Next Article in Journal
Regulation of Immune Functions by Non-Neuronal Acetylcholine (ACh) via Muscarinic and Nicotinic ACh Receptors
Next Article in Special Issue
The Sirenic Links between Diabetes, Obesity, and Bladder Cancer
Previous Article in Journal
Protective Effect of Water-Soluble C60 Fullerene Nanoparticles on the Ischemia-Reperfusion Injury of the Muscle Soleus in Rats
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Roles of the IGF Axis in the Regulation of the Metabolism: Interaction and Difference between Insulin Receptor Signaling and IGF-I Receptor Signaling

1
Department of Endocrinology and Metabolism, Graduate School of Medicine, Yokohama City University, Yokohama 236-0004, Japan
2
Laboratory of Diabetes and Metabolic Disorders, Institute for Molecular and Cellular Regulation (IMCR), Gunma University, Maebashi 371-8510, Japan
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2021, 22(13), 6817; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22136817
Submission received: 30 May 2021 / Revised: 22 June 2021 / Accepted: 23 June 2021 / Published: 24 June 2021
(This article belongs to the Special Issue The Role of the IGF Axis in Disease 2.0)

Abstract

:
It has been well established that insulin-like growth factors (IGFs) mainly mediate long-term actions in cell fates, whereas insulin predominantly exerts its role on metabolic activity. Indeed, insulin mediates multiple anabolic biological activities in glucose and amino acid transport, lipid and protein synthesis, the induction of glycogen, the inhibition of gluconeogenesis, lipolysis, and protein degradation. The interactions and differences between insulin receptor signaling and IGF-I receptor signaling in the metabolism and the cell fates are quite complicated. Because of the overlapping actions of IGF-I singling with insulin signaling, it has been difficult to distinguish the role of both signaling mechanisms on the metabolism. Furthermore, comprehensive information on the IGF-I function in respective tissues remains insufficient. Therefore, we need to clarify the precise roles of IGF-I signaling on the metabolism separate from those of insulin signaling. This review focuses on the metabolic roles of IGFs in the respective tissues, especially in terms of comparison with those of insulin, by overviewing the metabolic phenotypes of tissue-specific IGF-I and insulin receptor knockout mice, as well as those in mice treated with the dual insulin receptor/IGF-I receptor inhibitor OSI-906.

1. Introduction

Insulin-like growth factors (IGFs: IGF-I and IGF-II) have been shown to possess a variety of bioactivities in growth, cell differentiation, cell survival, and the maintenance of cell function. IGFs are expressed in multiple cell types. Most of the circulating IGF-I/II is secreted from hepatocytes [1,2]. IGF signaling is predominantly modulated by the growth hormone (GH), while multiple other factors, such as growth factors, tropic hormones, steroid hormones, erythropoietin, or the extracellular matrix, also regulate the IGF-I secretion from the liver [3].
The structure of insulin is well known to be highly homologous to those of IGFs [4]. Insulin and IGF-I act through their respective receptors. IGFs and insulin have been shown to share common signaling pathways, even though the physiological functions of insulin and IGFs are not identical. The IGF-I receptor (IGF-IR) is expressed highly in embryos and various tissues, except the liver and adipocytes of adults, whereas the insulin receptor (IR) is dominantly expressed in the liver, adipose tissue, and muscle [5]. The relative expression levels of IGF-IR and IR vary between tissues and across development stages, and these differences serve to dictate the relative tissue-specific role of IGF-I or insulin in regulating metabolic function. IGF-I is the principal ligand for IGF-IR and IR/IGF-IR hybrid receptors. IGF-I and insulin are able to activate each other’s receptors only at high concentrations. On the other hand, IGF-II can bind to IGF-IR as well as hybrid receptors and IR-A (a splicing variant of IR). Since IGFs potently bind to IGF-IR and activate pathways associated with cellular proliferation, the IGF axis has been recognized for its contribution to cancer growth and a propensity for metastasis. It is well known that cancer tissues express high levels of both IR and IGF-IR. However, in many common solid tumors, IGF axis deregulation is not itself the driver but occurs secondary to another molecular event that influences the expression of the ligands and/or receptors [6]. Since insulin receptor signaling also stimulates cell growth, hyperinsulinemia can potentially induce pathological disorders, including cancer and cardiovascular disease. Numerous in vitro and in vivo studies have now clearly established that insulin may affect the tumor progression by acting through its own receptor and not by crosstalk with IGF-IR [7].
IGF-IR is a disulfide bond-conjugated tetrameter that consists of two α subunits that are extracellular proteins binding to specific ligands and two β subunits that are transmembrane proteins possessing tyrosine kinase activity [8]. IGF-IR is highly homologous to IR, particularly in the cytoplasmic domain. Both splice variants of the IR (IR-A and IR-B) use signaling pathways that are similar to that of IGF-IR [9]. IGF-IR possesses a high affinity for IGF-I and IGF-II, but it can also bind to insulin with a lower affinity (Figure 1). In contrast, insulin receptor type B (IR-B), which is an alternative splicing variant of the insulin receptor gene, binds to insulin with high affinity and binds to IGFs with low affinity [9]. Another splicing variant of IR (IR-A) has a high affinity for both insulin and IGF-II. These differences in specificity play important roles in the distinct bioactivities of IGFs and insulin. It has also been shown that IGF-IR and IR can interact by forming hybrid receptors or by cross-talking in downstream intracellular signaling pathways, enabling them to modulate the amplitude of each signaling via the multiplicity of ligand-receptor interactions [10,11].
IGF-IR and IR are activated and auto-phosphorylated, followed by the binding of their ligands to the receptor. Subsequently, insulin receptor substrate (IRS)-1 and IRS-2 bind to the activated IGF-IR and IR through its phosphotyrosine binding (PTB) domain [12]. The activated IGF-IR or IR phosphorylates specific substrates, such as IRS-1, IRS-2, and Shc. The phosphorylated tyrosine residues of IRS-1, IRS-2, and Shc are recognized by various signaling molecules that contain an Src homology 2 (SH2) domain, including Grb2 and the 85 kDa regulatory subunit of phosphatidylinositol 3-kinase (PI3-kinase), leading to the activation of multiple downstream cascades (Figure 1). Grb2 is involved in the activation of a Ras-MAPK pathway. PI3-kinase, which consists of a regulatory subunit responsible for binding to IRSs and a catalytic subunit responsible for the phosphorylation of phosphatidylinositols, is a lipid kinase and a key element in the pathway leading to the metabolic effects of insulin. Thus, IGF-I has also been suggested to mediate insulin-like bioactivities.
To understand the roles of IGF-I/IGF-IR signaling in the regulation of metabolism and to distinguish the action of IGF-I/IGF-IR signaling from that of insulin/IR signaling, several reports have taken advantage of the Cre recombinase/loxP system to manipulate IGF-I and IGF-IR in a tissue-specific fashion (Table 1). In addition, we previously reported the role of IGF-IR and IR signaling in multiple metabolic organs, including the liver, adipose tissue, and pancreas, by using an orally bioavailable dual IGF-IR/IR inhibitor OSI-906 [13,14,15,16].

