- Review
- Open access
- Published:
Scaffold-free endocrine tissue engineering: role of islet organization and implications in type 1 diabetes
BMC Endocrine Disorders volume 25, Article number: 107 (2025)
Abstract
Type 1 diabetes (T1D) is a chronic hyperglycemia disorder emerging from beta-cell (insulin secreting cells of the pancreas) targeted autoimmunity. As the blood glucose levels significantly increase and the insulin secretion is gradually lost, the entire body suffers from the complications. Although various advances in the insulin analogs, blood glucose monitoring and insulin application practices have been achieved in the last few decades, a cure for the disease is not obtained. Alternatively, pancreas/islet transplantation is an attractive therapeutic approach based on the patient prognosis, yet this treatment is also limited mainly by donor shortage, life-long use of immunosuppressive drugs and risk of disease transmission. In research and clinics, such drawbacks are addressed by the endocrine tissue engineering of the pancreas. One arm of this engineering is scaffold-free models which often utilize highly developed cell-cell junctions, soluble factors and 3D arrangement of islets with the cellular heterogeneity to prepare the transplant formulations. In this review, taking T1D as a model autoimmune disease, techniques to produce so-called pseudoislets and their applications are studied in detail with the aim of understanding the role of mimicry and pointing out the promising efforts which can be translated from benchside to bedside to achieve exogenous insulin-free patient treatment. Likewise, these developments in the pseudoislet formation are tools for the research to elucidate underlying mechanisms in pancreas (patho)biology, as platforms to screen drugs and to introduce immunoisolation barrier-based hybrid strategies.
Graphical Abstract

Background
Pancreas is an organ in the abdominal cavity with a dual function through hormones to regulate certain aspects by the endocrine part and through the digestive pancreatic juice by the exocrine part [1]. Damage to this organ assisted by genetic factors under particular conditions is tied to or triggers health problems such as pancreatitis, neoplasms, pancreatic exocrine insufficiency, cystic fibrosis, Shwachman-Diamond Syndrome and diabetes associated with the exocrine and/or endocrine components [2]. Among this diverse disease set, diabetes is associated with chronic hyperglycemia which is the main consequence of lack of insulin (type 1 diabetes, T1D) or loss of the insulin production correlated with the insulin resistance (type 2 diabetes, T2D). As a complex disease originating from a dynamic interplay of causes contributed by distinct genetic and environmental factors, T1D is driven by the loss of non-self and self-discrimination for the beta-cells leading to an aggressive immune attack against these cells [3, 4].
In the genetic and environmental background of T1D pathogenesis, genetic criteria include human leukocyte antigen (HLA) and non-HLA loci. HLA genes are encoded in chromosome 6p21 and are grouped into class I and class II. Among these, especially, certain class II genes in HLA-DR and HLA-DQ have prominent roles to link T1D to the heredity [3]. For example, it has been demonstrated that HLA-DR3-DQ2 and/or HLA-DR4-DQ8 presence is the greatest risk factor, as the case study by Undlien et al. revealed that only 10% of the T1D patients did not have these haplotypes [5]. Additionally, the risk for islet autoimmunity highly correlates with siblings carrying HLA-DR3/4-DQ8 haplotype shared with proband siblings. The risk for these siblings was accounted for 85% by age 15 and lowered to 20% when the siblings did not share the same haplotype [6]. Patients carrying HLA-DR3-DQ2/DR3-DQ2 or DR3-DQ2/DR4-DQ8 genotypes were reported to have an increased risk of T1D and other diseases such as celiac disease concurrently [7]. Conversely, some of the HLA family members such as HLA-DR15-DQ6 were discovered as protective against T1D [8, 9], but cases with the failure of protection are reported [10]. Non-HLA genes associated with T1D involve in the immune system cell development and differentiation, as well as various signaling pathways with altered expression rates of soluble molecules or receptors such as insulin, protein tyrosine phosphatase non-receptor type 22 (PTPN22), cytotoxic T-lymphocyte antigen 4, interleukin (IL)-10 and interleukin 2 receptor subunit alpha. All these factors drive the loss of non-self and self-discrimination for the beta-cells leading to an aggressive immune attack [3]. Moreover, environmental factors such as gluten in the diet [11], obesity [12, 13], vitamin D deficiency [14, 15], infections including COVID-19 [16,17,18,19], gut microbiota [20, 21] and drugs [22, 23] are suggested to contribute to the pathophysiology of T1D.
After diagnosis, the main goal of the diabetes treatment is shaped around achieving normoglycemia. Due to destruction of the beta-cells, T1D treatment heavily depends on multiple daily combinatorial insulin injections [24,25,26,27]. In severe cases, beta-cell replacement therapy is applied where only pancreas/islet, simultaneous pancreas/islet and kidney or pancreas/islet after kidney transplantation is implemented into the treatment [24]. Among the pancreas and the islet transplantation, whole organ transplantation is the clinically applied strategy for the insulin independence for T1D patients whereas islet transplantation is a method which is still in clinical trials. However, donor source and eligibility criteria, surgical complications, contraindications to surgery (e.g. viral infections, arterial diseases, high body-mass index, alcohol usage, smoking) and life-long use of immunosuppressive drugs influence the distribution and the success of pancreas transplants [28, 29]. Alternatively, islet transplantation can be the solution to deal with some limitations of pancreas transplant. In this procedure, isolation of individual islets from the donor pancreas is required before transplantation to the patients. However, graft survival depends on the protection of the transplanted cells [30]. To address this bottleneck, an immunosuppressive regime is highly utilized in organ transplantations, yet this strategy of immunosuppression causes adverse effects on the transplanted cells as well as the body. To eliminate this, in 2000s, Edmonton protocol was established by Shapiro et al. that employs glucocorticoid-free immunosuppressive protocol of sirolimus, low-dose tacrolimus, and daclizumab (a monoclonal antibody against the IL-2 receptor) and, was put into the practice in clinical trials [31]. However, donor shortage and source, and related issues such as disease transmission [32], requirement of multiple transplants [33] and risks of transplantation [34] limit widespread application of the islet transplantation in the clinics.
To the clinics as well as to the research, artificial islets can be a possible choice to eliminate several of these drawbacks. Constructed with recent technologies, these systems are widely applied to the endocrine tissue research, especially as an intermediate between preclinic and clinics [35]. This review focuses on the most frequent scaffold-free design called “pseudoislet” and their current applications in T1D research. The ability of these 3D cell assemblies to resemble certain native pancreas features makes them promising candidates for the beta-cell-based studies. Here, as the first step, cellular heterogeneity and cell-cell contact in the pancreas are described to provide the basis for the cell types and the interactions in in vitro endocrine mimics of the pancreas. Next step involves an overview of the current field of the pseudoislets with the assembly techniques and their applications in pancreas biology and related fields.
Cellular environment and communications of the pancreas
Endocrine cells
An analysis of the pancreatic tissue architecture yields two compartments with distinct metabolic functions: Endocrine and exocrine parts. Endocrine part is crucial for the regulation of blood glucose and related metabolic activities [36]. In this section of the pancreas, islets of Langerhans (pancreatic islets) are the main cell clusters which are associated with hormone secretion. Cells in these clusters are mainly divided into five types (with core secretions): Alpha (α)-cells (glucagon), beta (β)-cells (insulin, C-peptide and amylin), delta (δ)-cells (somatostatin), epsilon (ε)-cells (ghrelin) and gamma (γ)-cells (pancreatic polypeptide) (Fig. 1A) [1, 37, 38].
Organization in pancreas. (A) Pancreas, its compartments and microenvironment. Pancreas has two compartments: Endocrine and exocrine pancreas. Endocrine pancreas contains islets wrapped by peri-islet basement membrane. In islets, 5 different endocrine cells are present: Alpha-cells, beta-cells, delta-cells, epsilon-cells and gamma-cells. Exocrine pancreas is formed by acinar cells and ducts that contain ductal cells and centroacinar cells. In addition, resident-immune system cells and PSCs, as well as nerve fibers are found in the pancreas tissue. The pancreas also accommodates blood vessels. In figure as example, capillaries are shown. In general, capillaries are made of endothelial cells facing the blood and they are covered by pericytes and basement membrane. The acellular part of the pancreas is ECM that is rich in proteins, glycosaminoglycans and proteoglycans. (B) Polarity determinants for beta-cells. Beta-cells express Dlg and scribble in different regions, but their expression is enriched in lateral surfaces. Par3 is abundant in the apical domain away from vasculature. The capillary interface/vascular face is marked with the presynaptic scaffold proteins of liprin, RIM2, ELKS and piccolo. (C) Proteins on beta-cells used to form junctions for cell-cell interactions. ZO-1, claudins and occludin containing tight junctions, connexin-36-based gap junctions provide intercellular communication between beta-cells. Adherens junctions established between beta-cell and beta-cell and, between beta-cell and alpha-cell contain several cadherin family members and N-CAM. Apart from these junctions, EphA receptor tyrosine kinases and their ephrin-A ligands are located on beta-cells and alpha-cells for cellular communication and insulin secretion. Transcellular neuroligin-2 is an important protein of beta-cells to control insulin secretion. Dlg: Discs large protein, ECM: Extracellular matrix, N-CAM: Neural cell adhesion molecule, Par3: Partitioning defective 3 homologue, PSC: Pancreatic stellate cell, RIM2: Rab3-interacting protein. In labeling of endocrine cells for (A), the shapes and sizes of the cell representations are not relevant to the actual morphology and size of the cells. Figure 1 contains images from Servier Medical Art, which were modified in color and size where necessary. Servier Medical Art (smart.servier.com) is licensed under a Creative Commons Attribution 4.0 International License (CC BY 4.0) (https://creativecommons.org/licenses/by/4.0/)
Among these secretions, insulin has the central role in the glucose homeostasis and is reported to participate in the diverse mechanisms ranging from the bone growth to the endothelial cell function [39]. Besides to the role of insulin-copackaged amylin in the suppression of the glucagon secretion and reduced food intake, potential consequences of it in the adiposity, neurogenesis, nociception and maternal behaviors are suggested [40]. Although C-peptide is the one of the processing products of the preproinsulin, it has capacity to reduce certain adverse effects of the diabetes on the cardiovascular system, nervous system and kidneys [41]. Furthermore, it is possible to conclude that C-peptide is a promising target to reverse some severe side effect of the T1D, as the improvement of the beta-cell responsiveness to the hyperglycemia and the alpha-cell responsiveness to the hypoglycemia were recorded in T1D patients with the large residual C-peptide levels [42]. The other major hormone of the pancreas, glucagon, cooperates with insulin for the management of circulating blood glucose level, but its effect is opposite to insulin. It has actions such as the increase in the glomerular filtration, reduced food intake and rise in lipid oxidation in the muscles [43]. Somatostatin is another essential factor in the proper pancreas function that manipulates various processes in the central and peripheral tissues [44], as Li et al. demonstrated that the ablation of the somatostatin producing cells ends with the severe hypoglycemia and the neonatal death of the mice [45].
Overall, glycemic control is achieved by central and peripheral mechanisms. Postprandial blood glucose level activates glucose stimulated insulin secretion (GSIS) that inhibits hepatic glucose production through insulin receptor activation on hepatocytes and indirect inhibition of glucagon secretion from alpha-cells. GSIS also activates glucose uptake by tissues including muscle and adipose tissue. Activated by the nutrients, gastrointestinal intestinal tract endocrine cells release incretins that increase insulin release. In addition, the central nervous system contributes to the blood glucose management by hypothalamic-pituitary-adrenal axis and autonomic nervous system. In this mechanism, cortisol stimulates insulin secretion and sympathetic nervous system inhibits GSIS, stimulates glucagon release and hepatic glucose release. In contrast, parasympathetic nervous system can increase insulin and glucagon secretions, and activation of this system in postprandial conditions advances glucose uptake and storage [46,47,48,49]. Another hormone of islets, somatostatin, exhibits increased secretion when glucose level is elevated in the environment [50, 51].
An examination of the epsilon-cell derived ghrelin activity reveals distinct actions in the physiological processes, most notably on the glucose and energy metabolism [52]. In the body, the primary source of ghrelin is stomach, and it acts as “hunger hormone” by regulating appetite and controls various functions including sleep-wake cycle, lipogenesis and cardiac activities [53,54,55]. As a secretion of the islets, pancreatic polypeptide serves as a peptide hormone with the purposes such as the inhibition of the gastric emptying and sensitizing the liver to the insulin [56]. In addition to their systemic tasks, these islet hormones are destined for paracrine signaling. Somatostatin inhibits the insulin and glucagon release, and insulin negatively regulates glucagon secretion. Yet, glucagon is a known inducer for the insulin secretion and the insulin activates the somatostatin release [57, 58]. There is also evidence for observing no effect of insulin on somatostatin secretion. This discrepancy can be due to experimental models and species-dependent islet-cell organization [59, 60]. Expression of the ghrelin appears to be inhibited by the pancreatic polypeptide [56], but one of the key features of this hormone is the inhibition of the insulin and somatostatin [52].
Within the islets, localization and number of each cell type vary, as for instance, it has been previously reported that in the mouse islets, alpha-cells reside in the periphery whereas beta-cells occupy the central region. In contrast, human islets have alpha-cells localized throughout the islets, even making contacts with the intra-islet blood vessels [61]. Bosco et al. revealed that this cellular distribution is unique to the large human islets and, the smaller islets had similar core-mantle distribution as in the rodent islets, yet in cultured human islets, central localization of the beta-cells increases [62]. The dominant cell type for the human and the rodent islets is the beta-cell that accounts for at least 50% of the total population. Other abundant cell type is the alpha-cell with 38% and 18% of the endocrine cells in the human and mouse islets [61]. Within the population of the endocrine cells, transcriptional and functional heterogeneity exist which are reflected in the hormone secretion, protein profile, cell-cell communication, and response to the stimulators [63,64,65,66,67,68,69]. Soluble factor and ion-based cellular communication is universal for the islets, but also islet morphology is a key to achieve the proper insulin response to the glucose stimulation, as the islets dispersed into the single cells show an impairment of the insulin secretion due to destroyed direct cell-cell interactions [70].
Despite, in the T1D, beta-cells are destroyed by the immune attack, thus, at the time of diagnosis, up to 70–95% of the beta-cell reduction is accounted for the clinical symptoms. This beta-cell mass distribution is heterogenous among the patients and may not correspond to the severity of the disease, but significant loss occurs in T1D. In contrast, meta-analysis results indicate that the loss as low as 40% is sufficient to manifest the symptoms [71,72,73]. After beta-cell loss, changes in the alpha-cell behavior have been reported where controversial evidence is present in decrease/increase of alpha-cell number and glucagon levels [74, 75]. Basically, this injury of the beta-cells enriches alpha-cells and causes functional impairment. These alpha-cells lose the expression of alpha-cell markers of MAFB and ARX, as well as they start to express the beta-cell marker NKX6.1. Furthermore, the events leading to the alpha-cell related complications depend on the changes in the expression of ion channels, cAMP-mediated pathways, and vesicle trafficking proteins. Thus, electrical activity as well as glucagon release weaken [76, 77]. To counteract the effects of the lowered beta-cell abundance, a recent study has shown that immunostained alpha-cells are positive for the insulin and the glucagon with PDX transcription factor expression in the diabetic mice [78]. In contrast, another possibility includes that MAFB and ARX marker transcriptions are preserved, and the genes for the glucagon biosynthesis and processing are upregulated in the T1D patients [79].
Although low in number, delta-cells are central to regulate the endocrine dynamics of the pancreas. Equipped with filopodia, these cells can reach out not only the beta-cells in close-proximity, but also distant beta-cells in the islets and exocytose the somatostatin from the soma and the filopodia. In an attempt to contact with the beta-cells, filopodia length increases in the prediabetic delta-cells [80]. In the childhood and adult onset diabetes, the number of alpha- and delta-cells are reported to be constant while due to a major loss of the beta-cells in the head and tail regions, islet size was measured to be smaller [81]. The islet size and size distribution can also remain similar to the healthy pancreas, but significant reduction in the islet density of recent onset T1D and long-standing T1D occur [82]. Previously, a significant improvement in the gamma-cell % in the islets was noted in the T1D patients [83]. Consequently, alpha-cell and delta-cell content may vary in the T1D that eventually alters other endocrine activities of the pancreas.
Cell-cell contact based signaling in the islets
Beta-cells are polarized to develop appropriate responses in the islets, as they express basolateral polarity determinants of discs large protein (Dlg) and scribble in different regions, but these proteins are enriched in lateral regions and, partitioning defective 3 homologue (Par3) is localized in the apical domain away from the vasculature. The capillary interface/vascular face is marked with the presynaptic scaffold proteins of liprin, Rab3-interacting protein (RIM2), ELKS and piccolo (Fig. 1B) [84,85,86,87,88]. These proteins guide beta-cells to position themselves towards the vasculature and in fact, the polarization of beta-cells is important for the secretory behavior. For instance, glucose transporter (GLUT)-2 is laterally positioned and zona occludens (ZO)-1 containing tight junctions are in the apical domain showing that the polarization can be a directing clue for the secretion [84]. Experiments on the mice and in vitro tests show that ELKS participates in the polarized insulin secretion via L-type voltage-dependent calcium channel (VDCC) opening and first-phase insulin secretion through direct binding to the VDCC-β subunit. These events occur at the vascular face of the beta-cells where ELKS is contained [85].
