Open access peer-reviewed chapter

Retinal Organoids over the Decade

Written By

Jing Yuan and Zi-Bing Jin

Submitted: 29 January 2022 Reviewed: 04 March 2022 Published: 16 April 2022

DOI: 10.5772/intechopen.104258

From the Edited Volume

Organoid Bioengineering - Advances, Applications and Challenges

Edited by Manash K. Paul

Chapter metrics overview

336 Chapter Downloads

View Full Metrics

Abstract

Retinal organoids (ROs) are 3D tissue structures derived from embryonic stem cells (ESCs) or induced pluripotent stem cells (iPSCs) in vitro, which characterize the structure and function of retina to a certain extent. Since 2011, mouse and human retinal organoids have been available, opening up new avenues for retinal development, disease and regeneration research. Over the decade, great progress has been made in the development of retinal organoids, which is reflected in the improvement of differentiation efficiency and development degree. At the same time, retinal organoids also show broad application prospects, which are widely used in the construction of disease models. On this basis, the mechanism of disease, drug screening and retinal regeneration therapy have been explored. Although retinal organoids have a bright future, the deficiency of their structure and function, the limitations of differentiation and culture, and the difference compared with embryonic retina still remain to be solved.

Keywords

  • retinal organoids
  • retinal differentiation
  • disease models
  • retinal degenerative diseases
  • transplant

1. Introduction

Located in the back of eyeball, the retina is a soft and transparent membrane attached to the inner surface of the choroid and forms part of the central nervous system. The retina can sense light stimuli, convert the light signals it receives into electrical signals, and then transmit them to the cerebral cortex through the optic nerve to form vision [1]. The retina is mainly composed of pigment epithelial cells, photoreceptor cells, bipolar cells, horizontal cells, amacrine cells, ganglion cells, and Müller glial cells [2]. Different neuron types form different layers, and the orderly arranged nuclei and synaptic regions are alternately arranged, forming a complex and orderly layered structure of the retina [3].

Our early research on the retina, derived only from human fetal retinal tissue [4, 5], faced significant challenges due to access difficulties and ethical issues [6]. Beyond that, most of what we know about the retina comes from studying animal retinas, but human and animal retinas differ in composition and function. For example, most mammals have only two types of cone photoreceptors that express S-opsin or M-opsin [7, 8], while humans have a third type that expresses L-opsin [9]; mice, the main subjects of our study, have a higher proportion of rods than humans [10], whose vision is determined by the density of cones in the macula and fovea [11, 12]. Therefore, it is of great significance to develop appropriate human retinal models to supplement animal models.

The establishment of human embryonic stem cell (ESC) lines [13] and the emergence of induced pluripotent stem cell (iPSC) technology [14] have turned our attention to cell research. Early 2D differentiation protocols used exogenous signaling molecules, Wnt antagonist DKK1 and bone morphogenetic protein (BMP) antagonist Noggin, to guide pluripotent stem cells to an anterior neural fate [15, 16, 17] and to differentiate into various types of retinal cells, including retinal pigment epithelial (RPE) cells, photoreceptors, and ganglion cells [18, 19, 20, 21, 22, 23, 24, 25]. However, 2D differentiation is far from interpreting retinal development in vivo. Retinal development and maturation are regulated by a series of interacting signal networks, such as factors secreted by RPE that promote photoreceptor maturation. Early retinal differentiation produced a single cell type [19, 20, 21, 22] and lacked the necessary interaction between cells. Therefore, we still need to find a more perfect model of human retina.

This breakthrough was achieved by constructing a 3D differentiation procedure. Through 3D differentiation, we can obtain retinal organoids that are highly reducible to the development process and complex structure of the retina, which we vividly call it “retina in a dish.” In this chapter, we review the development of retinal organoids and show their application in today’s life science research.

Advertisement

2. Overview of retinal organoids

In 2011, Sassi’s team used mouse embryonic stem cells (mESCs) to construct the first true retinal organoid through a 3D differentiation procedure [26]. In the following year, human retinal organoids were created [27], which is of epoch-making significance, meaning that human research on retinal development and retinal diseases has entered a new stage, and retinal organoids also provide a new and most potential tool for the treatment of retinal degeneration diseases.

During neurogenesis in vertebrates, the development of the retina can be roughly divided into two stages, the appearance of the optic cup structure and the orderly differentiation of seven types of retinal cells. In the first stage, the forebrain splits to form two secondary brain vesicles: telencephalon and diencephalon. In the diencephalon, eye field region first bulges outward to form the optic vesicle, and the distal vesicle invaginates to form the double-layer optic cup, which further develops into the outer retinal pigment epithelium and the inner neural retina (NR) (Figure 1a) [28, 29, 30, 31, 32]. In the second stage, the inner pluripotent retinal progenitor cells (RPCs) sequentially differentiate into retinal ganglion cells (RGCs), cone photoreceptors, horizontal cells, amacrine cells, rod photoreceptors, bipolar cells and Müller glia cells (Figure 1b) [33]. The cone and rod are connected to the retinal pigment epithelium and together form the outer nuclear layer (ONL). After extending to the outer plexiform layer (OPL), they form synapses with bipolar cells and horizontal cells in the inner nuclear layer (INL). On the other side of the inner nuclear layer, bipolar cells, amacrine cells, and ganglion cells form the synaptic networks of the inner plexiform layer (IPL). Müller glial cells span the whole layer of the retina, from the retinal pigment epithelium to the ganglion cell layer (GCL) (Figure 1c) [34, 35].

Figure 1.

Overview of retina. (a) the first stage of retinal development: The formation of double—Layer optic cup structure. (b) the second stage of retinal development: Retinal progenitor cells (RPCs) differentiate into seven types of retinal cells. (c) Structure of the retina.

Retinal development in vivo is regulated by a series of transcription factors, signal transduction factors and cell surface factors. In vitro, differentiation of retinal organoids is also a programmed process that mimics development in vivo by adding various signaling molecules in stages. First, stem cells proliferate and aggregate (Figure 2A), inducing the formation of embryoid body (EB) (Figure 2B) and neuroepithelium (Figure 2C), which appear as translucent bright rings under a microscope (Figure 2D). And then, they develop into optic vesicles (Figure 2E), followed by neuroretinas (Figure 2F), which in turn differentiate into seven types of retinal cells. The sequence of retinal cell types is consistent with in vivo development [36]. After differentiation, the cells undergo spontaneous nuclear migration, forming pinnacles and finally arranged into layered structures, in which the ganglion cells are located in the inner layer of the retinal organoid and the photoreceptors are located in the outer layer of the retinal organoid [26, 27]. Since RPE is usually a mass of cells not adjacent to the neuroretina and is not derived from the floating culture of optic vesicles, we do not consider it to be part of the retinoid organoid in this paper. With the continuous development of differentiation technology, photoreceptors in organoids become more and more mature, which is manifested by the appearance of outer segments and photosensitivity [37, 38].

Figure 2.

The differentiation of retinal organoids. (A) Growing human embryonic stem cells (H9). (B) Embryoid bodies (EB) at day 9 of differentiation. (C) Neuroepithelium appear at day 12. (D) Neuroepithelium appear as translucent bright rings at day 12. (E) Optic vesicle/cup at day 21. (F) Neural retinal (NR) region at day 26. Scale bars: 1000 μm (A, D, and E); 400 μm (B and C); 200 μm (F). All photos are provided by Dr. Ze-Hua Xu.

Advertisement

3. Differentiation of retinal organoids

The development of retinal organoid technology is the result of continuous attempts and innovations by a large number of researchers. Here, we try to review the progress of retinal organoids differentiation in the past ten years (Figure 3).

Figure 3.

Progress in retinal organoid differentiation over the decade.

3.1 Diversity of differentiation methods

There are various differentiation methods for retinal organoids, but in terms of differentiation steps, there are mainly two differentiation schemes (Figure 4). The first is a classic 3D differentiation protocol from Sassi’s team [26, 27]. The stem cells were dissociated and reassembled in a serum-free and low-growth factor medium (SFEBq culture, or serum-free culture of embryoid body-like aggregates with quick aggregation), and forced to form an embryoid body (EB) in a 96-well V-shaped plate. They were then stimulated by the addition of Matrigel to differentiate into neuroepithelial cells and subsequently into retinal progenitor cells and double-layer optic cup structures [27]. The cells were in suspension culture during the whole process of differentiation, and the formation of optic cups and the differentiation of neuroretina were self-organized [27]. 3D differentiation protocol is complicated in the early stage of differentiation, but it has a higher degree of reduction in the retinal development process, including the occurrence of optic cups invagination, the appearance of ciliary marginal stem cells at the NR-RPE boundary [39], and the establishment of dorsal-ventral (D-V) polarity [40].

Figure 4.

Two main methods of retinal organoid differentiation.

The second differentiation method combines 2D culture and 3D culture (2D/3D) [37, 41, 42, 43, 44, 45, 46], and the difference is mainly reflected in the early stage of neural induction. It has been reported that pluripotent stem cells can differentiate into the retina even when they are simply fused together [41, 42]. In this differentiation scheme, the stem cells were divided into small pieces by enzymatic hydrolysis [37] or mechanical methods [41, 43, 45] to form aggregates. The aggregates were cultured on a plate coated with Matrigel or floated in medium in the form of lumps of Matrigel/PSCs [43, 45]. After it differentiated into neuroepithelium and optic vesicles, the latter were separated for suspension culture and further differentiated into retinal organoids. This approach bypasses EB formation stage and induces optic vesicle formation by endogenous production of inducer molecules from aggregated cells, avoiding the aggregation step of SFEBq method and the need of Wnt/BMP antagonist [47]. These studies suggest that cell-cell and cell-extracellular matrix interactions are key to inducing retinal organoids differentiation in the early stage of stem cell differentiation.