2. IGF-I and Insulin Signaling in Metabolic Organs

2.1. IGF-I and Insulin Signaling in the Liver

Insulin is directly secreted from pancreatic β-cells into the portal system, and the liver degrades about half of the insulin that reaches the liver [17]. Thus, the liver is the major target tissue of insulin action. IR- and IGF-IR-mediated signaling play a crucial role in hepatic insulin action, which regulates glucose and fatty acid metabolism. Yakar et al. generated liver-specific IGF-I-deficient (LID) mice and reported their metabolic phenotype [18]. The LID mice showed a 75% reduction in circulating IGF-I and elevated GH levels [18]. The LID mice also showed a 4-fold increase in the serum insulin levels and abnormal glucose clearance after insulin injection. In contrast, the fasting blood glucose levels and those after a glucose tolerance test were similar between LID mice and their control littermates. Thus, liver-specific IGF-I deficiency induced insulin resistance and marked hyperinsulinemia. Since liver-specific IGF-I deletion resulted in a marked reduction in the circulating total IGF-I levels and elevated GH levels, a reduction in insulin sensitivity was observed specifically in the muscle tissues of LID mice [18]. Thus, circulating IGF-I acted as a component of overall insulin action in peripheral tissues.
In contrast, hepatocyte-specific insulin receptor knockout (LIRKO) mice exhibited severe insulin resistance and glucose intolerance, leading to a significant increase in β-cell mass as well as the failure of insulin to suppress hepatic glucose production. Glycogen storage was also impaired in juvenile mice. LIRKO mice also showed an age-dependent progression of liver dysfunction without steatosis [19].
Collectively, these reports indicate the significance of IR and IGF-IR signaling in hepatic insulin action and the metabolism of glucose and fatty acids through inter-organ networks.

2.2. IGF-I and Insulin Signaling in Skeletal Muscle

Insulin has an anabolic effect on the muscle, controlling muscle protein synthesis and inhibiting muscle protein breakdown [20,21]. Because IR and IGF-IR are highly homologous receptors, IGF-I also mediates the regulation of skeletal muscle hypertrophy and atrophy. IGF-I increases the skeletal muscle protein synthesis via PI3K/Akt/mTOR and PI3K/Akt/GSK3β pathways and potentiates skeletal muscle regeneration via the activation of skeletal muscle stem (satellite) cells [22]. IGF-IR is expressed in undifferentiated myoblasts, whereas IR is predominantly expressed in differentiated muscles, compared with IGF-IR [23]. In addition, insulin activates glucose uptake into skeletal muscle, which contributes to the maintenance of glucose homeostasis. On the other hand, the effects of IGF-IR signaling in the muscle on glucose metabolism might be limited.
Laustsen et al. [24] generated muscle-specific IGF-IR knockout mice by crossing IGF- IR-floxed mice with muscle creatine kinase (MCK) promoter-Cre mice. Although MCK is expressed in the heart as well as the skeletal muscle, these mice had a comparable ratio of heart weight to body weight and cardiac performance. There were also no differences in body weight and glucose tolerance at the age of 6 months. O’Neill et al. [25] also generated mice with IGF-IR knockout restricted to the striated muscle by crossing IGF-IR-floxed mice with transgenic Cre mice under the human skeletal muscle actin promoter (ACTA1-Cre) [26], which is active in differentiated muscle cells [22]. These mice had a normal body weight and skeletal muscle mass up to the age of 52 weeks. Glucose intolerance was not observed even after high-fat-diet feeding, compared with the controls. Mavalli et al. [27] also generated mice with IGF-IR knockout in skeletal muscle by crossing IGF-IR-floxed mice with transgenic Cre mice under the control of the Mef2c skeletal muscle promoter. Mef2c is expressed in the skeletal muscle during development and postnatally [28]. These mice exhibited an abnormal muscle development characterized by a reduction in the myofiber number and area, and they failed to attain body weight and lean/muscle mass similar to that of the control mice after aging. However, no differences in serum glucose or triglycerides levels were seen between the two genotypes. These data suggested that IGF-IR signaling in skeletal muscle mainly contributes to muscle development and not to metabolic regulation.
Although there were some phenotypic differences depending on the Cre recombinase that was used, the absence of IR-induced peripheral adiposity and metabolic disorders, including glucose intolerance, was assumed to have been caused by the reduction in insulin-stimulated glucose uptake into the skeletal muscle. Laustsen et al. [24] and Brüning et al. [29] generated muscle-specific IR knockout mice by crossing IR-floxed mice with MCK-Cre mice. These mice were comparable for body weight, whereas had a decreased ratio of heart weight to body weight and tended to have impaired cardiac performance. They also had an increased fat mass and elevated serum triglyceride and free fatty acids levels, whereas their glucose tolerance was normal. O’Neill et al. [25] also generated striated muscle-specific IR knockout mice by crossing ACTA1-Cre mice [26] with IR-floxed mice. There were no differences in body weight and skeletal muscle mass up to 52 weeks of age. A high-fat diet did not induce glucose intolerance in these mice. Additionally, mice overexpressing IR specifically in skeletal muscles had an increased lean mass, were protected from high-fat-diet-induced obesity, and had an improved glucose tolerance, all of which were associated with the reduction of inflammation in fat and liver tissue [30]. On the other hand, insulin-stimulated Akt phosphorylation was reduced in the muscles of those mice. It was suggested that the chronic overstimulation of IR led to post-receptor insulin resistance in muscle, indicating that a balance in the activity of IR signaling is critical.
Furthermore, Laustsen et al. [24] created the muscle-specific IR and IGF-IR double knockout mice using MCK promoter in transgenic mice, which had a reduced heart weight and developed heart failure, whereas had normal blood glucose and insulin levels. O’Neill et al. [25] also created the muscle-specific IR and IGF-IR double knockout mice using ACTA1 promoter in transgenic mice, which exhibited severe muscle atrophy but normal glucose tolerance.
In summary, IR/IGF-IR signaling in muscle plays an important role in muscle growth and glucose metabolism. In particular, IGF-IR mainly regulates muscle development.