Gap junctions (Fig. 1C) are exceptional intercellular channels which provide coupling of two cells metabolically and electrically. Islets express gap junction subunits called connexins and among all, connexin-36 has outstanding effects on the beta-cells [89, 90]. These gap junctions connect beta-cells, especially connexin-36 based junctions are functional in the electrical coupling and promoting the protection of the beta-cells against cytokines and streptozotocin (STZ). Recently, Farnsworth et al. demonstrated that the gap junction coupling decreases at the T1D onset and this change along with the modifications of Ca2+ amplifies the susceptibility to apoptosis. Yet, connexin-36 gap junction coupling in NOD mouse can be protective against the early onset of the T1D without progress in the long-term survival, islet function and beta-cell death [91]. In the cell-cell contact, gap junction coupling of the beta- and delta-cells is reported to have behavioral consequences, as the beta-cell signaling through these junctions controls the delta-cell [Ca2+]i fluctuations and somatostatin secretion [92]. Beta-cells also contain channels called pannexin-1 which are similar to the gap junction hemichannels, but they release ATP [93]. This extracellular ATP binds to receptors such as P2X purinoceptor 7 (P2X7R) mobilizing Ca2+ towards the intracellular pathway to secrete the insulin [94].
Adherens junctions established with E- and N-cadherins contribute to the beta-cell functionality with several potential mechanisms (Fig. 1C) [95]. In polarized beta-cells, catenins and cadherins control the insulin granule pool, as the cadherin or p120 catenin knockdown shows a trend towards increased basal insulin levels. This is achieved by the cadherin mediated downstream signaling of the catenins where a possible role of p120 catenin in the actin network remodeling along with the α- and β-catenins involved, and that the glucose increases the expression of cadherin and p120 catenins. Thus, E- and N-cadherins promote GSIS from the single beta-cells [96, 97]. Human islets express E- and N-cadherin, which are mostly present on the beta-cells with some positivity on the alpha-cells for N-cadherin. This cadherin expression pattern in the cellular interactions can sustain the islet viability. In fact, Parnaud et al. proved that the adhesion of the single beta-cells to the cadherin substrate is directly involved in the protection against cytokine induced and general apoptosis [98]. Desmoglein-2 is another member of the cadherin family that shows expression on various cells within the pancreas, including the cells within the islet (highest in the beta-cells), the vasculature and the exocrine tissue. Similar to E-cadherin, it can participate in the protection against STZ-induced cell death in vivo and cytokine induced apoptosis in vitro [99]. As a member of Ig superfamily, neural cell adhesion molecule (N-CAM) is expressed in the islets for the control of cell segregation, insulin and glucagon exocytosis management through the actin reorganization coupled with the cadherin activity [100,101,102].
Cadherins also contribute to the native islet organization in the pancreas organogenesis. In pancreatic development, E-cadherins of beta-cells are essential to cluster these cells into islets. For instance, expression of truncated E-cadherin disrupts assembly of beta-cell into aggregates and alpha-cells organize into islet-like clusters without beta-cells [103, 104]. Besides, this direct cell-cell contact provided by E-cadherin lowers beta-cell proliferation [105, 106]. In islets, the endocrine cell segregation is affected by expression of cadherins and N-CAM. It is suggested that these cell surface molecules create alterations in surface tension leading to localization of each cell type into distinct regions. Loss of one cadherin can be compensated by others, but loss of N-CAM for instance, results in alpha-cell and beta-cell mixing and cadherin redistribution [107]. Further, at ∼E10.5, cap cells contain less E-cadherin and this lowered amount is associated with the start of pancreatic branching and tip formation [108].
Various EphA receptor tyrosine kinases and their ephrin-A ligands are expressed by the pancreatic islets and the beta-cells (Fig. 1C), and EphA forward signaling inhibits the insulin secretion while ephrin-A reverse signaling stimulates the insulin secretion [109]. In T1D, some of the juxtacrine signaling such as ephrin-A ̶ EphA4/7 between the beta-cells and the alpha-cells are most probably lost relieving the F-actin based inhibitory effect of this interaction on the alpha-cells, further contributing to the alpha-cell dysfunction in the T1D [110]. Finally, claudins, occludin, and ZO-1 as the tight junction units expressed in the islets (111, 112]are important. For example, claudin-4 is associated with the maturation and dedifferentiation of the beta-cells [111].
Some proteins in the inhibitory synapses of the central nervous system have been discovered to be expressed in the islets with the capacity to manipulate insulin secretion. Transcellular neuroligin-2 is reported to be an insulin release stimulator as indicated in a co-culture system experiment where beta cells/islet cells were seeded on HEK293 cells transfected to express full-length neuroligin -2. Underlying mechanism for this effect includes reduction of the insulin granule docking as shown in the pancreatic islets of neuroligin-2 knock-out mice (33% decrease). Its common partner, neurexin, is shown to be dispensable from this process raising the question of the presence of other unassessed interaction partners on the beta-cells [113].
Islet capillaries and cell-cell communication
Pancreatic islets are rich in blood vessels to deliver the hormones to target cells as well as to sustain themselves [114, 115]. The intra-islet capillaries are formed by the blood contacting endothelial cells with a cover of pericytes and a vascular basement membrane. Further, the whole islet is wrapped with the peri-islet basement membrane isolating the islet from the surrounding tissue [116,117,118]. This morphological assembly as well as its organization within the islets are useful tools to deliver insulin to the endothelial cells.
One of the primary controllers of insulin secretion from beta-cells is the high glucose concentration in the bloodstream. Following glucose transport into the pancreatic beta-cells through glucose transporters (GLUT-1 and 3 human beta-cells and GLUT-2 for rodent beta-cells), glucose is phosphorylated by glucokinase and is delivered to glycolysis pathway [119]. Energy level as ATP amount generated by this pathway as well as TCA cycle with upcoming oxidative phosphorylation increase ATP/ADP ratio intracellularly. This stimulus activates ATP-sensitive potassium (KATP) channels to close leading to membrane depolarization. In turn, this electrical alteration in the membrane charge opens voltage-gated Ca2+ channels of beta-cells resulting in high Ca2+ influx into the cell. Next stage involves exocytosis of insulin granules from readily releasable pool with profound insulin increase in the blood. Another internal storage pool of beta-cells is also induced to sustain insulin level postprandially resulting in biphasic release. Anaplerotic pathways for TCA cycle intermediates, amino acids (e.g. lysine, glutamine, alanine), pentose monophosphate shunt and fatty acid metabolism further fuel the granule release [37, 120,121,122,123,124]. The released insulin travels through fenestration of the endothelium and by caveolae-mediated transendothelial transport and reaches blood circulation [125]. To facilitate the transfer process, density of the fenestrae in the islet capillaries are much higher than the exocrine tissue to deliver the hormones and other molecules to the blood circulation [126, 127].
At physiological concentration of insulin, endothelial cells uptake this hormone from vessel lumen through receptor mediated endocytosis with the insulin receptors [128,129,130]. Other possible routes also exist for insulin transport by endothelial cells such as receptor-mediated endocytosis (insulin receptor and insulin-like growth factor-1 receptor), fluid-phase endocytosis [129] and transcytosis [131, 132]. For this reason, the beta-cells mostly contact with the capillaries [133].
In the blood vessels, endothelial cells are connected to each other through transmembrane proteins called junctions. These endothelial cell junctions provide contact points between two cells and regulate processes including vessel permeability and stability, endothelial cell continuity, mechanosensation, leukocyte diapedesis, and inhibitory signaling for the proliferation [134,135,136,137,138,139]. The capillary junctions are mainly classified into two groups: Adherens and tight junctions. At adherens junctions, vascular endothelial cadherins (VE-cadherins) originating from the distinct membranes link two endothelial cells. In contrast, at tight junctions, adhesion is controlled by the claudin family (claudins 1, 5 and 12), occludin, junctional adhesion molecules (JAMs) and endothelial cell selective adhesion molecule (ESAM). Nectin–afadin complex is reported to contribute to both junctions. Platelet endothelial cell adhesion molecule-1 (PECAM-1) is another junction unit that connects the endothelial cells [139, 140].
Additionally, endothelial cells and beta-cells are coupled physically and through soluble factors. Connexin-43 of the endothelial cells and connexin-36 of the beta-cells are suggested to form the connexon connecting these cells [102]. Endothelin-1, hepatocyte growth factor, thrombospondin-1 and connective tissue growth factor secreted by the endothelial cells stimulate insulin release, serve in the development of the pancreas and support the beta-cell viability. Similarly, factors such as insulin, vascular endothelial growth factor A (VEGF-A) and angiopoietin (ANG)-1 of the beta-cells positively regulate the endothelial cell survival and physiological activities [141,142,143,144]. As an effective way to communicate with the neighboring cells, islets also produce extracellular vesicles. The surfaceome of these vesicles include alpha6- and beta1-integrin, CD44, ICAM-1 (intercellular adhesion molecule-1), CD31 and glucagon-like peptide-1 receptor (GLP1R) and the constituent molecules of these vesicles are such as insulin, C-peptide, CD63, programmed death-ligand 1 (PD-L1), Argonaute-2 and proteins of cell proliferation. Besides, distinctive cellular metabolic alterations occur in the endothelial cells, as these vesicles are equipped with various mRNAs for the endothelial cell activation and angiogenesis, insulin production and signal transduction, glycogen/lipid/protein synthesis and peroxisome proliferator-activated receptor pathway. Moreover, they are enriched with miRNAs such as miR-375, miR-126 and miR-296. Thus, upon transfer to the endothelial cells, pro-angiogenic (ANG-1, VEGF-alpha, vascular endothelial growth factor receptor (VEGFR) -1, -2) and anti-apoptotic factors (Bcl-2) are upregulated, and anti-angiogenic (thrombospondin1) and pro-apoptotic molecules (BAD) are downregulated pointing out the anti-apoptotic and angiogenic effects of the islets on the endothelial cells [143, 145,146,147].
Pericytes as mural vascular cells in islet capillaries cover 40% of the vasculature and, under sympathetic adrenergic control, they contract the blood vessels altering capillary diameter and flow. Conversely, high glucose in the circulation inhibits pericytes, thus dilation and high blood flow occur. These cells also aid beta-cell function and density [148,149,150]. They control beta-cell activity as well as the establishment of both islet basement membrane and interstitial matrix, as they secrete collagen IV, laminins (especially α2 and α4 laminins), proteoglycans, fibronectin, nidogen, and hyaluronan [118]. During T1D, as these cells become dysfunctional and they presume a myofibroblastic phenotype, coverage of the vessels by these cells decreases. These events trigger the impairment of the balance between vasoconstriction and vasodilation in the islet capillaries which affects the islet perfusion, possibly creating a hypoxic environment that activates the expression of the certain vascular modeling genes such as the vasodilator peptide adrenomedullin, migration and metabolism genes in the islets [150]. Additionally, islet vascular density has been significantly built-up in the T1D patients, as upregulation of the genes associated with the vasculature and angiogenesis whereas downregulation of the epithelial-to-mesenchymal transition genes have been reported. Interestingly, these islet blood vessels are shorter and smaller in diameter compared to the healthy pancreas capillaries [151, 152].
In general, as pericytes wrap around the capillaries, endothelial cells and pericytes present cell-cell contact mediated interactions and soluble factor signaling to regulate the healthy vessel activity. These cells interact with the extracellular matrix (ECM) and, through the regions called adhesion plaques and peg-and-socket interactions to manipulate the mechanical and chemical sensing in different regions in the body. At the cell-cell contact intersection, adhesion plaques and peg-and-socket regions closely connect two cells. Adhesion plaques are rich in fibronectin and anchor the pericytes to the bottom epithelium whereas connexin-43 based gap junctions and N-cadherin mediated adherens junctions are in the peg-and-socket regions. In these regions, various soluble factors and their corresponding receptors are also localized. For instance, VEGF-A and VEGFR-1, platelet-derived growth factor (PDGF)-BB and platelet-derived growth factor receptor-β (PDGFRβ), ANG-1 or ANG-2 and Tie2, Notch3 and Jagged1, Eph-B2 and Ephrin-B2 and, transforming growth factor (TGF)-β and transforming growth factor-β receptor (TGF-βR) communications develop for the diverse applications from the quiescence to the perfusion [153,154,155,156].
Exocrine pancreas and cell-cell contacts
Exocrine pancreas is responsible for the digestion process, since it secretes enzymes such as amylase, pancreatic lipase and trypsin into the pancreatic ducts [1, 38]. In this part of the pancreas, acinar cells produce and secrete digestive enzymes with a NaCl supplied fluid. Ductal cells are arranged as an epithelial lining to form intercalating, intralobular and interlobular ducts directing the enzymes to the duodenum. Ductal cells are also the sources of the protective layer of the mucus. An extension of the duct lined by the centroacinar cells reaches the center of the acini at the end of the intercalating duct. The ducts are responsible for the secretion of bicarbonate along proximal to distal ducts to prevent the enzyme aggregation while Cl− concentration decreases [157,158,159,160]. Unlike the healthy pancreas, T1D exocrine pancreas is damaged due to the abnormal metabolic transformation. Pancreatic amylase, lipase, retinol binding protein and prealbumin decrease in the long-standing T1D patients reflecting the outcome of this process [161]. The level of exocrine enzymatic activity such as lipase and trypsinogen are proposed to be serological markers in the staging of T1D and the size changes in the pancreas [162]. Significantly lower number of the acinar cells occupy the T1D patient pancreas [163] whereas no acinar atrophy is observed in some other patients, yet a possible link to the pancreatic cancer is established [164]. Further, these cells can be reprogrammed into beta-cells. On the basis of the recent observations, the findings on the ductal cell genetic reprogramming through Neurog3, MAFA, PDX1 and PAX6 for sourcing beta-cells showed encouraging therapeutic implications for the insulin-secreting cell transplants [165].
Primary gap junction hemichannels on the acinar cells are based on the connexin-32 and -26 [166, 167]. The evidence regarding the necessity of these junctions for the physiological function is obtained from, for instance, homozygous connexon-32 deficient mice which displayed increased enzymatic output (e.g. amylase) compared to the controls. This model paved the possibility of the inhibitory effect of the gap junctions on the number of mobilized acinar cells for the enzyme secretion, as upon release of this effect through the loss of connexon-32 channels, the acinar cell number and the amylase release are improved [168]. Similar to the endocrine pancreas, E-cadherin is expressed in the exocrine part and the loss of it severely disturbs the exocrine pancreas architecture and the downstream pathways such as Wnt signaling and Hippo/YAP signal transduction pathway, and the mutant acinar cells undergo the acinar-to-ductal metaplasia [169]. The presence of the tight junctions of the ductal cells regulates the ductal polarity and the paracellular transport which are crucial for the functional state of the ductal cells. Molecular composition of the ductal and acinar tight junctions includes the claudin family members such as claudin -1, -4, -7, occludin, JAM-A, ZO-1, ZO-2, and tricellulin [170,171,172].
Immune system cells and others
Remaining space available for the cells in the pancreas is occupied by other cell types such as pancreatic stellate cells (PSCs). PSCs are quiescent, vitamin A-rich lipid storing cells in the pancreas. They are precursors for the cancer-associated fibroblasts (CAFs) and in their activated state, they presume myofibroblast-like phenotype. Along with the CAFs, PSCs secrete ECM proteins (e.g. fibronectin, collagen I, hyaluronan) and other soluble factors (e.g. TGF-β, IL-1β, IL-6, leukaemia inhibitory factor (LIF), PDGF, VEGF), as well as release extracellular vesicles with the nucleic acid and other molecules (e.g. miR-5703, miR616-3p, miR-4456, lactate, glutamine, threonine, phenylalanine, stearate, palmitate) in the diverse stages of the cancer [173,174,175,176]. For instance, fibronectin deposited by the PSCs promotes gemcitabine resistance through ERK1/2 phosphorylation [177]. During T2D, insulin fuels activated PSC growth and ECM deposition resulting in the massive fibrosis [178]. In diabetic conditions, hyperglycemia, hyperinsulinemia and renin–angiotensin system activate PSC proliferation and fibrotic activity. Interference of these pathways is suggested to improve the treatment outcomes in the clinics [179, 180]. Furthermore, activated PSCs can cause beta-cell exhaustion [181]. However, PSCs are reported to act as a determinant to sustain the islet cell phenotype and survival in vitro as well as promote islet graft activity in vivo. In a recent study, it is proposed that activated PSCs can return to the quiescent state after transplantation and, the nonfibrogenic effect of these cells should be considered. Direct co-culture experiments where islets were seeded on top of PSCs suggest that interactions with PSCs improved islet function and viability, slowed down morphology loss and sustained islet markers of insulin, PDX and glucagon. In the same study, PSCs transplantation with syngeneic islets showed small but significant decrease in blood glucose level, and oral glucose tolerance tests showed improved glycemia and insulin secretion function of islets showing their potential to reach normoglycemia [182]. Activated PSCs were reported to support exocrine cell differentiation in human pancreas [183]. In autoantibody+ patients, fibronectin, periostin and collagen subunit encoding genes are expressed largely in the stellate cells, yet T1D patient stellate cells are negatively affected, and their phenotype shifts to the immunomodulatory one [150]. Contribution of these cells to the development of pathologies such as pancreatic cancer and pancreatitis is also well-documented with soluble factors, extracellular vesicles, ECM remodeling and synthesis, and resistance to chemotherapeutic drugs and inhibition of anti-tumor response of immune system can be built-up by PSCs [184,185,186,187,188,189,190,191,192,193,194,195]. Moreover, among PSCs, functional heterogeneity promotes pancreatic cancer development as both subgroups, myofibroblastic and inflammatory PSCs, complement each other to progress pancreatic cancer [193].