With the improvement of differentiation methods, the structure of retinal organoids has been improved. Photoreceptors can reach advanced maturity, characterized by the formation of the inner and outer segments and connecting cilia of photoreceptors, the appearance of photosensitivity [37, 44], the expression of photoreceptor neurotransmitters, and the formation of synaptic bands [38, 44]. By adjusting the differentiation method, we can also change the proportion of cells in organoids, such as retinal organoids rich in cones or RGCs [45, 46], which is good for cell transplantation. Oxygen is also an important factor in regulating the differentiation of retinal organoids, and hypoxic conditions (5%) effectively produce vesicles and cups as well as more mature neuroretinas [48]. Another study showed that high oxygen (40%) promoted the formation of NR in EB, as well as the generation, migration and maturation of retinal ganglion cells during metaphase differentiation [49]. The co-culture of RPE with retinal organoids promoted the differentiation of photoreceptors [50], while the co-culture with brain organoids promoted the axon extension of RGCs [51]. More encouragingly, researchers have differentiated human brain organoids with bilaterally symmetric vesicles [52].

3.2 Modulation of signaling molecules

Retinal development requires the regulation of a series of signaling molecules. Similarly, by adding different signaling molecules, retinal organoids differentiation can be regulated in vitro. Dickkopf-related protein 1 (DKK-1), a Wnt signaling pathway antagonist, salvages the self-organizing ability of stem cells to differentiate into retinal progenitor cells [53]. Insulin-like growth factor 1 (IGF-1) regulates the formation of retinal organoids and promotes the formation of the correct retinal lamellar structure by various retinal cells [54, 55]. In the absence of IGF-1, retinal lamination was absent at the early stage of differentiation, while photoreceptors decreased and retinal ganglion cells increased at the late stage of differentiation [55]. Addition of docosahexaenoic acid and fibroblast growth factor 1 can specifically promote the maturation of photoreceptors including cones [56]. Replacement of widely used all-trans retinoic acid with 9-cis-retinoic acid in culture medium promoted the expression of rod photoreceptors rhodopsin and the maturation of mitochondrial morphology [57, 58]. COCO protein can block BMP/TGFβ/Wnt signaling pathway, enhance photoreceptor precursors, and promote s-cone differentiation and inner segment protuberances formation [59, 60]. During retinal development, s-cone appear first, followed by L/M-cones. This time transition from the designation of the s-cone to the production of the L/M-cone is controlled by thyroid hormone (TH) signaling [61].

3.3 Combination of organoid technology and tissue engineering technology

There is also a lot of innovative research that combines retinal organoid technology with emerging materials technology. The use of bioreactors improved retinal stratification and increased the production of photoreceptors with cilia and new outer segments [62]. In static culture, the development of retinal organoids may be limited by oxygen and nutrient diffusion, and rotating-wall vessel (RWV) bioreactors can accelerate and improve the growth and differentiation of retinal organoids [63]. The spherical structure of retinal organoids limits its interaction with host RPE and the remaining neuroretinas during transplantation. In order to create a planar retinal organoid, a biodegradable scaffold was developed that mimics the extracellular matrix of neuroretinas [64]. Retina-on-a-chip is a new microphysiological model of the human retina that integrates seven different basic retinal cell types and provides vascular-like perfusion to retinal organoids [65]. Arrayed bottom-lined micropores composed of bionic hydrogels, facilitated rapid retinoid tissue formation from mESCs aggregates in an efficient and routine manner [66]. Automated microfluidic devices with significantly reduced shear stress can maintain the long-term survival of retinal organoids [67]. For details of some other differentiation improvements [68, 69, 70, 71], please refer to Figure 3.

Advertisement

4. Applications of retinal organoids

As a three-dimensional multicellular structure formed by self-organization in vitro, retinal organoids can reproduce the development process of retina in vivo to some extent, and can be used to summarize some structural and functional characteristics of human retina. Meanwhile, they are the most promising tools for retinal disease research (Figure 5).

Figure 5.

Applications of retinal organoids.

4.1 Retinal organoids as disease models

The reprogramming technique enables iPSCs-derived retinal organoids to retain the patient’s genetic characteristics, allowing us to study a variety of retinal diseases in detail in a dish. To date, retinitis pigmentosa (RP), Laber congenital amaurosis (LCA), retinoblastoma (RB) and some other retinal diseases (Table 1) have been reproduced in dishes using retinal organoid technology [47, 98].

Disease modelsGenetic mutationsPhenotypes of retinal organoidsReference
RPRp11(PRPF31 gene mutation)Gradually degenerating photoreceptors; disrupted cilia morphology[72]
RPRPGR gene frameshift mutationPhotoreceptors have significant defects in morphology, localization, transcription profile and electrophysiological activity; shorted cilium was found in patient retinal organoids[73]
RPRP17 structural variantsIncreased GDPD1 expression may lead to dysregulation of lipid metabolism, thus affecting photoreceptor function[74]
RPUSH2A gene mutationAberrant organoids polarization, defective neuroepithelium, and abnormal RPCs and photoceptors differentiation[75]
RPCrb1 gene mutationSmall but frequent disruptions of CRB complex members at the outer limiting membrane[76]
RPPDE6B gene mutationIncreased cGMP levels[77]
RPTRNT1 gene mutationReduced levels of full-length TRNT1 protein and expression of a truncated smaller protein; autophagy was defective, with abnormal accumulation of LC3-II and elevated oxidative stress levels[78]
RPUSH2A gene mutationPost-developmental photoreceptor degeneration[79]
RPREEP6 gene mutationThe expression of a retina-specific isoform REEP6.1 changes destabilize the protein[80]
XLRPRP2 gene mutationRod cell death peaked at day 150 and the outer nuclear layer thinned at day 180[81]
LCACRX gene mutationDefective photoreceptor maturation with diminished expression of visual opsins[82]
LCARPE65 gene mutationLower expression of RPE65, but similar phagocytic activity and VEGF secretion level[83]
LCAAIPL1 gene mutationPatient-derived organoids maintained retinal cell cytoarchitecture despite significantly reduced levels of AIPL1[84]
LCACEP290 gene mutationCilia defects were evident in photoreceptors[85]
LCACEP290 gene mutationA high level of aberrant splicing and cilia defects[86]
LCACRX gene mutationImmature and dysfunctional photoreceptor cells[87]
LCACEP290 gene mutationCEP290 aberrant splicing[88]
RBRB1 double allele deletionHomozygous deletion of RB1 did not affect the maturation and proliferation statuses of human iPSCs (no ROs phenotype)[89]
RBRB1 double allele mutation or deletionRb originated from ARR3-positive maturing cone precursors during development; tumorigenesis in retinal organoids[90]
RBRB1 germline mutationRetinoblastomas formed from retinal organoids have molecular, cellular and genomic features indistinguishable from human retinoblastomas[91]
S-cone syndromeNRL lossTwo distinct populations of s-opsin expressing photoreceptors emerge; one population more representative of typical cones, and the other of rod/cone intermediates[92]
Rod-cone dystrophyCRB1 gene mutationA novel CRB1 transcript displaying skipping of exon 6[93]
Macular telangiectasia type 2Normal organoids were treated with deoxyDead photoreceptors[94]
MicrophthalmiaVSX2(R200Q) gene mutationBipolar cells are absent; delayed photoreceptor maturation[95]
Stargardt diseaseABCA4 gene mutationABCA4 splicing defect[96]
X-linked juvenile cleft retina (XLRS)RS1 gene mutationRetinal splitting, defective retinoschisin production, outer-segment defects, abnormal paxillin turnover, and impaired ER-Golgi transportation[97]

Table 1.

Retinal organoids as disease models.

RP is a relatively common hereditary retinal disorder characterized by night blindness and progressive loss of visual field [99]. LCA, the main disease leading to congenital blindness in infants, accounts for more than 5% of hereditary retinopathy, with complete loss of binocular cone and rod function within 1 year after birth [100]. Both diseases have been reported to be associated with multiple pathogenic genes. By differentiating different genetically-mutated stem cell lines into retinal organoids [101], we can observe their disease phenotypes in dishes, including photoreceptor degeneration, ciliary morphology disorder, and various functional impairment at molecular levels. Retinoblastoma is the most common intraocular malignancy in children [102]. The main cause of retinoblastoma is the loss of RB1 gene expression [103]. RB1 gene is a tumor suppressor gene, but the mechanism of RB1 deletion leading to retinal cancer is not clear, one of the key questions is the origin of RB cancer. By constructing RB models based on retinal organoids [104], we successfully observed tumorigenesis in retinal organoids and demonstrated that RB originates from ARR3 positive precursors of mature cones during development [90]. Other disease models, such as s-cone syndrome, rod-cone dystrophy, Macular telangiectasia type 2, microphthalmia, Stargardt disease, X-linked juvenile cleft retina, have also contributed to our understanding of retinal diseases.

4.2 Retinal organoids as tools for therapeutic research

4.2.1 Gene therapy

Identification of pathogenic genes promotes the generation of animal models and elucidates the physiological functions of gene products to a certain extent, thus promoting the development of gene therapy. So far, most research has focused on saving retinal organoid disease phenotypes through gene editing of patient-specific induced pluripotent stem cells [72, 73, 87, 97, 105, 106]. However, this strategy cannot be applied to patients. Adeno-associated virus (AAV) show great promise as a gene therapy vector for a wide range of retinal diseases. For example, AAV-mediated gene augmentation has successfully treated LCA caused by RPE65 mutations [107]. AAV-mediated gene therapy based on retinal organoids has also shown promising results in the laboratory [81, 82, 108]. In addition, gene therapies such as antisense morpholino and antisense oligonucleotides (AONs) have also been reported (Table 2).

