Skip to main content

Stem cells differentiation into insulin-producing cells (IPCs): recent advances and current challenges

A Correction to this article was published on 15 November 2022

This article has been updated

Abstract

Type 1 diabetes mellitus (T1D) is a chronic disease characterized by an autoimmune destruction of insulin-producing β-pancreatic cells. Although many advances have been achieved in T1D treatment, current therapy strategies are often unable to maintain perfect control of glycemic levels. Several studies are searching for new and improved methodologies for expansion of β-cell cultures in vitro to increase the supply of these cells for pancreatic islets replacement therapy. A promising approach consists of differentiation of stem cells into insulin-producing cells (IPCs) in sufficient number and functional status to be transplanted. Differentiation protocols have been designed using consecutive cytokines or signaling modulator treatments, at specific dosages, to activate or inhibit the main signaling pathways that control the differentiation of induced pluripotent stem cells (iPSCs) into pancreatic β-cells. Here, we provide an overview of the current approaches and achievements in obtaining stem cell-derived β-cells and the numerous challenges, which still need to be overcome to achieve this goal. Clinical translation of stem cells-derived β-cells for efficient maintenance of long-term euglycemia remains a major issue. Therefore, research efforts have been directed to the final steps of in vitro differentiation, aiming at production of functional and mature β-cells and integration of interdisciplinary fields to generate efficient cell therapy strategies capable of reversing the clinical outcome of T1D.

Introduction

Diabetes mellitus

Diabetes mellitus (DM) is a metabolic disease, which arises from a complete deficiency of insulin production—type 1 diabetes (T1D)—or inability to utilize this hormone, as occurs in type 2 diabetes (T2D). It is among the top 10 causes of death in adults, being estimated to have caused 4.2 million deaths globally in 2019 [1]. According to the International Diabetes Federation (IDF) [1], approximately 463 million adults aged 20–79 years old are living with diabetes. An additional number of 1.1 million children and adolescents under 20 years old live with T1D. The IDF [1] also estimates that, by 2045, 700 million adults will be living with DM globally. Therefore, it is crucial to discover and understand the underlying mechanisms of this disease, as well as searching for new and more efficient alternative therapy strategies.

T1D is characterized by the autoimmune destruction of insulin-producing β-pancreatic cells. Autoreactive T cells are key mediators of β-cell destruction, resulting in a complete depletion of insulin hormone, which is essential for carbohydrate metabolism and regulation of normal blood sugar (glycemic) levels [2]. The balance between activated autoreactive memory/effector T cells (Teffs) and activated regulatory T cells (Tregs) is critical for maintaining a healthy immune status. The mechanisms of autoimmunity in T1D are driven by activation of the Teffs, leading to initiation or exacerbation of a preexisting autoimmune process. The persistent activation of Teffs, uncontrolled by Tregs, leads to chronic inflammation and immune response to β-pancreatic cells [2, 3].

According to the World Health Organization (WHO) [4], T1D is responsible for approximately 10% of the totality of DM cases in the World. In contrast, T2D is characterized by the development of insulin resistance due to alterations in the cell insulin receptor or in intermediate mediators of the insulin signaling pathway. The prolonged dysregulation of glycemic levels can cause several chronic health complications, such as diabetic nephropathy, cardiovascular diseases (heart attack, stroke and peripheral artery disease), retinopathy and diabetic neuropathy, which can lead to amputation of inferior members and even death.

T1D therapies

Despite all efforts placed on T1D research throughout the years, the cure for this disease still remains more of an aspiration. Insulin therapy constitutes the main form of treatment for T1D patients; however, continuous administration of exogenous insulin requires an intensive therapeutic regimen and frequent monitoring of glycemic levels, with limited degrees of effectiveness. In addition, this regimen does not accurately mimic the endogenous insulin secretion kinetics; therefore, it is not able to efficiently prevent some of the deleterious effects of hyperglycemia. Moreover, even though insulin therapy slows down the development of secondary complications, it is not able to control glycemic levels in hyper-labile patients [5], who are subject to a wide variation in glycemic rates, showing severe and often fatal hypoglycemic episodes, even under the best conditions of glycemic monitoring and insulin administration [5].

Some advances were made in the field of insulin administration, with the creation of alternative administration routes, such as inhalable insulin preparations, which have become clinically feasible [6, 7], and in the field of glucose level monitoring, with the creation of devices that utilize capillary blood samples [8]. However, it is still necessary to search for other alternative therapeutic strategies to improve the patient’s quality of life and enable a less strict and stressful regimen. From a physiological point of view, restoration of β-pancreatic cell functions through transplantation of insulin-producing tissue (whole pancreas or isolated pancreatic islets) may be the best therapeutic option so far.

Therapeutic alternatives for T1D

According to Fioretto et al. [9], whole organ pancreas transplantation is a viable therapeutic option, since it improves the patient’s quality of life and promotes regression of some late complications associated with T1D. However, this procedure constitutes a major surgical intervention, which requires a strict immunosuppressive regimen and heavily depends on properly functioning of the donor pancreas for a successful treatment, being recommended only for patients with brittle/labile T1D who also need a kidney transplant [10]. Pancreatic islets transplantation, introduced in Brazil by our research group [11, 12], has been shown to be a promising alternative to whole organ pancreas transplantation, since it is a simpler and less invasive procedure. According to Hering et al. [13], transplantation of pancreatic islets is a safe and efficient treatment option for T1D patients with hypoglycemia. Nevertheless, there are still some factors that limit this procedure, such as the low availability of pancreas donors and the requirement for constant patient immunosuppression [10, 14].

Chronic usage of immunosuppressant medication becomes necessary for immunological acceptance of the islet allograft; however, this regimen is associated with various side effects, such as oral sores, gastrointestinal diseases, hypertension, dyslipidemia, anemia, increased infection susceptibility, cancer and systemic toxicity [15]. Therefore, encapsulation of pancreatic islets has emerged as a promising strategy to avoid the need for these immunosuppressive drugs. Production of semipermeable microcapsules for biological application, containing cells or proteins, was initially suggested in the 90’s [16], but considerable progress has been achieved in the field since then, with a major increase in application possibilities, including as an alternative for T1D treatment.

To avoid using steroid-based agents that damage β-cells and are known to be diabetogenic or induce peripheral insulin resistance, a glucocorticoid-free immunosuppressive protocol was developed by the Shapiro’s Group [17], for usage in islet transplantation trials. This protocol includes sirolimus, low dosage of tacrolimus and a monoclonal antibody against the interleukin-2 receptor (daclizumab). Their findings, in a study with T1D patients, indicate that islet transplantation alone is associated with minimal risks for the patient and results in good metabolic control, with normalization of glycated hemoglobin values and restricted requirement for exogenous insulin [17]. This protocol, known as the Edmonton Protocol, was considered as a breakthrough, becoming the standard procedure for islet transplantation, constituting a promising step toward the development of a cure for T1D [18]. However, the standard procedure for pancreatic islets transplantation is based on isolation and purification of islet cells from deceased donors, a process that requires two to four donors per patient, since the efficiency of islet isolation is well below 100% and, additionally, only about 50% of the implanted islets survive after transplantation [19]. In addition, several factors interfere with the viability of the graft after transplantation, such as quality of the donated organ, viability and functionality of the purified islets and the patient’s own immune response [20]. Although many advances have been reached in the field, the need for a large number of viable islets, along with the low availability of donors, is still an important factor that compromise the viability of this methodology.

Although progress has been made, pancreas and islet transplantation are still limited by the limited number of pancreas donors, chronic immunosuppression, which causes a number of adverse effects, and, also, by the recurrence of autoimmunity/onset of alloimmunity [21]. Therefore, a variety of T1D immunotherapy approaches have been developed aiming to prevent or delay T1D onset in predisposed individuals or preserve insulin production in T1D patients [22,23,24]. A hallmark of T1D is the emergence of β-cells destructive autoantibodies against endogenous antigens, which include proinsulin (biosynthetic precursor of insulin), proinsulin C-A junction (connection of C-peptide and A chain of proinsulin), glutamic acid decarboxylase 65 (GAD65, tolerogenic vaccine for T1DM prevention), islet antigen 2 (IA-2) and zinc transporter 8 (ZnT8) [25,26,27,28,29]. Therefore, the overarching goal of immune-focused therapies in T1D is to prevent or delay the loss of functional β-cell mass.

Immunotherapies directed to T1D can be classified into non-autoantigen-specific and autoantigen-specific interventions [30]. Non-antigen-specific treatments are based upon the premise that enhancing immune regulatory mechanisms can ameliorate the destructive autoreactive immune responses, including those against β-cells. A large clinical trial was carried out investigating the therapeutic utility of cyclosporin A in the late 80s. Although cyclosporin A treatment increased T1D remission, this was only for a short duration, since the studies reported progressive increase in daily insulin requirement [31]. Similarly, there have been many clinical interventional studies carried out using anti-CD3 and anti-CD2034 monoclonal antibodies [32]. However, only transient preservation in C-peptide levels was observed [33]. Furthermore, a study investigating safety and efficacy of anti-thymocyte globulin (ATG) failed to preserve β-cell function after two years [34].

Compared to non-autoantigen-specific immunomodulation, autoantigen-specific immunotherapy is expected to selectively modulate T1D-related autoimmunity while preserving the global immune homeostasis intact [35,36,37]. There are studies related to modulation of autoantigen-specific T cell, such as Santamaria et al., 2016 that developed nanoparticles coated with autoantigen-related MHC-II/peptide complex molecules (pMHCII) [36]. There are also trials related to autoantigen-specific B lymphocyte modulation, which have been shown to be more promising than non-specific inhibition of B lymphocytes, for example, by depleting insulin-reactive B cells [37]. Significant progress has already been made through either non-autoantigen-specific immune modulation or T1DM autoantigen-specific immunotherapy. Nevertheless, so far, no T1DM immunotherapy is yet available to replace the standard insulin replacement therapy [30].

Another possible alternative for T1D cell therapy is based on using human mesenchymal stem cells (MSCs) due to their ability to release immunomodulatory molecules that may interrupt the early β-cell destruction by the patient’s own immune system [38]. This may be achieved by infusion of MSCs, which may be obtained from various tissues, directly into the patient’s bloodstream or by apheresis, followed by ex vivo stem cell Educator Therapy, in which the patient's blood passes through a closed-loop system that separates white blood cells, which are momentarily co-cultured with stem cells, before returning them to the patient's bloodstream [39].

Transdifferentiation has also become a potential method to produce functional β-cells. Some findings indicate that, under certain conditions, pancreatic cells, such as acinar and ductal cells, can transdifferentiate into β-cells, following viral transduction [40] or in response to soluble factors [41,42,43]. Nevertheless, further research is required to understand this transdifferentiation of non-β cells into insulin-producing cells (IPCs), since it remains unclear how similar reprogrammed cells are with respect to endogenous β-cells [44].

Several studies have been directed at new and improved methodologies for expansion of β-cell cultures in vitro, aiming at increasing the supply of IPCs for pancreatic islets replacement therapy. Since the nature of T1D disease is a dysfunction of only one cell type, β-cell, differentiation of pluripotent stem cells in β-like cells or IPCs represents a promising approach for T1D cell replacement therapy [45]. Stem cells display two main characteristics: They are non-specialized cells that self-renew for long periods of time, through the cell division cycle, without differentiating into other cell types, while maintaining their capacity to differentiate into different cell types, according to the physiological and experimental conditions to which they are submitted [46].

Achieving economically and technologically viable stem cell-derived therapies still constitutes a great challenge, which requires strict rules for handling and production under appropriate current Good Manufacturing Practice (cGMP) conditions. A cGMP facility is a production facility that includes the manufacturing space, the storage warehouse for raw and finished product and support laboratory areas, also including quality control and quality assurance programs, establishing a Quality System approach [47]. Implementation of procedures and protocols adapted to cGMP requirements is critical to ensure robust and consistent high-quality stem cell manufacturing.

To ensure uniformity from batch-to-batch, manufacturers are required to keep Master Batch Records (MBRs) and Batch Production Records (BPRs) [48]. Detailed standard operating procedures (SOP) and MBRs for manufacturing of stem cell-derived β-cells lots suitable for clinical transplantation are key to ensure that a viable mass of insulin-producing cells can be safely infused into the recipients. These SOPs detail each procedural step, from stem cell expansion and differentiation in vitro to pre-transplantation, quality controls and product release criteria for transplantation [49] ensuring that the reproducibility of the final product is in accordance with established specifications [50]. Also, operation in a closed system and automation of the manual steps enable sterility, processing robustness and reproducibility [48]. The main requirements for stem cell clinical-grade manufacturing, product characterization, infrastructure and concerns related to therapeutic application are shown in Fig. 1. Importantly, investigational products must go through a thorough review process by a regulatory agency, such as FDA (Food and Drug Administration), EMA (European Medicines Agency) and ANVISA (Brazilian Health Regulatory Agency), to determine the safety and effectiveness of products in a well-controlled clinical trial with human subjects.

Fig. 1
figure 1

Overview of the relevant requirements for institution of stem cell-derived therapy clinically, including β-like cells. cGMP: Current good manufacturing practice. iPSC: Induced pluripotent stem cell; ESC: embryonic stem cell

Based on ongoing clinical trials using stem cell-derived β-cells, the eligibility criteria for participating in a clinical study usually include age 18–65 years, clinical history of T1D with > 5 years of duration, episodes of severe hypoglycemia and stable diabetes treatment [51]. During the clinical study, a protocol for outcomes measurement is established in order to evaluate the effectiveness of the transplantation, with most clinical studies having one primary outcome measure, but some have more than one. Graft function depends on the complex physiologic relationship between the graft and the recipient, with several metabolic tests being necessary to monitor graft function and the success of the transplantation. Similar to islet transplantation, the primary endpoints for stem cell-derived β-cells should consist of normal HbA1c level (HbA1c ≤ 6.5%), absence of hypoglycemic episodes and graft durability. The major secondary endpoints include insulin independence, stimulatory test using meal tolerance test (MTT) and oral glucose tolerance test (OGT), continuous glucose monitoring and patient quality of life [52, 53].