In a healthy body, immune system protection against the pathogen invasion is indispensable from the survival and this network also assists homeostasis. To accomplish these, various tissues maintain resident immune system cells and upon requirement, migrating members strengthen the defense, thus physiological function. One of the resident constituents of the immune division in the pancreas is macrophages which are observed in both endocrine and exocrine sections. They make up 98% of the islet-resident CD45+ cells in the mice [196, 197]. Pancreatic macrophages are derived from the yolk sac and the bone marrow and play dual roles in physiological situations and transition into the pathological state. Polarization of the macrophages is the fundamental concept to determine their action in the beneficial-detrimental spectrum. Polarized macrophages can presume M1 or M2 phenotype depending on the factors they are exposed to. In a healthy pancreas, both sets are present and in fact, these macrophages do not strictly follow the definition of M1 and M2 macrophages. Macrophages in the pancreas are main suppliers of IL-1β, IL-6, IL-12, TNF (tumor necrosis factor)-α, IFN (interferon)-γ and nitric oxide (NO) which are largely responsible for the beta-cell dysfunction/death and the insulin resistance. Nevertheless, other macrophages generally express large amounts of IL-10, TGF-β1, epidermal growth factor (EGF), hepatocyte growth factor (HGF), PDGF, insulin-like growth factor (IGF)-1, matrix metalloproteinases (MMPs) and Arg1 to protect and regenerate the beta-cells, for the pancreas organogenesis and the insulin sensitivity [197,198,199,200,201,202]. In return, islets modulate macrophage monitoring by shaping the pseudopodia formation [203].
In the diseased state, macrophage derived factors such as RANTES and TNF-α genetically reprogram acinar cells contributing to the progression of the acinar-to-ductal metaplasia [204]. Another aspect of paracrine signaling involving macrophages activates PSCs during fibrotic growth [205]. Besides to their role in the T cell activation and autoantigen presentation [198], macrophages secrete CXCR2 (CXC chemokine receptor 2) ligands as well as induce beta-cells by IL-1β to secrete CXCL (C-X-C motif chemokine ligand)1 and CXCL2 to recruit the neutrophils to the pancreas in T1D [206]. Although macrophages have detrimental effects on the onset and progression of diabetes, they exercise beneficial activities for the pancreas. As described by Xiao et al., M2 macrophages interact with the beta-cells through TGF-β1 and EGF signaling. Signaling by these soluble factors triggers SMAD7-CyclinD axis and exclusion of p27 from the nucleus which switches on the beta-cell proliferation. However, another branch of TGF-β1 pathway that activates SMAD2 is inhibitory on the beta-cells. The countering route mediated by EGF activation inhibits SMAD2 activity [207]. Further, macrophages are advantageous actions for the beta-cells such as islet remodeling, vascularization and insulin secretion [208]. All these data indicate that support and attack are both modulated by macrophage-derived factors and their signaling pathways in T1D.
Dendritic cells are antigen presenting cell arm of the hematopoietic cells between innate and adaptive immune system in the multicellular organisms. Tissue-resident subgroup of the dendritic cells in nonlymphoid tissues often migrate to the lymph node to induce T-cell activity [209]. In the pancreas, they are localized adjacent to the intra-islet blood vessels, mostly between the vessel wall and the beta-cells. The dendrites of these cells are highly active moving within the beta-cells for screening. Dendritic cells engage with the beta-cell derived antigens for MHC-dependent presentation to the T cells in the healthy pancreas suggesting role of this dendritic cell activity requiring further investigation in the development of the autoimmune diseases [210]. In the acute pancreatitis of mouse, dendritic cells secrete proinflammatory cytokines, activate adaptive immune system as well as participate debris and by-product removal to support the pancreatic viability [211].
Among lymphoid lineage, CD3+ T cells are the most abundant being 79% of this population and, CD8+ cells constitute 78% of T cells most of which are memory cells in the islets [212]. As main players in the defense against pathogens, establishment of memory functions, immunoregulation and screening for cancer cells, T cells regulate homeostasis in the body [213] and during ageing, T cells accumulate in the non-diabetic mouse pancreas [214]. In the pancreas, PD-L1 (Programmed death-ligand 1) ̶ Programmed cell death protein-1 (PD-1) binding between macrophages and CD8+ memory T cells regulate immune homeostasis [215] as well as these cells represent 43% of CD8+ T cells in the insulitis lesions in the T1D [216]. 8–20% of CD4+ T cells in the pancreas are Tregs which are the suppressive cells to allow immune-tolerance and, have distinct subsets secreting IL-10 and TGF-β, preventing inflammation and suppressing the Th1 inflammation. Tregs in the pancreatic lymph nodes secrete the same factors and perform the same tasks, but also secretion of IFN-γ is noteworthy [217]. Occasionally, natural killer cells and B cells are found in the pancreas [212]. These group of cells (dendritic cells, monocytes, B cells, T cells and innate-lymphoid cells) also exist in the exocrine pancreas [197].
Finally, the involvement of the nervous system to the pancreas activity should not be underestimated. The enrichment of the parasympathetic afferent fibers in the pancreas participates in the mechanical and chemical sensing in homeostasis, pain and inflammation. Besides, sympathetic afferent nerves are used for mechanical and chemical sensing by transmitting the signals from vasculature, acini, islets and intrapancreatic ganglia. The efferent network is directly involved in the pancreas activity. For instance, factors such as vasoactive intestinal polypeptide and gastrin-releasing peptide secreted by the nerve fibers manipulate the insulin secretion [218, 219]. Moreover, sympathetic neurons cooperate for the proper islet assembly, GSIS and the beta-cell migration during the development [220].
Scaffold-free strategies for the design of an artificial islet
Considering the etiology of T1D, cell-based therapy and experimental models aim to simulate the native environment as well as seek strategies to advance the insulin secretion, lower cell/transplant number, find alternative insulin-secreting cell sources and provide companion cells to stimulate the viability and the vascularization. These can be achieved by the co-transplantation [221], cellular coating [222, 223], differentiation/alternative beta-cell sources [224, 225], development of ECM mimics [226,227,228] and immunoisolation barriers [229,230,231,232]. One of the other largely studied alternative strategies is multicellular assemblies which can achieve the targeted aims with only cells and when necessary, in combination with genetic engineering. This innovative approach unlocks various applications from disease modeling to cell/drug based T1D therapy (Table 1) [233, 234].
In general, artificial islets are produced from single beta-cells or with the combinations of the endocrine and non-endocrine cells of the pancreas. In literature, spheroid [249], islet spheroid [237], mixed (multicellular) spheroid [241, 248], spheroid cluster [246], homospheroid and heterospheroid [250], organoid [260], pseudoislet [274], SC-islet (organoid) [262], engineered pseudoislets [285], heterotypic pseudoislet [286], islet-like cell aggregate [292], aggregate [298], composite aggregate [299], coaggregate [297], cluster [306], islet cell cluster [303] and islet-like cluster [305] are common terminologies for the assemblies of islet-like 3D structures. In Table 1, for simplicity, spheroid, pseudoislet, cluster and aggregate terms are used, and for the terms which do not fit these classification, the nomenclature in the original paper has been used such as vascularized islet [309], islet-like structure [310], SC-islet [311], stem-cell derived islet [312] and islet microtissue [313]. Nevertheless, resulting 3D systems are fundamental for T1D research and clinics. In this review, they will collectively be referred to as pseudoislets for further simplification. It is noteworthy to mention that scaffold-free systems in the diabetes are not limited with the pseudoislet systems, but also cell-sheet engineering is another branch of this technology [314,315,316], yet here, focus will be on the pseudoislet systems. Various techniques are reported for the formation of pseudoislets (Table 1and Fig. 2). Although each has unique benefits and drawbacks [233, 250,251,252, 255, 263, 270, 278, 288, 302, 304, 311, 317,318,319,320], efficient formation of the pseudoislets is routinely performed for the diabetes research worldwide.
The most common pseudoislet assembly techniques. Although less widely applied, techniques such as magnetic levitation, assembly through layer-by-layer coated single cells, encapsulation and bioprinted molds are also reported (not shown in the figure). Figure 2 contains images from Servier Medical Art, which were modified in color, shape and size where necessary. Servier Medical Art (smart.servier.com) is licensed under a Creative Commons Attribution 4.0 International License (CC BY 4.0) (https://creativecommons.org/licenses/by/4.0/)
Applications of artificial Islets
Islet transplantation has the capacity to extend the number of treated diabetic patients, but inadequacy of cell source still limits its widespread application as cell-based therapy. One of the factors involved in the insulin-free treatment is 3D configuration which influences stimulation and secretion of insulin and islet viability, as cell-cell interactions regulate diverse beta-cell aspects including phenotype maintenance, insulin expression and secretion, and viability. As a result, 3D culture systems are increasingly employed in beta-cell replacement therapy in vitro and in vivo [238, 251, 294, 317, 321]. In literature, mainly combination of transformed cells has been studied for the development of pseudoislets (Table 1) which are intended for research rather than clinics. In research, pseudoislet strategy controls the cell number to reach the euglycemia [268] and prevents unlimited growth after transplantation [235, 238] which are essential in the clinical applications. Ability to be produced in the large scale up to 1 L [255] and incorporation of not only insulin secreting cells, but also other endocrine cells which secrete glucagon and somatostatin [239] is achievable with the pseudoislets. The islet size in the resulting beta-cell culture is another indicator to achieve optimum viable pseudoislets with the endocrine response. Established pseudoislet formation techniques can also result in hypoxia-induced cell death in especially large pseudoislets [268]. However, it can be controlled in the pseudoislets, thus hypoxia-induced death can be minimized [303].
A unique feature of these pseudoislets is their ability to resemble native islet organization. Walker et al. reported that reaggregated human islet cells maintain cellular identity, as the beta-cell (PDX1, NKX6.1) and the alpha cell transcription factor markers (MAFB, ARX) are preserved. Pseudoislet proliferation, apoptosis, architecture and dynamic insulin secretion profile were analogous to the native human islets which made them ideal candidates, for example, to investigate Gi and Gq receptors on the alpha- and beta-cell activities [277]. Also, human islet cell pseudoislets formed via hanging drop (Fig. 3A) reflect the uniform size of pseudoislets in contrast to human islets, as DNA content distribution is in narrow range and improved responsiveness to glucose can be achieved compared to native islets (8–13-fold stimulation vs. 2-fold stimulation) (Fig. 3B) [274]. Alternatively, microwells with specific geometry can allow spherical assembly of beta-cells with uniform pseudoislet sizes in contrast to native islets (Fig. 3C-D) [272]. Additionally, rearrangements as 3D pseudoislets successfully strengthen the physiological processes in the cells such as human clonal EndoC-βH1 and EndoC-βH5 cells, and overexpression of gap junction units of connexin 36 in the rodent INS-1 832/13 cells positively alters stimulation index [321].
Examples of several methods to form pseudoislets and contributions to artificial islet research. (A) Design of human pseudoislets by hanging drop technique with different initial cell numbers. (B) (Left) Comparison of DNA content of native islets and human pseudoislets (cultured for 2 days and initial cell number: 2000 cells). (Middle) Fold stimulation of native islets and human pseudoislets in GSIS. (Right) Insulin secretion (% of insulin content) of native islets and human pseudoislets in 2.8 mM and 12 mM glucose. (C) Human pseudoislets formed via centrifugation of a suspension of single cells in square-pyramidal microwells for 0–72 h. (D) Controlled, consistent pseudoislet size in square-pyramidal wells with initial cell numbers of 1000, 750 and 500 human islet cells per well. Scale bars: 200 μm. (E) Differentiation protocol for SC-islets from human embryonic stem cell line, H1. (F) Immunostaining of SC-islets obtained via protocol in (E) for insulin, glucagon, somatostatin and SLC18A1, and DNA staining at stage 7 culture. Scale bars: 100 μm. (G) (Left) Brightfield and immunofluorescence images. (Right) Quantification of CD31+ cells of native human islets, pseudoislets (EndoC-βH3 cells and HUVECs) by spontaneous aggregation and magnetic levitation. Scale bars: 50 μm for native tissue and 200 μm for pseudoislets. GSIS: Glucose stimulated insulin secretion, HUVEC: Human umbilical vein endothelial cells. (A) and (B) are reprinted from Lorza-Gil et al. [274] published by Springer Nature and are licensed under Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/). Labels (A, B, C, D) of original article are removed. (C) and (D) are reprinted from Yu et al. [272] published Springer Nature and are licensed under Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/). Labels (a, b, c, d) of original article are removed. (E) and (F) are reprinted from Balboa et al. [311] published by Springer Nature and are licensed under Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/). Labels (a, b) of original article are removed. (G) is reprinted from Urbanczyk et al. [286] published by Mary Ann Liebert, Inc. and is licensed under Creative Commons CC BY (http://creativecommons.org/licenses/by/4.0/)
Although pseudoislets can be obtained with similar endocrine cell ratios, there is a possibility that the distribution of the endocrine cells is not adjusted accordingly, and the basal insulin levels and insulin secretion are strong in contrast to the native islets. Still, they have a potential to normalize the glucose levels in the STZ-induced diabetic mice [275]. However, donor-based changes in GSIS and mainly higher insulin secretion are reported when human islet-cell based pseudoislets were treated with stimulators such as palmitate and forskolin at 12 mM glucose. In some cases, lack of endocrine cells and exocrine cells can be another deviation from native human islets [274]. The proinflammatory cytokine gene (IL1B, CCL2, CXCL8) and the ECM marker (ASPN, COL1A1, COL4A1) expressions are other altered parameters compared to the native islets [271], and loss of endocrine activity and display of the pancreatic duct characteristics in the culture were reported [322]. Even with such differences, pseudoislets are central to diabetes research and possibly to clinical transplantation. They, for instance, show localization in the target organ, can control blood glucose levels and are positive for insulin in the transplants [323]. As multicellular mimics of the islets, they can be used to establish main hallmarks in the T1D development. As in vitro tools in the cytokine mediated injury, in the activated peripheral blood mononuclear cell and HLA-A2-restricted preproinsulin-specific cytotoxic T lymphocyte attack, they can be manipulated to design separate conditions for each case, and the individual or combinatorial destructive activity on the beta-cell insulin release and the survival can be analyzed. This platform enables testing compounds such as liraglutide, gliclazide and exendin-4 to fight against the inflammatory attack [313]. Moreover, they can be studied to assess the effects of lysophosphatidylcholines on the Gpr40, Gpr55, and Gpr119 receptor stimulation and the insulin release [276], as well as be transferred to the microfluidic devices to establish a pancreas-on-chip for the disease modeling and drug screening [253]. Their applications to analyze the pancreatic cancer cell-beta-cell communication [288], as model beta-cell systems for the insulin detection system development [262], to design nanoencapsulation by the layer-by-layer assembly [264] or to assemble the pseudoislets by the layer-by-layer ECM protein coating [243] were recorded.
Equally, the source for the beta-cells is also a critical factor for the beta-cell replacement therapy in T1D. As donation is limited and for the research, continuous isolation is needed, other options as beta-cell sources are investigated. So far, reprogrammed islet alpha- and gamma-cells [273], pancreatic exocrine cells [324], transdifferentiated human liver cells [251], stem cells such as dental pulp stem cells [291], adipose-tissue derived stem cells [292], human pluripotent stem cells [306], omentum derived stem cells [256], liver stem-like cells [310] and human umbilical cord mesenchymal stem cells (MSC) [325] were tested as the (patient-specific) cell-based therapeutics. To direct the cellular fate, differentiation protocols are established by combining, for example, adherent culture, chemical treatment and induction of aggregate formation (Fig. 3E). By this way, for example, embryonic stem cells-derived pseudoislets with glucagon, somatostatin and insulin positive cells can be assembled and up to 6 weeks, proportion of insulin-positive monohormonal cells can be retained (~ 40%) (Fig. 3F) [311]. In one of the recent studies, the differentiated cells form final clusters with sizes of 172 ± 34 μm in diameter in suspension culture showing consistent sizes in contrast to islets and are positive for dithizone (DTZ) stain (zinc-chelating dye for beta-cell staining). Although these pseudoislets are positive for beta-cell markers (C-peptide, PDX1 and NKX6-1) indicating the beta-cell fate, small amount of glucagon positive and polyhormonal cells were recorded. Challenge of these pseudoislet cells in static and dynamic GSIS showed that simulation index (3.0 vs. 3.2) and dynamic insulin secretion profiles are similar, but lower insulin release was evident in different phases of release pointing an important limitation [304]. Although EndoC-βH1 pseudoislets are similar to islet assembly and this organization improved stimulation index (4.5 vs. 2.6 in monolayer culture), their level of insulin secretion awaits for improvement (mouse islet stimulation index of 16.4) [321].
The ability of the cells to interact with each other further opened the possibility to incorporate non-endocrine cells which are beneficial to reduce transplanted cell number, improve insulin secretion, contribute to ECM deposition and provide immunoprotection against cytokine attack. Among these, endothelial cells [241, 299], fibroblasts [241] and MSCs [250] have notable effects. In design of pseudoislets for the beta-cell replacement therapy, beta-cell heterogeneity in the pancreas is a useful strategy to obtain useful transplants, as FAC-sorting of the beta-cells based on CD63 expression showed that CD63high beta-cells secrete more insulin, and alternative manipulation might be required to enhance the functionality of the other subsets [283]. As these systems are required to be scaled-up, several factors such as metabolic alterations, growth kinetics, cell seeding density, yield and stirrer types should be considered [255, 305, 326]. Further, commercial beta-cells used for the formation of pseudoislets are derived from for instance, mouse insulinoma and they display differences in insulin expression and content, as well as can show features that belong to both alpha-cells and beta-cells [249]. Other drawbacks include functional impairment at high passages [327], diversity in the amount of secreted insulin between cell types [276] and beta-cell dedifferentiation due to continuous stimulation which restrict the use of cells such as MIN6 cells [328].