ROsGene therapyResultReference
CEP290-LCA-Optic CupsAntisense morpholinoEffectively blocked aberrant splicing and restored expression of full-length CEP290, restoring normal cilia-based protein trafficking[86]
RP11 (PRPF31)-RP-ROsCRISPR/Cas9-mediated gene correctionRescued protein expression and key cellular phenotypes in RPE and photoreceptors[72]
RPGR-RP-ROsCRISPR/Cas9-mediated gene correctionRescued photoreceptor structure and electrophysiological property, reversed the observed ciliopathy[73]
CEP290-LCA10-ROsAntisense oligonucleotidesRestored wild-type CEP290 mRNA and protein expression levels[88]
CEP290-Rd16-mROsAAV-mediated CEP290 fragmentImproved cilia phenotype[108]
RS1-XLRS-ROsCRISPR/Cas9-mediated base-editingNormalized the splitting phenotype, outer-segment defects, paxillin dynamics, ciliary marker expression, and transcriptome profiles[97]
RP2KO-XLRP-ROsAAV-mediated gene augmentationRescued the degeneration phenotype of the RP2 KO organoids, to prevent ONL thinning and restore rhodopsin expression[81]
ABCA4-Stargardt-ROsAntisense oligonucleotides(AONs)Saved the splicing defect[96]
G56R-ADRP-ROsCRISPR/Cas9 mediated gene knockoutROs differentiation and NR2E3 expression were normal[106]
CRX-LCA-ROsAAV-mediated CRX gene augmentation therapyPartially restored photoreceptor phenotype and expression of phototransduction-related genes; Reduced the loss of opsin expression[82]
CLN3-RP-ROsGene correctionRestored CLN3 mRNA and protein expression and prevented SCMAS and inner segment vacuolization[105]
CRX-LCA7-ROsCRISPR/Cas9-based gene knockoutRescued the photoreceptor phenotypes in organoids[87]

Table 2.

Gene therapy based on retinal organoids.

4.2.2 Cell replacement therapy

Hereditary retinal degenerative diseases such as RP, Stargardt’s disease and LCA are the leading cause of incurable blindness. The vision loss associated with these diseases is caused by the death of photoreceptors in the retina. Existing treatments, including neuroprotection and gene therapy, require the presence of endogenous photoreceptors. In addition, due to the complex mechanism of retinal degeneration diseases, especially RP, it has been found that there are multiple genes with multiple mutation modes, and treatment methods focusing on a single mutation are extremely difficult technically and economically. Thus, transplant-based photoreceptor cell replacement becomes an attractive therapeutic strategy for restoring visual function and, if successful, could be applied to a wide range of retinal degenerative diseases.

Research on retinal cell transplantation dates back to 2006 [109]. Mice were able to effectively integrate rod photoreceptor precursor cells isolated from juvenile mice retinas into the ONL. These cells can further differentiate in the host retina and exhibit morphological characteristics typical of mature photoreceptors, such as inner and outer segments, while expressing molecules necessary for light transduction, forming synaptic connections with downstream cells, generating light responses and promoting visual function [109, 110, 111, 112, 113, 114, 115]. These results demonstrate the feasibility of photoreceptor transplantation as a therapeutic strategy for restoring visual acuity after retinal degeneration.

However, this cannot be applied to the treatment of retinal diseases in humans. There are ethical challenges to primary photoreceptors transplantation, but stem cell-based photoreceptors can avoid this problem. It has been shown that photoreceptor cells derived from stem cells can be integrated into mouse retinas, restoring the animal’s response to light [116, 117]. This is far from enough, until the appearance of retinal organoids, retinal cell transplantation and clinical transformation have made a breakthrough.

Transplantation of retinal organoids, mainly photoreceptors, is also a process of constant exploration [118]. The safety and effectiveness of transplantation, the enrichment and purification of transplanted cells, the effects of retinal organoids at different stages of development and host retinas with different degrees of degeneration on the efficiency of transplantation, and the evaluation of cell integration and function after transplantation are all issues that need to be explored.

Table 3 gives a brief summary of some retinal organoid transplantation cases in recent years. There are two transplantation methods: one is to digest the retinal organ into a single cell, from which the photoreceptor cells are purified and enriched, and the transplantation is carried out in the form of cell suspension. Another method is to strip the photoreceptor layer from the retinal organ and transplant it in thin slices. This method is more difficult to operate, because it is difficult to maintain the correct shape and polarity of the retina when it is transplanted into the eye of the host, and appropriate transplantation instruments need to be designed. The implant may contain some interneurons that block the connection between photoreceptor cells and the remaining inner layer of the retina in the host, and there are eye size requirements in the host animal. The implant is usually performed in rats, cats and non-human primates. In general, we have gained a lot from the exploration of retinal organ transplantation. A number of studies have shown that transplanted cells or tissues can survive in the host eye for a long time, migrate and integrate into the correct location. Integrated cells can further differentiate and mature in vivo, presenting typical cell structures, such as internal and external segments, and expressing corresponding cell markers and synaptic markers. In some studies, the formation of synaptic connections between host and graft and improvement of host visual function were also observed. In the host, transplanted cells or tissues are electrically excitatory [136], indicating their potential for restoring visual function. Through behavioral and electrophysiological experiments, we found that the host can not only slow down the progressive visual loss to some extent, but also show a relatively significant recovery of visual function [127, 130, 131, 134, 135, 137, 138]. These are exciting results and suggest that cell replacement therapy based on retinal organoids is a promising treatment that will bring light to patients with retinal diseases.

Graft/HostTransplantation methodTransplantation resultReference
Rhodopsin-GFP-mESC-ROs-rod precursors/adult mice with retinal degenerationCell suspensionTransplanted cells integrate within the degenerated retinas of mice and mature into outer segment-bearing photoreceptors[119]
CRX-GFP-mESC-ROs-photoreceptors/adult NOD/SCID recipient miceCell suspensionAfter transplantation, the integrated cells showed typical mature rod structures with outer segments and banded synapses[120]
mESC or miPSC-derived 3D retinal tissue/advanced retinal degeneration model (rd1) that lacked ONLRetinal tissueA structured outer nuclear layer (ONL) with complete inner and outer segments was developed; host-graft synaptic connections were observed[121]
hESC-retina sheet/Two focal selective photoreceptor degeneration monkey modelsRetinal tissueTransplanted retinal tissue differentiated into a series of retinal cell types, including rod and cone photoreceptors that formed structured outer nuclear layers; formation of host-graft synaptic connections[122]
Rhodopsin-GFP-mESC-ROs-(CD73-MACS)-photoreceptors/mouse models with mild or severe cone-rod degenerationCell suspensionSome cells integrated into mouse retinas and acquired a mature morphology, expressing rod and synaptic markers in close proximity to secondary neurons[123]
Mesc-ROs-cone/adult Aipl1−/− miceCell suspensionTransplanted cells showed capacity to survive and mature in the subretinal space[124]
hPSC-ROs-L/M-opsin+ cones/Nrl−/−mice or advanced retinal degeneration miceCell suspensionHuman cones can become incorporated within an adult mammalian retina[44]
hiPSC-ROs(cGMP)-photoreceptors/immunodeficient mouseCell suspensionRetinal cells successfully integrated into the photoreceptor layer and developed into mature photoreceptors[125]
hESC-ROs-retinal sheets/mice of end-stage retinal degeneration with immunodeficiencyRetinal tissueLong-term survival and well-structured graft photoreceptor layer maturation without rejection or tumor formation; formation of host-graft synaptic connections[126]
hESC-ROs(30–65 days of differentiation)-retinal sheets/immunodeficient rho S334ter-3 ratsRetinal tissueThe transplanted sheets differentiated, integrated, and produced functional photoreceptors and other retinal cells; maturation of the transplanted retinal cells created visual improvements; the donor cells were synaptically active[127]
CRX-hiPSC-ROs-photoreceptors precursors (CD73-MACS)/P23H ratsCell suspensionCD73+ photoreceptor precursors can be isolated in large numbers and transplanted into rat eyes, showing capacity to survive and mature in close proximity to host inner retina (hiPSC-derived retinal cells did not appear to migrate to host ONL)[128]
CRX-hiPSC-ROs-photoreceptors precursors/Pde6brd1 miceCell suspensionThe CRX+ cells settled next to the inner nuclear layer and made connections with the inner neurons of the host retina, and approximately one-third of them expressed the pan cone marker, Arrestin 3, indicating further maturation upon integration into the host retina[129]
hESC-ROs-retinal progenitor cells (RPCs)/RD models of rats and miceCell suspensionTransplanted cells significantly improve vision and preserve the retinal structure[130]
hiPSC-ROs-Jaws-expressing PRs/blind mice lacking the photoreceptor layerCell suspensionLight-driven responses at the photoreceptor and ganglion cell levels were observed[131]
hiPSCs-ROs-Müller glia/rats depleted of retinal ganglion cells by NMDACell suspensionTransplanted cells can partially restore visual function[132]
hESC-ROs-retinal tissue/catRetinal tissueLarge number of graft-derived fibers connecting the graft and the host; presence of human-specific synaptophysin puncta in the cat retina[133]
hESC-Ros-retinal sheet/immunodeficient RCS ratsRetinal tissueThe transplanted organoids survived more than 7 months; developed photoreceptors with inner and outer segments, and other retinal cells; and were well-integrated within the host[134]
hiPSC-ROs-cones/mice with retinal degenerationCell suspensionRestoration of surprisingly complex light-evoked retinal ganglion cell responses and improved light-evoked behaviors in treated animals[135]
hiPSC-ROs-retinal sheets with PLGA scaffolds/rhesus monkeyRetinal tissueWith sufficient graft-host contact provided by the scaffold, the transplanted tissues survived for up to 1 year without tumorigenesis; Histological examinations indicated survival, further maturation, and migration[136]
mESC-ROs-retinal progenitor cells (RPC)/mice with retinal degenerationCell suspensionRPC grafts form active synaptic networks within sites of ADR that functionally integrate with the retinal neuron populations and that resemble physiological patterns of neural circuits to the normal retina[137]
ROs and polarized RPE sheets were made into a co-graft using bio-adhesives/immunodeficient RCS ratsTotal retina patchCo-grafts grew, generated new photoreceptors and developed neuronal processes that were integrated into the host retina[138]

Table 3.