2.3. IGF-I and Insulin Signaling in Adipose Tissue

Insulin and IGF-I action through their respective receptors are essential for the development and function of brown adipose tissue (BAT) and white adipose tissue (WAT) [31,32]. To investigate the direct impact of IGF-I on adipocytes, Kloting et al. crossed aP2-Cre mice with IGF-IR floxed mice [33]. Although these mice had an increased fat mass and enlarged adipocyte size, the recombination of the floxed alleles was not specific to the adipose tissue as it was also observed in the peripheral nervous system of aP2 promoter mice. Subsequently, adipose-tissue-specific IGF-IR knockout mice were established by crossbreeding adiponectin-Cre mice with IGF-IR floxed mice [34]. These adipocyte-specific IGF-IR knockout mice (F-IGFRKO) showed a reduction in fat pad mass (~25%) both in subcutaneous and brown adipose tissue, decreased lipogenic gene expression, and lower plasma leptin and adiponectin levels. In contrast, glucose tolerance and insulin sensitivity were not influenced by the adipocyte-specific IGF-IR deletion.
Mice with adipocyte-specific knockout for both insulin and IGF-I receptors (FIGIRKO) had markedly decreased white and brown fat mass [32]. FIGIRKO mice were unable to maintain their body temperature when placed in an environment at 4 °C. However, basal energy expenditure was increased in the FIGIRKO mice, despite a more than 85% reduction in brown fat mass. These FIGIRKO mice were also protected against high-fat-diet-induced obesity and age- and high-fat-diet-induced glucose intolerance. The brown fat activity was markedly decreased in the FIGIRKO mice but was responsive to β3-receptor stimulation. Thus, insulin/IGF-I signaling has a crucial role in the control of brown and white adipose tissue development, and a deficiency of both signaling pathways leads to abnormal thermogenesis and a paradoxical increase in the basal metabolic rate. In contrast, mice with a fat-specific disruption of the insulin receptor gene (FIRKO mice) had a 95% reduction in white adipose tissue but a paradoxical 50% increase in brown adipose tissue with the accumulation of large unilobular lipid droplets [34]. Consistent with a reduction in fat mass, the plasma leptin levels were decreased in FIRKO mice. In addition, FIRKO mice were protected against age-related and hypothalamic lesion-induced obesity and obesity-related glucose intolerance. FIRKO mice also exhibited a polarization of adipocytes into populations of large and small cells, which differed in the expression of fatty acid synthase, C/EBP alpha, and SREBP-1. Both FIGIRKO and FIRKO mice were unable to maintain a body temperature under cold exposure, and they exhibited ectopic lipid accumulation in the liver and muscle as well as pancreatic islet hyperplasia. FIRKO and FIGIRKO mice also developed overt diabetes, severe insulin resistance, and significant dyslipidemia.
Thus, insulin signaling in adipocytes is critical for the development of obesity and its associated metabolic abnormalities, and the abrogation of insulin signaling in fat unmasks a heterogeneity in the adipocyte response in terms of gene expression and triglyceride storage. Taken together, insulin and IGF-I action through IR and IGF-IR coordinately play a crucial role in adipose tissue development, its function, energy metabolism, and glucose metabolism.