Although life-long immunosuppression is another aspect of organ/cell transplantation to prevent rejection, it has adverse effects even on the beta-cells transplanted. Replacement of these drugs in the endocrine tissue engineering is suggested to be mainly with the development of the immunoisolation barriers [270, 329]. Additionally, a cell-based immunosuppressive regimen is proven to be useful to eliminate the use of the drugs and, consequently, can still protect the cells against immune reaction in the alloimmunity and even against the autoimmunity. One group of such cells are Treg cells which have the capacity to suppress the proliferation and activation of diverse immune system cells such as lymphocytes, NK cells and antigen-presenting cells. To support this strategy, Takemoto et al. transplanted pseudoislets of Tregs and islet cells intraportally to the STZ-induced diabetic mice and observed that this approach held a potential to reach the normoglycemia and that some insulin positive cells were retained in the animals after 120 days whereas massive destruction was evident in the pseudoislets without Tregs. This study points out the conclusion that Treg can provide immunoprotection to the transplanted islets to some extent [329]. Similarly, the transplantation of the pseudoislets composed of Sertoli cells and islet cells yielded encouraging outcomes, since Sertoli cells act as barriers to immune cell migration and Sertoli cell-derived factors contribute to shape the immunoprivileged testis environment as the unique cues for the immunoprotection of the cell transplants [297]. It is also possible to genetically engineer cells such as stellate cells to express Treg recruiting factor of C-C motif chemokine ligand 22 (CCL22). Incorporation of these cells into the pseudoislets can attract more Tregs to the area of graft, therefore, it has a potential to support the survival and capacity of the transplant to reach proper blood glucose and insulin levels [287].
Oxygen levels positively regulates the insulin secretion, thus, hypoxia adversely affects GSIS [279]. This effect is also valid in the developing and prolonged culture of the pseudoislets, as hypoxia will adversely affect the existence and the function [330]. For this purpose, a system that controls the oxygen transfer has utmost importance. For instance, Lee et al. designed a polydimethylsiloxane (PDMS) based microwell array with 10–1050 μm film thickness to manipulate the oxygen transfer and were able to demonstrate the effect of more oxygenated system on the insulin secretion and survival [244]. Incorporation of the microspheres to provide a route for the oxygen and nutrient transfer [298], or the microspheres loaded with drugs such as curcumin as anti-apoptotic agents and for the increase of the insulin release [308], pseudoislets with MSCs to fight the hypoxia by the soluble signals and the angiogenesis [307], human amniotic epithelial cells to resist the hypoxia and sustain the endocrine cell response [258] thereby, can serve for the desired outcome. Alternatively, size of the pseudoislets can be controlled with specifically designed microwells so that possibility of cell death and hypoxic core formation decreases. In one such studies, Ichihara et al. fabricated agarose-based microwells and rat-islet cell pseudoislets showed high level of cell death and development of hypoxia in larger pseudoislets (formed in 500 μm diameter wells) whereas in smaller ones, cell death was minimal and even lower than native islets [268].
Oxidative stress for the islets is also associated with the loss of the vessels during the isolation and with the lack of vascularization initially after transplantation. These processes have unfavorable impact on the islets, thus consequences of the transplantation and number of the transplants per patient. Therefore, supplementation of the transplants with compounds reducing the oxidative stress is beneficial for the patients. In the discovery of such compounds, pseudoislets assembled with the alpha- and beta-cells in the low attachment microplates provide example data. Sulfisoxazole, hydroxyzine dihydrochloride and tribenoside are some compounds reported by using pseudoislets as screening platforms [282]. Instead, vascularization-based guidance to reduce the effects of oxidative stress is also applicable. Although islets are 3D cell assemblies that significantly model beta-cell contacts, modulate calcium signaling and maintain ECM proteins, similarly vascular cells support beta-cells and are natural components in the pancreas [266]. Islets in the pancreas are typically highly vascularized to assist the supply of the metabolic requirements of the cells as well as for the endocrine function. They occupy less than 1% of pancreas mass, but they receive 15–20% of the pancreatic blood flow. Within such an organization, endothelial cells and beta-cells interact, secrete basement membrane ECM and, endothelial cells are important to export the insulin into the blood circulation. Thus, loss of beta-cell transplants due to the lack of vascularization becomes inevitable [285, 286]. Considering the organization in the pancreas, a 3D system of beta-cells and endothelial cells can be superior over beta-cell monolayer culture in the insulin secretion and ECM synthesis (collagen-IV, laminin), and further viability can be sustained [266]. A co-culture of these two types of cells as stable pseudoislets under controlled conditions can be a concept to reduce the impact of the multiple transplantation and can be unlimited source in the disease study. In the human pancreas, endothelial cell and beta-cell localization occur as heterogenous distribution (i.e. beta-cells covered by endothelial cells and, vice versa of the latter). These unique patterns are probable indicators which alter GSIS and eventually transplant success. As a result, research comparing these distributions and cell-cell interactions with human beta-cells offers a unique opportunity to design an ideal beta-cell response in the transplants. For example, a study by Urbanczyk et al. mimics three possible localization conditions created by magnetic levitation allowing desired cell distribution with improved endothelial cell incorporation (Fig. 3G) and, derived stable pseudoislets revealed that all three conditions have similar cell death, yet beta-cells covered by endothelial cells present higher insulin secretion and E-cadherin (a stimulator for the insulin expression, used in the cell-cell interactions of beta-cells) expression [286]. As pseudoislets with endothelial cell components can facilitate beta-cell engraftment, research on beta-cell and endothelial cell pseudoislets in a biomaterial such as fibrin [285] or incorporation of microvascular fragments can rescue beta-cell from hypoxia [260]. As third component, amniotic epithelial cells can be combined to regulate endothelial cell-based vascularization by angiogenic factors such as VEGF-A and to remodel ECM with MMPs. Additionally, these cells influence islet-cell PDX-1 and GLP1R expression by EGF which eventually contributes to the glycolytic gene and VEGF-A secretion improving not only beta-cell survival and insulin metabolism, but also vascularization [289].
After formation, pseudoislets are generally studied in static cultures, but a dynamic culture with a flow system might provide unique differences in the pseudoislet transcriptomics and metabolomics by simulating the flow experienced in the body. In one of such studies, Essaouiba et al. demonstrated that 3D biochips containing human induced pluripotent stem cell (hiPSC) derived beta-cell pseudoislets increased, for instance, hypoxia-inducible factor activity, glucose metabolism and enriched insulin as well as glucagon pathways, activated SOX17 and MAFB motifs for the maintenance of the differentiated state of beta-cells and their insulin secretion [254]. On-chip systems are also advantageous for the control of the environment, nutrient delivery, pseudoislet loading, time-dependent analysis and imaging, as well as end-point studies [278]. In recent reports, this dynamic system is used as a platform to build pseudoislets [259], to test drug efficacy [115, 281] as well as to model the effects of the flow-induced mechanical stress, chronic glycemia and lipidemia on the beta-cell gene expression in the diabetes [245].
The quality of the islet population, loss of the islets after transplantation due to lack of vascularization and post-transplantation immune activity restrict the patients receiving the islet transplants. Genetic manipulation of the islets is one strategy to overcome these barriers, as angiogenetic factor expression, control of the insulin production and transplant-mediated suppression of immune system are encouraging. This route involves efficient transfection of single cells which are reassembled into the pseudoislets during gene delivery. For instance, Voznesenskaya et al. reached the conclusion that transduction with AAV8 during pseudoislet formation yields 2.5-fold higher expression than intact isolated islets with improved transfer to the pseudoislet core [261]. Van Krieken et al. transduced the islet cells during pseudoislet aggregation by adenovirus carrying V1b receptor gene that can initiate GSIS by phospholipase C pathway and altering Ca2+ levels. [Arg8]-vasopressin induced beta-cell activity in transduced pseudoislets and supported insulin secretion in vitro [275]. Genetic engineering of islet cells to express exendin-4 (a GLP-1 analogue) to increase the insulin secretion and decrease the apoptosis [301], and poly(ethylenimine) (PEI)-mediated IL-10 gene delivery before aggregation [237] can protect the beta-cells and advance the insulin secretion.
In addition to applications in vitro, islets and islet-like tissues have been in clinical trials for the last couple decades. In clinical trials database (https://clinicaltrials.gov/), search on “Type 1 Diabetes” and “Islet Graft” shows 33 different studies and “Type 1 Diabetes” and “islet transplantation” as keywords show 142 studies. “Type 1 Diabetes” and “Cell Therapy” results in 151 results. Among the studies mentioned, in addition to islet transplantation, stem cells are the major focus due to their immunomodulatory and differentiation potential, as well as due to stem-cell derived factors for beta-cell regeneration. Stem cells can be used in as a part of cell therapy derived from allogenic sources to protect beta-cells (NCT04061746), to manage C-peptide levels (NCT01121029), to educate immune system cells (NCT02624804, NCT04011020, NCT01996228, NCT01350219), to determine safety and/or efficiency in T1D patients (NCT01143168, NCT00807651, NCT00703599, NCT05308836), to regenerate beta-cells and even for re-differentiation into local tissues (NCT01374854). There are also trials on allogeneic stem cell-derived, insulin-producing islet cell therapy (NCT04786262) and encapsulated allogeneic stem cell-derived, insulin-producing islet therapy (NCT05791201).
Another group includes hiPSCs and MSCs where these cell classes are investigated as sources of islet-like tissues. In these reports, pilot/clinical studies showed glycemic control within 3–4 months showing the promise of this technology [331, 332].
Conclusions and future directions
As T1D is increasingly diagnosed worldwide, development of new treatments is of interest for current efforts. Insulin treatment is the leading strategy in the clinics followed by the beta-cell replacement therapy to achieve glucose homeostasis, both of which are associated with serious drawbacks. Recent technologies enabled researchers to design scaffold-free artificial islets as endocrine models for purposes ranging from drug screening to pancreas physiology. A practical application is the utilization of so-called pseudoislets in beta-cell replacement therapy instead of exogenous insulin supply. These new systems are directed to address the challenges such as native organization, immunoprotection and functional improvements and, the results reported are more realistic compared to 2D monolayer culture and single cell containing transplants. Although various artificial islet models and their potential applications are established so far, the field remains to be in the preclinical stage due to factors such as standardization, beta-cell heterogeneity, requirement of complete mimic of cellular and acellular environment, donor shortage and lack of sufficient clinical trials. Thus, future research faces multiple challenges to translate the islet mimics into the clinics. Moreover, today, the majority of the pseudoislet research is shaped around incorporation of beta-cells and other endocrine cell types as well as endothelial cells with different techniques. Further, cells such as Tregs, Sertoli cells and stem cells to provide immunoprotection, and fibroblasts and stem cells to improve insulin secretion are added and efficiently islet-look alike and function-like cell assemblies are produced. However, cell-cell contact, and cell-soluble factor mediated control of beta-cell metabolic capacity are not limited to only these types of cells. In the future, with more genetically stable cells, contribution of nervous system to pseudoislet electrophysiology should be analyzed and flow-based systems rather than assay based dynamic perifusion systems should consistently be used to fully elucidate the promise of these 3D assemblies. An important point will also be addressing the issue of autologous islet-derived systems for personalized medicine.
Data availability
No datasets were generated or analysed during the current study.
Abbreviations
- ANG:
-
Angiopoietin
- CAF:
-
Cancer-associated fibroblast
- CCL22:
-
C-C motif chemokine ligand 22
- CHPOA:
-
Acrylate modified, cholesterol bearing pullulan
- CXCL:
-
C-X-C motif chemokine ligand
- CXCR2:
-
CXC chemokine receptor 2
- Dlg:
-
Discs large protein
- DTZ:
-
Dithizone
- ECM:
-
Extracellular matrix
- EGF:
-
Epidermal growth factor
- ESAM:
-
Endothelial cell selective adhesion molecule
- GLP-1:
-
Glucagon-like peptide-1
- GLP1R:
-
Glucagon-like peptide-1 receptor
- GLUT:
-
Glucose transporter
- GPCR:
-
G-protein coupled receptor
- GSIS:
-
Glucose stimulated insulin secretion
- HGF:
-
Hepatocyte growth factor
- hiPSC:
-
Human induced pluripotent stem cell
- HLA:
-
Human leukocyte antigen
- HUVEC:
-
Human umbilical vein endothelial cell
- ICAM-1:
-
Intercellular adhesion molecule-1
- IFN:
-
Interferon
- IGF:
-
Insulin-like growth factor
- IL:
-
Interleukin
- JAM:
-
Junctional adhesion molecule
- LIF:
-
Leukaemia inhibitory factor
- MBP-FGF2:
-
Maltose-binding protein-basic fibroblast growth factor 2
- MMP:
-
Matrix metalloproteinase
- MSC:
-
Mesenchymal stem cell
- N-CAM:
-
Neural cell adhesion molecule
- NO:
-
Nitric oxide
- P2X7R:
-
P2X purinoceptor 7
- Par3:
-
Partitioning defective 3 homologue
- PD-1:
-
Programmed cell death protein-1
- PDGF:
-
Platelet-derived growth factor
- PDGFRβ:
-
Platelet-derived growth factor receptor-β
- PD-L1:
-
Programmed death-ligand 1
- PDMS:
-
Polydimethylsiloxane
- PECAM-1:
-
Platelet endothelial cell adhesion molecule-1
- PEG:
-
Poly(ethylene glycol)
- PEI:
-
Poly(ethyleneimine)
- PMMA:
-
Poly(methyl methacrylate)
- PNIPAAm:
-
Poly(N-isopropylacrylamide)
- PSC:
-
Pancreatic stellate cell
- PTPN22:
-
Protein tyrosine phosphatase non-receptor type 22
- RIM2:
-
Rab3-interacting protein
- STZ:
-
Streptozotocin
- T1D:
-
Type 1 diabetes
- T2D:
-
Type 2 diabetes
- TGF:
-
Transforming growth factor
- TGF-βR:
-
Transforming growth factor β receptor
- TNF:
-
Tumor necrosis factor
- VDCC:
-
Voltage-dependent calcium channel
- VE-cadherin:
-
Vascular endothelial cadherin
- VEGF-A:
-
Vascular endothelial growth factor A
- VEGFR:
-
Vascular endothelial growth factor receptor
- ZO:
-
Zona occludens
References
Röder PV, Wu B, Liu Y, Han W. Pancreatic regulation of glucose homeostasis. Exp Mol Med. 2016;48(3):e219.
Karpińska M, Czauderna M. Pancreas—its functions, disorders, and physiological impact on the mammals’ organism. Front Physiol. 2022;13.
Pociot F, Lernmark Å. Genetic risk factors for type 1 diabetes. Lancet. 2016;387(10035):2331–9.
Hörber S, Achenbach P, Schleicher E, Peter A. Harmonization of immunoassays for biomarkers in diabetes mellitus. Biotechnol Adv. 2020;39:107359.
Undlien DE, Kockum I, Rønningen KS, Lowe R, Saanjeevi CB, Graham J, et al. HLA associations in type 1 diabetes among patients not carrying high-risk DR3-DQ2 or DR4-DQ8 haplotypes. Tissue Antigens. 1999;54(6):543–51.
Aly TA, Ide A, Jahromi MM, Barker JM, Fernando MS, Babu SR, et al. Extreme genetic risk for type 1A diabetes. Proc Natl Acad Sci. 2006;103(38):14074–9.
Smigoc Schweiger D, Mendez A, Kunilo Jamnik S, Bratanic N, Bratina N, Battelino T, et al. High-risk genotypes HLA-DR3-DQ2/DR3-DQ2 and DR3-DQ2/DR4-DQ8 in co-occurrence of type 1 diabetes and Celiac disease. Autoimmunity. 2016;49(4):240–7.
Kantárová D, Buc M. Genetic susceptibility to type 1 diabetes mellitus in humans. Physiol Res. 2007;56(3):255–66.
Thomas NJ, Dennis JM, Sharp SA, Kaur A, Misra S, Walkey HC, et al. DR15-DQ6 remains dominantly protective against type 1 diabetes throughout the first five decades of life. Diabetologia. 2021;64(10):2258–65.
Simmons KM, Mitchell AM, Alkanani AA, McDaniel KA, Baschal EE, Armstrong T, et al. Failed genetic protection: type 1 diabetes in the presence of HLA-DQB1*06:02. Diabetes. 2020;69(8):1763–9.
Antvorskov JC, Josefsen K, Engkilde K, Funda DP, Buschard K. Dietary gluten and the development of type 1 diabetes. Diabetologia. 2014;57(9):1770–80.
Oboza P, Ogarek N, Olszanecka-Glinianowicz M, Kocelak P. Can type 1 diabetes be an unexpected complication of obesity? Front Endocrinol. 2023;14.
Van der Schueren B, Ellis D, Faradji RN, Al-Ozairi E, Rosen J, Mathieu C. Obesity in people living with type 1 diabetes. Lancet Diabetes Endocrinol. 2021;9(11):776–85.
Liu C, Wang J, Wan Y, Xia X, Pan J, Gu W, et al. Serum vitamin D deficiency in children and adolescents is associated with type 1 diabetes mellitus. Endocr Connections. 2018;7(12):1275–9.
Infante M, Ricordi C, Sanchez J, Clare-Salzler MJ, Padilla N, Fuenmayor V, et al. Influence of vitamin D on islet autoimmunity and beta-cell function in type 1 diabetes. Nutrients. 2019;11(9):2185.
Nyalwidhe JO, Jurczyk A, Satish B, Redick S, Qaisar N, Trombly MI, et al. Proteomic and transcriptional profiles of human stem cell-derived Β cells following enteroviral challenge. Microorganisms. 2020;8(2):295.
Marchand L, Pecquet M, Luyton C. Type 1 diabetes onset triggered by COVID-19. Acta Diabetol. 2020;57(10):1265–6.