Research on transplantation of retinal organoids.

4.3 Retinal organoids for drug screening

Drug development focuses on screening, a process that requires cell models. The closer the cell model was to the physiological state, the more accurate the screening was. Therefore, organoids are undoubtedly a better choice for drug screening. Some ocular supplements, vitamin E, lutein, astaxanthin, and anthocyanin, have been reported to protect retinal photoreceptor degeneration induced by 4-hydroxytamoxifen (4-OHT) and light [139]. Few studies have successfully screened effective drugs using retinal organoids. In addition, there are some studies using retinal organoids as screening tools to explore the membrane transport effects of some microbial opsin [140]. These results suggest that retinal organoids can be used to validate the effectiveness of some therapeutic products and drug molecules.

Advertisement

5. Limitations and deficiencies

In recent years, although the technology of retinal organoids has made great progress, it is still beyond the reach of our existing tools and technologies to construct retinal organoids with the same structure and physiological functions as the mature retina in vivo. Our research on retinal organoids is still in its infancy and there are some limitations to overcome.

Long-term maintenance of retinal organoids depends on oxygen and nutrients, and in our existing culture system, oxygen diffusion limits the size of retinal organoids and the development of their internal cells, especially ganglion cells. Currently, we are trying to introduce a combination of tissue engineering techniques to solve this problem, such as the use of bioreactors and retinal chips [62, 63, 65]. The absence of vascular system also limits the long-term maintenance of organoids. Microglia are the resident immune cells of the central nervous system and are particularly important for the development of the retina, which can regulate the survival of neurons and prune synapses [141]. Co-differentiation of retinal organoids and vascular tissues or microglia in a dish is challenging because they come from different germ layers. The retina develops from the ectoderm, while vascular tissue-associated cells and microglia originate from the mesoderm. Therefore, we usually choose to achieve the complexity of retinal organoids through co-culture. In recent years, the realization of vascular structure in human brain has made some progress. After transplantation of human cerebral organs into the cerebral cortex of mice, the growth of blood vessels in mice was induced to increase the survival and maturation of cells [142]. In vitro, a study found that ectopic expression of human ETS variant 2 (ETV2) in hESCs can form a complex vascular-like network in human cortical organs and promote the maturation of organoid function [143]. In the future, we also expect that 3D printing of vascular tissue [144] and co-culture with mesodermal progenitor cells [145] will make the differentiation system of retinal organoids more perfect. For the retina in a petri dish to function, the most important point is to establish synaptic connections and form functional circuits. While our differentiated retina can form synaptic connections now, it’s not nearly as good as the complex network of synapses in the retina in vivo. Even retinal organoids derived from normal stem cell lines respond poorly to light. This may be due to the gradual degeneration of ganglion cells during late differentiation and lack of connection to the brain, which hindrance our assessment of retinal functional circuits. It may also be associated with limited growth of the outer segment due to the lack of direct interaction with RPE. Retinal organoid technology has solved the problem of cell-cell interaction, but in organisms, tissue-to-tissue and organ-to-organ interactions remain important for development. For example, the relationship of the retina to the lens, ciliary body and cornea, and the relationship of the retina to the brain.

Advertisement

6. Conclusions

It’s an exciting time, and technological advances have made a lot of things possible. Retinal organoids are our research tools for overcoming retinal diseases. It allows us to further understand the development and maturation of the retina, reproduce disease pathology and phenotypes in vitro, and explore the feasibility of gene therapy. In addition, it provides us with cells for cell transplantation and drug screening. We have enjoyed the great benefits brought by retinal organoids. However, their defects and deficiencies are also prominent, which is the direction we need to work towards. There is still a long way to go in the development of retinal organoids, and we expect that technological breakthroughs will enable us to advance to the next level in this field in the future.

Advertisement

Acknowledgments

We thank Dr. Ze-Hua Xu for providing photos of retinal organoid differentiation.