2.4. IGF-I and Insulin Signaling in Pancreatic β-Cells

IGF-IR/IR signaling in β-cells is essential for their adaptive proliferation and the maintenance of their function as insulin secretory cells. Impaired IGF-IR/IR signaling in β-cells contributes to a decrease in insulin secretion and a mass reduction in β-cells.
Kulkarni et al. [35] generated β-cell-specific IGF-IR knockout (βKO) mice by crossing IGF-IR-floxed mice with β-cell-specific rat-insulin-2 promoter (RIP)-Cre mice. The βIGF-IRKO mice showed normal growth and β-cell development and did not have an altered β-cell mass, but they had a reduced expression of GLUT2 and glucokinase in β-cells, resulting in defective glucose-stimulated insulin secretion and impaired glucose tolerance. Xuan et al. [36] also generated βIGF-IRKO mice using different strains of RIP-Cre mice; these mice exhibited defective glucose- and arginine-dependent insulin secretion and impaired glucose tolerance. However, no difference in β-cell mass or proliferation was seen. Thus, IGF-I in β-cells is crucial for insulin secretion but not for the regulation of β-cell proliferation.
IGF-II, as well as IGF-I and insulin, is a potent ligand of IGF-IR. IGF-I is expressed at a low level or is not detectable, whereas IGF-II is highly expressed in the β-cells of adult mice [37] and humans [38]. Honey Modi et al. [39] generated β-cell-specific IGF-II knockout (βIGFIIKO) mice by crossing IGF-II-floxed mice with Ins-1-Cre mice. Inoel Sandovici et al. [40] also generated βIGFIIKO mice by crossing IGF-II-floxed mice with RIP-Cre mice. They suggested that the autocrine actions of IGF-II impacted β-cell plasticity under conditions of increased metabolic demand, especially in females. Several reports have indicated that the overexpression of IGF-II increases β-cell expansion and replication in rodents [41,42,43]. In addition, Marion Cornu et al. [37] revealed that GLP-1 stimulated IGF-IR expression and IGF-IR signaling via autocrine IGF-II in β-cells, leading to the protection of β-cells from apoptosis. On the other hand, Yarong Lu et al. [44] generated mice with pancreatic-specific IGF-I gene deficiency (PID) by crossing Pdx1-Cre mice with IGF-I-floxed mice. The PID mice had an enlarged islet cell mass and were resistant to streptozotocin-induced diabetes. Thus, locally produced IGF-I within the pancreas might inhibit islet cell growth. Furthermore, IGFBPs (insulin-like growth factor binding proteins), which comprise a family of six proteins (1–6), bind to IGF-I and IGF-II in extracellular fluid. Jing Lu et al. [45] revealed that IGFBP-1 increased β-cell regeneration, at least in part, by inhibiting IGF-IR signaling. In addition, Nuria Palau et al. [46] revealed that the blockage of IGFBP-3 led to an increase in β-cell proliferation during obesity, probably through the suppression of IGF-IR signaling.
Kulkarni et al. [47] generated β-cell-specific IR knockout (βIRKO) mice by crossing IR-floxed mice with RIP-Cre mice. The βIRKO mice were born normally and showed no difference in islet morphology, but they exhibited impaired glucose tolerance, with decreased glucose-stimulated insulin secretion and a progressive reduction in the β-cell mass. Okada et al. [48] revealed that βIRKO mice failed to exhibit compensatory islet proliferation and hyperplasia in response to high-fat-diet feeding. Otani et al. [49] reported that a modest reduction in the glucose-sensing associated gene, GLUT2, and glucokinase gene expression in the islets of βIRKO mice, resulting in defective glucose-stimulated insulin secretion and impaired glucose tolerance in βIRKO mice. In addition, Shirakawa et al. [50] revealed that IR signaling promoted adaptive β-cell proliferation in response to pregnancy, acute/chronic insulin resistance, and aging through the FoxM1/PLK1 (polo-like kinase-1)/CENP-A (centromere protein A) pathway-mediated mitotic cell-cycle progression.
In addition, Ueki et al. [51] created the β-cell-specific IR and IGF-IR double knockout (βDKO) mice by breeding the RIP-Cre transgenic βIRKO mice and βIGF-IRKO mice. The βDKO mice were born normally, and no differences in β-cell mass or islet architectures were seen between the βDKO mice and their controls at birth. However, the βDKO mice had a reduced β-cell mass at 2 weeks of age and developed overt diabetes at 3 weeks of age, representing a worse diabetic phenotype than that seen in single knockout models of IR or IGF-IR in β-cells. βIR−/−βIGF-IR+/− mice exhibited severe hyperglycemia with a reduced β-cell mass at 2 weeks of age, whereas βIR+/−βIGF-IR−/− mice exhibited mild hyperglycemia with no change in the β-cell mass. These results suggested that IR participated in the maintenance of β-cell mass rather than IGF-IR.
Recently, Ansarullah et al. [52] reported an insulin inhibitory receptor, inceptor. Inceptor calibrates IR and IGF-IR internalization via clathrin-mediated endocytosis, thereby regulating the ligand sensitization of IR and IGF-IR in β-cells. β-cell-specific inceptor knockout mice exhibited an increase in IR and IGF-IR signaling, enhanced β-cell proliferation, and improved glucose tolerance. Furthermore, treatment with an anti-inceptor neutralizing monoclonal antibody resulted in the sustained activation of IR and IGF-IR in β-cells. The expression of IR-mediated signaling molecules is known to be reduced in islets from type 2 diabetes patients [53]. Thus, treatments targeted at IR/IGF-IR signaling via inceptor might contribute to the development of radical treatments to recover β-cell function in type 2 diabetes patients.
In addition, whether IR/IGF-IR itself is needed in β-cells remains controversial. Several reports of mice deficient for molecules involved in the IR/IGF-IR-mediated signaling pathway in β-cells have been made. Kubota et al. [54] generated β-cell-specific IRS-2 knockout mice by crossing IRS-2-floxed mice with RIP-Cre mice, which exhibited impaired glucose-induced insulin secretion and reduced β-cell proliferation. Hashimoto et al. [55] generated β-cell-specific PDK-1 knockout mice by crossing PDK-1-floxed mice with RIP-Cre mice, which reduced the β-cell mass and resulted in hyperglycemia with a reduction in plasma insulin levels. Thus, the downstream signaling of IR/IGF-IR is obviously critical for β-cell maintenance, whereas it is unclear whether the interaction of ligands with IR/IGF-IR is required for β-cell function. Since IR/IGF-IR-mediated signaling molecules, such as IRS proteins, form complexes by binding to IR/IGF-IR, those receptors might act as anchors for signaling complexes in a manner that is independent of their ligands.
In summary, IGF-IR signaling in β-cells contributed to β-cell functions and β-cell proliferation. IR and IGF-IR signaling compensate for each other’s actions in β-cells.

