full, not half-empty, and still filling. Lab Anim (NY) 46(10):378–387.10.1038/laban.1345Suche in Google Scholar PubMed

[54] Kalueff, A. V., Stewart, A. M., et al. (2014) Zebrafish as an emerging model for studying complex brain disorders. Trends Pharmacol Sci 35(2):63–75.10.1016/j.tips.2013.12.002Suche in Google Scholar PubMed PubMed Central

[55] Murillo-Cuesta, S., Artuch, R., et al. (2020) The Value of Mouse Models of Rare Diseases: A Spanish Experience. Front Genet 11:583932.10.3389/fgene.2020.583932Suche in Google Scholar PubMed PubMed Central

[56] Li, H., and Auwerx, J. (2020) Mouse Systems Genetics as a Prelude to Precision Medicine. Trends Genet 36(4):259–272.10.1016/j.tig.2020.01.004Suche in Google Scholar PubMed PubMed Central

[57] International Mouse Knockout, C., Collins, F. S., et al. (2007) A mouse for all reasons. Cell 128(1):9–13.10.1016/j.cell.2006.12.018Suche in Google Scholar PubMed

[58] Meehan, T. F., Conte, N., et al. (2017) Disease model discovery from 3328 gene knockouts by The International Mouse Phenotyping Consortium. Nat Genet 49(8):1231–1238.10.1038/ng.3901Suche in Google Scholar PubMed PubMed Central

[59] Hall, B., Cho, A., et al. (2018) Genome Editing in Mice Using CRISPR/Cas9 Technology. Curr Protoc Cell Biol 81(1):e57.10.1002/cpcb.57Suche in Google Scholar PubMed PubMed Central

[60] Lupianez, D. G., Kraft, K., et al. (2015) Disruptions of topological chromatin domains cause pathogenic rewiring of gene-enhancer interactions. Cell 161(5):1012–1025.10.1016/j.cell.2015.04.004Suche in Google Scholar PubMed PubMed Central

[61] Rotaru, D. C., Mientjes, E. J., et al. (2020) Angelman Syndrome: From Mouse Models to Therapy. Neuroscience 445:172–189.10.1016/j.neuroscience.2020.02.017Suche in Google Scholar PubMed

[62] Judson, M. C., Shyng, C., et al. (2021) Dual-isoform hUBE3A gene transfer improves behavioral and seizure outcomes in Angelman syndrome model mice. JCI Insight 6(20).10.1172/jci.insight.144712Suche in Google Scholar PubMed PubMed Central

[63] Lee, D., Chen, W., et al. (2023) Antisense oligonucleotide therapy rescues disturbed brain rhythms and sleep in juvenile and adult mouse models of Angelman syndrome. Elife 12.10.7554/eLife.81892Suche in Google Scholar PubMed PubMed Central

[64] Alcaraz, W. A., Chen, E., et al. (2011) Modifier genes and non-genetic factors reshape anatomical deficits in Zfp423-deficient mice. Hum Mol Genet 20(19):3822–3830.10.1093/hmg/ddr300Suche in Google Scholar PubMed PubMed Central

[65] Sayed, N., Liu, C., et al. (2016) Translation of Human-Induced Pluripotent Stem Cells: From Clinical Trial in a Dish to Precision Medicine. J Am Coll Cardiol 67(18):2161–2176.10.1016/j.jacc.2016.01.083Suche in Google Scholar PubMed PubMed Central

[66] Jacquemont, S., Berry-Kravis, E., et al. (2014) The challenges of clinical trials in fragile X syndrome. Psychopharmacology (Berl) 231(6):1237–1250.10.1007/s00213-013-3289-0Suche in Google Scholar PubMed PubMed Central

[67] Mocho, J. P. (2020) 3Rs – Reduce Reuse Recycle. Lab Anim 54(2):194.10.1177/0023677220909272Suche in Google Scholar PubMed

[68] Nittari, G., Roy, P., et al. (2023) Rodent Models of Huntington’s Disease: An Overview. Biomedicines 11(12).10.3390/biomedicines11123331Suche in Google Scholar PubMed PubMed Central

[69] Hornyik, T., Rieder, M., et al. (2022) Transgenic rabbit models for cardiac disease research. Br J Pharmacol 179(5):938–957.10.1111/bph.15484Suche in Google Scholar PubMed