Honeyman MC, Coulson BS, Stone NL, Gellert SA, Goldwater PN, Steele CE, et al. Association between rotavirus infection and pancreatic islet autoimmunity in children at risk of developing type 1 diabetes. Diabetes. 2000;49(8):1319–24.
Müller JA, Groß R, Conzelmann C, Krüger J, Merle U, Steinhart J, et al. SARS-CoV-2 infects and replicates in cells of the human endocrine and exocrine pancreas. Nat Metabolism. 2021;3(2):149–65.
Yuan X, Wang R, Han B, Sun C, Chen R, Wei H, et al. Functional and metabolic alterations of gut microbiota in children with new-onset type 1 diabetes. Nat Commun. 2022;13(1):6356.
Vatanen T, Franzosa EA, Schwager R, Tripathi S, Arthur TD, Vehik K, et al. The human gut microbiome in early-onset type 1 diabetes from the TEDDY study. Nature. 2018;562(7728):589–94.
Lin C, Li X, Qiu Y, Chen Z, Liu J. PD-1 inhibitor-associated type 1 diabetes: a case report and systematic review. Front Public Health. 2022;10:885001.
de Filette JMK, Pen JJ, Decoster L, Vissers T, Bravenboer B, Van der Auwera BJ, et al. Immune checkpoint inhibitors and type 1 diabetes mellitus: a case report and systematic review. Eur J Endocrinol. 2019;181(3):363–74.
American Diabetes Association Professional Practice Committee. 9. Pharmacologic approaches to glycemic treatment: standards of medical care in diabetes-2022. Diabetes Care. 2022;45(Suppl 1):S125-s43.
Wong EY, Kroon L. Ultra-rapid-acting insulins: how fast is really needed? Clin Diabetes. 2021;39(4):415–23.
Rodbard HW, Rodbard D. Biosynthetic human insulin and insulin analogs. Am J Ther. 2020;27(1):e42–51.
Hirsch IB, Juneja R, Beals JM, Antalis CJ, Wright EE. Jr. The evolution of insulin and how it informs therapy and treatment choices. Endocr Rev. 2020;41(5):733–55.
Dholakia S, Royston E, Quiroga I, Sinha S, Reddy S, Gilbert J, et al. The rise and potential fall of pancreas transplantation. Br Med Bull. 2017;124(1):171–9.
Marfil-Garza BA, Shapiro AMJ, Kin T. Clinical islet transplantation: current progress and new frontiers. J Hepato-Biliary-Pancreat Sci. 2021;28(3):243–54.
Wang Q, Huang Y-x, Liu L, Zhao X-h, Sun Y, Mao X et al. Pancreatic islet transplantation: current advances and challenges. Front Immunol. 2024;15.
Shapiro AMJ, Lakey JRT, Ryan EA, Korbutt GS, Toth E, Warnock GL, et al. 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.
Matsumoto S, Shimoda M. Current situation of clinical islet transplantation from allogeneic toward xenogeneic. J Diabetes. 2020;12(10):733–41.
Hering BJ, Clarke WR, Bridges ND, Eggerman TL, Alejandro R, Bellin MD, 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.
Shapiro AMJ, Pokrywczynska M, Ricordi C. Clinical pancreatic islet transplantation. Nat Reviews Endocrinol. 2017;13(5):268–77.
Patel SN, Mathews CE, Chandler R, Stabler CL. The foundation for engineering a pancreatic islet niche. Front Endocrinol. 2022;13.
Noguchi GM, Huising MO. Integrating the inputs that shape pancreatic islet hormone release. Nat Metabolism. 2019;1(12):1189–201.
Campbell JE, Newgard CB. Mechanisms controlling pancreatic islet cell function in insulin secretion. Nat Rev Mol Cell Biol. 2021;22(2):142–58.
Ellis C, Ramzy A, Kieffer TJ. Regenerative medicine and cell-based approaches to restore pancreatic function. Nat Reviews Gastroenterol Hepatol. 2017;14(10):612–28.
Rahman MS, Hossain KS, Das S, Kundu S, Adegoke EO, Rahman MA, et al. Role of insulin in health and disease: an update. Int J Mol Sci. 2021;22(12):6403.
Hay DL, Chen S, Lutz TA, Parkes DG, Roth JD. Amylin: pharmacology, physiology, and clinical potential. Pharmacol Rev. 2015;67(3):564–600.
Wahren J. C-peptide and the pathophysiology of microvascular complications of diabetes. J Intern Med. 2017;281(1):3–6.
Rickels MR, Evans-Molina C, Bahnson HT, Ylescupidez A, Nadeau KJ, Hao W, et al. High residual C-peptide likely contributes to glycemic control in type 1 diabetes. J Clin Investig. 2020;130(4):1850–62.
Hædersdal S, Andersen A, Knop FK, Vilsbøll T. Revisiting the role of glucagon in health, diabetes mellitus and other metabolic diseases. Nat Reviews Endocrinol. 2023;19(6):321–35.
Kumar U, Singh S. Role of somatostatin in the regulation of central and peripheral factors of satiety and obesity. Int J Mol Sci. 2020;21(7):2568.
Li N, Yang Z, Li Q, Yu Z, Chen X, Li J-C, et al. Ablation of somatostatin cells leads to impaired pancreatic islet function and neonatal death in rodents. Cell Death Dis. 2018;9(6):682.
Mirzadeh Z, Faber CL, Schwartz MW. Central nervous system control of glucose homeostasis: a therapeutic target for type 2 diabetes? Annu Rev Pharmacol Toxicol. 2022;62:55–84.
Alonge KM, D’Alessio DA, Schwartz MW. Brain control of blood glucose levels: implications for the pathogenesis of type 2 diabetes. Diabetologia. 2021;64(1):5–14.
Hyun U, Sohn J-W. Autonomic control of energy balance and glucose homeostasis. Exp Mol Med. 2022;54(4):370–6.
Carnagarin R, Matthews VB, Herat LY, Ho JK, Schlaich MP. Autonomic regulation of glucose homeostasis: a specific role for sympathetic nervous system activation. Curr Diab Rep. 2018;18(11):107.
Hoffman EG, D’Souza NC, Liggins RT, Riddell MC. Pharmacologic Inhibition of somatostatin receptor 2 to restore glucagon counterregulation in diabetes. Front Pharmacol. 2024;14.
Denwood G, Tarasov A, Salehi A, Vergari E, Ramracheya R, Takahashi H et al. Glucose stimulates somatostatin secretion in pancreatic δ-cells by cAMP-dependent intracellular Ca2 + release. J Gen Physiol. 2019;151(9).
Davis TR, Pierce MR, Novak SX, Hougland JL. Ghrelin octanoylation by ghrelin O-acyltransferase: protein acylation impacting metabolic and neuroendocrine signalling. Open Biology. 2021;11(7):210080.
Alamri BN, Shin K, Chappe V, Anini Y. The role of Ghrelin in the regulation of glucose homeostasis. Horm Mol Biol Clin Investig. 2016;26(1):3–11.
Colldén G, Tschöp MH, Müller TD. Therapeutic potential of targeting the Ghrelin pathway. Int J Mol Sci. 2017;18(4):798.
Garcia-Rendueles MER, Varela L, Horvath TL, Ghrelin. Trends Endocrinol Metabolism. 2024;35(11):1021–2.
Banerjee A, Onyuksel H. Human pancreatic polypeptide in a phospholipid-based micellar formulation. Pharm Res. 2012;29(6):1698–711.
Vergari E, Knudsen JG, Ramracheya R, Salehi A, Zhang Q, Adam J, et al. Insulin inhibits glucagon release by SGLT2-induced stimulation of somatostatin secretion. Nat Commun. 2019;10(1):139.
Svendsen B, Larsen O, Gabe MBN, Christiansen CB, Rosenkilde MM, Drucker DJ, et al. Insulin secretion depends on intra-islet glucagon signaling. Cell Rep. 2018;25(5):1127–e342.
Hauge-Evans AC, Anderson RL, Persaud SJ, Jones PM. Delta cell secretory responses to insulin secretagogues are not mediated indirectly by insulin. Diabetologia. 2012;55(7):1995–2004.
Svendsen B, Holst JJ. Paracrine regulation of somatostatin secretion by insulin and glucagon in mouse pancreatic Islets. Diabetologia. 2021;64(1):142–51.
Cabrera O, Berman DM, Kenyon NS, Ricordi C, Berggren PO, Caicedo A. The unique cytoarchitecture of human pancreatic islets has implications for islet cell function. Proc Natl Acad Sci USA. 2006;103(7):2334–9.
Bosco D, Armanet M, Morel P, Niclauss N, Sgroi A, Muller YD, et al. Unique arrangement of α- and β-cells in human islets of Langerhans. Diabetes. 2010;59(5):1202–10.
Benninger RKP, Kravets V. The physiological role of β-cell heterogeneity in pancreatic islet function. Nat Reviews Endocrinol. 2022;18(1):9–22.
Perez-Frances M, van Gurp L, Abate MV, Cigliola V, Furuyama K, Bru-Tari E, et al. Pancreatic Ppy-expressing γ-cells display mixed phenotypic traits and the adaptive plasticity to engage insulin production. Nat Commun. 2021;12(1):4458.
Nasteska D, Fine NHF, Ashford FB, Cuozzo F, Viloria K, Smith G, et al. PDX1low MAFAlow β-cells contribute to islet function and insulin release. Nat Commun. 2021;12(1):674.
Miranda MA, Macias-Velasco JF, Lawson HA. Pancreatic β-cell heterogeneity in health and diabetes: classes, sources, and subtypes. Am J Physiology-Endocrinology Metabolism. 2021;320(4):E716–31.
Feng Y, Qiu W-L, Yu X-X, Zhang Y, He M-Y, Li L-C, et al. Characterizing pancreatic β-cell heterogeneity in the streptozotocin model by single-cell transcriptomic analysis. Mol Metabolism. 2020;37:100982.
Trogden KP, Lee J, Bracey KM, Ho K-H, McKinney H, Zhu X, et al. Microtubules regulate pancreatic β-cell heterogeneity via spatiotemporal control of insulin secretion hot spots. eLife. 2021;10:e59912.
Pettway YD, Saunders DC, Brissova M. The human α cell in health and disease. J Endocrinol. 2023;258(1):e220298.
Santos-Silva JC, Carvalho CP, de Oliveira RB, Boschero AC, Collares-Buzato CB. Cell-to-cell contact dependence and junctional protein content are correlated with in vivo maturation of pancreatic beta cells. Can J Physiol Pharmacol. 2012;90(7):837–50.
Cnop M, Welsh N, Jonas J-C, Jörns A, Lenzen S, Eizirik DL. Mechanisms of pancreatic β-cell death in type 1 and type 2 diabetes: many differences, few similarities. Diabetes. 2005;54(suppl2):S97–107.
Klinke DJ 2. Extent of beta cell destruction is important but insufficient to predict the onset of type 1 diabetes mellitus. PLoS ONE. 2008;3(1):e1374.
Oram RA, Sims EK, Evans-Molina C. Beta cells in type 1 diabetes: mass and function; sleeping or dead? Diabetologia. 2019;62(4):567–77.
Yosten GLC. Alpha cell dysfunction in type 1 diabetes. Peptides. 2018;100:54–60.
Panzer JK, Caicedo A. Targeting the pancreatic α-cell to prevent hypoglycemia in type 1 diabetes. Diabetes. 2021;70(12):2721–32.
Brissova M, Haliyur R, Saunders D, Shrestha S, Dai C, Blodgett DM, et al. α cell function and gene expression are compromised in type 1 diabetes. Cell Rep. 2018;22(10):2667–76.
Liu W, Kin T, Ho S, Dorrell C, Campbell SR, Luo P, et al. Abnormal regulation of glucagon secretion by human islet alpha cells in the absence of beta cells. eBioMedicine. 2019;50:306–16.
Bru-Tari E, Cobo-Vuilleumier N, Alonso-Magdalena P, Dos Santos RS, Marroqui L, Nadal A, et al. Pancreatic alpha-cell mass in the early-onset and advanced stage of a mouse model of experimental autoimmune diabetes. Sci Rep. 2019;9(1):9515.
Bosi E, Marchetti P, Rutter GA, Eizirik DL. Human alpha cell transcriptomic signatures of types 1 and 2 diabetes highlight disease-specific dysfunction pathways. iScience. 2022;25(10):105056.
Arrojo e Drigo R, Jacob S, García-Prieto CF, Zheng X, Fukuda M, Nhu HTT, et al. Structural basis for delta cell paracrine regulation in pancreatic islets. Nat Commun. 2019;10(1):3700.
Poudel A, Savari O, Striegel DA, Periwal V, Taxy J, Millis JM, et al. Beta-cell destruction and preservation in childhood and adult onset type 1 diabetes. Endocrine. 2015;49(3):693–702.
Seiron P, Wiberg A, Kuric E, Krogvold L, Jahnsen FL, Dahl-Jørgensen K, et al. Characterisation of the endocrine pancreas in type 1 diabetes: islet size is maintained but islet number is markedly reduced. J Pathology: Clin Res. 2019;5(4):248–55.
Wang YJ, Traum D, Schug J, Gao L, Liu C, Atkinson MA, et al. Multiplexed in situ imaging mass cytometry analysis of the human endocrine pancreas and immune system in type 1 diabetes. Cell Metabol. 2019;29(3):769–e834.
Cottle L, Gan WJ, Gilroy I, Samra JS, Gill AJ, Loudovaris T, et al. Structural and functional polarisation of human pancreatic beta cells in islets from organ donors with and without type 2 diabetes. Diabetologia. 2021;64(3):618–29.
Ohara-Imaizumi M, Aoyagi K, Yamauchi H, Yoshida M, Mori MX, Hida Y, et al. ELKS/voltage-sependent Ca2 + channel-β subunit module regulates polarized Ca2 + influx in pancreatic Β cells. Cell Rep. 2019;26(5):1213–e267.
Gan WJ, Zavortink M, Ludick C, Templin R, Webb R, Webb R, et al. Cell Polarity defines three distinct domains in pancreatic β-cells. J Cell Sci. 2017;130(1):143–51.
Jevon D, Cottle L, Hallahan N, Harwood R, Samra JS, Gill AJ, et al. Capillary contact points determine beta cell Polarity, control secretion and are disrupted in the Db/db mouse model of diabetes. Diabetologia. 2024;67(8):1683–97.
Ohara-Imaizumi M, Aoyagi K, Ohtsuka T. Role of the active zone protein, ELKS, in insulin secretion from pancreatic β-cells. Mol Metabolism. 2019;27:S81–91.
Serre-Beinier V, Bosco D, Zulianello L, Charollais A, Caille D, Charpantier E, et al. Cx36 makes channels coupling human pancreatic beta-cells, and correlates with insulin expression. Hum Mol Genet. 2009;18(3):428–39.
Farnsworth NL, Benninger RKP. New insights into the role of connexins in pancreatic islet function and diabetes. FEBS Lett. 2014;588(8):1278–87.
Farnsworth NL, Piscopio RA, Schleicher WE, Ramirez DG, Miranda JG, Benninger RKP. Modulation of gap junction coupling within the islet of Langerhans during the development of type 1 diabetes. Front Physiol. 2022;13.
Miranda C, Begum M, Vergari E, Briant LJB. Gap junction coupling and islet delta-cell function in health and disease. Peptides. 2022;147:170704.
Penuela S, Gehi R, Laird DW. The biochemistry and function of pannexin channels. Biochim Et Biophys Acta (BBA) - Biomembr. 2013;1828(1):15–22.
Tozzi M, Larsen AT, Lange SC, Giannuzzo A, Andersen MN, Novak I. The P2X7 receptor and pannexin-1 are involved in glucose-induced autocrine regulation in β-cells. Sci Rep. 2018;8(1):8926.
Dissanayake WC, Sorrenson B, Shepherd PR. The role of adherens junction proteins in the regulation of insulin secretion. Biosci Rep. 2018;38(2).
Dissanayake WC, Shepherd PR. β-cells retain a pool of insulin-containing secretory vesicles regulated by adherens junctions and the cadherin-binding protein p120 Catenin. J Biol Chem. 2022;298(8):102240.
Parnaud G, Lavallard V, Bedat B, Matthey-Doret D, Morel P, Berney T, et al. Cadherin engagement improves insulin secretion of single human β-cells. Diabetes. 2015;64(3):887–96.
Parnaud G, Gonelle-Gispert C, Morel P, Giovannoni L, Muller YD, Meier R, et al. Cadherin engagement protects human β-cells from apoptosis. Endocrinology. 2011;152(12):4601–9.
Myo Min KK, Rojas-Canales D, Penko D, DeNichilo M, Cockshell MP, Ffrench CB, et al. Desmoglein-2 is important for islet function and β-cell survival. Cell Death Dis. 2022;13(10):911.
Esni F, Täljedal IB, Perl AK, Cremer H, Christofori G, Semb H. Neural cell adhesion molecule (N-CAM) is required for cell type segregation and normal ultrastructure in pancreatic Islets. J Cell Biol. 1999;144(2):325–37.
Olofsson CS, Håkansson J, Salehi A, Bengtsson M, Galvanovskis J, Partridge C, et al. Impaired insulin exocytosis in neural cell adhesion molecule–/– mice due to defective reorganization of the submembrane F-actin network. Endocrinology. 2009;150(7):3067–75.
Wieland FC, van Blitterswijk CA, van Apeldoorn A, LaPointe VL. The functional importance of the cellular and extracellular composition of the islets of Langerhans. J Immunol Regenerative Med. 2021;13:100048.
Dahl U, Sjödin A, Semb H. Cadherins regulate aggregation of pancreatic β-cells in vivo. Development. 1996;122(9):2895–902.