Many of the previously mentioned therapeutic strategies, exemplified in Table 1, are currently being clinically tested. Strategies focused on immunomodulation by MSCs infusion are the most highly represented among cell therapies for T1D currently in clinical trials. The combination of immunomodulatory and regenerative properties of MSCs made these cells the most frequently investigated stem cells for clinical applications during the last couple of decades [54]. The immunoregulatory mechanism mediated by MSCs is based on inhibition of effector T cells and other immune cells, while inducing Tregs, reducing directly and indirectly the production of pro-inflammatory cytokines. Many immunosuppressive cells, such as Tregs, regulatory B cells (Bregs), endothelial progenitor cells (EPCs) and myeloid-derived suppressor cells (MDSCs), express TNFR2, TNFα receptor, in direct relationship to their immunosuppression efficiency [55]. In fact, Beldi et al. [56] showed that mouse TNFR2 KO-MSCs have significantly lower immunosuppressive and immunomodulatory effect against T cells. It was further demonstrated that TNFR2 blockade led to increased levels of IFNγ, TNFα and IL-6 pro-inflammatory and decreased IL-10 and TGFβ anti-inflammatory cytokines and nitric oxide production. Moreover, TNFR2 deficiency leads to the induction of Tregs with remarkably less immunosuppressive effect [54]. It has also been suggested that mast cells could confer resistance to T1D, by promoting increased Treg cells, and decreased IL-17-producing T cells in the pancreatic lymph nodes [57]. Considering the autoimmune nature of T1D, marked with a disbalance in Teff and Tregs, as previously described, the MSCs and other molecules that boost Tregs responses represent a therapeutic option for immunomodulation to improve T1D outcomes.

Table 1 Main therapeutic strategies for T1D in current clinical trials

Only two trials involve the usage of pluripotent stem cells fully or not differentiated into insulin-producing β-cells, which includes two companies, namely ViaCyte and Vertex. The ViaCyte initiative is considered as the first cell replacement therapy in clinical trials, with islet-like cells derived from stem cells, testing the safety and efficacy of pancreatic precursor cells incorporated into its encapsulation devices, namely PEC-Encap and PEC-direct. The biologically active component of the PEC-Encap and PEC-direct product candidate is stem cell-derived pancreatic islet cell progenitors, called PEC-01 cells. ViaCyte has shown that once implanted and engrafted, the cells mature into β cells and other islet cell types and are able to secrete insulin in a regulated manner. The PEC-Encap was developed with the purpose of eliminating the need for immunosuppression. The device was evaluated in a 24-month open-label, dose-escalating Phase 1/2 study in T1D patients with minimal insulin-producing β-cell function. The potential for prolonged cell survival has been demonstrated, for as long as 24 months, but has been inconsistent among subjects and primarily limited by a foreign body response to the device component which indicates the requirement for optimization of the device materials [58]. PEC-Direct is an islet cell replacement therapy comprised of stem cell-derived pancreatic islet progenitor cells in a pouch that allows direct vascularization of the implanted cells, thus requiring a concomitant immunosuppressant regimen. A report analysis of data from the first cohort of 15 patients showed that up to one year, patients had 20% reduced insulin requirements, spent 13% more time in target blood glucose range, had stable average HbA1c < 7.0% and had improved hypoglycemic awareness. Implantation of PEC-01 cells was well tolerated, and the serious adverse events that impacted two patients have been previously documented to be associated with the immunosuppression protocol. Only one patient had a > 50% reduction in insulin requirements within one year post-implantation, and no patients achieved insulin independence [59].

Recently, a report by the Vertex company announced positive day 90 data for the first patient from the Phase 1/2 clinical trial of VX-880, an investigational stem cell-derived, fully differentiated pancreatic islet-like cell replacement therapy. This patient had a 91% decrease in daily insulin requirement and simultaneous robust improvements in glucose control, indicating that treatment was generally well tolerated. This was the first demonstration of patient with T1D achieving robust restoration of insulin production from such a cell therapy. The patient was treated with a single infusion of VX-880 at half the target dose in conjunction with immunosuppressive therapy. There were no serious adverse events related to VX-880, and the majority of the adverse events were considered mild to moderate1 [60].

The other trials depicted in Table 1 involve pancreatic islet cell transplantation, based on the Edmonton Protocol or variation thereof, in combination with an immunosuppression regimen (NCT00133809; NCT00434811) or evaluation of different transplantation sites (NCT02402439; NCT02821026) or the combination with other non-endocrine tissues (NCT03977662). Considering the already available advances in the pluripotent stem cells area and the advantages that stem cells-derived IPCs could provide for T1D treatment, these data highlight the crucial necessity to establish efficient and reproducible protocols for stem cell differentiation into IPCs in order to enable their clinical applicability. Therefore, the aim of this review is to provide an overview of the current approaches and achievements in obtaining stem cells-derived IPCs in vitro and the challenges which still need to be overcome.

Stem cells as a source of insulin-producing cells

Stem cells

Stem cells (SCs) are non-specialized cells capable of both auto-renewal and differentiation into different cell types [61]. The cell differentiation process depends on the physiological or experimental conditions to which the cells are subjected, being induced, on the one hand, by intracellular factors, such as expression of key genes, and, on the other, by extracellular factors, such as differentiation-inducing molecules present in the cellular microenvironment [62].

SCs can be classified into three main types: embryonic SCs, adult SCs and induced pluripotent SCs. Embryonic stem cells (ESCs) comprise a class of stem cells derived from the inner cell mass of the blastocyst. ESCs are pluripotent cells that can generate cells from all three embryonic leaflets (endoderm, mesoderm and ectoderm); therefore, they have the greatest potential for cell differentiation [63, 64]. Adult stem cells (ASCs) are non-differentiated cells found in most specialized adult tissues, being able to generate only a selection of cell types of those which are present in that tissue, mainly due to their maintenance and self-renewal [41]. Although displaying a lower proliferation and differentiation potential, ASCs present the great advantage of enabling autologous transplantation [65, 66].

Induced pluripotent stem cells (iPSCs) are genetically modified and reprogrammed cells that originate from adult cells through cellular genetic modification mechanisms, generating cell products, which are similar to ESCs [67]. The reprogramming process is based on transfection of transcriptional factor genes (Oct4, Sox2, c-Myc and Klf4), which are highly expressed in ESCs, through retroviral transduction [68]. After introduction of these reprogramming factors, it is possible to obtain, from differentiated adult cells, groups of cells that are similar to human ESCs, regarding their morphology, cell proliferation rate, antigenic profile, gene expression profile, epigenetic profile, telomere activity and differentiation capacity.

In vitro stem cell differentiation into IPCs as a therapeutic strategy for T1D

In vertebrates, the embryonic pancreas originates from dorsal and ventral protrusions which branch out of the primitive gut. The two pancreatic buds then grow and merge to form the definitive pancreas [69]. The adult pancreas is a retroperitoneal gland divided into three parts: the head (proximal), body and tail (distal). The pancreatic gland has two main cellular compartments with distinct functions, namely the exocrine and the endocrine compartments. The exocrine pancreas, mainly constituted by acinar cells, is responsible for the production and secretion of digestive enzymes, such as proteases, lipases and nucleases, and corresponds to most of the pancreatic mass [70, 71]. In contrast, the endocrine pancreas represents only a small percentage (1–2%) of the entire organ, with cells being organized into cellular groups called islets of Langerhans, which are embedded into the exocrine tissue. The endocrine pancreas consists mainly of four cell types, namely , β, δ and PP cells, which produce, respectively, the glucagon hormone, insulin hormone, somatostatin hormone and the pancreatic polypeptide [69, 70].

iPSCs and ESCs are ideal candidates for differentiation into β-cells due to their outstanding renewal ability, which enables the generation of high numbers of cells that have long been sought in the clinic [71]. In general, the main objectives to be achieved during the differentiation process are: (a) identification of stem cells or progenitor lineages that are capable of self-renewal and differentiation; (b) identification of proliferative signals as well as instructive signals that induce the differentiation process; and (c) identification of molecular signals that maintain the correct physiological state and viability of the differentiated cells [69].

Different strategies have been adopted to obtain IPCs, namely spontaneous differentiation with further selection of Nestin + progenitor cells [72], inhibition of phosphatidylinositol-3-kinase (PI3K) [73], mimicking the in vivo developmental process by adding differentiation factors [74,75,76,77], co-culture with fetal pancreatic buds or culture in the presence of fetal pancreas conditioned medium [78] or transgenic expression of pancreas-specific transcription factors, such as foxa2, ptf1a, pdx1, hnf4a (hepatocyte nuclear factor 4 alpha), hnf6 (hepatocyte nuclear factor 6), ngn3, pax4, neuroD1 and nkx6.1 [71, 79]. Currently, differentiation protocols have been designed using consecutive cytokines or signaling modulators treatments, in specific doses, to activate or inhibit the main signaling pathways that control the differentiation of iPSCs into pancreatic β-cells, namely Wnt; Nodal/Activin A; BMPs; FGF; EGF (epidermal growth factor); Hedgehog; retinoid; and Notch (Fig. 1) [80]. Obtaining mature IPCs in vitro depends on a refined control of concentration, time and duration of treatment with the defined growth and differentiation factors.

Embryoid bodies (EBs)

One of the first steps of the differentiation protocol is the formation of embryoid bodies (EBs), which is necessary to mimic the in vivo embryonic stage of cellular organization. The EBs spontaneously differentiate into cell types of all three primary germ layers, namely ectoderm, mesoderm and endoderm. The EBs formation stage is described as being crucial for determination of the final cells differentiation potential to generate IPCs. Depending on the size of the EB, there is a greater probability of obtaining precursor cells of different cell types [81, 82]. Because the number of specifically differentiated cell types is relatively low after spontaneous differentiation, the following steps aim to induce different signaling pathways to promote cell differentiation and specification. On this basis, the subsequent stages are the formation of definitive endoderm, followed by pancreatic progenitors, pancreatic endocrine cells and, finally, β-cells. Differential gene expression analysis during this process should be useful to follow the in vitro differentiation stages (Table 2; Fig. 2).

Table 2 Function of the main genes involved in β-cell differentiation
Fig. 2
figure 2

Schematic representation of the signaling pathways that coordinate each step of β-cell differentiation and expression levels of the main transcription factor and functional proteins during β-cell differentiation and maturation. NGN3 and MAFB are transiently expressed, while the others remain expressed after maturation. BMP: Bone morphogenetic protein; EGF: Epidermal growth factor; FGF: Fibroblast growth factor; PKC: Protein Kinase C; SHH: Sonic hedgehog; T3: Triiodothyronine; and TGF-β: Transforming growth factor beta

Definitive endoderm (DE)

Initially, factors that lead to activation of the Nodal pathway are employed, since the signaling gradient of these factors leads to endoderm (high nodal) and mesoderm (low nodal) segregation, thus displaying a key function in endodermal formation [130]. Activin A, a member of the TGF-β superfamily (transforming growth factor β), is described as a crucial activation factor for the Nodal pathway [129]. Nodal-mediated signaling modulates the FGF, BMP and Wnt pathways, activating the gastrulation process [131]. Therefore, activin A may be used to mimic Nodal activity in vitro. Expression of Sonic hedgehog (SHH), a potent intercellular patterning signal, is strikingly absent from pancreatic endoderm. Hebrook et al. [132] showed that activin signaling, as a notochord factor, can decrease Shh expression, while inducing expression of Pdx1 and insulin by chick endoderm, thereby permitting pancreas development. However, some studies have shown that activin A may also induce neuronal cells [133]. Therefore, one of the most important parameters for efficient endoderm differentiation is definition of the activin A concentration [129]. Retinoic acid also plays a crucial role in endoderm development during a step between endoderm formation and pancreatic progenitors’ specification [134].

Once formed, definitive endoderm generates the gut tube, which is patterned into anterior and posterior fates by gradients of WNT, FGF and retinoic acid (RA) signaling [135]. WNT signaling is described to have a direct and multifaceted role for WNT signaling in intestinal specification and patterning. WNT signaling acts directly on definitive endoderm to induce Cdx2, a major regulator of intestine-specific genes involved in cell growth and differentiation [136]. Reports demonstrated the ability of WNT to cooperate with Activin signaling to promote definitive endoderm formation, where the optimal induction of differentiation in definitive endoderm was achieved in cells simultaneously treated with Wnt3a [74, 137,138,139,140]. However, Kunisada et al. [141] found that treatment with activin A plus CHIR99021 induced SOX17 and FOXA2 double-positive definitive endoderm more efficiently, when compared with activin A plus Wnt3a.

Pancreatic progenitors

The next step is to induce pancreatic precursor cells, which are cells that display the potential to give rise to all pancreatic lineages and originate all the functional endocrine and exocrine cell types. Considering that the mesenchymal tissues have a critical importance for growth of all pancreatic cell lineages, studies indicate that the FGF signaling pathway, derived from the surrounding mesenchymal tissue, is essential for the formation of specific cell domains. FGF10, as a mesenchymal factor, has an indispensable role in development of the pancreatic epithelium, acting as a mitogenic factor to stimulate proliferation and allowing amplification of pancreatic cells in vitro [142, 143]. It has been demonstrated that culture of dissociated endodermal cells at lower density, followed by longer retinoic acid and FGF10 signaling, results in a high yield of pancreatic progenitors expressing key markers, such as Pdx1 and Nkx6.1 [144]. Ostrom [145] also provides support for an intrinsic role for retinoic acid signaling in specified Ipf1/Pdx1 + pancreatic progenitor cells. FGF2 or basic FGF, known as a notochordal signal, can affect this phase, since it maintains pdx1 expression in the endoderm and potentiates β-cell differentiation [132].

KGF (keratinocyte growth factor), also known as FGF7, is a member of the fibroblast growth factor family that can stimulate ductal cell proliferation [146]. It has also been observed that in rats, KGF acts on ductal cells by activation of distinct signaling pathways to promote β-cell regeneration [141]. KGF is widely used in stepwise differentiation media as it can generate both PDX1 + and subsequent PDX1 + /NKX6.1 + pancreatic progenitors populations, respectively [75, 147,148,149]. Activation of protein kinase C (PKC) is reported to induce pancreatic precursors during β-cell differentiation protocols [75], PKC activation increases β-cell proliferation, size and mass in vivo and is required for growth factor-stimulated β-cell proliferation in vitro [150, 151].