References

  1. 1. Demb JB, Singer JH. Functional circuitry of the retina. Annual Review of Vision Science. 2015;1:263-289. DOI: 10.1146/annurev-vision-082114-035334
  2. 2. Reese BE. Development of the retina and optic pathway. Vision Research. 2011;51:613-632. DOI: 10.1016/j.visres.2010.07.010
  3. 3. Masland RH. The neuronal organization of the retina. Neuron. 2012;76:266-280. DOI: 10.1016/j.neuron.2012.10.002
  4. 4. Bumsted O’Brien KM, Cheng H, Jiang Y, Schulte D, Swaroop A, Hendrickson AE. Expression of photoreceptor-specific nuclear receptor NR2E3 in rod photoreceptors of fetal human retina. Investigative Ophthalmology & Visual Science. 2004;45:2807-2812. DOI: 10.1167/iovs.03-1317
  5. 5. McLeod DS, Hasegawa T, Prow T, Merges C, Lutty G. The initial fetal human retinal vasculature develops by vasculogenesis. Developmental Dynamics. 2006;235:3336-3347. DOI: 10.1002/dvdy.20988
  6. 6. Bell CM, Zack DJ, Berlinicke CA. Human organoids for the study of retinal development and disease. Annual Review of Vision Science. 2020;6:91-114. DOI: 10.1146/annurev-vision-121219-081855
  7. 7. Szatko KP, Korympidou MM, Ran Y, Berens P, Dalkara D, Schubert T, et al. Neural circuits in the mouse retina support color vision in the upper visual field. Nature Communications. 2020;11:3481. DOI: 10.1038/s41467-020-17113-8
  8. 8. Masuda T, Zhang X, Berlinicke C, Wan J, Yerrabelli A, Conner EA, et al. The transcription factor GTF2IRD1 regulates the topology and function of photoreceptors by modulating photoreceptor gene expression across the retina. The Journal of Neuroscience. 2014;34:15356-15368. DOI: 10.1523/JNEUROSCI.2089-14.2014
  9. 9. Ortin-Martinez A, Nadal-Nicolas FM, Jimenez-Lopez M, Alburquerque-Bejar JJ, Nieto-Lopez L, Garcia-Ayuso D, et al. Number and distribution of mouse retinal cone photoreceptors: Differences between an albino (Swiss) and a pigmented (C57/BL6) strain. PLoS One. 2014;9:e102392. DOI: 10.1371/journal.pone.0102392
  10. 10. Volland S, Esteve-Rudd J, Hoo J, Yee C, Williams DS. A comparison of some organizational characteristics of the mouse central retina and the human macula. PLoS One. 2015;10:e0125631. DOI: 10.1371/journal.pone.0125631
  11. 11. Peng YR, Shekhar K, Yan W, Herrmann D, Sappington A, Bryman GS, et al. Molecular classification and comparative Taxonomics of foveal and peripheral cells in primate retina. Cell. 2019;176(1222-1237):e1222. DOI: 10.1016/j.cell.2019.01.004
  12. 12. Provis JM, Dubis AM, Maddess T, Carroll J. Adaptation of the central retina for high acuity vision: Cones, the fovea and the avascular zone. Progress in Retinal and Eye Research. 2013;35:63-81. DOI: 10.1016/j.preteyeres.2013.01.005
  13. 13. Thomson JA, Itskovitz-Eldor J, Shapiro SS, Waknitz MA, Swiergiel JJ, Marshall VS, et al. Embryonic stem cell lines derived from human blastocysts. Science. 1998;282:1145-1147. DOI: 10.1126/science.282.5391.1145
  14. 14. Takahashi K, Tanabe K, Ohnuki M, Narita M, Ichisaka T, Tomoda K, et al. Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell. 2007;131:861-872. DOI: 10.1016/j.cell.2007.11.019
  15. 15. Banin E, Obolensky A, Idelson M, Hemo I, Reinhardtz E, Pikarsky E, et al. Retinal incorporation and differentiation of neural precursors derived from human embryonic stem cells. Stem Cells. 2006;24:246-257. DOI: 10.1634/stemcells.2005-0009
  16. 16. Lamba DA, Karl MO, Ware CB, Reh TA. Efficient generation of retinal progenitor cells from human embryonic stem cells. Proceedings of the National Academy of Sciences. 2006;103:12769-12774
  17. 17. Mukhopadhyay M, Shtrom S, Rodriguez-Esteban C, Chen L, Tsukui T, Gomer L, et al. Dickkopf1 is required for embryonic head induction and limb morphogenesis in the mouse. Developmental Cell. 2001;1:423-434
  18. 18. Zhou S, Flamier A, Abdouh M, Tetreault N, Barabino A, Wadhwa S, et al. Differentiation of human embryonic stem cells into cone photoreceptors through simultaneous inhibition of BMP. TGFbeta and Wnt Signaling. Development. 2015;142:3294-3306. DOI: 10.1242/dev.125385
  19. 19. Uygun BE, Sharma N, Yarmush M. Retinal pigment epithelium differentiation of stem cells: Current status and challenges. Critical Reviews in Biomedical Engineering. 2009;37:355-375. DOI: 10.1615/critrevbiomedeng.v37.i4-5.30
  20. 20. Sluch VM, Chamling X, Liu MM, Berlinicke CA, Cheng J, Mitchell KL, et al. Enhanced stem cell differentiation and Immunopurification of genome engineered human retinal ganglion cells. Stem Cells Translational Medicine. 2017;6:1972-1986. DOI: 10.1002/sctm.17-0059
  21. 21. Osakada F, Ikeda H, Mandai M, Wataya T, Watanabe K, Yoshimura N, et al. Toward the generation of rod and cone photoreceptors from mouse, monkey and human embryonic stem cells. Nature Biotechnology. 2008;26:215-224. DOI: 10.1038/nbt1384
  22. 22. Keller G. Embryonic stem cell differentiation: Emergence of a new era in biology and medicine. Genes & Development. 2005;19:1129-1155. DOI: 10.1101/gad.1303605
  23. 23. Jin Z-B, Takahashi M. Generation of retinal cells from pluripotent stem cells. Progress in Brain Research. 2012;201:171-181
  24. 24. Jin Z-B, Okamoto S, Mandai M, Takahashi M. Induced pluripotent stem cells for retinal degenerative diseases: A new perspective on the challenges. Journal of Genetics. 2009;88:417-424
  25. 25. Jin Z-B, Okamoto S, Osakada F, Homma K, Assawachananont J, Hirami Y, et al. Modeling retinal degeneration using patient-specific induced pluripotent stem cells. PloS One. 2011;6:e17084
  26. 26. Eiraku M, Takata N, Ishibashi H, Kawada M, Sakakura E, Okuda S, et al. Self-organizing optic-cup morphogenesis in three-dimensional culture. Nature. 2011;472:51-56. DOI: 10.1038/nature09941
  27. 27. Nakano T, Ando S, Takata N, Kawada M, Muguruma K, Sekiguchi K, et al. Self-formation of optic cups and storable stratified neural retina from human ESCs. Cell Stem Cell. 2012;10:771-785. DOI: 10.1016/j.stem.2012.05.009
  28. 28. Zuber ME, Gestri G, Viczian AS, Barsacchi G, Harris WA. Specification of the vertebrate eye by a network of eye field transcription factors. Development. 2003;130:5155-5167. DOI: 10.1242/dev.00723
  29. 29. O’Rahilly R, Muller F. Developmental stages in human embryos: Revised and new measurements. Cells, Tissues, Organs. 2010;192:73-84. DOI: 10.1159/000289817
  30. 30. Müller F, O’Rahilly RJA. The first appearance of the neural tube and optic primordium in the human embryo at stage. Anatomy and Embryology. 1985;172:157-169
  31. 31. Pearson AA. The development of the eyelids. Part I. External features. Journal of Anatomy. 1980;130:33
  32. 32. Perron M, Kanekar S, Vetter ML, Harris WA. The genetic sequence of retinal development in the ciliary margin of theXenopusEye. Developmental Biology. 1998;199:185-200
  33. 33. Cepko CL, Austin CP, Yang X, Alexiades M, Ezzeddine D. Cell fate determination in the vertebrate retina. Proceedings of the National Academy of Sciences. 1996;93:589-595
  34. 34. Blanks JC, Adinolfi AM, Lolley RN. Synaptogenesis in the photoreceptor terminal of the mouse retina. Journal of Comparative Neurology. 1974;156:81-93
  35. 35. Fisher LJ. Development of synaptic arrays in the inner plexiform layer of neonatal mouse retina. Journal of Comparative Neurology. 1979;187:359-372
  36. 36. O’Hara-Wright M, Gonzalez-Cordero A. Retinal organoids: A window into human retinal development. Development. 2020;147. DOI: 10.1242/dev.189746
  37. 37. Zhong X, Gutierrez C, Xue T, Hampton C, Vergara MN, Cao LH, et al. Generation of three-dimensional retinal tissue with functional photoreceptors from human iPSCs. Nature Communications. 2014;5:4047. DOI: 10.1038/ncomms5047
  38. 38. Wahlin KJ, Maruotti JA, Sripathi SR, Ball J, Angueyra JM, Kim C, et al. Photoreceptor outer segment-like structures in long-term 3D retinas from human pluripotent stem cells. Scientific Reports. 2017;7:766. DOI: 10.1038/s41598-017-00774-9
  39. 39. Kuwahara A, Ozone C, Nakano T, Saito K, Eiraku M, Sasai Y. Generation of a ciliary margin-like stem cell niche from self-organizing human retinal tissue. Nature Communications. 2015;6:6286. DOI: 10.1038/ncomms7286
  40. 40. Hasegawa Y, Takata N, Okuda S, Kawada M, Eiraku M, Sasai Y. Emergence of dorsal-ventral polarity in ES cell-derived retinal tissue. Development. 2016;143:3895-3906. DOI: 10.1242/dev.134601
  41. 41. Reichman S, Terray A, Slembrouck A, Nanteau C, Orieux G, Habeler W, et al. From confluent human iPS cells to self-forming neural retina and retinal pigmented epithelium. Proceedings of the National Academy of Sciences of the United States of America. 2014;111:8518-8523. DOI: 10.1073/pnas.1324212111
  42. 42. Phillips MJ, Wallace KA, Dickerson SJ, Miller MJ, Verhoeven AD, Martin JM, et al. Blood-derived human iPS cells generate optic vesicle-like structures with the capacity to form retinal laminae and develop synapses. Investigative Ophthalmology & Visual Science. 2012;53:2007-2019. DOI: 10.1167/iovs.11-9313