Tixi W, Maldonado M, Chang Y-T, Chiu A, Yeung W, Parveen N, et al. Coordination between ECM and cell-cell adhesion regulates the development of islet aggregation, architecture, and functional maturation. eLife. 2023;12:e90006.
Carvell M, Marsh P, Persaud S, Jones P. E-cadherin interactions regulate β-cell proliferation in islet-like structures. Cell Physiol Biochem. 2008;20(5):617–26.
Wakae-Takada N, Xuan S, Watanabe K, Meda P, Leibel RL. Molecular basis for the regulation of islet beta cell mass in mice: the role of E-cadherin. Diabetologia. 2013;56(4):856–66.
Adams MT, Blum B. Determinants and dynamics of pancreatic islet architecture. Islets. 2022;14(1):82–100.
Ahuja N, Cleaver O. The cell cortex as mediator of pancreatic epithelial development and endocrine differentiation. Curr Opin Genet Dev. 2022;72:118–27.
Konstantinova I, Nikolova G, Ohara-Imaizumi M, Meda P, Kucera T, Zarbalis K, et al. EphA-Ephrin-A-mediated beta cell communication regulates insulin secretion from pancreatic islets. Cell. 2007;129(2):359–70.
Hutchens T, Piston DW. EphA4 receptor forward signaling inhibits glucagon secretion from α-cells. Diabetes. 2015;64(11):3839–51.
Li H, Neelankal John A, Nagatake T, Hamazaki Y, Jiang FX. Claudin 4 in pancreatic Β cells is involved in regulating the functional state of adult islets. FEBS Open Bio. 2020;10(1):28–40.
Collares-Buzato CB, Carvalho CPF, Furtado AG, Boschero AC. Upregulation of the expression of tight and adherens junction-associated proteins during maturation of neonatal pancreatic islets in vitro. J Mol Histol. 2004;35(8):811–22.
Suckow AT, Zhang C, Egodage S, Comoletti D, Taylor P, Miller MT, et al. Transcellular neuroligin-2 interactions enhance insulin secretion and are integral to pancreatic Β cell function*. J Biol Chem. 2012;287(24):19816–26.
Muratore M, Santos C, Rorsman P. The vascular architecture of the pancreatic Islets: a homage to August Krogh. Comp Biochem Physiol A: Mol Integr Physiol. 2021;252:110846.
Jun Y, Lee J, Choi S, Yang JH, Sander M, Chung S, et al. In vivo–mimicking microfluidic perfusion culture of pancreatic islet spheroids. Sci Adv. 2019;5(11):eaax4520.
Jansson L, Barbu A, Bodin B, Drott CJ, Espes D, Gao X, et al. Pancreatic islet blood flow and its measurement. Ups J Med Sci. 2016;121(2):81–95.
Burganova G, Bridges C, Thorn P, Landsman L. The role of vascular cells in pancreatic beta-cell function. Front Endocrinol. 2021;12:667170.
Sakhneny L, Epshtein A, Landsman L. Pericytes contribute to the islet basement membranes to promote beta-cell gene expression. Sci Rep. 2021;11(1):2378.
Berger C, Zdzieblo D. Glucose transporters in pancreatic Islets. Pflügers Archiv - Eur J Physiol. 2020;472(9):1249–72.
Rorsman P, Braun M. Regulation of insulin secretion in human pancreatic Islets. Annu Rev Physiol. 2013;75(1):155–79.
Rorsman P, Ashcroft FM. Pancreatic β-cell electrical activity and insulin secretion: of mice and men. Physiol Rev. 2018;98(1):117–214.
Liu Z, Jeppesen PB, Gregersen S, Chen X, Hermansen K. Dose-and glucose-dependent effects of amino acids on insulin secretion from isolated mouse islets and clonal INS-1E beta-cells. Rev Diabet Stud. 2008;5(4):232.
Tokarz VL, MacDonald PE, Klip A. The cell biology of systemic insulin function. J Cell Biol. 2018;217(7):2273–89.
Omar-Hmeadi M, Idevall-Hagren O. Insulin granule biogenesis and exocytosis. Cell Mol Life Sci. 2021;78(5):1957–70.
Brissova M, Shostak A, Shiota M, Wiebe PO, Poffenberger G, Kantz J, et al. Pancreatic islet production of vascular endothelial growth factor-A is essential for islet vascularization, revascularization, and function. Diabetes. 2006;55(11):2974–85.
Hogan MF, Hull RL. The islet endothelial cell: a novel contributor to beta cell secretory dysfunction in diabetes. Diabetologia. 2017;60(6):952–9.
Jonsson A, Hedin A, Müller M, Skog O, Korsgren O. Transcriptional profiles of human islet and exocrine endothelial cells in subjects with or without impaired glucose metabolism. Sci Rep. 2020;10(1):22315.
Konishi M, Sakaguchi M, Lockhart SM, Cai W, Li ME, Homan EP et al. Endothelial insulin receptors differentially control insulin signaling kinetics in peripheral tissues and brain of mice. Proceedings of the National Academy of Sciences. 2017;114(40):E8478-E87.
Jaldin-Fincati JR, Pereira RVS, Bilan PJ, Klip A. Insulin uptake and action in microvascular endothelial cells of lymphatic and blood origin. Am J Physiology-Endocrinology Metabolism. 2018;315(2):E204–17.
Watson LS, Wilken-Resman B, Williams A, DiLucia S, Sanchez G, McLeod TL, et al. Hyperinsulinemia alters insulin receptor presentation and internalization in brain microvascular endothelial cells. Diabetes Vascular Disease Res. 2022;19(4):14791641221118626.
Azizi PM, Zyla RE, Guan S, Wang C, Liu J, Bolz S-S, et al. Clathrin-dependent entry and vesicle-mediated exocytosis define insulin transcytosis across microvascular endothelial cells. Mol Biol Cell. 2015;26(4):740–50.
Williams IM, Wasserman DH. Capillary endothelial insulin transport: the rate-limiting step for insulin-stimulated glucose uptake. Endocrinology. 2021;163(2).
Hoang Do O, Thorn P. Insulin secretion from beta cells within intact islets: location matters. Clin Exp Pharmacol Physiol. 2015;42(4):406–14.
Claesson-Welsh L, Dejana E, McDonald DM. Permeability of the endothelial barrier: identifying and reconciling controversies. Trends Mol Med. 2021;27(4):314–31.
Komarova YA, Kruse K, Mehta D, Malik AB. Protein interactions at endothelial junctions and signaling mechanisms regulating endothelial permeability. Circul Res. 2017;120(1):179–206.
Sjöberg E, Melssen M, Richards M, Ding Y, Chanoca C, Chen D et al. Endothelial VEGFR2-PLCγ signaling regulates vascular permeability and antitumor immunity through eNOS/Src. J Clin Investig. 2023;133(20).
Yeh Y-T, Serrano R, François J, Chiu J-J, Li Y-SJ, del Álamo JC, et al. Three-dimensional forces exerted by leukocytes and vascular endothelial cells dynamically facilitate diapedesis. Proc Natl Acad Sci. 2018;115(1):133–8.
Vestweber D. How leukocytes cross the vascular endothelium. Nat Rev Immunol. 2015;15(11):692–704.
Wallez Y, Huber P. Endothelial adherens and tight junctions in vascular homeostasis, inflammation and angiogenesis. Biochim Et Biophys Acta (BBA) - Biomembr. 2008;1778(3):794–809.
Dejana E. Endothelial cell–cell junctions: happy together. Nat Rev Mol Cell Biol. 2004;5(4):261–70.
Narayanan S, Loganathan G, Dhanasekaran M, Tucker W, Patel A, Subhashree V, et al. Intra-islet endothelial cell and β-cell crosstalk: implication for islet cell transplantation. World J Transplantation. 2017;7(2):117–28.
Peiris H, Bonder CS, Coates PTH, Keating DJ, Jessup CF. The β-Cell/EC axis: how do islet cells talk to each other? Diabetes. 2013;63(1):3–11.
Figliolini F, Cantaluppi V, De Lena M, Beltramo S, Romagnoli R, Salizzoni M, et al. Isolation, characterization and potential role in beta cell-endothelium cross-talk of extracellular vesicles released from human pancreatic Islets. PLoS ONE. 2014;9(7):e102521.
Johansson M, Mattsson Gr, Andersson A, Jansson L, Carlsson P-O. Islet endothelial cells and pancreatic β-cell proliferation: studies in vitro and during pregnancy in adult rats. Endocrinology. 2006;147(5):2315–24.
Ribeiro D, Andersson E-M, Heath N, Persson-kry A, Collins R, Hicks R, et al. Human pancreatic islet-derived extracellular vesicles modulate insulin expression in 3D-differentiating iPSC clusters. PLoS ONE. 2017;12(11):e0187665.
Mattke J, Vasu S, Darden CM, Kumano K, Lawrence MC, Naziruddin B. Role of exosomes in islet transplantation. Front Endocrinol. 2021;12.
Rao C, Cater DT, Roy S, Xu J, De Oliveira AG, Evans-Molina C et al. Beta cell extracellular vesicle PD-L1 as a novel regulator of CD8 + T cell activity and biomarker during the evolution of type 1 diabetes. Diabetologia. 2024.
Nwadozi E, Rudnicki M, Haas TL. Metabolic coordination of pericyte phenotypes: therapeutic implications. Front Cell Dev Biology. 2020;8.
Almaça J, Weitz J, Rodriguez-Diaz R, Pereira E, Caicedo A. The pericyte of the pancreatic islet regulates capillary diameter and local blood flow. Cell Metabol. 2018;27(3):630–e444.
Mateus Gonçalves L, Fahd Qadir MM, Boulina M, Makhmutova M, Pereira E, Almaça J. Pericyte dysfunction and impaired vasomotion are hallmarks of islets during the pathogenesis of type 1 diabetes. Cell Rep. 2023;42(8):112913.
Granlund L, Hedin A, Korsgren O, Skog O, Lundberg M. Altered microvasculature in pancreatic islets from subjects with type 1 diabetes. PLoS ONE. 2022;17(10):e0276942.
Canzano JS, Nasif LH, Butterworth EA, Fu DA, Atkinson MA, Campbell-Thompson M. Islet microvasculature alterations with loss of beta-cells in patients with type 1 diabetes. J Histochem Cytochemistry. 2019;67(1):41–52.
Nwadozi E, Haas TL. Emerging roles of pericytes in coordinating skeletal muscle functions: implications and therapeutic potential. Curr Tissue Microenvironment Rep. 2021;2(3):29–39.
Winkler EA, Bell RD, Zlokovic BV. Central nervous system pericytes in health and disease. Nat Neurosci. 2011;14(11):1398–405.
Brandt MM, van Dijk CGM, Maringanti R, Chrifi I, Kramann R, Verhaar MC, et al. Transcriptome analysis reveals microvascular endothelial cell-dependent pericyte differentiation. Sci Rep. 2019;9(1):15586.
Armulik A, Abramsson A, Betsholtz C. Endothelial/pericyte interactions. Circul Res. 2005;97(6):512–23.
Reichert M, Rustgi AK. Pancreatic ductal cells in development, regeneration, and neoplasia. J Clin Investig. 2011;121(12):4572–8.
Beer RL, Parsons MJ, Rovira M. Centroacinar cells: at the center of pancreas regeneration. Dev Biol. 2016;413(1):8–15.
Schnipper J, Dhennin-Duthille I, Ahidouch A, Ouadid-Ahidouch H. Ion channel signature in healthy pancreas and pancreatic ductal adenocarcinoma. Front Pharmacol. 2020;11:568993.
Balázs A, Balla Z, Kui B, Maléth J, Rakonczay Z Jr., Duerr J, et al. Ductal mucus obstruction and reduced fluid secretion are early defects in chronic pancreatitis. Front Physiol. 2018;9:632.
Dozio N, Indirli R, Giamporcaro GM, Frosio L, Mandelli A, Laurenzi A et al. Impaired exocrine pancreatic function in different stages of type 1 diabetes. BMJ Open Diabetes Res Care. 2021;9(1).
Ross JJ, Wasserfall CH, Bacher R, Perry DJ, McGrail K, Posgai AL, et al. Exocrine pancreatic enzymes are a serological biomarker for type 1 diabetes staging and pancreas size. Diabetes. 2021;70(4):944–54.
Wright JJ, Saunders DC, Dai C, Poffenberger G, Cairns B, Serreze DV, et al. Decreased pancreatic acinar cell number in type 1 diabetes. Diabetologia. 2020;63(7):1418–23.
Granlund L, Hedin A, Wahlhütter M, Seiron P, Korsgren O, Skog O et al. Histological and transcriptional characterization of the pancreatic acinar tissue in type 1 diabetes. BMJ Open Diabetes Res Care. 2021;9(1).
Lee J, Sugiyama T, Liu Y, Wang J, Gu X, Lei J, et al. Expansion and conversion of human pancreatic ductal cells into insulin-secreting endocrine cells. eLife. 2013;2:e00940.
Meda P, Pepper MS, Traub O, Willecke K, Gros D, Beyer E, et al. Differential expression of gap junction connexins in endocrine and exocrine glands. Endocrinology. 1993;133(5):2371–8.
Meda P, Bruzzone R, Chanson M, Bosco D, Orci L. Gap junctional coupling modulates secretion of exocrine pancreas. Proc Natl Acad Sci USA. 1987;84(14):4901–4.
Chanson M, Fanjul M, Bosco D, Nelles E, Suter S, Willecke K, et al. Enhanced secretion of amylase from exocrine pancreas of connexin32-deficient mice. J Cell Biol. 1998;141(5):1267–75.
Serrill JD, Sander M, Shih HP. Pancreatic exocrine tissue architecture and integrity are maintained by E-cadherin during postnatal development. Sci Rep. 2018;8(1):13451.
Yamaguchi H, Kojima T, Ito T, Kimura Y, Imamura M, Son S, et al. Transcriptional control of tight junction proteins via a protein kinase C signal pathway in human telomerase reverse transcriptase-transfected human pancreatic duct epithelial cells. Am J Pathol. 2010;177(2):698–712.
Kojima T, Yamaguchi H, Ito T, Kyuno D, Kono T, Konno T, et al. Tight junctions in human pancreatic duct epithelial cells. Tissue Barriers. 2013;1(4):e24894.
Borka K, Kaliszky P, Szabó E, Lotz G, Kupcsulik P, Schaff Z, et al. Claudin expression in pancreatic endocrine tumors as compared with ductal adenocarcinomas. Virchows Arch. 2007;450(5):549–57.
Barrera LN, Evans A, Lane B, Brumskill S, Oldfield FE, Campbell F, et al. Fibroblasts from distinct pancreatic pathologies exhibit disease-specific properties. Cancer Res. 2020;80(13):2861–73.
Rebelo R, Xavier CPR, Giovannetti E, Vasconcelos MH. Fibroblasts in pancreatic cancer: molecular and clinical perspectives. Trends Mol Med. 2023;29(6):439–53.
Han L, Wu Y, Fang K, Sweeney S, Roesner UK, Parrish M, et al. The splanchnic mesenchyme is the tissue of origin for pancreatic fibroblasts during homeostasis and tumorigenesis. Nat Commun. 2023;14(1):1.
Erkan M, Adler G, Apte MV, Bachem MG, Buchholz M, Detlefsen S, et al. StellaTUM: current consensus and discussion on pancreatic stellate cell research. Gut. 2012;61(2):172–8.
Amrutkar M, Aasrum M, Verbeke CS, Gladhaug IP. Secretion of fibronectin by human pancreatic stellate cells promotes chemoresistance to gemcitabine in pancreatic cancer cells. BMC Cancer. 2019;19(1):596.
Yang J, Waldron RT, Su H-Y, Moro A, Chang H-H, Eibl G, et al. Insulin promotes proliferation and fibrosing responses in activated pancreatic stellate cells. Am J Physiology-Gastrointestinal Liver Physiol. 2016;311(4):G675–87.
Yang Y, Kim JW, Park HS, Lee EY, Yoon KH. Pancreatic stellate cells in the islets as a novel target to preserve the pancreatic β-cell mass and function. J Diabetes Invest. 2020;11(2):268–80.
Kim J-W, Park S-Y, You Y-H, Ham D-S, Lee S-H, Yang HK, et al. Suppression of ROS production by exendin-4 in PSC attenuates the high glucose-induced islet fibrosis. PLoS ONE. 2016;11(12):e0163187.
Zang G, Sandberg M, Carlsson PO, Welsh N, Jansson L, Barbu A. Activated pancreatic stellate cells can impair pancreatic islet function in mice. Ups J Med Sci. 2015;120(3):169–80.
Paul PK, Das R, Drow TJ, de Souza AH, Balamurugan AN, Belt Davis D, et al. Pancreatic stellate cells prolong ex vivo islet viability and function and improve engraftment. Stem Cells Translational Med. 2022;11(6):630–43.
Li J, Chen B, Fellows GF, Goodyer CG, Wang R. Activation of pancreatic stellate cells is beneficial for exocrine but not endocrine cell differentiation in the developing human pancreas. Front Cell Dev Biology. 2021;9.
Wang Z, He R, Dong S, Zhou W. Pancreatic stellate cells in pancreatic cancer: as potential targets for future therapy. Front Oncol. 2023;13.
Amrutkar M, Berg K, Balto A, Skilbrei MG, Finstadsveen AV, Aasrum M, et al. Pancreatic stellate cell-induced gemcitabine resistance in pancreatic cancer is associated with LDHA- and MCT4-mediated enhanced Glycolysis. Cancer Cell Int. 2023;23(1):9.