Pancreatic endocrine cells

After induction of pancreatic precursors cells, the in vitro differentiation process must be centered on obtaining endocrine cell specification. Endocrine differentiation is initiated in PDX1 + /NKX6.1 + progenitor through inhibition of Notch signaling, allowing the expression of ngn3, as previously described [74, 152]. These authors reported that cells undergoing endocrine differentiation lose responsiveness to Notch, because Notch activation in Ngn3 + endocrine precursors prevents their differentiation. Ngn3 + cells are promising candidates for endocrine progenitor cells, since they display proliferative capacity and generate cells that express islets-specific transcription factors, such as NEUROD, NKX6.1 and PAX6 [110]. At this stage, using distinct combinations of transcription factors, a specific gene expression profile is initiated and maintained, allowing specification of multipotent progenitors toward the differentiated lineages [112]. Initially, D’Amour et al. [74] included DAPT (gamma secretase inhibitor), a Notch pathway inhibitor, to obtain NGN3 + cells, but it was later shown that it may have a slight beneficial effect on this differentiation step. Despite that, other studies demonstrated that DAPT could have an important role in inducing islet-like structures from embryonic pancreatic precursor cells [153] and several differentiation protocols included the Notch inhibitor γ-secretase inhibitor II [75, 154, 155]. Retinoic acid is proposed to expand the endocrine cell population and block the formation of exocrine cells in a dose-dependent manner [134, 137, 156]. Also, it has been shown that activin A enhances transcription of the ngn3 gene through Smad4 (TGF-β/Smad pathway), binding to the promoter region of ngn3 [157].

It is well known that thyroid hormones have several effects in the development of many endocrine glands, including pancreas [158]. Aiello et al. [158] showed that the thyroid receptors TRα1 and TRβ1 mRNAs were differentially expressed at different phases of embryonic murine pancreas development. These authors found increased mRNA levels of the pro-endocrine gene ngn3 (and increased number of β-cells in cultures previously treated with Triiodothyronine (T3)). The mechanism of T3 action was found to be induction of acinar reprogramming into ductal-like cells that subsequently will differentiated into endocrine cells [112]. Some studies indicate that the Pdx1 + progenitor cells differentiation process requires two major events after establishment of the definitive endoderm, in order to mimic the in vivo process, namely blocking liver differentiation induction by BMP antagonism and induction of pancreatic progenitors by the retinoic signaling pathway [159]. The alternative hepatic lineage differentiation path can be inhibited by treating the cells with different types of inhibitors, such as NOGGIN (BMP antagonist) [152, 160]. It has been demonstrated that the combination of EGF and nicotinamide signaling, together with inhibition of the BMP pathways, promotes an efficient development of NKX6.1 + progenitors from hiPSC lines [143]. The BMP antagonism requirement should be reversed after induction of the pancreatic cell lineages, since BMP signaling is necessary for maintenance of pdx1 expression and additional cell differentiation later on, which can be complicated to implement during in vitro differentiation [161, 162].

Nostro et al. and Chen et al. [135, 163] showed that the inhibition of the TGFβ/activin/nodal and BMP pathways by adding the small molecule ALK4/5/7 inhibitor SB431542 (SB) and Noggin immediately following PDX1 induction had an additive effect, resulting in a sixfold increase in INS expression over that observed in untreated cultures. The results indicated that inhibition of TGFβ/activin/nodal and BMP signaling following induction of pancreatic progenitors does promote differentiation to the endocrine lineage. Rezania et al. [154] and Pagliuca et al. [75] reported that ALK5i and T3 played a significant role at late stages of the differentiation protocol to generate stem cell-derived β-cells. However, Velazco-Cruz et al. [164] identified that inhibiting TGF-β signaling during the last stage of the protocol greatly reduces the function of these differentiated cells, while treatment with Alk5i during the previous stage is necessary for a robust β-like cell phenotype.

Mature β-cells

The final stage of β-pancreatic differentiation aims at β-cells specification and maturation to obtain high levels of cells displaying glucose-stimulated insulin secretion capacity. Mature adult pancreatic β-cells are functionally defined by their rapid response to elevated glucose [154]. To reach this cellular profile, frequently, factors and molecules that are known to act in adult pancreas are used. Betacellulin, a member of the epidermal growth factor family, is produced by proliferating pancreatic β cells [135] and can increase Pdx1 and insulin production [165]. At this stage, nicotinamide supplementation is usually added to the culture medium, to enhance the in vitro differentiation of cultured human pancreatic cells, favoring the expression of insulin, glucagon and somatostatin [166]. Nicotinamide has historically been used to augment pancreatic β-cell differentiation and to protect islet cells from toxic insults, due to its antioxidant properties. Studies showed that cells exposure to nicotinamide is essential for robust nkx6.1 expression in hiPSC differentiation to pancreatic endocrine progenitors, acting predominantly through PARP (poly-ADP-ribose polymerases) inhibition [167, 168]. Thowfeequ et al. [169] showed that the addition of betacellulin and nicotinamide to the modified differentiation protocol sustained PDX1 expression and induced pancreatic β-cell differentiation in human ES cell line.

Glucagon-like peptide-1 (GLP-1) is an intestinal incretin hormone that binds to specific G protein-coupled receptors on pancreatic β-cells to stimulate insulin secretion via cAMP-dependent pathways. Consequently, GLP-1 plays a crucial role in β-cell mass regeneration [170]. Exogenous GLP-1 increases islet cell proliferation in Ins-1 cells via a PI3-kinase-dependent pathway [171]. Exendin 4, a long-acting GLP-1 analogue, is resistant to dipeptidyl peptidase IV (DPP-IV) cleavage, being more useful clinically, and can also be used to promote β-cell proliferation. Considering the importance of mafA expression in β-cells, representing an important indicator of β-cell maturity, studies identified that thyroid hormone is also a physiological regulator of β-cell maturation through direct interaction with the mafA promoter [172]. Therefore, thyroid hormone may improve in vitro functional maturation of immature stem cells-derived insulin-expressing cells. Moreover, it is believed that VEGF (vascular endothelial growth factor) is predominantly secreted by β-cells in adult pancreas, affecting islet function and physiology [173]. Consequently, exogenous supplementation with VEGF has been associated with reduction in β-cell apoptosis and maintenance of β-cell mass [174].

Another important component that is crucial and should be provided during β-cell differentiation and maturation is the major components of the extracellular matrix (ECM) of islet cells, including laminin and collagen. The islet ECM has been shown to regulate survival, insulin secretion, proliferation and islet morphology. Moreover, laminin and type IV collagen were identified to be beneficial for β-cell function in vitro [175]. Laminins were shown to induce expression of islet-specific transcription factors and hormones, such as Pdx1, insulin1, insulin2, glucagon and Glut2 [176]. In in vitro experiments, collagen has been associated with provide the desired mechanical properties of transplanted grafts, to improve the performance of scaffolds and, in combination with other ECM proteins, such as laminin, to enhance glucose-stimulated insulin secretion in pancreatic islets [177, 178]. Therefore, providing islet matrix proteins to the in vitro differentiation process is a key determinant for presentation of matrix-bound signals, warranting a microenvironment which is closer to the native in vivo situation, thereby sustaining the maintenance of cellular viability.

Challenges and achievements

Although several factors are important for successful generation of IPCs from iPSCs or ESCs, careful handling of cell culture conditions stands out as one of the most critical factors [161]. Table 3 highlights the main growth and differentiation factors used during the four critical steps of β-cell differentiation from hiPSC or hESC, described in major reports found in the literature. Numerous efforts have been employed to obtain hPSC-derived β-cells since Lumelsky et al. [72] first described a protocol to enrich IPC from ESCs by selecting NESTIN + cells, but only in 2014 two different research groups [75, 154] published a protocol showing the differentiation of human embryonic stem cells (hESCs) into β-cells that resemble cadaveric β-cells with respect to both gene expression and function. It was quite a breakthrough in developing stem cell-derived β-cells, and currently, the protocol developed by the Melton Lab [75] is the basis for the Vertex clinical trial therapy.

Table 3 Differentiation factors from major reports in the literature regarding stem cell-derived β-cell

Acquisition of dynamic insulin secretion upon glucose stimulation is a key feature of β-cells. This dynamic function is represented by a pulsatile behavior of two-phase insulin secretion: The first phase has a period of 10–15 min following stimulation by glucose, comprising a high amplitude but with short duration, while the second phase has a lower amplitude and a longer duration of 1–2 h [186, 187]. According to Table 3, many of the protocols generated immature mixed populations of cells at different developmental stages, displaying polyhormonal properties and, additionally, IPCs-transplanted mice usually maintain euglycemia for only a short period of time or present a compromised GSIS dynamics. Indeed, the majority of β-like cells derived from stem cells differentiation resemble fetal β-cells regarding their maturity [188]. Velasco-Cruz et al. [164] first reported robust dynamic insulin secretion of SC-β cells. It was further shown that manipulation of the polymerization state of actin cytoskeleton influences NEUROG3-dependent endocrine induction. The results obtained allowed overcoming the requirement for three-dimensional culture in stem cell-derived β-cell differentiation and creating a fully planar protocol [189]. These findings enable simplifying the differentiation methodology, requiring only basic stem cell culture experience, as well as familiarity with assessment techniques which are commonly used in biology laboratories [190].

Nair et al. [191] optimized the Russ et al. protocol [152] to increase β-like cells maturity through reaggregation of INS+ β-like cells isolated by fluorescence-activated cell sorting (FACS); however, these cells presented a marked first phase response to glucose but failed to sustain the second phase of insulin secretion. Studies by Yoshihara et al. [192] demonstrated that stem cells-derived β-cells could acquire adult insulin secretion behavior through overexpression of estrogen-related receptor γ (ERRγ), which is hypothesized to regulate mitochondrial metabolic pathways required for GSIS. In the attempt to characterize the protocol of in vitro differentiation, single-cell transcriptome has been undertaken to visualize populations and pathways regulated during the stages [193].

It is important to highlight that native pancreatic islet is highly vascularized cellular aggregates, consisting of, approximately, 10% of blood vessels, which are essential to allow networking between glucose concentration sensing and insulin secretion by β-cells and, also, to provide proper islet oxygenation [194]. The lack of these vasculature interactions is one of the main reasons for the low survival rate of transplanted islets [195]. In this context, studies have hypothesized that in vitro interaction between ESC-derived EBs and endothelial cells may augment the differentiation toward pancreatic endocrine progenitors and IPCs [196]. Weizman et al. [197] also proposed a 3D architecture system using polymeric scaffolds to culture hESC-derived pancreatic cells embedded in a vascular niche composed of endothelial cells and/or fibroblasts. Therefore, endothelial cells may provide key factors that lead to the endocrine cell fate during in vitro differentiation. In general, incorporation of endothelial cells and other important cells normally present in the β-pancreatic niche may be beneficial for improving IPCs differentiation and functionality.

Pancreatic islets also receive complex neural inputs, and β-cells present a phenotypically diverse population, with a mosaic of metabolic and electrical activity patterns [198]. Although adult β-cells populations are totally differentiated, they are heterogeneous with respect to their insulin secretory abilities, mitochondrial function, calcium signaling and proliferative properties. For this reason, maturity is not defined only by the expression of major molecular markers, such as PDX1, NKX6.1 and MAFA, or by high insulin expression levels [41]. Johnston and colleagues [199] have reported that a 5–8% subset of β-cells forms “super-connected hubs” within an interconnected islet cellular network. It has also been shown that these cells serve as pacemakers that can synchronize the calcium and insulin secretory responses across the whole islet. In addition, β-cells can be divided into two major populations: One comprised of cells that are capable of proliferation and the other one comprised of mature β-cells that are marked by the expression of Fltp (also known as Flattop or Cfap126), a Wnt/PCP (planar cell polarity) effector. FLTP + cells represent the subpopulation of mature β-cells, while Fltp-negative cells comprise immature and proliferative cells [200]. However, Dorrell et al. [200] demonstrated that human β-cells have at least four different cellular subtypes, which may be classified based on their cell surface markers expression. This suggests a functional heterogeneity among β-cells and illustrates the degree of complexity of the insulin release kinetics that stem cells-derived IPCs should probably achieve.

Typically, a patient requires two transplants, each of which with at least 10.000 islet “equivalents” (IEQs) per kilogram of body weight, to achieve insulin independence [19]. Proportionately, a single 70 kg patient requires approximately 700 million of transplanted IEQs [19]. This poses important challenges related to manufacturing sufficiently pure and potent cells, at scale, for clinical use and, also, protecting these cells from immune rejection following transplantation. Some strategies to address these limitations have already been described. Schulz et al. [149] reported a process that allows scaled production of hESC and, subsequently, of pancreatic progenitors. These authors developed a feeder-free culture system for expansion of the CyT49 hESC line and generation of large-scale single-cell master and working banks of CyT49 under good manufacturing practices (cGMP) [149].

Although autologous transplantation of patient-specific IPCs derived from iPSCs emerged as an attractive strategy, it still requires suppression of the preexisting autoimmunity [201, 202]. The negative effects of some immunosuppressants in human β-cell transplantation patients have been widely reported, being associated with complications at new onset DM upon transplantation [203,204,205]. Another interesting approach is to mediate genetic manipulation in order to control the expression of HLA class I and II genes, allowing the graft to escape from immune recognition and destruction [206]. Furthermore, Yoshihara et al. [185] showed that human islet-like organoids (HILOs) generated from iPSCs overexpressing PD-L1, a known determinant of immune tolerance in β-cells, are protected from xenograft and allogenic rejection and maintain glucose homeostasis in diabetic mice.

The encapsulation strategies are currently the most promising approach, representing the most adequate alternative, when compared to adoption of the immunosuppressive regimen. Cell encapsulation creates a physical barrier for the transplanted IPCs, providing a 3D architecture that may attenuate the deleterious impact of the host immune system on newly transplanted cells [207]. Encapsulation of pancreatic islets with artificial membranes allows preservation of their physical characteristics and functional integrity. Furthermore, studies carried out by our group demonstrated that incorporation of polylaminin into the microcapsule polymer attenuated the post-transplantation immunological response against microcapsules grafted in mice, suggesting an improved maintenance of the grafted encapsulated pancreatic islets in the recipient organism [14]. Vegas et al. [208] carried out an experiment of long-term evaluation of encapsulated SC-derived β cells in immune-competent mice. They showed that stem cell-derived β-cells can promote long-term glycemic correction (174 days) in an immune-competent diabetic animal in the absence of immunosuppressive therapy, using a modified alginate capable of mitigating the innate immune-mediated foreign body responses, with euglycemic mice still being present at the end of the experiment. Subsequently, the same chemically modified alginate, called Z1-Y15, was shown to prevent pericapsular fibrotic overgrowth and maintain encapsulated islets function after four months of allogenic transplantation in non-human primates, in the absence of immunosuppression, in a pre-clinical study [209]. These authors also suggest an alternative transplantation site into the bursa omentalis, which can support nutritional exchange for long-term islet viability. This technology was incorporated by the Sigilon Therapeutics company and has already been tested in clinical trials (phase 1/2) to assess the safety, tolerability and preliminary efficacy of SIG-001, which is composed of human cells that are engineered to produce FVIII, in adults with severe or moderately severe hemophilia A [210].