  43. 43. Lowe A, Harris R, Bhansali P, Cvekl A, Liu W. Intercellular adhesion-dependent cell survival and ROCK-regulated Actomyosin-driven forces mediate self-formation of a retinal organoid. Stem Cell Reports. 2016;6:743-756. DOI: 10.1016/j.stemcr.2016.03.011
  44. 44. Gonzalez-Cordero A, Kruczek K, Naeem A, Fernando M, Kloc M, Ribeiro J, et al. Recapitulation of human retinal development from human pluripotent stem cells generates transplantable populations of cone photoreceptors. Stem Cell Reports. 2017;9:820-837. DOI: 10.1016/j.stemcr.2017.07.022
  45. 45. Kim S, Lowe A, Dharmat R, Lee S, Owen LA, Wang J, et al. Generation, transcriptome profiling, and functional validation of cone-rich human retinal organoids. Proceedings of the National Academy of Sciences of the United States of America. 2019;116:10824-10833. DOI: 10.1073/pnas.1901572116
  46. 46. Rabesandratana O, Chaffiol A, Mialot A, Slembrouck-Brec A, Joffrois C, Nanteau C, et al. Generation of a transplantable population of human iPSC-derived retinal ganglion cells. Frontiers in Cell and Development Biology. 2020;8:585675. DOI: 10.3389/fcell.2020.585675
  47. 47. Zhang X, Wang W, Jin ZB. Retinal organoids as models for development and diseases. Cell Regeneration. 2021;10:33. DOI: 10.1186/s13619-021-00097-1
  48. 48. Chen HY, Kaya KD, Dong L, Swaroop A. Three-dimensional retinal organoids from mouse pluripotent stem cells mimic in vivo development with enhanced stratification and rod photoreceptor differentiation. Molecular Vision. 2016;22:1077
  49. 49. Gao L, Chen X, Zeng Y, Li Q , Zou T, Chen S, et al. Intermittent high oxygen influences the formation of neural retinal tissue from human embryonic stem cells. Scientific Reports. 2016;6:29944. DOI: 10.1038/srep29944
  50. 50. Akhtar T, Xie H, Khan MI, Zhao H, Bao J, Zhang M, et al. Accelerated photoreceptor differentiation of hiPSC-derived retinal organoids by contact co-culture with retinal pigment epithelium. Stem Cell Research. 2019;39:101491. DOI: 10.1016/j.scr.2019.101491
  51. 51. Fligor CM, Lavekar SS, Harkin J, Shields PK, VanderWall KB, Huang KC, et al. Extension of retinofugal projections in an assembled model of human pluripotent stem cell-derived organoids. Stem Cell Reports. 2021;16:2228-2241. DOI: 10.1016/j.stemcr.2021.05.009
  52. 52. Gabriel E, Albanna W, Pasquini G, Ramani A, Josipovic N, Mariappan A, et al. Human brain organoids assemble functionally integrated bilateral optic vesicles. Cell Stem Cell. 2021;28(1740-1757):e1748. DOI: 10.1016/j.stem.2021.07.010
  53. 53. Luo Z, Zhong X, Li K, Xie B, Liu Y, Ye M, et al. An optimized system for effective derivation of three-dimensional retinal tissue via Wnt signaling regulation. Stem Cells. 2018;36:1709-1722. DOI: 10.1002/stem.2890
  54. 54. Mellough CB, Collin J, Khazim M, White K, Sernagor E, Steel DH, et al. IGF-1 signaling plays an important role in the formation of three-dimensional laminated neural retina and other ocular structures from human embryonic stem cells. Stem Cells. 2015;33:2416-2430. DOI: 10.1002/stem.2023
  55. 55. Zerti D, Molina MM, Dorgau B, Mearns S, Bauer R, Al-Aama J, et al. IGFBPs mediate IGF-1’s functions in retinal lamination and photoreceptor development during pluripotent stem cell differentiation to retinal organoids. Stem Cells. 2021;39:458-466. DOI: 10.1002/stem.3331
  56. 56. Brooks MJ, Chen HY, Kelley RA, Mondal AK, Nagashima K, De Val N, et al. Improved retinal organoid differentiation by modulating signaling pathways revealed by comparative transcriptome analyses with development In vivo. Stem Cell Reports. 2019;13:891-905. DOI: 10.1016/j.stemcr.2019.09.009
  57. 57. Ueda K, Onishi A, Ito SI, Nakamura M, Takahashi M. Generation of three-dimensional retinal organoids expressing rhodopsin and S- and M-cone opsins from mouse stem cells. Biochemical and Biophysical Research Communications. 2018;495:2595-2601. DOI: 10.1016/j.bbrc.2017.12.092
  58. 58. Kaya KD, Chen HY, Brooks MJ, Kelley RA, Shimada H, Nagashima K, et al. Transcriptome-based molecular staging of human stem cell-derived retinal organoids uncovers accelerated photoreceptor differentiation by 9-cis retinal. Molecular Vision. 2019;25:663
  59. 59. Pan D, Xia XX, Zhou H, Jin SQ , Lu YY, Liu H, et al. COCO enhances the efficiency of photoreceptor precursor differentiation in early human embryonic stem cell-derived retinal organoids. Stem Cell Research & Therapy. 2020;11:366. DOI: 10.1186/s13287-020-01883-5
  60. 60. Zhang X, Jin ZB. Directed induction of retinal organoids from human pluripotent stem cells. Journal of Visualized Experiments. 2021;170:e62298. DOI: 10.3791/62298
  61. 61. Eldred KC, Hadyniak SE, Hussey KA, Brenerman B, Zhang PW, Chamling X, et al. Thyroid hormone signaling specifies cone subtypes in human retinal organoids. Science. 2018;362:eaau6348. DOI: 10.1126/science.aau6348
  62. 62. Ovando-Roche P, West EL, Branch MJ, Sampson RD, Fernando M, Munro P, et al. Use of bioreactors for culturing human retinal organoids improves photoreceptor yields. Stem Cell Research & Therapy. 2018;9:156. DOI: 10.1186/s13287-018-0907-0
  63. 63. DiStefano T, Chen HY, Panebianco C, Kaya KD, Brooks MJ, Gieser L, et al. Accelerated and improved differentiation of retinal organoids from pluripotent stem cells in Rotating-Wall vessel bioreactors. Stem Cell Reports. 2018;10:300-313. DOI: 10.1016/j.stemcr.2017.11.001
  64. 64. Singh D, Wang SB, Xia T, Tainsh L, Ghiassi-Nejad M, Xu T, et al. A biodegradable scaffold enhances differentiation of embryonic stem cells into a thick sheet of retinal cells. Biomaterials. 2018;154:158-168. DOI: 10.1016/j.biomaterials.2017.10.052
  65. 65. Achberger K, Probst C, Haderspeck J, Bolz S, Rogal J, Chuchuy J, et al. Merging organoid and organ-on-a-chip technology to generate complex multi-layer tissue models in a human retina-on-a-chip platform. eLife. 2019;8:e46188. DOI: 10.7554/eLife.46188
  66. 66. Decembrini S, Hoehnel S, Brandenberg N, Arsenijevic Y, Lutolf MP. Hydrogel-based milliwell arrays for standardized and scalable retinal organoid cultures. Scientific Reports. 2020;10:1-10. DOI: 10.1038/s41598-020-67012-7
  67. 67. Xue Y, Seiler MJ, Tang WC, Wang JY, Delgado J, McLelland BT, et al. Retinal organoids on-a-chip: A micro-millifluidic bioreactor for long-term organoid maintenance. Lab on a Chip. 2021;21:3361-3377. DOI: 10.1039/d1lc00011j
  68. 68. Eldred MK, Charlton-Perkins M, Muresan L, Harris WA. Self-organising aggregates of zebrafish retinal cells for investigating mechanisms of neural lamination. Development. 2017;144:1097-1106. DOI: 10.1242/dev.142760
  69. 69. Reichman S, Slembrouck A, Gagliardi G, Chaffiol A, Terray A, Nanteau C, et al. Generation of storable retinal organoids and retinal pigmented epithelium from adherent human iPS cells in Xeno-free and feeder-free conditions. Stem Cells. 2017;35:1176-1188. DOI: 10.1002/stem.2586
  70. 70. Kobayashi W, Onishi A, Tu HY, Takihara Y, Matsumura M, Tsujimoto K, et al. Culture Systems of Dissociated Mouse and Human Pluripotent Stem Cell-Derived Retinal Ganglion Cells Purified by two-step Immunopanning. Investigative Ophthalmology & Visual Science. 2018;59:776-787. DOI: 10.1167/iovs.17-22406
  71. 71. Zilova L, Weinhardt V, Tavhelidse T, Schlagheck C, Thumberger T, Wittbrodt J. Fish primary embryonic pluripotent cells assemble into retinal tissue mirroring in vivo early eye development. eLife. 2021;10. DOI: 10.7554/eLife.66998
  72. 72. Buskin A, Zhu L, Chichagova V, Basu B, Mozaffari-Jovin S, Dolan D, et al. Disrupted alternative splicing for genes implicated in splicing and ciliogenesis causes PRPF31 retinitis pigmentosa. Nature Communications. 2018;9:4234. DOI: 10.1038/s41467-018-06448-y
  73. 73. Deng WL, Gao ML, Lei XL, Lv JN, Zhao H, He KW, et al. Gene correction reverses ciliopathy and photoreceptor loss in iPSC-derived retinal organoids from retinitis Pigmentosa patients. Stem Cell Reports. 2018;10:1267-1281. DOI: 10.1016/j.stemcr.2018.02.003
  74. 74. de Bruijn SE, Fiorentino A, Ottaviani D, Fanucchi S, Melo US, Corral-Serrano JC, et al. Structural variants create new topological-associated domains and ectopic retinal enhancer-gene contact in dominant retinitis Pigmentosa. The American Journal of Human Genetics. 2020;107:802-814. DOI: 10.1016/j.ajhg.2020.09.002
  75. 75. Guo Y, Wang P, Ma JH, Cui Z, Yu Q , Liu S, et al. Modeling retinitis Pigmentosa: Retinal organoids generated from the iPSCs of a patient with the USH2A mutation show early developmental abnormalities. Frontiers in Cellular Neuroscience. 2019;13:361. DOI: 10.3389/fncel.2019.00361
  76. 76. Quinn PM, Buck TM, Mulder AA, Ohonin C, Alves CH, Vos RM, et al. Human iPSC-derived retinas recapitulate the fetal CRB1 CRB2 complex formation and demonstrate that photoreceptors and Muller glia are targets of AAV5. Stem Cell Reports. 2019;12:906-919. DOI: 10.1016/j.stemcr.2019.03.002