Chang M, Chen W, Xia R, Peng Y, Niu P, Fan H. Pancreatic stellate cells and the targeted therapeutic strategies in chronic pancreatitis. Molecules. 2023;28(14):5586.
Hamada S, Matsumoto R, Masamune A. Pancreatic stellate cells and metabolic alteration: physiology and pathophysiology. Front Physiol. 2022;13.
Shi X, Wang M, Zhang Y, Guo X, Liu M, Zhou Z et al. Hypoxia activated HGF expression in pancreatic stellate cells confers resistance of pancreatic cancer cells to EGFR Inhibition. eBioMedicine. 2022;86.
Liu H, Zhang H, Liu X, Guo W, Liu Q, Chen L, et al. Pancreatic stellate cells exploit Wnt/β-catenin/TCF7-mediated glutamine metabolism to promote pancreatic cancer cells growth. Cancer Lett. 2023;555:216040.
Ng B, Viswanathan S, Widjaja AA, Lim W-W, Shekeran SG, Goh JWT, et al. IL11 activates pancreatic stellate cells and causes pancreatic inflammation, fibrosis and atrophy in a mouse model of pancreatitis. Int J Mol Sci. 2022;23(7):3549.
Sarkar R, Xu Z, Perera CJ, Apte MV. Emerging role of pancreatic stellate cell-derived extracellular vesicles in pancreatic cancer. Sem Cancer Biol. 2023;93:114–22.
Jin G, Hong W, Guo Y, Bai Y, Chen B. Molecular mechanism of pancreatic stellate cells activation in chronic pancreatitis and pancreatic cancer. J Cancer. 2020;11(6):1505–15.
Zhang Z, Zhang H, Liu T, Chen T, Wang D, Tang D. Heterogeneous pancreatic stellate cells are powerful contributors to the malignant progression of pancreatic cancer. Front Cell Dev Biology. 2021;9.
Li C, Cui L, Yang L, Wang B, Zhuo Y, Zhang L, et al. Pancreatic stellate cells promote tumor progression by promoting an immunosuppressive microenvironment in murine models of pancreatic cancer. Pancreas. 2020;49(1):120–7.
Xiao Y, Qin T, Sun L, Qian W, Li J, Duan W, et al. Resveratrol ameliorates the malignant progression of pancreatic cancer by inhibiting hypoxia-induced pancreatic stellate cell activation. Cell Transplant. 2020;29:0963689720929987.
Ma Z, Ruedl C. Turnover kinetics of pancreatic macrophages in lean and obese diabetic mice. Front Endocrinol. 2022;13.
Calderon B, Carrero JA, Ferris ST, Sojka DK, Moore L, Epelman S, et al. The pancreas anatomy conditions the origin and properties of resident macrophages. J Exp Med. 2015;212(10):1497–512.
Van Gassen N, Staels W, Van Overmeire E, De Groef S, Sojoodi M, Heremans Y, et al. Concise review: macrophages: versatile gatekeepers during pancreatic β-cell development, injury, and regeneration. Stem Cells Translational Med. 2015;4(6):555–63.
Cruz AF, Rohban R, Esni F. Macrophages in the pancreas: villains by circumstances, not necessarily by actions. Immun Inflamm Dis. 2020;8(4):807–24.
Bijnen M, Bajénoff M. Gland macrophages: reciprocal control and function within their niche. Trends Immunol. 2021;42(2):120–36.
Jensen DM, Hendricks KV, Mason AT, Tessem JS. Good cop, bad cop: the opposing effects of macrophage activation state on maintaining or damaging functional β-cell mass. Metabolites. 2020;10(12):485.
Orliaguet L, Dalmas E, Drareni K, Venteclef N, Alzaid F. Mechanisms of macrophage polarization in insulin signaling and sensitivity. Front Endocrinol. 2020;11.
Weitz JR, Makhmutova M, Almaça J, Stertmann J, Aamodt K, Brissova M, et al. Mouse pancreatic islet macrophages use locally released ATP to monitor beta cell activity. Diabetologia. 2018;61(1):182–92.
Liou G-Y, Döppler H, Necela B, Krishna M, Crawford HC, Raimondo M, et al. Macrophage-secreted cytokines drive pancreatic acinar-to-ductal metaplasia through NF-κB and MMPs. J Cell Biol. 2013;202(3):563–77.
Aghdassi AA, Mayerle J, Christochowitz S, Weiss FU, Sendler M, Lerch MM. Animal models for investigating chronic pancreatitis. Fibrogenesis Tissue Repair. 2011;4(1):26.
Diana J, Lehuen A. Macrophages and β-cells are responsible for CXCR2-mediated neutrophil infiltration of the pancreas during autoimmune diabetes. EMBO Mol Med. 2014;6(8):1090–104.
Xiao X, Gaffar I, Guo P, Wiersch J, Fischbach S, Peirish L, et al. M2 macrophages promote beta-cell proliferation by up-regulation of SMAD7. Proc Natl Acad Sci. 2014;111(13):E1211–20.
Chittezhath M, Gunaseelan D, Zheng X, Hasan R, Tay VSY, Lim ST, et al. Islet macrophages are associated with islet vascular remodeling and compensatory hyperinsulinemia during diabetes. Am J Physiology-Endocrinology Metabolism. 2019;317(6):E1108–20.
Merad M, Manz MG. Dendritic cell homeostasis. Blood. 2009;113(15):3418–27.
Calderon B, Suri A, Miller MJ, Unanue ER. Dendritic cells in Islets of Langerhans constitutively present Β cell-derived peptides bound to their class II MHC molecules. Proc Natl Acad Sci. 2008;105(16):6121–6.
Bedrosian AS, Nguyen AH, Hackman M, Connolly MK, Malhotra A, Ibrahim J, et al. Dendritic cells promote pancreatic viability in mice with acute pancreatitis. Gastroenterology. 2011;141(5):1915–e2614.
Radenkovic M, Uvebrant K, Skog O, Sarmiento L, Avartsson J, Storm P, et al. Characterization of resident lymphocytes in human pancreatic islets. Clin Exp Immunol. 2016;187(3):418–27.
Kumar BV, Connors TJ, Farber DL. Human T cell development, localization, and function throughout life. Immunity. 2018;48(2):202–13.
Denroche HC, Miard S, Sallé-Lefort S, Picard F, Verchere CB. T cells accumulate in non-diabetic islets during ageing. Immun Ageing. 2021;18(1):8.
Weisberg SP, Carpenter DJ, Chait M, Dogra P, Gartrell-Corrado RD, Chen AX, et al. Tissue-resident memory T cells mediate immune homeostasis in the human pancreas through the PD-1/PD-L1 pathway. Cell Rep. 2019;29(12):3916–e325.
Kuric E, Seiron P, Krogvold L, Edwin B, Buanes T, Hanssen KF, et al. Demonstration of tissue resident memory CD8 T cells in insulitic lesions in adult patients with recent-onset type 1 diabetes. Am J Pathol. 2017;187(3):581–8.
Lu J, Zhang C, Li L, Xue W, Zhang C, Zhang X. Unique features of pancreatic-resident regulatory T cells in autoimmune type 1 diabetes. Front Immunol. 2017;8.
Lkhagvasuren B, Mee-inta O, Zhao Z-W, Hiramoto T, Boldbaatar D, Kuo Y-M. Pancreas-brain crosstalk. Front Neuroanat. 2021;15.
Hampton RF, Jimenez-Gonzalez M, Stanley SA. Unravelling innervation of pancreatic islets. Diabetologia. 2022;65(7):1069–84.
Borden P, Houtz J, Leach Steven D, Kuruvilla R. Sympathetic innervation during development is necessary for pancreatic islet architecture and functional maturation. Cell Rep. 2013;4(2):287–301.
Forbes S, Bond AR, Thirlwell KL, Burgoyne P, Samuel K, Noble J, et al. Human umbilical cord perivascular cells improve human pancreatic islet transplant function by increasing vascularization. Sci Transl Med. 2020;12(526):eaan5907.
Lebreton F, Bellofatto K, Wassmer CH, Perez L, Lavallard V, Parnaud G, et al. Shielding islets with human amniotic epithelial cells enhances islet engraftment and revascularization in a murine diabetes model. Am J Transplant. 2020;20(6):1551–61.
Teramura Y, Ekdahl KN, Barbu A. A hybrid of cells and pancreatic islets toward a new bioartificial pancreas. Regenerative Therapy. 2016;3:68–74.
Aloy-Reverté C, Moreno-Amador JL, Nacher M, Montanya E, Semino CE. Use of RGD-functionalized sandwich cultures to promote redifferentiation of human pancreatic beta cells after in vitro expansion. Tissue Eng Part A. 2018;24(5–6):394–406.
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.
Seo H, Son J, Park J-K. Controlled 3D co-culture of beta cells and endothelial cells in a micropatterned collagen sheet for reproducible construction of an improved pancreatic pseudo-tissue. APL Bioeng. 2020;4(4).
Lee K-M, Jung G-S, Park J-K, Choi S-K, Jeon WB. Effects of Arg–Gly–Asp-modified elastin-like polypeptide on pseudoislet formation via up-regulation of cell adhesion molecules and extracellular matrix proteins. Acta Biomater. 2013;9(3):5600–8.
Kim J, Shim IK, Hwang DG, Lee YN, Kim M, Kim H, et al. 3D cell printing of islet-laden pancreatic tissue-derived extracellular matrix bioink constructs for enhancing pancreatic functions. J Mater Chem B. 2019;7(10):1773–81.
Dong H, Fahmy TM, Metcalfe SM, Morton SL, Dong X, Inverardi L, et al. Immuno-isolation of pancreatic islet allografts using pegylated nanotherapy leads to long-term normoglycemia in full MHC mismatch recipient mice. PLoS ONE. 2012;7(12):e50265.
Kepsutlu B, Nazli C, Bal T, Kizilel S. Design of bioartificial pancreas with functional micro/nano-based encapsulation of Islets. Curr Pharm Biotechnol. 2014;15(7):590–608.
Stock AA, Manzoli V, De Toni T, Abreu MM, Poh Y-C, Ye L, et al. Conformal coating of stem cell-derived islets for Β cell replacement in type 1 diabetes. Stem Cell Rep. 2020;14(1):91–104.
Teramura Y, Iwata H. Bioartificial pancreas: microencapsulation and conformal coating of islet of Langerhans. Adv Drug Deliv Rev. 2010;62(7):827–40.
Wassmer C-H, Lebreton F, Bellofatto K, Bosco D, Berney T, Berishvili E. Generation of insulin-secreting organoids: a step toward engineering and transplanting the bioartificial pancreas. Transpl Int. 2020;33(12):1577–88.
Beydag-Tasöz BS, Yennek S, Grapin-Botton A. Towards a better understanding of diabetes mellitus using organoid models. Nat Reviews Endocrinol. 2023;19(4):232–48.
Kinoshita N, Echigo Y, Shinohara S, Gu Y, Miyazaki J, Inoue K, et al. Regulation of cell proliferation using tissue engineering in MIN6 cells. Cell Transplant. 2001;10(4–5):473–7.
Jun Y, Kang AR, Lee JS, Jeong GS, Ju J, Lee DY, et al. 3D co-culturing model of primary pancreatic islets and hepatocytes in hybrid spheroid to overcome pancreatic cell shortage. Biomaterials. 2013;34(15):3784–94.
Kim HJ, Alam Z, Hwang JW, Hwang YH, Kim MJ, Yoon S et al. Optimal formation of genetically modified and functional pancreatic islet spheroids by using hanging-drop strategy. Transplantation Proceedings. 2013;45(2):605– 10.
Tanaka H, Tanaka S, Sekine K, Kita S, Okamura A, Takebe T, et al. The generation of pancreatic β-cell spheroids in a simulated microgravity culture system. Biomaterials. 2013;34(23):5785–91.
Jo YH, Jang IJ, Nemeno JG, Lee S, Kim BY, Nam BM et al. Artificial islets from hybrid spheroids of three pancreatic cell lines. Transplantation Proceedings. 2014;46(4):1156-60.
Kusamori K, Nishikawa M, Mizuno N, Nishikawa T, Masuzawa A, Shimizu K, et al. Transplantation of insulin-secreting multicellular spheroids for the treatment of type 1 diabetes in mice. J Controlled Release. 2014;173:119–24.
Kusamori K, Nishikawa M, Mizuno N, Nishikawa T, Masuzawa A, Tanaka Y, et al. Increased insulin secretion from insulin-secreting cells by construction of mixed multicellular spheroids. Pharm Res. 2016;33(1):247–56.
Suzuki T, Kanamori T, Inouye S. Quantitative visualization of synchronized insulin secretion from 3D-cultured cells. Biochem Biophys Res Commun. 2017;486(4):886–92.
Fukuda Y, Akagi T, Asaoka T, Eguchi H, Sasaki K, Iwagami Y, et al. Layer-by-layer cell coating technique using extracellular matrix facilitates rapid fabrication and function of pancreatic β-cell spheroids. Biomaterials. 2018;160:82–91.
Lee G, Jun Y, Jang H, Yoon J, Lee J, Hong M, et al. Enhanced oxygen permeability in membrane-bottomed concave microwells for the formation of pancreatic islet spheroids. Acta Biomater. 2018;65:185–96.
Lee SH, Hong S, Song J, Cho B, Han EJ, Kondapavulur S, et al. Microphysiological analysis platform of pancreatic islet β-cell spheroids. Adv Healthc Mater. 2018;7(2):1701111.
Hashim M, Yokoi N, Takahashi H, Gheni G, Okechi OS, Hayami T, et al. Inhibition of SNAT5 induces incretin-responsive state from incretin-unresponsive state in pancreatic β-cells: study of β-cell spheroid clusters as a model. Diabetes. 2018;67(9):1795–806.
Gao B, Jing C, Ng K, Pingguan-Murphy B, Yang Q. Fabrication of three-dimensional islet models by the geometry-controlled hanging-drop method. Acta Mech Sin. 2019;35(2):329–37.
Mizukami Y, Takahashi Y, Shimizu K, Konishi S, Takakura Y, Nishikawa M. Regulation of the distribution of cells in mixed spheroids by altering migration direction. Tissue Eng Part A. 2019;25(5–6):390–8.
Myasnikova D, Osaki T, Onishi K, Kageyama T, Zhang Molino B, Fukuda J. Synergic effects of oxygen supply and antioxidants on pancreatic β-cell spheroids. Sci Rep. 2019;9(1):1802.
Bal T, Inceoglu Y, Karaoz E, Kizilel S. Sensitivity study for the key parameters in heterospheroid preparation with insulin-secreting β-cells and mesenchymal stem cells. ACS Biomaterials Sci Eng. 2019;5(10):5229–39.
Lee YN, Yi HJ, Goh H, Park JY, Ferber S, Shim IK, et al. Spheroid fabrication using concave microwells enhances the differentiation efficacy and function of insulin-producing cells via cytoskeletal changes. Cells. 2020;9(12):2551.
Wassmer C-H, Bellofatto K, Perez L, Lavallard V, Cottet-Dumoulin D, Ljubicic S, et al. Engineering of primary pancreatic islet cell spheroids for three-dimensional culture or transplantation: a methodological comparative study. Cell Transplant. 2020;29:0963689720937292.
Essaouiba A, Jellali R, Shinohara M, Scheidecker B, Legallais C, Sakai Y, et al. Analysis of the behavior of 2D monolayers and 3D spheroid human pancreatic beta cells derived from induced pluripotent stem cells in a microfluidic environment. J Biotechnol. 2021;330:45–56.
Essaouiba A, Jellali R, Poulain S, Tokito F, Gilard F, Gakière B, et al. Analysis of the transcriptome and metabolome of pancreatic spheroids derived from human induced pluripotent stem cells and matured in an organ-on-a-chip. Mol Omics. 2022;18(8):791–804.
Petry F, Salzig D. Large-scale production of size-adjusted β-cell spheroids in a fully controlled stirred-tank reactor. Processes. 2022;10(5):861.
Jeong JH, Park KN, Kim JH, Noh K, Hur SS, Kim Y, et al. Self-organized insulin-producing β-cells differentiated from human omentum-derived stem cells and their in vivo therapeutic potential. Biomaterials Res. 2023;27(1):82.
Candiello J, Grandhi TSP, Goh SK, Vaidya V, Lemmon-Kishi M, Eliato KR, et al. 3D heterogeneous islet organoid generation from human embryonic stem cells using a novel engineered hydrogel platform. Biomaterials. 2018;177:27–39.
Lebreton F, Lavallard V, Bellofatto K, Bonnet R, Wassmer CH, Perez L, et al. Insulin-producing organoids engineered from islet and amniotic epithelial cells to treat diabetes. Nat Commun. 2019;10(1):4491.
Tao T, Wang Y, Chen W, Li Z, Su W, Guo Y, et al. Engineering human islet organoids from iPSCs using an organ-on-chip platform. Lab Chip. 2019;19(6):948–58.
Nalbach L, Roma LP, Schmitt BM, Becker V, Körbel C, Wrublewsky S, et al. Improvement of islet transplantation by the fusion of islet cells with functional blood vessels. EMBO Mol Med. 2021;13(1):e12616.
Voznesenskaya A, Berggren P-O, Ilegems E. Sustained heterologous gene expression in pancreatic islet organoids using adeno-associated virus serotype 8. Front Bioeng Biotechnol. 2023;11.
Olsen C, Wang C, Abadpour S, Lundanes E, Hansen AS, Skottvoll FS, et al. Determination of insulin secretion from stem cell-derived islet organoids with liquid chromatography-tandem mass spectrometry. J Chromatogr B. 2023;1215:123577.