An important concern with stem cells-derived therapeutic products is the presence of undifferentiated or partially differentiated cells that may not only interfere with the desired cell types activity, but, also, be tumorigenic. For this reason, optimization of the in vitro differentiation process is fundamental to minimize the formation of unwanted cell types and, consequently, validate this technology for clinical use [44]. Additional approaches to eliminate non-differentiated cells include the use of antibody-toxin molecules or conjugates that selectively kill non-differentiated cells [211]. Despite the existing risks, many different strategies have been employed to promote in vivo maturation of transplanted progenitor cells [137, 212]. However, the use of encapsulation devices that provide their precise location in the body and the possibility to be recovered in case of graft failure or other complications is a promising approach to allow safe progenitor cells transplantation [182, 213]. In general, the choice of cells at different stages of maturation has many safety-related implications, with mature differentiated cells being the safest ones since they display low levels of residual plasticity [214].

Many advances have been made with respect to the establishment of differentiation protocols capable of generating homogeneous cell masses at early stages of development. Also, many efforts have been made to generate better functioning β-cells by introducing some features that could favor the differentiation process, such as promoting clustering of immature β-like cells into endocrine-enriched niches [191], assembly of islet-like organoids onto hydrogel slabs [82, 215], engineering human islet organoids using an organ-on-a-chip platform [216] and culturing in decellularized pancreatic scaffolds [217]. However, a standardized differentiation protocol is still lacking, and the final differentiation stages also need to be better understood. To address this challenge, understanding the whole transcriptome, epigenome and proteome of the differentiation process could help to obtain insights into the pathways that lead to the process of mature and functional β-cells generation.

Conclusions

In conclusion, generation of pancreatic β-cells from pluripotent stem cells constitutes a very promising therapeutic approach to provide insulin independence to millions of diabetic patients. Differentiation protocols, cell culture methodology and encapsulation protocols are being developed to optimize β-cells production and provide protection against the autoimmune response displayed by T1D patients. Although several previously mentioned challenges still need to be overcome, a great deal of efforts has been employed combining several interdisciplinary fields, such as stem cell biology, embryology, immunology, cell encapsulation and tissue bioengineering, to enable the development of effective cellular therapies.

Availability of data and materials

Not applicable.

Change history

Abbreviations

ANVISA:

Brazilian Health Regulatory Agency

ASCs:

Adult stem cells

ATG:

Anti-thymocyte globulin

BPRs:

Batch production records

cGMP:

Current good manufacturing practices

DM:

Diabetes mellitus

EBs:

Embryoid bodies

ECM:

Extracellular matrix

EMA:

European medicines agency

ESCs:

Embryonic stem cells

FACS:

Fluorescence-activated cell sorting

FDA:

Food and drug administration

GAD65:

Glutamic acid decarboxylase 65

HILOs:

Human islet-like organoids

IA-2:

Islet antigen 2

IDF:

International diabetes federation

IEQs:

Islet “equivalents”

IPCs:

Insulin-producing cells

iPSCs:

Induced pluripotent stem cells

MBRs:

Master batch records

MSCs:

Mesenchymal stem cells

MTT:

Meal tolerance test

OGT:

Oral glucose tolerance test

pMHCII:

MHC-II/peptide complex molecules

SCs:

Stem cells

SOP:

Standard operating procedures

T1D:

Type 1 diabetes mellitus

T2D:

Type 2 diabetes

Teffs:

Effector T cells

Tregs:

Activated regulatory T cells

WHO:

World Health Organization

ZnT8:

Zinc transporter 8

References

  1. International Diabetes Federation. IDF diabetes atlas, 9th ed. Brussels: Belgium. 2019. https://www.diabetesatlas.org. Accessed 14 Jun 2020.

  2. Pugliese A. Autoreactive T cells in type 1 diabetes. J Clin Invest. 2017;127(8):2881–91. https://doi.org/10.1172/JCI94549.

    Article  PubMed  PubMed Central  Google Scholar 

  3. Grinberg-Bleyer Y, et al. Pathogenic T cells have a paradoxical protective effect in murine autoimmune diabetes by boosting Tregs. J Clin Investig. 2010;120(12):4558–68. https://doi.org/10.1172/JCI42945.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. World Health Organization (WHO). Diabetes: key facts. https://www.who.int/news-room/fact-sheets/detail/diabetes. Accessed 14 Jun 2020.

  5. The Diabetes Control and Complications Trial Research Group. Hypoglycemia in the diabetes control and complications trial. Diabetes. 1997;46(2):271–86. https://doi.org/10.2337/diab.46.2.271 (PMID: 9000705).

    Article  Google Scholar 

  6. Klonoff DC. Afrezza inhaled insulin: the fastest-acting FDA-approved insulin on the market has favorable properties. J Diabetes Sci Technol. 2014;8(6):1071–3. https://doi.org/10.1177/1932296814555820.

    Article  PubMed  PubMed Central  Google Scholar 

  7. Mohanty RR. Inhaled insulin - current direction of insulin research. J Clin Diagn Res. 2017. https://doi.org/10.7860/JCDR/2017/23626.9732.

    Article  PubMed  PubMed Central  Google Scholar 

  8. Bailey T, Bode BW, Christiansen MP, Klaff LJ, Alva S. The performance and usability of a factory-calibrated flash glucose monitoring system. Diabetes Technol Ther. 2015;17(11):787–94. https://doi.org/10.1089/dia.2014.0378.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Fioretto P, Steffes MW, Sutherland DE, Goetz FC, Mauer M. Reversal of lesions of diabetic nephropathy after pancreas transplantation. N Engl J Med. 1998;339(2):69–75. https://doi.org/10.1056/NEJM199807093390202.

    Article  CAS  PubMed  Google Scholar 

  10. Leal-Lopes, C. Terapias alternativas para o Diabetes Mellitus tipo 1: caracterização funcional do gene Txnip na diferenciação β-pancreática e desenvolvimento de biomaterial inovador para microencapsulamento celular. 203f. Tese de Doutorado – Universidade de São Paulo, Jun. 2018. https://doi.org/10.11606/T.46.2018.tde-24082018-083447

  11. Eliaschewitz FG, Aita CAM, Genzini T, Noronha IL, Lojudice FH, Labriola L, Krogh K, Oliveira EMC, Silva IC, Mendonça Z, Franco D, Miranda MP, Noda E, de Castro LA, Andreolli M, Goldberg AC, Sogayar MC. First Brazilian pancreatic islet transplantation in a patient with type 1 diabetes mellitus. Transplant Proc. 2004;36(4):1117–8. https://doi.org/10.1016/j.transproceed.2004.04.065.

    Article  CAS  PubMed  Google Scholar 

  12. Eliaschewitz FG, Franco DR, Mares-Guia TR, Noronha IL, Labriola L, Sogayar MC. Arq bras endocrinol metabol. Rev Port. 2009;53(1):15–23. https://doi.org/10.1016/j.transproceed.2004.04.065.

    Article  Google Scholar 

  13. Hering BJ, Clarke WR, Bridges ND, et al. Phase 3 trial of transplantation of human islets in type 1 diabetes complicated by severe hypoglycemia. Diabetes Care. 2016;39(7):1230–40. https://doi.org/10.2337/dc15-1988.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Leal-Lopes C, Grazioli G, Mares-Guia TR, Coelho-Sampaio T, Sogayar MC. Polymerized laminin incorporation into alginate-based microcapsules reduces pericapsular overgrowth and inflammation. J Tissue Eng Regen Med. 2019;13(10):1912–22. https://doi.org/10.1002/term.2942.

    Article  CAS  PubMed  Google Scholar 

  15. Shapiro AM, Pokrywczynska M, Ricordi C. Clinical pancreatic islet transplantation. Nat Rev Endocrinol. 2017;13(5):268–77. https://doi.org/10.1038/nrendo.2016.178.

    Article  CAS  PubMed  Google Scholar 

  16. Chang TMS. Semipermeable microcapsules. Science. 1964;146(3643):524–5. https://doi.org/10.1126/science.146.3643.524.

    Article  CAS  PubMed  Google Scholar 

  17. James Shapiro AM, Lakey JRT, Ryan EA, Korbutt GS, Toth E, Warnock GL, Kneteman NM, Rajotte RV. Islet transplantation in seven patients with type 1 diabetes mellitus using a glucocorticoid-free immunosuppressive regimen. N Engl J Med. 2000;343(4):230–8. https://doi.org/10.1056/NEJM200007273430401.

    Article  Google Scholar 

  18. Juvenile Diabetes Cure Alliance. The Edmonton Protocol Turns 20: What Have We Learned? http://thejdca.org/2018-edmonton-protocol. Accessed 6 Dec 2018.

  19. Shapiro AM. State of the art of clinical islet transplantation and novel protocols of immunosuppression. Curr Diab Rep. 2011;11(5):345–54. https://doi.org/10.1007/s11892-011-0217-8.

    Article  PubMed  Google Scholar 

  20. Rother KI, Harlan DM. Challenges facing islet transplantation for the treatment of type 1 diabetes mellitus. J Clin Investig. 2004;114(7):877–83. https://doi.org/10.1172/JCI23235.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. American Diabetes Association. Diagnosis and classification of diabetes mellitus. Diabetes Care. 2009;32(Suppl. 1):62–7.

    Article  Google Scholar 

  22. Smith EL, Peakman M. Peptide immunotherapy for type 1 diabetes—clinical advances. Front Immunol. 2018;9:392.

    Article  PubMed  PubMed Central  Google Scholar 

  23. Itoh A, Ridgway WM. Targeting innate immunity to downmodulate adaptive immunity and reverse type 1 diabetes. ImmunoTargets and therapy. 2017;6:31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Aghazadeh Y, Nostro MC. Cell therapy for type 1 diabetes: current and future strategies. Curr DiabRep. 2017;17(6):1–9.

    CAS  Google Scholar 

  25. Lernmark Å, Larsson HE. Immune therapy in type 1 diabetes mellitus. Nat Rev Endocrinol. 2013;9(2):92–103.

    Article  CAS  PubMed  Google Scholar 

  26. Roep BO, Solvason N, Gottlieb PA, Abreu JRF, Harrison LC, Eisenbarth GS, Liping Yu, Leviten M, Hagopian WA, Buse JB, von Herrath M, Quan J, King RS, Robinson WH, Utz PJ, Garren H, Steinman L. Plasmid-encoded proinsulin preserves C-peptide while specifically reducing proinsulin-specific CD8+T cells in type 1 diabetes. Sci Transl Med. 2013. https://doi.org/10.1126/scitranslmed.3006103.

    Article  PubMed  PubMed Central  Google Scholar 

  27. Ludvigsson J, Krisky D, Casas R, Battelino T, Castaño L, Greening J, Kordonouri O, Otonkoski T, Pozzilli P, Robert J-J, Veeze HJ, Palmer J. GAD65 antigen therapy in recently diagnosed type 1 diabetes mellitus. N Engl J Med. 2012;366(5):433–42. https://doi.org/10.1056/NEJMoa1107096.

    Article  CAS  PubMed  Google Scholar 

  28. Williams AJK, Lampasona V, Schlosser M, Mueller PW, Pittman DL, Winter WE, Akolkar B, Wyatt R, Brigatti C, Krause S, Achenbach P. Detection of antibodies directed to the N-terminal region of GAD is dependent on assay format and contributes to differences in the specificity of GAD autoantibody assays for type 1 diabetes. Diabetes. 2015;64(9):3239–46. https://doi.org/10.2337/db14-1693.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Fabris M, Zago S, Liguori M, Trevisan MT, Zanatta M, Comici A, Zanette G, Carlin E, Curcio F, Tonutti E. Anti-zinc transporter protein 8 autoantibodies significantly improve the diagnostic approach to type 1 diabetes: an Italian multicentre study on paediatric patients. Autoimmun Highlights. 2015;6(1):17–22.

    Article  CAS  Google Scholar 

  30. Frumento D, Ben Nasr M, El Essawy B, D’Addio F, Zuccotti GV, Fiorina P. Immunotherapy for type 1 diabetes. J Endocrinol Invest. 2017;40(8):803–14.

    Article  PubMed  Google Scholar 

  31. Martin S, Pawlowski B, Greulich B, Ziegler AG, Mandrup-Poulsen T, Mahon J. Natural course of remission in IDDM during 1st yr after diagnosis. Diabetes Care. 1992;15(1):66–74.

    Article  CAS  PubMed  Google Scholar 

  32. Keymeulen B, Walter M, Mathieu C, Kaufman L, Gorus F, Hilbrands R, Vandemeulebroucke E, Van de Velde U, Crenier L, De Block C, Candon S, Waldmann H, Ziegler AG, Chatenoud L, Pipeleers D. Four-year metabolic outcome of a randomised controlled CD3-antibody trial in recent-onset type 1 diabetic patients depends on their age and baseline residual beta cell mass. Diabetologia. 2010;53(4):614–23. https://doi.org/10.1007/s00125-009-1644-9.

    Article  CAS  PubMed  Google Scholar 

  33. Ludvigsson J. Therapies to preserve β-cell function in type 1 diabetes. Drugs. 2016;76(2):169–85.

    Article  CAS  PubMed  Google Scholar 

  34. Gitelman SE, Gottlieb PA, Felner EI, Willi SM, Fisher LK, Moran A, Gottschalk M, Moore WV, Pinckney A, Keyes-Elstein L, Harris KM, Kanaparthi S, Phippard D, Ding L, Bluestone JA, Ehlers M. Antithymocyte globulin therapy for patients with recent-onset type 1 diabetes: 2 year results of a randomised trial. Diabetologia. 2016;59(6):1153–61. https://doi.org/10.1007/s00125-016-3917-4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Miller SD, Turley DM, Podojil JR. Antigen-specific tolerance strategies for the prevention and treatment of autoimmune disease. Nat Rev Immunol. 2007;7(9):665–77.

    Article  CAS  PubMed  Google Scholar 

  36. Clemente-Casares X, Blanco J, Ambalavanan P, Yamanouchi J, Singha S, Fandos C, Tsai S, Wang J, Garabatos N, Izquierdo C, Agrawal S, Keough MB, Wee Yong V, James E, Moore A, Yang Y, Stratmann T, Serra P, Santamaria P. Expanding antigen-specific regulatory networks to treat autoimmunity. Nature. 2016;530(7591):434–40. https://doi.org/10.1038/nature16962.