  77. 77. Gao ML, Lei XL, Han F, He KW, Jin SQ , Zhang YY, et al. Patient-specific retinal organoids recapitulate disease features of late-onset retinitis Pigmentosa. Frontiers in Cell and Development Biology. 2020;8:128. DOI: 10.3389/fcell.2020.00128
  78. 78. Sharma TP, Wiley LA, Whitmore SS, Anfinson KR, Cranston CM, Oppedal DJ, et al. Patient-specific induced pluripotent stem cells to evaluate the pathophysiology of TRNT1-associated retinitis pigmentosa. Stem Cell Research. 2017;21:58-70. DOI: 10.1016/j.scr.2017.03.005
  79. 79. Tucker BA, Mullins RF, Streb LM, Anfinson K, Eyestone ME, Kaalberg E, et al. Patient-specific iPSC-derived photoreceptor precursor cells as a means to investigate retinitis pigmentosa. eLife. 2013;2:e00824. DOI: 10.7554/eLife.00824
  80. 80. Arno G, Agrawal SA, Eblimit A, Bellingham J, Xu M, Wang F, et al. Mutations in REEP6 cause autosomal-recessive retinitis Pigmentosa. American Journal of Human Genetics. 2016;99:1305-1315. DOI: 10.1016/j.ajhg.2016.10.008
  81. 81. Lane A, Jovanovic K, Shortall C, Ottaviani D, Panes AB, Schwarz N, et al. Modeling and rescue of RP2 retinitis Pigmentosa using iPSC-derived retinal organoids. Stem Cell Reports. 2020;15:67-79. DOI: 10.1016/j.stemcr.2020.05.007
  82. 82. Kruczek K, Qu Z, Gentry J, Fadl BR, Gieser L, Hiriyanna S, et al. Gene therapy of dominant CRX-Leber congenital Amaurosis using patient stem cell-derived retinal organoids. Stem Cell Reports. 2021;16:252-263. DOI: 10.1016/j.stemcr.2020.12.018
  83. 83. Li G, Gao G, Wang P, Song X, Xu P, Xie B, et al. Generation and characterization of induced pluripotent stem cells and retinal organoids from a Leber’s congenital Amaurosis patient with novel RPE65 mutations. Frontiers in Molecular Neuroscience. 2019;12:212. DOI: 10.3389/fnmol.2019.00212
  84. 84. Lukovic D, Artero Castro A, Kaya KD, Munezero D, Gieser L, Davo-Martinez C, et al. Retinal organoids derived from hiPSCs of an AIPL1-LCA patient maintain Cytoarchitecture despite reduced levels of mutant AIPL1. Scientific Reports. 2020;10:5426. DOI: 10.1038/s41598-020-62047-2
  85. 85. Shimada H, Lu Q , Insinna-Kettenhofen C, Nagashima K, English MA, Semler EM, et al. In vitro modeling using ciliopathy-patient-derived cells reveals distinct cilia dysfunctions caused by CEP290 mutations. Cell Reports. 2017;20:384-396. DOI: 10.1016/j.celrep.2017.06.045
  86. 86. Parfitt DA, Lane A, Ramsden CM, Carr AF, Munro PM, Jovanovic K, et al. Identification and correction of mechanisms underlying inherited blindness in human iPSC-derived optic cups. Cell Stem Cell. 2016;18:769-781. DOI: 10.1016/j.stem.2016.03.021
  87. 87. Chirco KR, Chew S, Moore AT, Duncan JL, Lamba DA. Allele-specific gene editing to rescue dominant CRX-associated LCA7 phenotypes in a retinal organoid model. Stem Cell Reports. 2021;16:2690-2702. DOI: 10.1016/j.stemcr.2021.09.007
  88. 88. Dulla K, Aguila M, Lane A, Jovanovic K, Parfitt DA, Schulkens I, et al. Splice-modulating oligonucleotide QR-110 restores CEP290 mRNA and function in human c.2991+1655A>G LCA10 models. Molecular Therapy Nucleic Acids. 2018;12:730-740. DOI: 10.1016/j.omtn.2018.07.010
  89. 89. Deng X, Iwagawa T, Fukushima M, Watanabe S. Characterization of human-induced pluripotent stem cells carrying homozygous RB1 gene deletion. Genes to Cells. 2020;25:510-517. DOI: 10.1111/gtc.12771
  90. 90. Liu H, Zhang Y, Zhang YY, Li YP, Hua ZQ , Zhang CJ, et al. Human embryonic stem cell-derived organoid retinoblastoma reveals a cancerous origin. Proceedings of the National Academy of Sciences of the United States of America. 2020;117:33628-33638. DOI: 10.1073/pnas.2011780117
  91. 91. Norrie JL, Nityanandam A, Lai K, Chen X, Wilson M, Stewart E, et al. Retinoblastoma from human stem cell-derived retinal organoids. Nature Communications. 2021;12:4535. DOI: 10.1038/s41467-021-24781-7
  92. 92. Kallman A, Capowski EE, Wang J, Kaushik AM, Jansen AD, Edwards KL, et al. Investigating cone photoreceptor development using patient-derived NRL null retinal organoids. Communications Biology. 2020;3:82. DOI: 10.1038/s42003-020-0808-5
  93. 93. Zhang X, Thompson JA, Zhang D, Charng J, Arunachalam S, McLaren TL, et al. Characterization of CRB1 splicing in retinal organoids derived from a patient with adult-onset rod-cone dystrophy caused by the c.1892A>G and c.2548G>A variants. Molecular Genetics & Genomic Medicine. 2020;8:e1489. DOI: 10.1002/mgg3.1489
  94. 94. Gantner ML, Eade K, Wallace M, Handzlik MK, Fallon R, Trombley J, et al. Serine and lipid metabolism in macular disease and peripheral neuropathy. The New England Journal of Medicine. 2019;381:1422-1433. DOI: 10.1056/NEJMoa1815111
  95. 95. Phillips MJ, Perez ET, Martin JM, Reshel ST, Wallace KA, Capowski EE, et al. Modeling human retinal development with patient-specific induced pluripotent stem cells reveals multiple roles for visual system homeobox 2. Stem Cells. 2014;32:1480-1492. DOI: 10.1002/stem.1667
  96. 96. Khan M, Arno G, Fakin A, Parfitt DA, Dhooge PPA, Albert S, et al. Detailed phenotyping and therapeutic strategies for Intronic ABCA4 variants in Stargardt disease. Molecular Therapy Nucleic Acids. 2020;21:412-427. DOI: 10.1016/j.omtn.2020.06.007
  97. 97. Huang KC, Wang ML, Chen SJ, Kuo JC, Wang WJ, Nhi Nguyen PN, et al. Morphological and molecular defects in human three-dimensional retinal organoid model of X-linked juvenile Retinoschisis. Stem Cell Reports. 2019;13:906-923. DOI: 10.1016/j.stemcr.2019.09.010
  98. 98. Zhang XH, Jin ZB. Patient iPSC-derived retinal organoids: Observable retinal diseases in-a-dish. Histology and Histopathology. 2021;36:705-710. DOI: 10.14670/HH-18-307
  99. 99. Ran X, Cai WJ, Huang XF, Liu Q , Lu F, Qu J, et al. ‘RetinoGenetics’: A comprehensive mutation database for genes related to inherited retinal degeneration. Database: The Journal of Biological Databases and Curation. 2014;2014. DOI: 10.1093/database/bau047
  100. 100. Tsang SH, Sharma T. Leber congenital Amaurosis. In: Tsang SH, Sharma T, editors. Atlas of Inherited Retinal Diseases. Cham: Springer International Publishing; 2018. pp. 131-137
  101. 101. Li YP, Deng WL, Jin ZB. Modeling retinitis pigmentosa through patient-derived retinal organoids. STAR Protocols. 2021;2:100438. DOI: 10.1016/j.xpro.2021.100438
  102. 102. Dimaras H, Kimani K, Dimba EA, Gronsdahl P, White A, Chan HS, et al. Retinoblastoma. The Lancet. 2012;379:1436-1446
  103. 103. Friend SH, Bernards R, Rogelj S, Weinberg RA, Rapaport JM, Albert DM, et al. A human DNA segment with properties of the gene that predisposes to retinoblastoma and osteosarcoma. Nature. 1986;323:643-646
  104. 104. Liu H, Hua ZQ , Jin ZB. Modeling human retinoblastoma using embryonic stem cell-derived retinal organoids. STAR Protocols. 2021;2:100444. DOI: 10.1016/j.xpro.2021.100444
  105. 105. Zhang X, Zhang D, Thompson JA, Chen SC, Huang Z, Jennings L, et al. Gene correction of the CLN3 c.175G>a variant in patient-derived induced pluripotent stem cells prevents pathological changes in retinal organoids. Molecular Genetics & Genomic Medicine. 2021;9:e1601. DOI: 10.1002/mgg3.1601
  106. 106. Diakatou M, Dubois G, Erkilic N, Sanjurjo-Soriano C, Meunier I, Kalatzis V. Allele-specific knockout by CRISPR/Cas to treat autosomal dominant retinitis Pigmentosa caused by the G56R mutation in NR2E3. International Journal of Molecular Sciences. 2021;22:2607. DOI: 10.3390/ijms22052607
  107. 107. Russell S, Bennett J, Wellman JA, Chung DC, Yu Z-F, Tillman A, et al. Efficacy and safety of voretigene neparvovec (AAV2-hRPE65v2) in patients with RPE65-mediated inherited retinal dystrophy: A randomised, controlled, open-label, phase 3 trial. The Lancet. 2017;390:849-860. DOI: 10.1016/s0140-6736(17)31868-8
  108. 108. Mookherjee S, Chen HY, Isgrig K, Yu W, Hiriyanna S, Levron R, et al. A CEP290 C-terminal domain complements the mutant CEP290 of Rd16 mice In trans and rescues retinal degeneration. Cell Reports. 2018;25(611-623):e616. DOI: 10.1016/j.celrep.2018.09.043
  109. 109. MacLaren RE, Pearson RA, MacNeil A, Douglas RH, Salt TE, Akimoto M, et al. Retinal repair by transplantation of photoreceptor precursors. Nature. 2006;444:203-207. DOI: 10.1038/nature05161
  110. 110. Bartsch U, Oriyakhel W, Kenna PF, Linke S, Richard G, Petrowitz B, et al. Retinal cells integrate into the outer nuclear layer and differentiate into mature photoreceptors after subretinal transplantation into adult mice. Experimental Eye Research. 2008;86:691-700. DOI: 10.1016/j.exer.2008.01.018
  111. 111. Lakowski J, Baron M, Bainbridge J, Barber AC, Pearson RA, Ali RR, et al. Cone and rod photoreceptor transplantation in models of the childhood retinopathy Leber congenital amaurosis using flow-sorted Crx-positive donor cells. Human Molecular Genetics. 2010;19:4545-4559. DOI: 10.1093/hmg/ddq378
  112. 112. Eberle D, Schubert S, Postel K, Corbeil D, Ader M. Increased integration of transplanted CD73-positive photoreceptor precursors into adult mouse retina. Investigative Ophthalmology & Visual Science. 2011;52:6462-6471. DOI: 10.1167/iovs.11-7399