[70] Okamoto, K., Matsunari, H., et al. (2023) Phenotypic features of genetically modified DMD-X(KO)X(WT) pigs. Regen Ther 24:451–458.10.1016/j.reth.2023.09.010Suche in Google Scholar PubMed PubMed Central

[71] Roberts, J. H., and Halper, J. (2021) Connective Tissue Disorders in Domestic Animals. Adv Exp Med Biol 1348:325–335.10.1007/978-3-030-80614-9_15Suche in Google Scholar PubMed

[72] Begemann, A., Acuna, M. A., et al. (2019) Further corroboration of distinct functional features in SCN2A variants causing intellectual disability or epileptic phenotypes. Mol Med 25(1):6.10.1186/s10020-019-0073-6Suche in Google Scholar PubMed PubMed Central

[73] Fang, Z., Xie, L., et al. (2022) Severe epilepsy phenotype with SCN1A missense variants located outside the sodium channel core region: Relationship between functional results and clinical phenotype. Seizure 101:109–116.10.1016/j.seizure.2022.07.018Suche in Google Scholar PubMed

[74] Adhish, M., and Manjubala, I. (2023) Effectiveness of zebrafish models in understanding human diseases-A review of models. Heliyon 9(3):e14557.10.1016/j.heliyon.2023.e14557Suche in Google Scholar PubMed PubMed Central

[75] Signorini, C., Leoncini, S., et al. (2014) Redox imbalance and morphological changes in skin fibroblasts in typical Rett syndrome. Oxid Med Cell Longev 2014:195935.10.1155/2014/195935Suche in Google Scholar PubMed PubMed Central

[76] Kremer, L. S., Bader, D. M., et al. (2017) Genetic diagnosis of Mendelian disorders via RNA sequencing. Nat Commun 8:15824.10.1038/ncomms15824Suche in Google Scholar PubMed PubMed Central

[77] Yepez, V. A., Gusic, M., et al. (2022) Clinical implementation of RNA sequencing for Mendelian disease diagnostics. Genome Med 14(1):38.10.1186/s13073-022-01019-9Suche in Google Scholar PubMed PubMed Central

[78] DeRoin, L., Cavalcante de Andrade Silva, M., et al. (2022) Feasibility and limitations of cultured skin fibroblasts for germline genetic testing in hematologic disorders. Hum Mutat 43(7):950–962.10.1002/humu.24374Suche in Google Scholar PubMed PubMed Central

[79] Rouhani, F. J., Zou, X., et al. (2022) Substantial somatic genomic variation and selection for BCOR mutations in human induced pluripotent stem cells. Nat Genet 54(9):1406–1416.10.1038/s41588-022-01147-3Suche in Google Scholar PubMed PubMed Central

[80] Takahashi, K., Tanabe, K., et al. (2007) Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell 131(5):861–872.10.1016/j.cell.2007.11.019Suche in Google Scholar PubMed

[81] Raab, S., Klingenstein, M., et al. (2014) A Comparative View on Human Somatic Cell Sources for iPSC Generation. Stem Cells Int 2014:768391.10.1155/2014/768391Suche in Google Scholar PubMed PubMed Central

[82] Tabar, V., and Studer, L. (2014) Pluripotent stem cells in regenerative medicine: challenges and recent progress. Nat Rev Genet 15(2):82–92.10.1038/nrg3563Suche in Google Scholar PubMed PubMed Central

[83] Tapia, N., and Scholer, H. R. (2016) Molecular Obstacles to Clinical Translation of iPSCs. Cell Stem Cell 19(3):298–309.10.1016/j.stem.2016.06.017Suche in Google Scholar PubMed

[84] Doss, M. X., and Sachinidis, A. (2019) Current Challenges of iPSC-Based Disease Modeling and Therapeutic Implications. Cells 8(5).10.3390/cells8050403Suche in Google Scholar PubMed PubMed Central

[85] Khattak, S., Brimble, E., et al. (2015) Human induced pluripotent stem cell derived neurons as a model for Williams-Beuren syndrome. Mol Brain 8(1):77.10.1186/s13041-015-0168-0Suche in Google Scholar PubMed PubMed Central

[86] Kinnear, C., Chang, W. Y., et al. (2013) Modeling and rescue of the vascular phenotype of Williams-Beuren syndrome in patient induced pluripotent stem cells. Stem Cells Transl Med 2(1):2–15.10.5966/sctm.2012-0054Suche in Google Scholar PubMed PubMed Central