2.5. IGF-I and Insulin Signaling in the Brain

IR and IGF-IR are also expressed in the central nervous system. Those two receptors have different physiological functions in the brain, as shown by genetic knockout in mice. Mice with the genetic deletion of IGF-IR in the brain had a reduced brain size, generalized growth retardation, and behavioral changes. In contrast, mice with the genetic deletion of IR in the brain (neuron-specific IR knockout; NIRKO) had a normal brain size and development, but they exhibited metabolic phenotypes, including mild obesity and insulin resistance [56]. In fact, female NIRKO mice showed an increased food intake, and both male and female mice developed diet-sensitive obesity with increases in body fat and plasma leptin levels, mild insulin resistance, elevated plasma insulin levels, and hypertriglyceridemia. NIRKO mice also exhibited impaired spermatogenesis and ovarian follicle maturation because of the hypothalamic dysregulation of the luteinizing hormone. Thus, IR signaling in the CNS plays an important role in the regulation of energy disposal, fuel metabolism, and reproduction. Interestingly, the heterozygous deletion of IGF-IR in the brain did not affect the brain growth, but it resulted in a complex phenotype that included reduced body weight, increased fat mass, impaired glucose tolerance, and an extended lifespan [57].

3. Metabolic Phenotype in OSI-906 Treated Mice

OSI-906 (linsitinib), an anti-tumor drug, specifically inhibits the autophosphorylation of both IR and IGF-IR, resulting in the induction of insulin resistance [58]. OSI-906 has been shown to block the ligand-induced activation of downstream pathways, including the phosphorylation of Akt, Erk1/2, and p70S6K, in the 3T3/huIGF-IR cell line (LISN) [13]. The inhibition of IR/IGF-IR signaling by OSI-906 in a two-stage BALB/c-3T3 cell transformation assay (BALB-CTA) diminished the phosphorylation of Akt, p70S6K, S6 protein, 4EBP-1, GSK3β, and AMPK, but not the phosphorylation of Erk1/2 [14]. The daily administration of OSI-906 to mice for a week induced increased β-cell mass and β-cell proliferation, accompanied by hyperinsulinemia, glucose intolerance, lipodystrophy, and liver steatosis [13,14] (Figure 2).
We previously reported that the systemic inhibition of IR and IGF-IR with OSI-906 induced acute insulin resistance in mice [13]. The oral administration of OSI-906 evoked glucose intolerance in a dose-dependent manner in wild-type mice. The serum glucose and insulin levels were immediately elevated an hour after OSI-906 administration, and continuous hyperglycemia and hyperinsulinemia were observed during the treatment with OSI-906 until day 7. The mice treated with OSI-906 for a week showed a significantly increased β-cell mass and β-cell proliferation. These results suggested that β-cell proliferation was mediated by an insulin signaling-independent pathway, as no change in the cyclin D2 expression was observed in the mouse islets [14]. Notably, the withdrawal of OSI-906 after a week of administration reversed the hyperglycemia and hyperinsulinemia, as well as the β-cell proliferation, although the increased β-cell mass was restored [14].
Mice treated with OSI-906 for 7 days showed lipoatrophy in BAT and WAT with elevated serum free fatty acid levels (Figure 2). The expression of leptin, but not adiponectin, in fat was significantly lower in the OSI-906-treated mice, compared with that in control mice [14]. Conditional IGF-IR inactivation was previously shown to result in an increased WAT mass and did not affect the BAT mass or its activity [33]. These findings imply differences in the roles of IGF-IR and IR signaling on the regulation of adipose tissue mass.
Insulin signaling in the liver was impaired in mice treated with OSI-906, resulting in elevated gluconeogenic gene expressions. Mice treated with OSI-906 for 7 days showed hepatic steatosis (Figure 2). The liver weight, as well as both liver triglyceride and liver glycogen contents, were significantly increased by the administration of OSI-906. On the other hand, the gene expressions involved in lipogenesis, such as Srebp1c, Fas, and Scd1, were not altered by treatment with OSI-906 in the liver [14,16]. Notably, leptin supplementation on OSI-906-treated mice reduced hepatic triglyceride accumulation and improved steatosis without improving lipodystrophy. Interestingly, liver steatosis was ameliorated after the withdrawal of OSI-906, which coincided with the recovery from lipodystrophy [14] (Figure 2). These findings indicate the inter-organ network between the liver and adipose tissues in the onset or remission of IR and IGF-IR signaling-mediated hepatic steatosis.