Luther MJ, Davies E, Muller D, Harrison M, Bone AJ, Persaud SJ, et al. Cell-to-cell contact influences proliferative marker expression and apoptosis in MIN6 cells grown in islet-like structures. Am J Physiology-Endocrinology Metabolism. 2005;288(3):E502–9.
Zhi Z-l, Liu B, Jones PM, Pickup JC. Polysaccharide multilayer nanoencapsulation of insulin-producing β-cells grown as pseudoislets for potential cellular delivery of insulin. Biomacromolecules. 2010;11(3):610–6.
Bhaiji T, Zhi Z-L, Pickup JC. Improving cellular function and immune protection via layer-by-layer nanocoating of pancreatic islet β-cell spheroids cocultured with mesenchymal stem cells. J Biomedical Mater Res Part A. 2012;100A(6):1628–36.
Spelios MG, Kenna LA, Wall B, Akirav EM. In vitro formation of Β cell pseudoislets using islet-derived endothelial cells. PLoS ONE. 2013;8(8):e72260.
Green AD, Vasu S, McClenaghan NH, Flatt PR. Implanting 1.1B4 human β-cell pseudoislets improves glycaemic control in diabetic severe combined immune deficient mice. World J Diabetes. 2016;7(19):523–33.
Ichihara Y, Utoh R, Yamada M, Shimizu T, Uchigata Y. Size effect of engineered islets prepared using microfabricated wells on islet cell function and arrangement. Heliyon. 2016;2(6):e00129.
Lecomte M-J, Pechberty S, Machado C, Da Barroca S, Ravassard P, Scharfmann R, et al. Aggregation of engineered human β-cells into pseudoislets: insulin secretion and gene expression profile in normoxic and hypoxic milieu. Cell Med. 2016;8(3):99–112.
Bal T, Oran DC, Sasaki Y, Akiyoshi K, Kizilel S. Sequential coating of insulin secreting beta cells within multilayers of polysaccharide nanogels. Macromol Biosci. 2018;18(5):1800001.
Harata M, Liu S, Promes JA, Burand AJ, Ankrum JA, Imai Y. Delivery of ShRNA via lentivirus in human pseudoislets provides a model to test dynamic regulation of insulin secretion and gene function in human Islets. Physiological Rep. 2018;6(20):e13907.
Yu Y, Gamble A, Pawlick R, Pepper AR, Salama B, Toms D, et al. Bioengineered human pseudoislets form efficiently from donated tissue, compare favourably with native Islets in vitro and restore normoglycaemia in mice. Diabetologia. 2018;61(9):2016–29.
Furuyama K, Chera S, van Gurp L, Oropeza D, Ghila L, Damond N, et al. Diabetes relief in mice by glucose-sensing insulin-secreting human α-cells. Nature. 2019;567(7746):43–8.
Lorza-Gil E, Gerst F, Oquendo MB, Deschl U, Häring H-U, Beilmann M, et al. Glucose, adrenaline and palmitate antagonistically regulate insulin and glucagon secretion in human pseudoislets. Sci Rep. 2019;9(1):10261.
van Krieken PP, Voznesenskaya A, Dicker A, Xiong Y, Park JH, Lee JI, et al. Translational assessment of a genetic engineering methodology to improve islet function for transplantation. eBioMedicine. 2019;45:529–41.
Drzazga A, Cichońska E, Koziołkiewicz M, Gendaszewska-Darmach E. Formation of βTC3 and MIN6 pseudoislets changes the expression pattern of Gpr40, Gpr55, and Gpr119 receptors and improves lysophosphatidylcholines-potentiated glucose-stimulated insulin secretion. Cells. 2020;9(9):2062.
Walker JT, Haliyur R, Nelson HA, Ishahak M, Poffenberger G, Aramandla R et al. Integrated human pseudoislet system and microfluidic platform demonstrate differences in GPCR signaling in islet cells. JCI Insight. 2020;5(10).
Zbinden A, Marzi J, Schlünder K, Probst C, Urbanczyk M, Black S et al. Non-invasive marker-independent high content analysis of a microphysiological human pancreas-on-a-chip model. Matrix Biol. 2020;85–6:205–20.
Hart NJ, Weber C, Price N, Banuelos A, Schultz M, Huey B, et al. Insulinoma-derived pseudo-islets for diabetes research. Am J Physiology-Cell Physiol. 2021;321(2):C247–56.
Wieland FC, Sthijns M, Geuens T, van Blitterswijk CA, LaPointe VLS. The role of alpha cells in the self-assembly of bioengineered Islets. Tissue Eng Part A. 2021;27(15–16):1055–63.
Sokolowska P, Jastrzebska E, Dobrzyn A, Brzozka Z. Investigation of the therapeutic potential of new antidiabetic compounds using islet-on-a-chip microfluidic model. Biosensors. 2022;12(5):302.
Rademakers T, Sthijns MMJPE, Paulino da Silva Filho O, Joris V, Oosterveer J, Lam TW, et al. Identification of compounds protecting pancreatic islets against oxidative stress using a 3D pseudoislet screening platform. Adv Biology. 2023;7(12):2300264.
Rubio-Navarro A, Gómez-Banoy N, Stoll L, Dündar F, Mawla AM, Ma L, et al. A beta cell subset with enhanced insulin secretion and glucose metabolism is reduced in type 2 diabetes. Nat Cell Biol. 2023;25(4):565–78.
Wang L, Wan J, Xu Y, Huang Y, Wang D, Zhu D et al. Endothelial cells promote pseudo-islet function through BTC-EGFR-JAK/STAT signaling pathways. Ann Biomed Eng. 2024.
Hospodiuk M, Dey M, Ayan B, Sosnoski D, Moncal KK, Wu Y, et al. Sprouting angiogenesis in engineered pseudo islets. Biofabrication. 2018;10(3):035003.
Urbanczyk M, Zbinden A, Layland SL, Duffy G, Schenke-Layland K. Controlled heterotypic pseudo-islet assembly of human β-cells and human umbilical vein endothelial cells using magnetic levitation. Tissue Eng Part A. 2020;26(7–8):387–99.
Oran DC, Lokumcu T, Inceoglu Y, Akolpoglu MB, Albayrak O, Bal T, et al. Engineering human stellate cells for beta cell replacement therapy promotes in vivo recruitment of regulatory T cells. Mater Today Bio. 2019;2:100006.
Kumar S, Jach D, Macfarlane W, Crnogorac-Jurcevic T. A 3-dimensional coculture model to visualize and monitor interaction between pancreatic cancer and islet Β cells. Pancreas. 2021;50(7):982–9.
Wassmer C-H, Lebreton F, Bellofatto K, Perez L, Cottet-Dumoulin D, Andres A et al. Bio-engineering of pre-vascularized islet organoids for the treatment of type 1 diabetes. Transpl Int. 2022;35.
O’Sullivan ES, Johnson AS, Omer A, Hollister-Lock J, Bonner-Weir S, Colton CK, et al. Rat islet cell aggregates are superior to Islets for transplantation in microcapsules. Diabetologia. 2010;53(5):937–45.
Govindasamy V, Ronald VS, Abdullah AN, Nathan KRG, Ab. Aziz ZAC, Abdullah M, et al. Differentiation of dental pulp stem cells into islet-like aggregates. J Dent Res. 2011;90(5):646–52.
Chandra V, Muthyala GS, Jaiswal S, Bellare AK, Nair JR. Islet-like cell aggregates generated from human adipose tissue derived stem cells ameliorate experimental diabetes in mice. PLoS ONE. 2011;6(6):e20615.
Bernard AB, Lin CC, Anseth KS. A microwell cell culture platform for the aggregation of pancreatic β-cells. Tissue Eng Part C Methods. 2012;18(8):583–92.
Liu X, Yan F, Yao H, Chang M, Qin J, Li Y, et al. Involvement of RhoA/ROCK in insulin secretion of pancreatic β-cells in 3D culture. Cell Tissue Res. 2014;358(2):359–69.
Hilderink J, Spijker S, Carlotti F, Lange L, Engelse M, van Blitterswijk C, et al. Controlled aggregation of primary human pancreatic islet cells leads to glucose-responsive pseudoislets comparable to native Islets. J Cell Mol Med. 2015;19(8):1836–46.
Kim SM, Lee EJ, Jung HS, Han N, Kim YJ, Kim TK, et al. Co-culture of α TC-6 cells and Β TC-1 cells: morphology and function. Endocrinol Metabolism (Seoul). 2015;30(1):92–7.
Takemoto N, Liu X, Takii K, Teramura Y, Iwata H. Transplantation of co-aggregates of Sertoli cells and islet cells into liver without immunosuppression. Transplantation. 2014;97(3):287–93.
Inoo K, Bando H, Tabata Y. Enhanced survival and insulin secretion of Insulinoma cell aggregates by incorporating gelatin hydrogel microspheres. Regenerative Therapy. 2018;8:29–37.
Skrzypek K, Barrera YB, Groth T, Stamatialis D. Endothelial and beta cell composite aggregates for improved function of a bioartificial pancreas encapsulation device. Int J Artif Organs. 2018;41(3):152–9.
Chen H, Zeng T, Yoshitomi T, Kawazoe N, Komatsu H, Yang Y, et al. Porous microwell scaffolds for 3D culture of pancreatic beta cells to promote cell aggregation and insulin secretion. Mater Adv. 2024;5(5):2019–26.
Yook S, Jeong J-H, Jung YS, Hong SW, Im BH, Seo JW, et al. Molecularly engineered islet cell clusters for diabetes mellitus treatment. Cell Transplant. 2012;21(8):1775–89.
Dang LT-T, Bui AN-T, Pham VM, Phan NK, Van Pham P. Production of islet-like insulin-producing cell clusters in vitro from adiposederived stem cells. Biomedical Res Therapy. 2015;2(1):3.
Pathak S, Regmi S, Gupta B, Pham TT, Yong CS, Kim JO, et al. Engineered islet cell clusters transplanted into subcutaneous space are superior to pancreatic islets in diabetes. FASEB J. 2017;31(11):5111–21.
Velazco-Cruz L, Song J, Maxwell KG, Goedegebuure MM, Augsornworawat P, Hogrebe NJ, et al. Acquisition of dynamic function in human stem cell-derived Β cells. Stem Cell Rep. 2019;12(2):351–65.
Suenaga R, Konagaya S, Yamaura J, Ito R, Tanaka S, Ishizaki Y, et al. Microwell bag culture for large-scale production of homogeneous islet-like clusters. Sci Rep. 2022;12(1):5221.
Liang S, Zhao J, Baker RK, Tran E, Zhan L, Kieffer TJ. Differentiation of stem cell-derived pancreatic progenitors into insulin-secreting islet clusters in a multiwell-based static 3D culture system. Cell Rep Methods. 2023;3(5):100466.
Shin JY, Jeong JH, Han J, Bhang SH, Jeong GJ, Haque MR, et al. Transplantation of heterospheroids of islet cells and mesenchymal stem cells for effective angiogenesis and antiapoptosis. Tissue Eng Part A. 2015;21(5–6):1024–35.
Pathak S, Regmi S, Gupta B, Poudel BK, Pham TT, Kim J-R, et al. Hybrid congregation of islet single cells and curcumin-loaded polymeric microspheres as an -interventional strategy to overcome apoptosis associated with pancreatic Islets transplantation. ACS Appl Mater Interfaces. 2016;8(39):25702–13.
Takahashi Y, Sekine K, Kin T, Takebe T, Taniguchi H. Self-condensation culture enables vascularization of tissue fragments for efficient therapeutic transplantation. Cell Rep. 2018;23(6):1620–9.
Navarro-Tableros V, Gai C, Gomez Y, Giunti S, Pasquino C, Deregibus MC, et al. Islet-like structures generated in vitro from adult human liver stem cells revert hyperglycemia in diabetic SCID mice. Stem Cell Reviews Rep. 2019;15(1):93–111.
Balboa D, Barsby T, Lithovius V, Saarimäki-Vire J, Omar-Hmeadi M, Dyachok O, et al. Functional, metabolic and transcriptional maturation of human pancreatic islets derived from stem cells. Nat Biotechnol. 2022;40(7):1042–55.
Verhoeff K, Cuesta-Gomez N, Maghera J, Dadheech N, Pawlick R, Smith N, et al. Scalable bioreactor-based suspension approach to generate stem cell-derived islets from healthy donor-derived iPSCs. Transplantation. 2025;109(1):e22–35.
Yesildag B, Mir-Coll J, Neelakandhan A, Gibson CB, Perdue NR, Rufer C, et al. Liraglutide protects β-cells in novel human islet spheroid models of type 1 diabetes. Clin Immunol. 2022;244:109118.
Hermanns C, da Silva Filho OP, Vaithilingam V, van Apeldoorn A. The potential of cell sheet technology for beta cell replacement therapy. Curr Transplantation Rep. 2022;9(3):199–208.
Homma J, Sekine H, Shimizu T. Tricultured cell sheets develop into functional pancreatic islet tissue with a vascular network. Tissue Eng Part A. 2023;29(7–8):211–24.
Fujita I, Utoh R, Yamamoto M, Okano T, Yamato M. The liver surface as a favorable site for islet cell sheet transplantation in type 1 diabetes model mice. Regenerative Therapy. 2018;8:65–72.
Diane A, Mohammed LI, Al-Siddiqi HH. Islets in the body are never flat: transitioning from two-dimensional (2D) monolayer culture to three-dimensional (3D) spheroid for better efficiency in the generation of functional hPSC-derived pancreatic Β cells in vitro. Cell Communication Signal. 2023;21(1):151.
Katt ME, Placone AL, Wong AD, Xu ZS, Searson PC. In vitro tumor models: advantages, disadvantages, variables, and selecting the right platform. Front Bioeng Biotechnol. 2016;4.
Tevlek A, Kecili S, Ozcelik OS, Kulah H, Tekin HC. Spheroid engineering in microfluidic devices. ACS Omega. 2023;8(4):3630–49.
Lin R-Z, Chang H-Y. Recent advances in three-dimensional multicellular spheroid culture for biomedical research. Biotechnol J. 2008;3(9–10):1172–84.
Puginier E, Leal-Fischer K, Gaitan J, Lallouet M, Scotti P-A, Raoux M, et al. Extracellular electrophysiology on clonal human β-cell spheroids. Front Endocrinol. 2024;15:1402880.
Sakata N, Yoshimatsu G, Kawakami R, Nakano K, Yamada T, Yamamura A, et al. The Porcine islet-derived organoid showed the characteristics as pancreatic duct. Sci Rep. 2024;14(1):6401.
Zhou M, Linn T, Petry SF. EndoC-βH3 pseudoislets are suitable for intraportal transplantation in diabetic mice. Islets. 2024;16(1):2406041.
Zhou Q, Brown J, Kanarek A, Rajagopal J, Melton DA. In vivo reprogramming of adult pancreatic exocrine cells to β-cells. Nature. 2008;455(7213):627–32.
Huang F, Lai J, Qian L, Hong W, Li L-c. Differentiation of Uc-MSCs into insulin secreting islet-like clusters by trypsin through TGF-beta signaling pathway. Differentiation. 2024;135:100744.
Iworima DG, Baker RK, Ellis C, Sherwood C, Zhan L, Rezania A, et al. Metabolic switching, growth kinetics and cell yields in the scalable manufacture of stem cell-derived insulin-producing cells. Stem Cell Res Ther. 2024;15(1):1.
Cheng K, Delghingaro-Augusto V, Nolan CJ, Turner N, Hallahan N, Andrikopoulos S, et al. High passage MIN6 cells have impaired insulin secretion with impaired glucose and lipid oxidation. PLoS ONE. 2012;7(7):e40868.
Lawson R, Maret W, Hogstrand C. Prolonged stimulation of insulin release from MIN6 cells causes zinc depletion and loss of β-cell markers. J Trace Elem Med Biol. 2018;49:51–9.
Takemoto N, Konagaya S, Kuwabara R, Iwata H. Coaggregates of regulatory T cells and islet cells allow long-term graft survival in liver without immunosuppression. Transplantation. 2015;99(5):942–7.
Vlahos AE, Kinney SM, Kingston BR, Keshavjee S, Won S-Y, Martyts A, et al. Endothelialized collagen based pseudo-islets enables tuneable subcutaneous diabetes therapy. Biomaterials. 2020;232:119710.
Wang S, Du Y, Zhang B, Meng G, Liu Z, Liew SY, et al. Transplantation of chemically induced pluripotent stem-cell-derived islets under abdominal anterior rectus sheath in a type 1 diabetes patient. Cell. 2024;187(22):6152–e6418.
Wu J, Li T, Guo M, Ji J, Meng X, Fu T, et al. Treating a type 2 diabetic patient with impaired pancreatic islet function by personalized endoderm stem cell-derived islet tissue. Cell Discovery. 2024;10(1):45.
Acknowledgements
The author would like to thank Uskudar University and Bioengineering Research Group. Graphical abstract, Fig. 1 and Fig. 2 contain images from Servier Medical Art, which were modified in color, shape and size where necessary. Servier Medical Art (smart.servier.com) is licensed under a Creative Commons Attribution 4.0 International License (CC BY 4.0) (https://creativecommons.org/licenses/by/4.0/).
Funding
The author received no funding.
Author information
Authors and Affiliations
Contributions
TB performed the literature research and analyzed the information, drafted and edited the manuscript.
Corresponding author
Ethics declarations
Ethical approval and consent to participate
Not applicable.
Consent for publication
Not applicable.
Competing interests
The authors declare no competing interests.
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/.
About this article
Cite this article
Bal, T. Scaffold-free endocrine tissue engineering: role of islet organization and implications in type 1 diabetes. BMC Endocr Disord 25, 107 (2025). https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12902-025-01919-y
Received:
Accepted:
Published:
DOI: https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12902-025-01919-y