    Article  CAS  PubMed  Google Scholar 

  37. Henry RA, Kendall PL, Thomas JW. Autoantigen-specific B-cell depletion overcomes failed immune tolerance in type 1 diabetes. Diabetes. 2012;61(8):2037–44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Cho J, D’Antuono M, Glicksman M, Wang J, Jonklaas J. A review of clinical trials: mesenchymal stem cell transplant therapy in type 1 and type 2 diabetes mellitus. Am J Stem Cells. 2018;7(4):82.

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Zhao Y, Jiang Z, Zhao T, et al. Reversal of type 1 diabetes via islet β cell regeneration following immune modulation by cord blood-derived multipotent stem cells. BMC Med. 2012;10:3. https://doi.org/10.1186/1741-7015-10-3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Zhou Q, Brown J, Kanarek A, et al. In vivo reprogramming of adult pancreatic exocrine cells to β-cells. Nature. 2008;455:627–32. https://doi.org/10.1038/nature07314.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Baeyens L, Lemper M, Leuckx G, et al. Transient cytokine treatment induces acinar cell reprogramming and regenerates functional beta cell mass in diabetic mice. Nat Biotechnol. 2014;32:76–83. https://doi.org/10.1038/nbt.2747.

    Article  CAS  PubMed  Google Scholar 

  42. Klein D, Álvarez-Cubela S, Lanzoni G, et al. BMP-7 induces adult human pancreatic exocrine-to-endocrine conversion. Diabetes. 2015;64:4123–34. https://doi.org/10.2337/db15-0688.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Qadir MMF, Álvarez-Cubela S, Klein D, et al. P2RY1/ALK3-expressing cells within the adult human exocrine pancreas are BMP-7 expandable and exhibit progenitor-like characteristics. Cell Rep. 2018;22:2455–68. https://doi.org/10.1016/j.celrep.2018.02.006.

    Article  CAS  Google Scholar 

  44. Nair GG, Tzanakakis ES, Hebrok M. Emerging routes to the generation of functional β-cells for diabetes mellitus cell therapy. Nat Rev Endocrinol. 2020;16(9):506–18. https://doi.org/10.1038/s41574-020-0375-3.

    Article  PubMed  PubMed Central  Google Scholar 

  45. Lojudice FH, Sogayar MC. Células-tronco no tratamento e cura do diabetes mellitus. Cien Saude Colet. 2008;13(1):19–21. https://doi.org/10.1590/S1413-81232008000100005.

    Article  PubMed  Google Scholar 

  46. McCulloch EA, Till JE. Perspectives on the properties of stem cells. Nat Med. 2005;11(10):1026–8. https://doi.org/10.1038/nm1005-1026.

    Article  CAS  PubMed  Google Scholar 

  47. Giancola R, Bonfini T, Iacone A. Cell therapy: cGMP facilities and manufacturing. Muscles Ligaments Tendons J. 2012;2(3):243.

    PubMed  PubMed Central  Google Scholar 

  48. U. S. Food and Drug Administration. Center for drug evaluation and research. code of federal regulations title 21, volume 4: current good manufacturing practice for finished pharmaceuticals. 2022.

  49. Leal-Lopes C, da Cunha Mantovani M, Sogayar MC. Advanced therapy medicinal products in type I diabetes mellitus: technological and regulatory challenges. Vigilância Sanitária em Debate. 2018;6(1):41–55.

    Article  Google Scholar 

  50. Unger C, Skottman H, Blomberg P, Sirac Dilber M, Hovatta O. Good manufacturing practice and clinical-grade human embryonic stem cell lines. Hum Mol Genet. 2008;17(R1):R48–53.

    Article  CAS  PubMed  Google Scholar 

  51. Vertex Pharmaceuticals Incorporated. A safety, tolerability, and efficacy study of VX-880 in participants with type 1 diabetes. 2021. (ClinicalTrials.gov Identifier: NCT04786262).

  52. U. S. Food and Drug Administration. Guidance for industry. Consideration for allogeneic pancreatic islet cell products. 2009.

  53. Welsch CA, Rust WL, Csete M. Concise review: lessons learned from islet transplant clinical trials in developing stem cell therapies for type 1 diabetes. Stem Cells Transl Med. 2019;8(3):209–14.

    Article  PubMed  Google Scholar 

  54. Beldi G, Bahiraii S, Lezin C, Nouri Barkestani M, Abdelgawad ME, Uzan G, Naserian S. TNFR2 is a crucial hub controlling mesenchymal stem cell biological and functional properties. Front Cell Dev Biol. 2020;8:596831.

    Article  PubMed  PubMed Central  Google Scholar 

  55. Shamdani S, Uzan G, Naserian S. TNFα-TNFR2 signaling pathway in control of the neural stem/progenitor cell immunosuppressive effect: Different experimental approaches to assess this hypothetical mechanism behind their immunological function. Stem Cell Res Ther. 2020;11(1):307. https://doi.org/10.1186/s13287-020-01816-2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Beldi G, Khosravi M, Abdelgawad ME, Salomon BL, Uzan G, Haouas H, Naserian S. TNFα/TNFR2 signaling pathway: an active immune checkpoint for mesenchymal stem cell immunoregulatory function. Stem Cell Res Ther. 2020;11(1):1–15.

    Article  Google Scholar 

  57. Henry RR, Pettus J, Wilensky JON, Shapiro AJ, Senior PA, Roep B, Wang R, Kroon EJ, Scott M, D’Amour KE, Foyt HL. Initial clinical evaluation of VC-01TM combination product—a stem cell–derived islet replacement for type 1 diabetes (T1D). Diabetes. 2018. https://doi.org/10.2337/db18-138-OR.

    Article  Google Scholar 

  58. Carlos D, Yaochite JN, Rocha FA, Toso VD, Malmegrim KC, Ramos SG, et al. Mast cells control insulitis and increase Treg cells to confer protection against STZ-induced type 1 diabetes in mice. Eur J Immunol. 2015;45(10):2873–85.

    Article  CAS  PubMed  Google Scholar 

  59. Ramzy A, Thompson DM, Ward-Hartstonge KA, Ivison S, Cook L, Garcia RV, Loyal J, Kim PT, Warnock GL, Levings MK, Kieffer TJ. Implanted pluripotent stem-cell-derived pancreatic endoderm cells secrete glucose-responsive C-peptide in patients with type 1 diabetes. Cell Stem Cell. 2021;28(12):2047–61.

    Article  CAS  PubMed  Google Scholar 

  60. Vertex. Vertex announces positive day 90 data for the first patient in the phase 1/2 clinical trial dosed with VX-880, a Novel investigational stem cell-derived therapy for the treatment of type 1 diabetes. https://news.vrtx.com/press-release/vertex-announces-positive-day-90-data-first-patient-phase-12-clinical-trial-dosed-vx?_ga=2.53361578.345811804.1646342387-705593813.1646342387. Accessed 10 Mar 2022.

  61. Arrighi N. Definition and classification of stem cells. Stem Cells. 2018;2018:1–45.

    Google Scholar 

  62. Hwang NS, Varghese S, Elisseeff J. Controlled differentiation of stem cells. Adv Drug Deliv Rev. 2008;60(2):199–214.

    Article  CAS  PubMed  Google Scholar 

  63. Thomson JA, Itskovitz-Eldor J, Shapiro SS, Waknitz MA, Swiergiel JJ, Marshall VS, Jones JM. Embryonic stem cell lines derived from human blastocysts. Science. 1998;282(5391):1145–7.

    Article  CAS  PubMed  Google Scholar 

  64. Rippon HJ, Bishop AE. Embryonic stem cells. Cell Prolif. 2004;37(1):23–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Poulsom R, Alison MR, Forbes SJ, Wright NA. Adult stem cell plasticity. J Pathol J Pathol Soc G B Irel. 2002;197(4):441–56.

    Google Scholar 

  66. Brignier AC, Gewirtz AM. Embryonic and adult stem cell therapy. J Allergy Clin Immunol. 2010;125(2):S336–44.

    Article  PubMed  Google Scholar 

  67. Takahashi K, Tanabe K, Ohnuki M, et al. Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell. 2007;131(5):861–72. https://doi.org/10.1016/j.cell.2007.11.019.

    Article  CAS  PubMed  Google Scholar 

  68. Takahashi K, Yamanaka S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell. 2006;126(4):663–76. https://doi.org/10.1016/j.cell.2006.07.024.

    Article  CAS  PubMed  Google Scholar 

  69. Edlund H. Pancreatic organogenesis–developmental mechanisms and implications for therapy. Nat Rev Genet. 2002;3(7):524–32. https://doi.org/10.1038/nrg841.

    Article  CAS  PubMed  Google Scholar 

  70. Sakhneny L, Khalifa-Malka L, Landsman L. Pancreas organogenesis: approaches to elucidate the role of epithelial-mesenchymal interactions. In: Seminars in cell & developmental biology, vol. 92, Academic Press; 2019. p. 89–96. https://doi.org/10.1016/j.semcdb.2018.08.012

  71. Schroeder IS. Potential of pluripotent stem cells for diabetes therapy. Curr Diabetes Rep. 2012;12(5):490–8. https://doi.org/10.1007/s11892-012-0292-5.

    Article  Google Scholar 

  72. Lumelsky N, Blondel O, Laeng P, Velasco I, Ravin R, McKay R. Differentiation of embryonic stem cells to insulin-secreting structures similar to pancreatic islets. Science. 2001;292(5520):1389–94.

    Article  CAS  PubMed  Google Scholar 

  73. Ptasznik A, Beattie GM, Mally MI, Cirulli V, Lopez A, Hayek A. Phosphatidylinositol 3-kinase is a negative regulator of cellular differentiation. J Cell Biol. 1997;137(5):1127–36. https://doi.org/10.1083/jcb.137.5.1127.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. D’Amour KA, Bang AG, Eliazer S, Kelly OG, Agulnick AD, Smart NG, Moorman MA, Kroon E, Carpenter MK, Baetge EE. Production of pancreatic hormone–expressing endocrine cells from human embryonic stem cells. Nat Biotechnol. 2006;24(11):1392–401. https://doi.org/10.1038/nbt1259.

    Article  CAS  PubMed  Google Scholar 

  75. Pagliuca FW, Millman JR, Gürtler M, Segel M, Van Dervort A, Ryu JH, Peterson QP, Greiner D, Melton DA. Generation of Functional human pancreatic β cells in vitro. Cell. 2014;159(2):428–39. https://doi.org/10.1016/j.cell.2014.09.040.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Zhang D, Jiang W, Liu M, Sui X, Yin X, Chen S, Shi Y, Deng H. Highly efficient differentiation of human ES cells and iPS cells into mature pancreatic insulin-producing cells. Cell Res. 2009;19(4):429–38. https://doi.org/10.1038/cr.2009.28.

    Article  CAS  PubMed  Google Scholar 

  77. Schroeder IS, Rolletschek A, Blyszczuk P, Kania G, Wobus AM. Differentiation of mouse embryonic stem cells to insulin-producing cells. Nat Protoc. 2006;1(2):495–507. https://doi.org/10.1038/nprot.2006.71.

    Article  CAS  PubMed  Google Scholar 

  78. Vaca P, Martín F, Vegara-Meseguer JM, Rovira JM, Berná G, Soria B. Induction of differentiation of embryonic stem cells into insulin-secreting cells by fetal soluble factors. Stem Cells. 2006;24(2):258–65. https://doi.org/10.1634/stemcells.2005-0058.

    Article  CAS  PubMed  Google Scholar 

  79. Servitja JM, Ferrer J. Transcriptional networks controlling pancreatic development and beta cell function. Diabetologia. 2004;47(4):597–613. https://doi.org/10.1007/s00125-004-1368-9.

    Article  CAS  PubMed  Google Scholar 

  80. Al-Khawaga S, Memon B, Butler AE, Taheri S, Abou-Samra AB, Abdelalim EM. Pathways governing development of stem cell-derived pancreatic β cells: lessons from embryogenesis: development of pancreatic β cells. Biol Rev. 2018;93(1):364–89. https://doi.org/10.1111/brv.12349.

    Article  PubMed  Google Scholar 

  81. Kossugue PM. Diferenciação de células-tronco embrionárias murinas (mESCs) em células produtoras de insulina (IPCs) e caracterização funcional do gene Purkinje Cell Protein 4 (Pcp4) neste processo. 142f. Dissertação de Doutorado – Universidade de São Paulo, mai. 2013.

  82. Candiello J, Grandhi TSP, Goh SK, Vaidya V, Lemmon-Kishi M, Eliato KR, Ros R, Kumta PN, Rege K, Banerjee I. 3D heterogeneous islet organoid generation from human embryonic stem cells using a novel engineered hydrogel platform. Biomaterials. 2018;177:27–39. https://doi.org/10.1016/j.biomaterials.2018.05.031.

    Article  CAS  PubMed  Google Scholar 

  83. Yu HB, Kunarso G, Hong FH, Stanton LW. Zfp206, Oct4, and Sox2 are integrated components of a transcriptional regulatory network in embryonic stem cells. J Biol Chem. 2009;284(45):31327–35. https://doi.org/10.1074/jbc.M109.016162.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Hayashi Y, Caboni L, Das D, et al. Structure-based discovery of NANOG variant with enhanced properties to promote self-renewal and reprogramming of pluripotent stem cells. Proc Natl Acad Sci U S A. 2015;112(15):4666–71. https://doi.org/10.1073/pnas.1502855112.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Weinstein DC, et al. The winged-helix transcription factor HNF-3b is required for notochord development in the mouse embryo. Cell. 1994;78:575–88. https://doi.org/10.1016/0092-8674(94)90523-1.

    Article  CAS  PubMed  Google Scholar 

  86. Ang SL, Rossant J. HNF-3b is essential for node and notochord formation in mouse development. Cell. 1994;78:561–74. https://doi.org/10.1016/0092-8674(94)90522-3.