4. Perspective

In this review, we focused on the phenotypes of conditional knockout models of the IR/IGF-IR and outlined the roles of the IGF-I axis on glucose metabolism in metabolic organs. A defect in IGF-IR signaling contributes to the development of diabetes. Thus, anti-diabetic drugs that target the IGF-I axis have the potential to be useful for human diabetes patients. On the other hand, the activation of IGF-IR signaling is generally considered to promote carcinogenesis. Diabetes patients are at a high risk of cancer because of the increase in IGF-IR activities arising from compensatory hyperinsulinemia for insulin resistance. Metformin, one of the most broadly used agents for the treatment of type 2 diabetes patients, has been shown to have anti-cancer effects because of its suppression of IGF-I/IGF-IR signaling activities by multiple mechanisms, such as the activation of AMP-activated protein kinase (AMPK) and the associated inhibition of mTOR (mechanistic target of rapamycin) signaling [59]. Metformin has been reported to have a tumor suppressor effect in many cancer types in large-scale clinical studies involving many human type 2 diabetes patients, in addition to analyses using cell lines or animal models. The development of drugs that have both anti-cancer and anti-diabetic effects to modulate IGF-I/IGF-IR signaling activity appropriately is expected in the future.
The elucidation of the role of the IGF-I axis in metabolism will also contribute to the development of the field of aging. IGF-IR heterozygous knockout mice live, on average, 26% longer than their wild-type littermates and display greater resistance to oxidative stress, a known determinant of aging [60]. In addition, the inhibition of the TOR (target of rapamycin) signaling pathway, which is a nutrient response signaling pathway, extends the lifespan of Caenorhabditis elegans [61] and Drosophila [62]. Caloric restriction is known to extend the lifespan in animal models, which might be involved in IGF-IR signaling through the suppression of mTOR signaling. Furthermore, the increase in systemic low-grade inflammation in the elderly is referred to as “inflammaging,” which contributes to the development of chronic diseases, including diabetes, and accelerates aging. GH receptor-deficient mice reduced the circulating levels of IGF-I, resulting in a prolonged lifespan, and protected them from age-related inflammasome activation and associated T cell senescence [63]. Additionally, in humans, the presence of heterozygous mutations in the IGF-IR gene [64] or GH1 (human growth hormone) gene SNPs (single nucleotide polymorphisms) [65] reportedly affect longevity, indicating an involvement between IGF-IR signaling and lifespan.
In this way, the IGF-I axis is a principal mechanism that links diabetes, cancer, and aging. If a drug that modulates the IGF-I axis is developed in the future, it could become a silver bullet that solves these clinically important issues simultaneously. Further elucidation of the roles of the IGF-I axis from various perspectives is expected.