  113. 113. Pearson RA, Barber AC, Rizzi M, Hippert C, Xue T, West EL, et al. Restoration of vision after transplantation of photoreceptors. Nature. 2012;485:99-103. DOI: 10.1038/nature10997
  114. 114. Barber AC, Hippert C, Duran Y, West EL, Bainbridge JW, Warre-Cornish K, et al. Repair of the degenerate retina by photoreceptor transplantation. Proceedings of the National Academy of Sciences of the United States of America. 2013;110:354-359. DOI: 10.1073/pnas.1212677110
  115. 115. Singh MS, Charbel Issa P, Butler R, Martin C, Lipinski DM, Sekaran S, et al. Reversal of end-stage retinal degeneration and restoration of visual function by photoreceptor transplantation. Proceedings of the National Academy of Sciences of the United States of America. 2013;110:1101-1106. DOI: 10.1073/pnas.1119416110
  116. 116. Lamba DA, Gust J, Reh TA. Transplantation of human embryonic stem cell-derived photoreceptors restores some visual function in Crx-deficient mice. Cell Stem Cell. 2009;4:73-79. DOI: 10.1016/j.stem.2008.10.015
  117. 117. West EL, Gonzalez-Cordero A, Hippert C, Osakada F, Martinez-Barbera JP, Pearson RA, et al. Defining the integration capacity of embryonic stem cell-derived photoreceptor precursors. Stem Cells. 2012;30:1424-1435. DOI: 10.1002/stem.1123
  118. 118. Jin ZB, Gao ML, Deng WL, Wu KC, Sugita S, Mandai M, et al. Stemming retinal regeneration with pluripotent stem cells. Progress in Retinal and Eye Research. 2019;69:38-56. DOI: 10.1016/j.preteyeres.2018.11.003
  119. 119. Gonzalez-Cordero A, West EL, Pearson RA, Duran Y, Carvalho LS, Chu CJ, et al. Photoreceptor precursors derived from three-dimensional embryonic stem cell cultures integrate and mature within adult degenerate retina. Nature Biotechnology. 2013;31:741-747. DOI: 10.1038/nbt.2643
  120. 120. Decembrini S, Koch U, Radtke F, Moulin A, Arsenijevic Y. Derivation of traceable and transplantable photoreceptors from mouse embryonic stem cells. Stem Cell Reports. 2014;2:853-865. DOI: 10.1016/j.stemcr.2014.04.010
  121. 121. Assawachananont J, Mandai M, Okamoto S, Yamada C, Eiraku M, Yonemura S, et al. Transplantation of embryonic and induced pluripotent stem cell-derived 3D retinal sheets into retinal degenerative mice. Stem Cell Reports. 2014;2:662-674. DOI: 10.1016/j.stemcr.2014.03.011
  122. 122. Shirai H, Mandai M, Matsushita K, Kuwahara A, Yonemura S, Nakano T, et al. Transplantation of human embryonic stem cell-derived retinal tissue in two primate models of retinal degeneration. Proceedings of the National Academy of Sciences of the United States of America. 2016;113:E81-E90. DOI: 10.1073/pnas.1512590113
  123. 123. Santos-Ferreira T, Volkner M, Borsch O, Haas J, Cimalla P, Vasudevan P, et al. Stem cell-derived photoreceptor transplants differentially integrate into mouse models of cone-rod dystrophy. Investigative Ophthalmology & Visual Science. 2016;57:3509-3520. DOI: 10.1167/iovs.16-19087
  124. 124. Kruczek K, Gonzalez-Cordero A, Goh D, Naeem A, Jonikas M, Blackford SJI, et al. Differentiation and transplantation of embryonic stem cell-derived cone photoreceptors into a mouse model of end-stage retinal degeneration. Stem Cell Reports. 2017;8:1659-1674. DOI: 10.1016/j.stemcr.2017.04.030
  125. 125. Zhu J, Reynolds J, Garcia T, Cifuentes H, Chew S, Zeng X, et al. Generation of transplantable retinal photoreceptors from a current good manufacturing practice-manufactured human induced pluripotent stem cell line. Stem Cells Translational Medicine. 2018;7:210-219. DOI: 10.1002/sctm.17-0205
  126. 126. Iraha S, Tu HY, Yamasaki S, Kagawa T, Goto M, Takahashi R, et al. Establishment of Immunodeficient retinal degeneration model mice and functional maturation of human ESC-derived retinal sheets after transplantation. Stem Cell Reports. 2018;10:1059-1074. DOI: 10.1016/j.stemcr.2018.01.032
  127. 127. McLelland BT, Lin B, Mathur A, Aramant RB, Thomas BB, Nistor G, et al. Transplanted hESC-derived retina organoid sheets differentiate, integrate, and improve visual function in retinal degenerate rats. Investigative Ophthalmology & Visual Science. 2018;59:2586-2603. DOI: 10.1167/iovs.17-23646
  128. 128. Gagliardi G, Ben M’Barek K, Chaffiol A, Slembrouck-Brec A, Conart JB, Nanteau C, et al. Characterization and transplantation of CD73-positive photoreceptors isolated from human iPSC-derived retinal organoids. Stem Cell Reports. 2018;11:665-680. DOI: 10.1016/j.stemcr.2018.07.005
  129. 129. Collin J, Zerti D, Queen R, Santos-Ferreira T, Bauer R, Coxhead J, et al. CRX expression in pluripotent stem cell-derived photoreceptors Marks a transplantable subpopulation of early cones. Stem Cells. 2019;37:609-622. DOI: 10.1002/stem.2974
  130. 130. Zou T, Gao L, Zeng Y, Li Q , Li Y, Chen S, et al. Organoid-derived C-kit (+)/SSEA4(−) human retinal progenitor cells promote a protective retinal microenvironment during transplantation in rodents. Nature Communications. 2019;10:1205. DOI: 10.1038/s41467-019-08961-0
  131. 131. Garita-Hernandez M, Lampic M, Chaffiol A, Guibbal L, Routet F, Santos-Ferreira T, et al. Restoration of visual function by transplantation of optogenetically engineered photoreceptors. Nature Communications. 2019;10:4524. DOI: 10.1038/s41467-019-12330-2
  132. 132. Eastlake K, Wang W, Jayaram H, Murray-Dunning C, Carr AJF, Ramsden CM, et al. Phenotypic and functional characterization of Muller glia isolated from induced pluripotent stem cell-derived retinal organoids: Improvement of retinal ganglion cell function upon transplantation. Stem Cells Translational Medicine. 2019;8:775-784. DOI: 10.1002/sctm.18-0263
  133. 133. Singh RK, Occelli LM, Binette F, Petersen-Jones SM, Nasonkin IO. Transplantation of human embryonic stem cell-derived retinal tissue in the subretinal space of the cat eye. Stem Cells and Development. 2019;28:1151-1166. DOI: 10.1089/scd.2019.0090
  134. 134. Lin B, McLelland BT, Aramant RB, Thomas BB, Nistor G, Keirstead HS, et al. Retina organoid transplants develop photoreceptors and improve visual function in RCS rats with RPE dysfunction. Investigative Ophthalmology & Visual Science. 2020;61:34. DOI: 10.1167/iovs.61.11.34
  135. 135. Ribeiro J, Procyk CA, West EL, O’Hara-Wright M, Martins MF, Khorasani MM, et al. Restoration of visual function in advanced disease after transplantation of purified human pluripotent stem cell-derived cone photoreceptors. Cell Reports. 2021;35:109022. DOI: 10.1016/j.celrep.2021.109022
  136. 136. Luo Z, Xian B, Li K, Li K, Yang R, Chen M, et al. Biodegradable scaffolds facilitate epiretinal transplantation of hiPSC-derived retinal neurons in nonhuman primates. Acta Biomaterialia. 2021;134:289-301. DOI: 10.1016/j.actbio.2021.07.040
  137. 137. He XY, Zhao CJ, Xu H, Chen K, Bian BS, Gong Y, et al. Synaptic repair and vision restoration in advanced degenerating eyes by transplantation of retinal progenitor cells. Stem Cell Reports. 2021;16:1805-1817. DOI: 10.1016/j.stemcr.2021.06.002
  138. 138. Thomas BB, Lin B, Martinez-Camarillo JC, Zhu D, McLelland BT, Nistor G, et al. Co-grafts of human embryonic stem cell derived retina organoids and retinal pigment epithelium for retinal reconstruction in Immunodeficient retinal degenerate Royal College of surgeons rats. Frontiers in Neuroscience. 2021;15:752958. DOI: 10.3389/fnins.2021.752958
  139. 139. Ito SI, Onishi A, Takahashi M. Chemically-induced photoreceptor degeneration and protection in mouse iPSC-derived three-dimensional retinal organoids. Stem Cell Research. 2017;24:94-101. DOI: 10.1016/j.scr.2017.08.018
  140. 140. Garita-Hernandez M, Guibbal L, Toualbi L, Routet F, Chaffiol A, Winckler C, et al. Optogenetic light sensors in human retinal organoids. Frontiers in Neuroscience. 2018;12:789. DOI: 10.3389/fnins.2018.00789
  141. 141. Rathnasamy G, Foulds WS, Ling E-A, Kaur C. Retinal microglia—A key player in healthy and diseased retina. Progress in Neurobiology. 2019;173:18-40
  142. 142. Mansour AA, Gonçalves JT, Bloyd CW, Li H, Fernandes S, Quang D, et al. An in vivo model of functional and vascularized human brain organoids. Nature Biotechnology. 2018;36:432-441
  143. 143. Cakir B, Xiang Y, Tanaka Y, Kural MH, Parent M, Kang Y-J, et al. Engineering of human brain organoids with a functional vascular-like system. Nature Methods. 2019;16:1169-1175
  144. 144. Richards D, Jia J, Yost M, Markwald R, Mei Y. 3D bioprinting for vascularized tissue fabrication. Annals of Biomedical Engineering. 2017;45:132-147. DOI: 10.1007/s10439-016-1653-z
  145. 145. Worsdorfer P, Dalda N, Kern A, Kruger S, Wagner N, Kwok CK, et al. Generation of complex human organoid models including vascular networks by incorporation of mesodermal progenitor cells. Scientific Reports. 2019;9:15663. DOI: 10.1038/s41598-019-52204-7

Written By

Jing Yuan and Zi-Bing Jin

Submitted: 29 January 2022 Reviewed: 04 March 2022 Published: 16 April 2022