    Article  CAS  PubMed  Google Scholar 

  87. Wu KL, Gannon M, Peshavaria M, et al. Hepatocyte nuclear factor 3beta is involved in pancreatic beta-cell-specific transcription of the pdx-1 gene. Mol Cell Biol. 1997;17(10):6002–13. https://doi.org/10.1128/mcb.17.10.6002.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Sund NJ, Vatamaniuk MZ, Casey M, et al. Tissue-specific deletion of Foxa2 in pancreatic beta cells results in hyperinsulinemic hypoglycemia. Genes Dev. 2001;15(13):1706–15. https://doi.org/10.1101/gad.901601.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Wang H, Gauthier BR, Hagenfeldt-Johansson KA, Iezzi M, Wollheim CB. Foxa2 (HNF3beta) controls multiple genes implicated in metabolism-secretion coupling of glucose-induced insulin release. J Biol Chem. 2002;277(20):17564–70. https://doi.org/10.1074/jbc.M111037200.

    Article  CAS  PubMed  Google Scholar 

  90. D’Amour KA, Agulnick AD, Eliazer S, Kelly OG, Kroon E, Baetge EE. Efficient differentiation of human embryonic stem cells to definitive endoderm. Nat Biotechnol. 2005;23(12):1534–41. https://doi.org/10.1038/nbt1163.

    Article  CAS  PubMed  Google Scholar 

  91. McGrath KE, Koniski AD, Maltby KM, McGann JK, Palis J. Embryonic expression and function of the chemokine SDF-1 and its receptor, CXCR4. Dev Biol. 1999;213(2):442–56. https://doi.org/10.1006/dbio.1999.9405.

    Article  CAS  PubMed  Google Scholar 

  92. Kanai-Azuma M, Kanai Y, Gad JM, et al. Depletion of definitive gut endoderm in Sox17-null mutant mice. Development. 2002;129(10):2367–79. https://doi.org/10.1242/dev.129.10.2367.

    Article  CAS  PubMed  Google Scholar 

  93. Spence JR, Lange AW, Lin SC, et al. Sox17 regulates organ lineage segregation of ventral foregut progenitor cells. Dev Cell. 2009;17(1):62–74. https://doi.org/10.1016/j.devcel.2009.05.012.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Cardinale V, Wang Y, Carpino G, et al. Multipotent stem/progenitor cells in human biliary tree give rise to hepatocytes, cholangiocytes, and pancreatic islets. Hepatology. 2011;54(6):2159–72. https://doi.org/10.1002/hep.24590.

    Article  CAS  PubMed  Google Scholar 

  95. Jonatan D, Spence JR, Method AM, Kofron M, Sinagoga K, Haataja L, Arvan P, Deutsch GH, Wells JM. Sox17 regulates insulin secretion in the normal and pathologic mouse beta cell. PLoS ONE. 2014;9(8): e104675. https://doi.org/10.1371/journal.pone.0104675.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Taylor BL, Liu F-F, Sander M. Nkx6.1 is essential for maintaining the functional state of pancreatic beta cells. Cell Rep. 2013;4(6):1262–75. https://doi.org/10.1016/j.celrep.2013.08.010.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Jonsson J, et al. Insulin-promoter-factor 1 is required for pancreas development in mice. Nature. 1994;371:606–9. https://doi.org/10.1038/371606a0.

    Article  CAS  PubMed  Google Scholar 

  98. Gao T, McKenna B, Li C, et al. Pdx1 maintains β cell identity and function by repressing an α cell program. Cell Metab. 2014;19:259–71. https://doi.org/10.1016/j.cmet.2013.12.002.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Kawaguchi Y, et al. The role of the transcriptional regulator Ptf1a in converting intestinal to pancreatic progenitors. Nat Genet. 2002;32:128–34. https://doi.org/10.1038/ng959.

    Article  CAS  PubMed  Google Scholar 

  100. Esni F, Ghosh B, Biankin AV, Lin JW, Albert MA, Yu X, MacDonald RJ, Civin CI, Real FX, Pack MA, Ball DW, Leach SD. Notch inhibits Ptf1 function and acinar cell differentiation in developing mouse and zebrafish pancreas. Development. 2004;131:4213–24. https://doi.org/10.1242/dev.01280.

    Article  CAS  PubMed  Google Scholar 

  101. Beres TM, Masui T, Swift GH, Shi L, Henke RM, MacDonald RJ. PTF1 is an organ-specific and Notch-independent basic helix–loop–helix complex containing the mammalian Suppressor of Hairless (RBP-J) or its paralogue, RBP-L. Mol Cell Biol. 2006;26:117–30. https://doi.org/10.1128/MCB.26.1.117-130.2006.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Masui T, Long Q, Beres TM, Magnuson MA, MacDonald RJ. Early pancreatic development requires the vertebrate suppressor of Hairless (RBPJ) in the PTF1 bHLH complex. Genes Dev. 2007;21:2629–43. https://doi.org/10.1101/gad.1575207.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Seymour PA, Freude KK, Tran MN, Mayes EE, Jensen J, Kist R, Scherer G, Sander M. SOX9 is required for maintenance of the pancreatic progenitor cell pool. Proc Natl Acad Sci U S A. 2007;104(6):1865–70. https://doi.org/10.1073/pnas.0609217104.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Seymour PA, Freude KK, Dubois CL, Shih HP, Patel NA, Sander M. A dosage-dependent requirement for Sox9 in pancreatic endocrine cell formation. Dev Biol. 2008;323(1):19–30. https://doi.org/10.1016/j.ydbio.2008.07.034.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Seymour PA, Shih HP, Patel NA, Freude KK, Xie R, Lim CJ, Sander M. A Sox9/Fgf feed-forward loop maintains pancreatic organ identity. Development. 2012;139(18):3363–72. https://doi.org/10.1242/dev.078733.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Shih HP, Kopp JL, Sandhu M, Dubois CL, Seymour PA, Grapin-Botton A, Sander M. A Notch-dependent molecular circuitry initiates pancreatic endocrine and ductal cell differentiation. Development. 2012;139(14):2488–99. https://doi.org/10.1242/dev.078634.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Shih HP, Seymour PA, Patel NA, Xie R, Wang A, Liu PP, Yeo GW, Magnuson MA, Sander M. A gene regulatory network cooperatively controlled by Pdx1 and Sox9 governs lineage allocation of foregut progenitor cells. Cell Rep. 2015;13(2):326–36. https://doi.org/10.1016/j.celrep.2015.08.082.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Portela-Gomes GM, Gayen JR, Grimelius L, Stridsberg M, Mahata SK. The importance of chromogranin A in the development and function of endocrine pancreas. Regul Pept. 2008;151(1–3):19–25. https://doi.org/10.1016/j.regpep.2008.07.005.

    Article  CAS  PubMed  Google Scholar 

  109. Tomita T. Significance of chromogranin A and synaptophysin in pancreatic neuroendocrine tumors. Bosn J Basic Med Sci. 2020;20(3):336.

    CAS  PubMed  PubMed Central  Google Scholar 

  110. Jensen J, Heller RS, Funder-Nielsen T, Pedersen EE, Lindsell C, Weinmaster G, Madsen OD, Serup P. Independent development of pancreatic alpha- and beta-cells from neurogenin3-expressing precursors: a role for the notch pathway in repression of premature differentiation. Diabetes. 2000;49:163–76. https://doi.org/10.2337/diabetes.49.2.163.

    Article  CAS  PubMed  Google Scholar 

  111. Rukstalis JM, Habener JF. Neurogenin3: a master regulator of pancreatic islet differentiation and regeneration. Islets. 2009;1(3):177–84. https://doi.org/10.4161/isl.1.3.9877.

    Article  PubMed  Google Scholar 

  112. Márquez-Aguirre AL, Canales-Aguirre AA, Padilla-Camberos E, Esquivel-Solis H, Díaz-Martínez NE. Development of the endocrine pancreas and novel strategies for β-cell mass restoration and diabetes therapy. Braz J Med Biol Res. 2015;48(9):765–76. https://doi.org/10.1590/1414-431x20154363.

    Article  PubMed  PubMed Central  Google Scholar 

  113. Huang HP, Liu M, El-Hodiri HM, Chu K, Jamrich M, Tsai MJ. Regulation of the pancreatic islet-specific gene BETA2 (neuroD) by neurogenin 3. Mol Cell Biol. 2000;20:3292–307. https://doi.org/10.1128/mcb.20.9.3292-3307.2000.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Breslin MB, Zhu M, Lan MS. NeuroD1/E47 regulates the E-box element of a novel zinc finger transcription factor, IA-1, in developing nervous system. J Biol Chem. 2003;278:38991–7. https://doi.org/10.1074/jbc.M306795200.

    Article  CAS  PubMed  Google Scholar 

  115. Marsich E, Vetere A, Di Piazza M, Tell G, Paoletti S. The PAX6 gene is activated by the basic helix–loop–helix transcription factor NeuroD/BETA2. Biochem J. 2003;376:707–15. https://doi.org/10.1042/bj20031021.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Gierl MS, Karoulias N, Wende H, Strehle M, Birchmeier C. The zinc-finger factor Insm1 (IA-1) is essential for the development of pancreatic beta cells and intestinal endocrine cells. Genes Dev. 2006;20:2465–78. https://doi.org/10.1101/gad.381806.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Sussel L, et al. Mice lacking the homeodomain transcription factor Nkx2.2 have diabetes due to arrested differentiation of pancreatic b cells. Development. 1998;125:2213–21. https://doi.org/10.1242/dev.125.12.2213.

    Article  CAS  PubMed  Google Scholar 

  118. Cissell MA, Zhao L, Sussel L, Henderson E, Stein R. Transcription factor occupancy of the insulin gene in vivo. Evidence for direct regulation by Nkx2.2. J Biol Chem. 2003;278:751–6.

    Article  CAS  PubMed  Google Scholar 

  119. Raum JC, Gerrish K, Artner I, Henderson E, Guo M, Sussel L, Schisler JC, Newgard CB, Stein R. FoxA2, Nkx2.2, and PDX-1 regulate islet beta-cell-specific mafA expression through conserved sequences located between base pairs-8118 and -7750 upstream from the transcription start site. Mol Cell Biol. 2006;26:5735–43. https://doi.org/10.1101/gad.173039.111.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Smith SB, Ee HC, Conners JR, German MS. Paired-homeodomain transcription factor PAX4 acts as a transcriptional repressor in early pancreatic development. Mol Cell Biol. 1999;19:8272–80. https://doi.org/10.1128/MCB.19.12.8272.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Ritz-Laser B, Estreicher A, Gauthier BR, Mamin A, Edlund H, Philippe J. The pancreatic beta-cell-specific transcription factor Pax-4 inhibits glucagon gene expression through Pax-6. Diabetologia. 2002;45(1):97–107. https://doi.org/10.1007/s125-002-8249-9.

    Article  CAS  PubMed  Google Scholar 

  122. Collombat P, Mansouri A, Hecksher-Sorensen J, Serup P, Krull J, Gradwohl G, Gruss P. Opposing actions of Arx and Pax4 in endocrine pancreas development. Genes Dev. 2003;17:2591–603. https://doi.org/10.1101/gad.269003.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Collombat P, Hecksher-Sorensen J, Broccoli V, Krull J, Ponte I, Mundiger T, Smith J, Gruss P, Serup P, Mansouri A. The simultaneous loss of Arx and Pax4 genes promotes a somatostatin-producing cell fate specification at the expense of the alpha- and beta-cell lineages in the mouse endocrine pancreas. Development. 2005;132:2969–80. https://doi.org/10.1242/dev.01870.

    Article  CAS  PubMed  Google Scholar 

  124. Wang Q, Elghazi L, Martin S, Martins I, Srinivasan RS, Geng X, Sleeman M, Collombat P, Houghton J, Sosa-Pineda B. Ghrelin is a novel target of Pax4 in endocrine progenitors of the pancreas and duodenum. Dev Dyn. 2008;237:51–61. https://doi.org/10.1002/dvdy.21379.

    Article  CAS  PubMed  Google Scholar 

  125. Irwin DM. Evolution of the insulin gene: changes in gene number, sequence, and processing. Front Endocrinol. 2021. https://doi.org/10.3389/fendo.2021.649255.

    Article  Google Scholar 

  126. Kataoka K, Han SI, Shioda S, Hirai M, Nishizawa M, Handa H. MafA is a glucose-regulated and pancreatic beta-cell-specific transcriptional activator for the insulin gene. J Biol Chem. 2002;277(51):49903–10. https://doi.org/10.1074/jbc.M206796200.

    Article  CAS  PubMed  Google Scholar 

  127. Artner I, Hang Y, Mazur M, et al. MafA and MafB regulate genes critical to beta-cells in a unique temporal manner. Diabetes. 2010;59(10):2530–9. https://doi.org/10.2337/db10-0190.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Artner I, Blanchi B, Raum JC, Guo M, Kaneko T, Cordes S, Sieweke M, Stein R. MafB is required for islet β cell maturation. Proc Natl Acad Sci. 2007;104(10):3853–8. https://doi.org/10.1073/pnas.0700013104.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Fukumoto H, Seino S, Imura H, et al. Sequence, tissue distribution, and chromosomal localization of mRNA encoding a human glucose transporter-like protein. Proc Natl Acad Sci U S A. 1988;85(15):5434–8. https://doi.org/10.1073/pnas.85.15.5434.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. McCracken KW, Wells JM. Molecular pathways controlling pancreas induction. In: Seminars in cell and developmental biology, vol. 23, Academic Press; 2012. p. 656–662. https://doi.org/10.1016/j.semcdb.2012.06.009

  131. Qadir MMF, Lanzoni G, Ricordi C, Domínguez-Bendala J. Human pancreatic progenitors. In: Transplantation, bioengineering, and regeneration of the endocrine pancreas. Elsevier; 2020. p. 183–200. https://doi.org/10.1016/B978-0-12-814831-0.00013-0.

    Chapter  Google Scholar 

  132. Hebrok M, Kim SK, Melton DA. Notochord repression of endodermal Sonic hedgehog permits pancreas development. Genes Dev. 1998;12(11):1705–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Rodríguez-Martínez G, Molina-Hernández A, Velasco I. Activin a promotes neuronal differentiation of cerebrocortical neural progenitor cells. PLoS ONE. 2012;7(8):e43797. https://doi.org/10.1371/journal.pone.0043797.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Stafford D, Prince VE. Retinoic acid signaling is required for a critical early step in zebrafish pancreatic development. Curr Biol. 2002;12(14):1215–20. https://doi.org/10.1016/S0960-9822(02)00929-6.