Author Contributions

Conceptualization, J.S.; writing—review and editing, T.O., M.K., Y.T. and J.S.; visualization, T.O., M.K. and J.S.; supervision, Y.T. and J.S. funding acquisition, T.O., M.K., Y.T. and J.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported in part by Grants-in-Aid for Scientific Research (B) 20H03733 from MEXT of Japan (to Y.T.); Grants-in-Aid for Scientific Research (C) 20K08866 from MEXT of Japan, from AMED-A*STAR SCICORP for research on cell therapy, from the Kamome Memorial Foundation of Yokohama City University, and from the Japan IDDM network (to J.S.); a Grant-in-Aid for Young Scientists (B) 18K16241 from MEXT of Japan, a Junior Scientist Development Grant provided by Novo Nordisk Pharma, the MSD Life Science Foundation, and the Japan Foundation for Applied Enzymology (FFDR; to T.O.); and a Grant-in-Aid for Young Scientists (B) 20K17516 from MEXT of Japan, a Junior Scientist Development Grant provided by Novo Nordisk Pharma, and the Japan Foundation for Applied Enzymology (FFDR; to M.K.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sjögren, K.; Liu, J.L.; Blad, K.; Skrtic, S.; Vidal, O.; Wallenius, V.; LeRoith, D.; Törnell, J.; Isaksson, O.G.; Jansson, J.O.; et al. Liver-derived insulin-like growth factor I (IGF-I) is the principal source of IGF-I in blood but is not required for postnatal body growth in mice. Proc. Natl. Acad. Sci. USA 1999, 96, 7088–7092. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Yakar, S.; Liu, J.L.; Stannard, B.; Butler, A.; Accili, D.; Sauer, B.; LeRoith, D. Normal growth and development in the absence of hepatic insulin-like growth factor I. Proc. Natl. Acad. Sci. USA 1999, 96, 7324–7329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Yakar, S.; Pennisi, P.; Kim, C.H.; Zhao, H.; Toyoshima, Y.; Gavrilova, O.; LeRoith, D. Studies involving the GH-IGF axis: Lessons from IGF-I and IGF-I receptor gene targeting mouse models. J. Endocrinol. Investig. 2005, 28 (Suppl. 5), 19–22. [Google Scholar]
  4. McRory, J.E.; Sherwood, N.M. Ancient divergence of insulin and insulin-like growth factor. DNA Cell Biol. 1997, 16, 939–949. [Google Scholar] [CrossRef] [PubMed]
  5. Pandini, G.; Frasca, F.; Mineo, R.; Sciacca, L.; Vigneri, R.; Belfiore, A. Insulin/insulin-like growth factor I hybrid receptors have different biological characteristics depending on the insulin receptor isoform involved. J. Biol. Chem. 2002, 277, 39684–39695. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Hartog, H.; Wesseling, J.; Boezen, H.M.; van der Graaf, W.T. The insulin-like growth factor 1 receptor in cancer: Old focus, new future. Eur. J. Cancer 2007, 43, 1895–1904. [Google Scholar] [CrossRef] [PubMed]
  7. Vigneri, R.; Goldfine, I.D.; Frittitta, L. Insulin, insulin receptors, and cancer. J. Endocrinol. Investig. 2016, 39, 1365–1376. [Google Scholar] [CrossRef] [PubMed]
  8. De Meyts, P.; Whittaker, J. Structural biology of insulin and IGF1 receptors: Implications for drug design. Nat. Rev. Drug Discov. 2002, 1, 769–783. [Google Scholar] [CrossRef]
  9. Moller, D.E.; Yokota, A.; Caro, J.F.; Flier, J.S. Tissue-specific expression of two alternatively spliced insulin receptor mRNAs in man. Mol. Endocrinol. 1989, 3, 1263–1269. [Google Scholar] [CrossRef] [Green Version]
  10. Belfiore, A.; Frasca, F.; Pandini, G.; Sciacca, L.; Vigneri, R. Insulin receptor isoforms and insulin receptor/insulin-like growth factor receptor hybrids in physiology and disease. Endocr. Rev. 2009, 30, 586–623. [Google Scholar] [CrossRef] [Green Version]
  11. Belfiore, A.; Malaguarnera, R.; Vella, V.; Lawrence, M.C.; Sciacca, L.; Frasca, F.; Morrione, A.; Vigneri, R. Insulin Receptor Isoforms in Physiology and Disease: An Updated View. Endocr. Rev. 2017, 38, 379–431. [Google Scholar] [CrossRef]
  12. Pullen, R.A.; Lindsay, D.G.; Wood, S.P.; Tickle, I.J.; Blundell, T.L.; Wollmer, A.; Krail, G.; Brandenburg, D.; Zahn, H.; Gliemann, J.; et al. Receptor-binding region of insulin. Nature 1976, 259, 369–373. [Google Scholar] [CrossRef]
  13. Shirakawa, J.; Okuyama, T.; Yoshida, E.; Shimizu, M.; Horigome, Y.; Tuno, T.; Hayasaka, M.; Abe, S.; Fuse, M.; Togashi, Y.; et al. Effects of the antitumor drug OSI-906, a dual inhibitor of IGF-1 receptor and insulin receptor, on the glycemic control, β-cell functions, and β-cell proliferation in male mice. Endocrinology 2014, 155, 2102–2111. [Google Scholar] [CrossRef] [Green Version]
  14. Tajima, K.; Shirakawa, J.; Togashi, Y.; Yamazaki, S.; Okuyama, T.; Kyohara, M.; Konishi, H.; Terauchi, Y. Metabolic recovery of lipodystrophy, liver steatosis, and pancreatic β cell proliferation after the withdrawal of OSI-906. Sci. Rep. 2017, 7, 4119. [Google Scholar]
  15. Shirakawa, J.; Tajima, K.; Okuyama, T.; Kyohara, M.; Togashi, Y.; De Jesus, D.F.; Basile, G.; Kin, T.; Shapiro, A.M.J.; Kulkarni, R.N.; et al. Luseogliflozin increases beta cell proliferation through humoral factors that activate an insulin receptor- and IGF-1 receptor-independent pathway. Diabetologia 2020, 63, 577–587. [Google Scholar] [CrossRef]
  16. Okuyama, T.; Shirakawa, J.; Tajima, K.; Ino, Y.; Vethe, H.; Togashi, Y.; Kyohara, M.; Inoue, R.; Miyashita, D.; Li, J.; et al. Linagliptin Ameliorates Hepatic Steatosis via Non-Canonical Mechanisms in Mice Treated with a Dual Inhibitor of Insulin Receptor and IGF-1 Receptor. Int. J. Mol. Sci. 2020, 21, 7815. [Google Scholar] [CrossRef]
  17. Satake, S.; Moore, M.C.; Igawa, K.; Converse, M.; Farmer, B.; Neal, D.W.; Cherrington, A.D. Direct and indirect effects of insulin on glucose uptake and storage by the liver. Diabetes 2002, 51, 1663–1671. [Google Scholar] [CrossRef] [Green Version]