[87] Wang, L., Liu, Y., et al. (2021) Bardet-Biedl syndrome proteins regulate intracellular signaling and neuronal function in patient-specific iPSC-derived neurons. J Clin Invest 131(8).10.1172/JCI146287Suche in Google Scholar PubMed PubMed Central

[88] Barabino, A., Flamier, A., et al. (2020) Deregulation of Neuro-Developmental Genes and Primary Cilium Cytoskeleton Anomalies in iPSC Retinal Sheets from Human Syndromic Ciliopathies. Stem Cell Reports 14(3):357–373.10.1016/j.stemcr.2020.02.005Suche in Google Scholar PubMed PubMed Central

[89] Berical, A., Lee, R. E., et al. (2022) A multimodal iPSC platform for cystic fibrosis drug testing. Nat Commun 13(1):4270.10.1038/s41467-022-31854-8Suche in Google Scholar PubMed PubMed Central

[90] Fleischer, A., Vallejo-Diez, S., et al. (2020) iPSC-Derived Intestinal Organoids from Cystic Fibrosis Patients Acquire CFTR Activity upon TALEN-Mediated Repair of the p.F508del Mutation. Mol Ther Methods Clin Dev 17:858–870.10.1016/j.omtm.2020.04.005Suche in Google Scholar PubMed PubMed Central

[91] Simsek, S., Zhou, T., et al. (2016) Modeling Cystic Fibrosis Using Pluripotent Stem Cell-Derived Human Pancreatic Ductal Epithelial Cells. Stem Cells Transl Med 5(5):572–579.10.5966/sctm.2015-0276Suche in Google Scholar PubMed PubMed Central

[92] Gattazzo, F., Urciuolo, A., et al. (2014) Extracellular matrix: a dynamic microenvironment for stem cell niche. Biochim Biophys Acta 1840(8):2506–2519.10.1016/j.bbagen.2014.01.010Suche in Google Scholar PubMed PubMed Central

[93] Passier, R., Orlova, V., et al. (2016) Complex Tissue and Disease Modeling using hiPSCs. Cell Stem Cell 18(3):309–321.10.1016/j.stem.2016.02.011Suche in Google Scholar PubMed

[94] Liu, C., Oikonomopoulos, A., et al. (2018) Modeling human diseases with induced pluripotent stem cells: from 2D to 3D and beyond. Development 145(5).10.1242/dev.156166Suche in Google Scholar PubMed PubMed Central

[95] Dutta, D., Heo, I., et al. (2017) Disease Modeling in Stem Cell-Derived 3D Organoid Systems. Trends Mol Med 23(5):393–410.10.1016/j.molmed.2017.02.007Suche in Google Scholar PubMed

[96] Lancaster, M. A., Renner, M., et al. (2013) Cerebral organoids model human brain development and microcephaly. Nature 501(7467):373–379.10.1038/nature12517Suche in Google Scholar PubMed PubMed Central

[97] Williams, J., Hurling, C., et al. (2023) Modelling renal defects in Bardet-Biedl syndrome patients using human iPS cells. Front Cell Dev Biol 11:1163825.10.3389/fcell.2023.1163825Suche in Google Scholar PubMed PubMed Central

[98] Takebe, T., Zhang, B., et al. (2017) Synergistic Engineering: Organoids Meet Organs-on-a-Chip. Cell Stem Cell 21(3):297–300.10.1016/j.stem.2017.08.016Suche in Google Scholar PubMed

[99] Wang, G., McCain, M. L., et al. (2014) Modeling the mitochondrial cardiomyopathy of Barth syndrome with induced pluripotent stem cell and heart-on-chip technologies. Nat Med 20(6):616–623.10.1038/nm.3545Suche in Google Scholar PubMed PubMed Central

[100] Andersen, J., Revah, O., et al. (2020) Generation of Functional Human 3D Cortico-Motor Assembloids. Cell 183(7):1913–1929 e1926.10.1016/j.cell.2020.11.017Suche in Google Scholar PubMed PubMed Central

[101] Goldrick, C., Guri, I., et al. (2023) 3D multicellular systems in disease modelling: From organoids to organ-on-chip. Front Cell Dev Biol 11:1083175.10.3389/fcell.2023.1083175Suche in Google Scholar PubMed PubMed Central

Published Online: 2024-06-06
Published in Print: 2024-06-05

© 2024 the author(s), published by De Gruyter.

This work is licensed under the Creative Commons Attribution 4.0 International License.

Heruntergeladen am 8.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/medgen-2024-2020/html
Button zum nach oben scrollen