    Article  CAS  PubMed  Google Scholar 

  135. Nostro MC, Sarangi F, Ogawa S, Holtzinger A, Corneo B, Li X, Micallef SJ, Park IH, Basford C, Wheeler MB, Daley GQ. Stage-specific signaling through TGFβ family members and WNT regulates patterning and pancreatic specification of human pluripotent stem cells. Development. 2011;138(5):861–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Sherwood RI, Maehr R, Mazzoni EO, Melton DA. Wnt signaling specifies and patterns intestinal endoderm. Mech Dev. 2011;128(7–10):387–400.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Kroon E, Martinson LA, Kadoya K, Bang AG, Kelly OG, Eliazer S, Young H, Richardson M, Smart NG, Cunningham J, Agulnick AD. Pancreatic endoderm derived from human embryonic stem cells generates glucose-responsive insulin-secreting cells in vivo. Nat Biotechnol. 2008;26(4):443–52.

    Article  CAS  PubMed  Google Scholar 

  138. Maehr R, Chen S, Snitow M, Ludwig T, Yagasaki L, Goland R, Leibel RL, Melton DA. Generation of pluripotent stem cells from patients with type 1 diabetes. Proc Natl Acad Sci. 2009;106(37):15768–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Johannesson M. Human embryonic stem cells: directed differentiation into posterior foregut endoderm and a functional assay for definitive endoderm. 2009.

  140. Ameri J, Ståhlberg A, Pedersen J, Johansson JK, Johannesson MM, Artner I, Semb H. FGF2 specifies hESC-derived definitive endoderm into foregut/midgut cell lineages in a concentration-dependent manner. Stem Cells. 2010;28(1):45–56.

    Article  CAS  PubMed  Google Scholar 

  141. Kunisada Y, Tsubooka-Yamazoe N, Shoji M, Hosoya M. Small molecules induce efficient differentiation into insulin-producing cells from human induced pluripotent stem cells. Stem Cell Res. 2012;8(2):274–84.

    Article  CAS  PubMed  Google Scholar 

  142. Ndlovu R, Deng L-C, Jin W, Li X-K, Zhang J-S. Fibroblast growth factor 10 in pancreas development and pancreatic cancer. Front Genet. 2018. https://doi.org/10.3389/fgene.2018.00482.

    Article  PubMed  PubMed Central  Google Scholar 

  143. Bhushan A, Itoh N, Kato S, Thiery JP, Czernichow P, Bellusci S, Scharfmann R. Fgf10 is essential for maintaining the proliferative capacity of epithelial progenitor cells during early pancreatic organogenesis. Development. 2001;128(24):5109–17. https://doi.org/10.1242/dev.128.24.5109.

    Article  CAS  PubMed  Google Scholar 

  144. Memon B, Karam M, Al-Khawaga S, Abdelalim EM. Enhanced differentiation of human pluripotent stem cells into pancreatic progenitors co-expressing PDX1 and NKX61. Stem Cell Res Ther. 2018;9(1):15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Öström M, Loffler KA, Edfalk S, Selander L, Dahl U, Ricordi C, Jeon J, Correa-Medina M, Diez J, Edlund H. Retinoic acid promotes the generation of pancreatic endocrine progenitor cells and their further differentiation into β-cells. PLoS ONE. 2008;3(7):e2841. https://doi.org/10.1371/journal.pone.0002841.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Uzan B, Figeac F, Portha B, Movassat J. Mechanisms of KGF mediated signaling in pancreatic duct cell proliferation and differentiation. PLoS ONE. 2009;4(3): e4734.

    Article  PubMed  PubMed Central  Google Scholar 

  147. Movassat J, Beattie GM, Lopez AD, Portha B, Hayek A. Keratinocyte growth factor and beta-cell differentiation in human fetal pancreatic endocrine precursor cells. Diabetologia. 2003;46(6):822–9.

    Article  CAS  PubMed  Google Scholar 

  148. Shahjalal HM, Abdal Dayem A, Lim KM, Jeon TI, Cho SG. Generation of pancreatic β cells for treatment of diabetes: advances and challenges. Stem Cell Res Ther. 2018;9(1):1–19.

    Article  Google Scholar 

  149. Schulz TC, Young HY, Agulnick AD, Babin MJ, Baetge EE, Bang AG, Bhoumik A, Cepa I, Cesario RM, Haakmeester C, Kadoya K. A scalable system for production of functional pancreatic progenitors from human embryonic stem cells. PLoS ONE. 2012;7(5): e37004.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Velazquez-Garcia S, Valle S, Rosa TC, Takane KK, Demirci C, Alvarez-Perez JC, Mellado-Gil JM, Ernst S, Scott DK, Vasavada RC, Alonso LC. Activation of protein kinase C-ζ in pancreatic β-cells in vivo improves glucose tolerance and induces β-cell expansion via mTOR activation. Diabetes. 2011;60(10):2546–59.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Vasavada RC, Wang L, Fujinaka Y, Takane KK, Rosa TC, Mellado-Gil JM, Friedman PA, Garcia-Ocaña A. Protein kinase C-ζ activation markedly enhances β-cell proliferation: an essential role in growth factor-mediated β-cell mitogenesis. Diabetes. 2007;56(11):2732–43.

    Article  CAS  PubMed  Google Scholar 

  152. Russ HA, Parent AV, Ringler JJ, Hennings TG, Nair GG, Shveygert M, Guo T, Puri S, Haataja L, Cirulli V, Blelloch R. Controlled induction of human pancreatic progenitors produces functional beta-like cells in vitro. EMBO J. 2015;34(13):1759–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Mason MN, Mahoney MJ. Inhibition of gamma-secretase activity promotes differentiation of embryonic pancreatic precursor cells into functional islet-like clusters in poly(ethylene glycol) hydrogel culture. Tissue Eng Part A. 2010;16(8):2593–603. https://doi.org/10.1089/ten.tea.2010.0015.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Rezania A, Bruin JE, Arora P, Rubin A, Batushansky I, Asadi A, O’Dwyer S, Mojibian NM, Tobias Albrecht Y, Yang HC, Johnson JD, Kieffer TJ. Reversal of diabetes with insulin-producing cells derived in vitro from human pluripotent stem cells. Nat Biotechnol. 2014;32(11):1121–33. https://doi.org/10.1038/nbt.3033.

    Article  CAS  PubMed  Google Scholar 

  155. Millman JR, Xie C, Van Dervort A, Gürtler M, Pagliuca FW, Melton DA. Generation of stem cell-derived β-cells from patients with type 1 diabetes. Nat Commun. 2016;7(1):1–9.

    Google Scholar 

  156. Chen Y, Pan FC, Brandes N, Afelik S, Sölter M, Pieler T. Retinoic acid signaling is essential for pancreas development and promotes endocrine at the expense of exocrine cell differentiation in Xenopus. Dev Biol. 2004;271(1):144–60. https://doi.org/10.1016/j.ydbio.2004.03.030.

    Article  CAS  PubMed  Google Scholar 

  157. Gao Y, Zhang R, Dai S, Zhang X, Li X, Bai C. Role of TGF-β/smad pathway in the transcription of pancreas-specific genes during beta cell differentiation. Front Cell Dev Biol. 2019. https://doi.org/10.3389/fcell.2019.00351.

    Article  PubMed  PubMed Central  Google Scholar 

  158. Aiello V, Moreno-Asso A, Servitja JM, Martin M. Thyroid hormones promote endocrine differentiation at expenses of exocrine tissue. Exp Cell Res. 2014;322(2):236–48.

    Article  CAS  PubMed  Google Scholar 

  159. Mfopou JK, Chen B, Sui L, Sermon K, Bouwens L. Recent advances and prospects in the differentiation of pancreatic cells from human embryonic stem cells. Diabetes. 2010;59(9):2094–101. https://doi.org/10.2337/db10-0439.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Abdelalim EM. Advances and challenges in the differentiation of pluripotent stem cells into pancreatic β cells. World J Stem Cells. 2015;7(1):174. https://doi.org/10.4252/wjsc.v7.i1.174.

    Article  PubMed  PubMed Central  Google Scholar 

  161. Wandzioch E, Zaret KS. Dynamic signaling network for the specification of embryonic pancreas and liver progenitors. Science. 2009;324(5935):1707–10. https://doi.org/10.1126/science.1174497.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  162. Sui L, Geens M, Sermon K, Bouwens L, Mfopou JK. Role of BMP signaling in pancreatic progenitor differentiation from human embryonic stem cells. Stem Cell Rev Rep. 2013;9(5):569–77. https://doi.org/10.1007/s12015-013-9435-6.

    Article  CAS  PubMed  Google Scholar 

  163. Chen C, Xie Z, Shen Y, Xia SF. The roles of thyroid and thyroid hormone in pancreas: physiology and pathology. Int J Endocrinol. 2018;2018:1–14. https://doi.org/10.1155/2018/2861034.

    Article  CAS  Google Scholar 

  164. Velazco-Cruz L, Song J, Maxwell KG, Goedegebuure MM, Augsornworawat P, Hogrebe NJ, Millman JR. Acquisition of dynamic function in human stem cell-derived β cells. Stem Cell Rep. 2019;12(2):351–65.

    Article  CAS  Google Scholar 

  165. Shing Y, Christofori G, Hanahan D, Ono Y, Sasada R, Igarashi K, Folkman J. Betacellulin-A novel mitogen from pancreatic-beta tumor-cells. In: Molecular biology of the cell, vol. 3. Publ office, 9650 Rockville Pike, Bethesda, MD 20814: Amer Soc Cell Biology; 1992, p. A28–A28

  166. Liu SH, Lee LT. Efficient differentiation of mouse embryonic stem cells into insulin-producing cells. Exp Diabetes Res. 2012. https://doi.org/10.1155/2012/201295.

    Article  PubMed  PubMed Central  Google Scholar 

  167. Woodford C, Yin T, Chang HH, Regeenes R, Vellanki RN, Mohan H, Zandstra PW. Nicotinamide promotes differentiation of pancreatic endocrine progenitors from human pluripotent stem cells through poly (ADP-ribose) polymerase inhibition. bioRxiv. 2020; https://doi.org/10.1101/2020.04.21.05295

  168. Ye DZ, Tai M-H, Linning KD, Csaba Szabo L, Olson K. MafA expression and insulin promoter activity are induced by nicotinamide and related compounds in INS-1 pancreatic β-cells. Diabetes. 2006;55(3):742–50. https://doi.org/10.2337/diabetes.55.03.06.db05-0653.

    Article  CAS  PubMed  Google Scholar 

  169. Thowfeequ S, Ralphs KL, Yu WY, Slack JMW, Tosh D. Betacellulin inhibits amylase and glucagon production and promotes beta cell differentiation in mouse embryonic pancreas. Diabetologia. 2007;50(8):1688–97.

    Article  CAS  PubMed  Google Scholar 

  170. Kieffer TJ, Habener JF. The glucagon-like peptides. Endocr Rev. 1999;20(6):876–913. https://doi.org/10.1210/edrv.20.6.0385.

    Article  CAS  PubMed  Google Scholar 

  171. Buteau J, Roduit R, Susini S, Prentki M. Glucagon-like peptide-1 promotes DNA synthesis, activates phosphatidylinositol 3-kinase and increases transcription factor pancreatic and duodenal homeobox gene 1 (PDX-1) DNA binding activity in beta (INS-1)-cells. Diabetologia. 1999;42(7):856–64. https://doi.org/10.1007/s001250051238.

    Article  CAS  PubMed  Google Scholar 

  172. Aguayo-Mazzucato C, Zavacki AM, Marinelarena A, Hollister-Lock J, El Khattabi I, Marsili A, Weir GC, Arun Sharma P, Larsen R, Bonner-Weir S. Thyroid hormone promotes postnatal rat pancreatic β-cell development and glucose-responsive insulin secretion through MAFA. Diabetes. 2013;62(5):1569–80. https://doi.org/10.2337/db12-0849.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Brissova M, Shostak A, Shiota M, Wiebe PO, Poffenberger G, Kantz J, Chen Z, Chad Carr W, Jerome G, Jin Chen H, Baldwin S, Nicholson W, Bader DM, Jetton T, Gannon M, Powers AC. Pancreatic islet production of vascular endothelial growth factor-a is essential for islet vascularization, revascularization, and function. Diabetes. 2006;55(11):2974–85. https://doi.org/10.2337/db06-0690.

    Article  CAS  PubMed  Google Scholar 

  174. Prasadan K, Shiota C, Xiangwei X, Ricks D, Fusco J, Gittes G. A synopsis of factors regulating beta cell development and beta cell mass. Cell Mol Life Sci. 2016;73(19):3623–37. https://doi.org/10.1007/s00018-016-2231-0.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Weber LM, Hayda KN, Anseth KS. Cell–matrix interactions improve β-cell survival and insulin secretion in three-dimensional culture. Tissue Eng Part A. 2008;14(12):1959–68. https://doi.org/10.1089/ten.tea.2007.0238.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Leite AR, Corrêa-Giannella ML, Dagli MLZ, Fortes MAZ, Vegas VMT, Giannella-Neto D. Fibronectin and laminin induce expression of islet cell markers in hepatic oval cells in culture. Cell Tissue Res. 2007;327(3):529–37. https://doi.org/10.1007/s00441-006-0340-z.

    Article  CAS  PubMed  Google Scholar 

  177. Alberto Llacua L, Faas MM, de Vos P. Extracellular matrix molecules and their potential contribution to the function of transplanted pancreatic islets. Diabetologia. 2018;61(6):1261–72. https://doi.org/10.1007/s00125-017-4524-8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Stendahl JC, Kaufman DB, Stupp SI. Extracellular matrix in pancreatic islets: relevance to scaffold design and transplantation. Cell Transp. 2009;18(1):1–12. https://doi.org/10.3727/096368909788237195.

    Article  Google Scholar 

  179. Jiang J, Au M, Lu K, Eshpeter A, Korbutt G, Fisk G, Majumdar AS. Generation of insulin-producing islet-like clusters from human embryonic stem cells. Stem cells. 2007;25(8):1940–53. https://doi.org/10.1634/stemcells.2006-0761.

    Article  CAS  PubMed  Google Scholar 

  180. Jiang W, Shi Y, Zhao D, Chen S, Yong J, Zhang J, Qing T, Sun X, Zhang P, Ding M, Li D, Deng H. In vitro derivation of functional insulin-producing cells from human embryonic stem cells. Cell Res. 2007;17(4):333–44. https://doi.org/10.1038/cr.2007.28.

    Article  CAS  PubMed  Google Scholar 

  181. Takeuchi H, Nakatsuji N, Suemori H. Endodermal differentiation of human pluripotent stem cells to insulin-producing cells in 3D culture. Sci Rep. 2014;4:4488.

    Article  PubMed  PubMed Central  Google Scholar 

  182. Agulnick AD, Ambruzs DM, Moorman MA, Bhoumik A, Cesario RM, Payne JK, Kelly JR, Haakmeester C, Srijemac R, Wilson AZ, Kerr J, Frazier MA, Kroon EJ, D’Amour KA. Insulin-producing endocrine cells differentiated in vitro from human embryonic stem cells function in macroencapsulation devices in vivo. Stem Cells Transl Med. 2015;4(10):1214–22. https://doi.org/10.5966/sctm.2015-0079.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  183. Kim Y, Kim H, Ko UH, Youjin O, Lim A, Sohn J-W, Shin JH, Kim H, Han Y-M. Islet-like organoids derived from human pluripotent stem cells efficiently function in the glucose responsiveness in vitro and in vivo. Sci Rep. 2016. https://doi.org/10.1038/srep35145.

    Article  PubMed  PubMed Central  Google Scholar 

  184. Southard SM, Kotipatruni RP, Rust WL. Generation and selection of pluripotent stem cells for robust differentiation to insulin-secreting cells capable of reversing diabetes in rodents. PLoS ONE. 2018;13(9):e0203126. https://doi.org/10.1371/journal.pone.0203126.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Yoshihara E, O’Connor C, Gasser E, Wei Z, Oh TG, Tseng TW, Wang D, Cayabyab F, Dai Y, Yu RT, Liddle C. Immune-evasive human islet-like organoids ameliorate diabetes. Nature. 2020. https://doi.org/10.1038/s41586-020-2631-z.

    Article  PubMed  PubMed Central  Google Scholar 

  186. Boland BB, Rhodes CJ, Grimsby JS. The dynamic plasticity of insulin production in β-cells. Mol Metab. 2017;6(9):958–73. https://doi.org/10.1016/j.molmet.2017.04.010.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Seino S, Shibasaki T, Minami K. Dynamics of insulin secretion and the clinical implications for obesity and diabetes. J Clin Investig. 2011;121(6):2118–25. https://doi.org/10.1172/JCI45680.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  188. Hrvatin S, O’Donnell CW, Deng F, Millman JR, Pagliuca FW, DiIorio P, Rezania A, Gifford DK, Melton DA. Differentiated human stem cells resemble fetal, not adult, β cells. Proc Nat Aca Sci. 2014;111(8):3038–43. https://doi.org/10.1073/pnas.1400709111.

    Article  CAS  Google Scholar 

  189. Hogrebe NJ, Augsornworawat P, Maxwell KG, Velazco-Cruz L, Millman JR. Targeting the cytoskeleton to direct pancreatic differentiation of human pluripotent stem cells. Nat Biotechnol. 2020;38(4):460–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  190. Hogrebe NJ, Maxwell KG, Augsornworawat P, Millman JR. Generation of insulin-producing pancreatic β cells from multiple human stem cell lines. Nat Protoc. 2021;16(9):4109–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  191. Nair GG, Liu JS, Russ HA, Tran S, Saxton MS, Chen R, et al. Recapitulating endocrine cell clustering in culture promotes maturation of human stem-cell-derived β cells. Nat Cell Biol. 2019;21(2):263–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  192. Yoshihara E, Wei Z, Lin CS, Fang S, Ahmadian M, Kida Y, Tseng T, Dai Y, Yu RT, Liddle C, Atkins AR, Downes M, Evans RM. ERRγ Is required for the metabolic maturation of therapeutically functional glucose-responsive β cells. Cell Metab. 2016;23(4):622–34. https://doi.org/10.1016/j.cmet.2016.03.005.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  193. Veres A, Faust AL, Bushnell HL, Engquist EN, Kenty JHR, Harb G, Poh YC, Sintov E, Gürtler M, Pagliuca FW, Peterson QP. Charting cellular identity during human in vitro β-cell differentiation. Nature. 2019;569(7756):368–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Jansson L, Barbu A, Bodin B, Drott CJ, Espes D, Gao X, Grapensparr L, Källskog Ö, Lau J, Liljebäck H, Palm F, Quach M, Sandberg M, Strömberg V, Ullsten S, Carlsson P-O. Pancreatic islet blood flow and its measurement. Upsala J Med Sci. 2016;121(2):81–95. https://doi.org/10.3109/03009734.2016.1164769.

    Article  PubMed  PubMed Central  Google Scholar 

  195. Olerud J, Johansson Å, Carlsson PO. Vascular niche of pancreatic islets. Exp Rev Endocrinol Metab. 2009;4(5):481–91.

    Article  CAS  Google Scholar 

  196. Talavera-Adame D, Wu G, He Y, Ng TT, Gupta A, Kurtovic S, Hwang JY, Farkas DL, Dafoe DC. Endothelial cells in co-culture enhance embryonic stem cell differentiation to pancreatic progenitors and insulin-producing cells through BMP signaling. Stem Cell Rev Rep. 2011;7(3):532–54395. https://doi.org/10.1007/s12015-011-9232-z.

    Article  CAS  PubMed  Google Scholar 

  197. Weizman A, Michael I, Wiesel-Motiuk N, Rezania A, Levenberg S. The effect of endothelial cells on hESC-derived pancreatic progenitors in a 3D environment. Biomater Sci. 2014;2(11):1706–14. https://doi.org/10.1039/C4BM00304G.

    Article  CAS  PubMed  Google Scholar 

  198. Salem V, Silva LD, Suba K, Georgiadou E, Gharavy SNM, Akhtar N, Carrat G. Leader β-cells coordinate Ca 2+ dynamics across pancreatic islets in vivo. Nat Metab. 2019;1(6):615–29.

    Article  CAS  PubMed  Google Scholar 

  199. Johnston NR, Mitchell RK, Haythorne E, Pessoa MP, Semplici F, Ferrer J, Berishvili E. Beta cell hubs dictate pancreatic islet responses to glucose. Cell Metab. 2016;24(3):389–401. https://doi.org/10.1016/j.cmet.2016.06.020.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  200. Bader E, Migliorini A, Gegg M, Moruzzi N, Gerdes J, Roscioni SS, Bakhti M, Brandl E, Irmler M, Beckers J, Aichler M, Feuchtinger A, Leitzinger C, Zischka H, Wang-Sattler R, Jastroch M, Tschöp M, Machicao F, Staiger H, Häring H-U, Chmelova H, Chouinard JA, Oskolkov N, Korsgren O, Speier S, Lickert H. Identification of proliferative and mature β-cells in the islets of Langerhans. Nature. 2016;535(7612):430–4. https://doi.org/10.1038/nature18624.

    Article  CAS  PubMed  Google Scholar 

  201. Dorrell C, Schug J, Canaday PS, Russ HA, Tarlow BD, Grompe MT, Horton T, Hebrok M, Streeter PR, Kaestner KH, Grompe M. Human islets contain four distinct subtypes of β cells. Nat Commun. 2016. https://doi.org/10.1038/ncomms11756.

    Article  PubMed  PubMed Central  Google Scholar 

  202. Millman JR, Pagliuca FW. Autologous pluripotent stem cell–derived β-like cells for diabetes cellular therapy. Diabetes. 2017;66(5):1111–20. https://doi.org/10.2337/db16-1406.

    Article  CAS  PubMed  Google Scholar 

  203. Elsie Zahr R, Molano D, Pileggi A, Ichii H, Jose SS, Bocca N, An W, Gonzalez-Quintana J, Fraker C, Ricordi C, Inverardi L. Rapamycin impairs in vivo proliferation of islet beta-cells. Transplantation. 2007;84(12):1576–83. https://doi.org/10.1097/01.tp.0000296035.48728.28.

    Article  CAS  PubMed  Google Scholar 

  204. Øzbay LA, Smidt K, Mortensen DM, Carstens J, Jørgensen KA, Rungby J. Cyclosporin and tacrolimus impair insulin secretion and transcriptional regulation in INS-1E beta-cells: calcineurin inhibitor induced diabetes. Br J Pharmacol. 2011;162(1):136–46. https://doi.org/10.1111/j.1476-5381.2010.01018.x.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  205. Uchizono Y, Iwase M, Nakamura U, Sasaki N, Goto D, Iida M. Tacrolimus impairment of insulin secretion in isolated rat islets occurs at multiple distal sites in stimulus-secretion coupling. Endocrinology. 2004;145(5):2264–72. https://doi.org/10.1210/en.2003-1152.

    Article  CAS  PubMed  Google Scholar 

  206. Sackett SD, Rodriguez A, Odorico JS. The nexus of stem cell-derived beta-cells and genome engineering. Rev Diabet Stud. 2017;14(1):39.

    Article  PubMed  PubMed Central  Google Scholar 

  207. Maria-Engler SS, Mares-Guia M, Correa MLC, Oliveira EMC, Aita CAM, Krogh K, Genzini T, Miranda MP, Ribeiro M, Vilela L, Noronha IL, Eliaschewitz FG, Sogayar MC. Microencapsulation and tissue engineering as an alternative treatment of diabetes. Braz J Med Biol Res. 2001;34(6):691–7. https://doi.org/10.1590/S0100-879X2001000600001.

    Article  CAS  PubMed  Google Scholar 

  208. Vegas AJ, Veiseh O, Gürtler M, Millman JR, Pagliuca FW, Bader AR, Doloff JC, Li J, Chen M, Olejnik K, Tam HH, Jhunjhunwala S, Langan E, Aresta-Dasilva S, Gandham S, McGarrigle JJ, Bochenek MA, Hollister-Lock J, Oberholzer J, Greiner DL, Weir GC, Melton DA, Langer R, Anderson DG. Long-term glycemic control using polymer-encapsulated human stem cell–derived beta cells in immune-competent mice. Nat Med. 2016;22(3):306–11. https://doi.org/10.1038/nm.4030.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  209. Bochenek MA, Veiseh O, Vegas AJ, McGarrigle JJ, Qi M, Marchese E, Omami M, Doloff JC, Mendoza-Elias J, Nourmohammadzadeh M, Khan A, Yeh C-C, Xing Y, Isa D, Ghani S, Li J, Landry C, Bader AR, Olejnik K, Chen M, Hollister-Lock J, Wang Y, Greiner DL, Weir GC, Strand BL, Anne MA, Rokstad IL, Langer R, Anderson DG, Oberholzer J. Alginate encapsulation as long-term immune protection of allogeneic pancreatic islet cells transplanted into the omental bursa of macaques. Nat Biomed Eng. 2018;2(11):810–21. https://doi.org/10.1038/s41551-018-0275-1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  210. Sigilon Therapeutics, Inc. Safety and Efficacy of Encapsulated Allogeneic FVIII Cell Therapy in Haemophilia A. 2020. (ClinicalTrials.gov Identifier: NCT04541628).

  211. Mayhew CN, Wells JM. Converting human pluripotent stem cells into β-cells: recent advances and future challenges. Curr Opin Organ Transp. 2010;15(1):54–60. https://doi.org/10.1097/MOT.0b013e3283337e1c.

    Article  Google Scholar 

  212. Rezania A, Bruin JE, Xu J, Narayan K, Fox JK, O’Neil JJ, Kieffer TJ. Enrichment of human embryonic stem cell-derived NKX6. 1-expressing pancreatic progenitor cells accelerates the maturation of insulin-secreting cells in vivo. Stem Cells. 2013;31(11):2432–42.

    Article  CAS  PubMed  Google Scholar 

  213. Bruin JE, Rezania A, Jean X, Narayan K, Fox JK, O’Neil JJ, Kieffer TJ. Maturation and function of human embryonic stem cell-derived pancreatic progenitors in macroencapsulation devices following transplant into mice. Diabetologia. 2013;56(9):1987–98. https://doi.org/10.1007/s00125-013-2955-4.

    Article  PubMed  Google Scholar 

  214. Cito M, Pellegrini S, Piemonti L, Sordi V. The potential and challenges of alternative sources of β cells for the cure of type 1 diabetes. Endocr Connect. 2018;7(3):R114–25. https://doi.org/10.1530/EC-18-0012.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  215. Augsornworawat P, Velazco-Cruz L, Song J, Millman JR. A hydrogel platform for in vitro three dimensional assembly of human stem cell-derived islet cells and endothelial cells. Acta Biomater. 2019;97:272–80. https://doi.org/10.1016/j.actbio.2019.08.031.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  216. Tao T, Wang Y, Chen W, Li Z, Su W, Guo Y, Deng P, Qin J. Engineering human islet organoids from iPSCs using an organ-on-chip platform. Lab Chip. 2019;19(6):948–58. https://doi.org/10.1039/C8LC01298A.

    Article  CAS  PubMed  Google Scholar 

  217. Wan J, Huang Y, Zhou P, Guo Y, Wu C, Zhu S, et al. Culture of iPSCs derived pancreatic β-like cells in vitro using decellularized pancreatic scaffolds: a preliminary trial. BioMed Res Int. 2017. https://doi.org/10.1155/2017/4276928.

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors would like to thank the Brazilian funding agencies FAPESP (São Paulo Research Foundation) and CNPq (National research Council of Science and technology), the Cell and Molecular Therapy Center (NUCEL) from the University of São Paulo and the Department of Biochemistry, Chemistry Institute of the University of São Paulo.

Funding

This work was supported by grants from FAPESP (São Paulo Research Foundation)—FAPESP Thematic Grant No. 2016/05311-2, pre-doctoral fellowship to IBBS (Process No. 2019/21935-4), undergraduate fellowship to CHK (Process No. 2019/22267-5) and the Brazilian National Council for Science and Technology (CNPq), with research Grant No. 465656/2014-5 to INCT-Regenera and a pre-doctoral fellowship to VPC (Process No. 140269/2021-5).

Author information

Authors and Affiliations

Authors

Contributions

IBBS and CHK performed the literature search and wrote the manuscript with the supervision of MCS. VPC assembled tables and searched for current therapeutic alternatives for Type 1 Diabetes. MSC reviewed all versions of the manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Mari Cleide Sogayar.

Ethics declarations

Consent for publication

Not applicable.

Human and animal rights and informed consent

This article does not contain any studies with human or animal subjects performed by any of the authors.

Competing interests

All authors certify that they have no affiliations with or involvement in any organization or entity with any financial interest or non-financial interest in the subject matter or materials discussed in this manuscript.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Silva, I.B.B., Kimura, C.H., Colantoni, V.P. et al. Stem cells differentiation into insulin-producing cells (IPCs): recent advances and current challenges. Stem Cell Res Ther 13, 309 (2022). https://doi.org/10.1186/s13287-022-02977-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13287-022-02977-y

Keywords