  18. Yakar, S.; Liu, J.L.; Fernandez, A.M.; Wu, Y.; Schally, A.V.; Frystyk, J.; Chernausek, S.D.; Mejia, W.; Le Roith, D. Liver-specific igf-1 gene deletion leads to muscle insulin insensitivity. Diabetes 2001, 50, 1110–1118. [Google Scholar] [CrossRef] [Green Version]
  19. Michael, M.D.; Kulkarni, R.N.; Postic, C.; Previs, S.F.; Shulman, G.I.; Magnuson, M.A.; Kahn, C.R. Loss of insulin signaling in hepatocytes leads to severe insulin resistance and progressive hepatic dysfunction. Mol. Cell 2000, 6, 87–97. [Google Scholar] [CrossRef]
  20. Meek, S.E.; Persson, M.; Ford, G.C.; Nair, K.S. Differential regulation of amino acid exchange and protein dynamics across splanchnic and skeletal muscle beds by insulin in healthy human subjects. Diabetes 1998, 47, 1824–1835. [Google Scholar] [CrossRef]
  21. Chow, L.S.; Albright, R.C.; Bigelow, M.L.; Toffolo, G.; Cobelli, C.; Nair, K.S. Mechanism of insulin’s anabolic effect on muscle: Measurements of muscle protein synthesis and breakdown using aminoacyl-tRNA and other surrogate measures. Am. J. Physiol. Endocrinol. Metab. 2006, 291, E729–E736. [Google Scholar] [CrossRef] [PubMed]
  22. Yoshida, T.; Delafontaine, P. Mechanisms of IGF-1-Mediated Regulation of Skeletal Muscle Hypertrophy and Atrophy. Cells 2020, 9, 1970. [Google Scholar] [CrossRef] [PubMed]
  23. O’Neill, B.T.; Lee, K.Y.; Klaus, K.; Softic, S.; Krumpoch, M.T.; Fentz, J.; Stanford, K.I.; Robinson, M.M.; Cai, W.; Kleinridders, A.; et al. Insulin and IGF-1 receptors regulate FoxO-mediated signaling in muscle proteostasis. J. Clin. Investig. 2016, 126, 3433–3446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Laustsen, P.G.; Russell, S.J.; Cui, L.; Entingh-Pearsall, A.; Holzenberger, M.; Liao, R.; Kahn, C.R. Essential role of insulin and insulin-like growth factor 1 receptor signaling in cardiac development and function. Mol. Cell. Biol. 2007, 27, 1649–1664. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. O’Neill, B.T.; Lauritzen, H.P.; Hirshman, M.F.; Smyth, G.; Goodyear, L.J.; Kahn, C.R. Differential Role of Insulin/IGF-1 Receptor Signaling in Muscle Growth and Glucose Homeostasis. Cell Rep. 2015, 11, 1220–1235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Miniou, P.; Tiziano, D.; Frugier, T.; Roblot, N.; Le Meur, M.; Melki, J. Gene targeting restricted to mouse striated muscle lineage. Nucleic Acids Res. 1999, 27, e27–e30. [Google Scholar] [CrossRef] [Green Version]
  27. Mavalli, M.D.; DiGirolamo, D.J.; Fan, Y.; Riddle, R.C.; Campbell, K.S.; van Groen, T.; Frank, S.J.; Sperling, M.A.; Esser, K.A.; Bamman, M.M.; et al. Distinct growth hormone receptor signaling modes regulate skeletal muscle development and insulin sensitivity in mice. J. Clin. Investig. 2010, 120, 4007–4020. [Google Scholar] [CrossRef] [Green Version]
  28. Anderson, C.M.; Hu, J.; Barnes, R.M.; Heidt, A.B.; Cornelissen, I.; Black, B.L. Myocyte enhancer factor 2C function in skeletal muscle is required for normal growth and glucose metabolism in mice. Skelet. Muscle 2015, 5, 7. [Google Scholar] [CrossRef] [Green Version]
  29. Brüning, J.C.; Michael, M.D.; Winnay, J.N.; Hayashi, T.; Hörsch, D.; Accili, D.; Goodyear, L.J.; Kahn, C.R. A muscle-specific insulin receptor knockout exhibits features of the metabolic syndrome of NIDDM without altering glucose tolerance. Mol. Cell 1998, 2, 559–569. [Google Scholar] [CrossRef]
  30. Wang, G.; Yu, Y.; Cai, W.; Batista, T.M.; Suk, S.; Noh, H.L.; Hirshman, M.; Nigro, P.; Li, M.E.; Softic, S.; et al. Muscle-Specific Insulin Receptor Overexpression Protects Mice from Diet-Induced Glucose Intolerance but Leads to Postreceptor Insulin Resistance. Diabetes 2020, 69, 2294–2309. [Google Scholar] [CrossRef]
  31. Blüher, M.; Michael, M.D.; Peroni, O.D.; Ueki, K.; Carter, N.; Kahn, B.B.; Kahn, C.R. Adipose tissue selective insulin receptor knockout protects against obesity and obesity-related glucose intolerance. Dev. Cell 2002, 3, 25–38. [Google Scholar] [CrossRef] [Green Version]
  32. Boucher, J.; Mori, M.A.; Lee, K.Y.; Smyth, G.; Liew, C.W.; Macotela, Y.; Rourk, M.; Bluher, M.; Russell, S.J.; Kahn, C.R. Impaired thermogenesis and adipose tissue development in mice with fat-specific disruption of insulin and IGF-1 signalling. Nat. Commun. 2012, 3, 902. [Google Scholar] [CrossRef]
  33. Klöting, N.; Koch, L.; Wunderlich, T.; Kern, M.; Ruschke, K.; Krone, W.; Brüning, J.C.; Blüher, M. Autocrine IGF-1 action in adipocytes controls systemic IGF-1 concentrations and growth. Diabetes 2008, 57, 2074–2082. [Google Scholar] [CrossRef] [Green Version]
  34. Boucher, J.; Softic, S.; El Ouaamari, A.; Krumpoch, M.T.; Kleinridders, A.; Kulkarni, R.N.; O’Neill, B.T.; Kahn, C.R. Differential Roles of Insulin and IGF-1 Receptors in Adipose Tissue Development and Function. Diabetes 2016, 65, 2201–2213. [Google Scholar] [CrossRef] [Green Version]
  35. Kulkarni, R.N.; Holzenberger, M.; Shih, D.Q.; Ozcan, U.; Stoffel, M.; Magnuson, M.A.; Kahn, C.R. Beta-cell-specific deletion of the Igf1 receptor leads to hyperinsulinemia and glucose intolerance but does not alter beta-cell mass. Nat. Genet. 2002, 31, 111–115. [Google Scholar] [CrossRef]
  36. Xuan, S.; Kitamura, T.; Nakae, J.; Politi, K.; Kido, Y.; Fisher, P.E.; Morroni, M.; Cinti, S.; White, M.F.; Herrera, P.L.; et al. Defective insulin secretion in pancreatic beta cells lacking type 1 IGF receptor. J. Clin. Investig. 2002, 110, 1011–1019. [Google Scholar] [CrossRef]
  37. Cornu, M.; Yang, J.Y.; Jaccard, E.; Poussin, C.; Widmann, C.; Thorens, B. Glucagon-like peptide-1 protects beta-cells against apoptosis by increasing the activity of an IGF-2/IGF-1 receptor autocrine loop. Diabetes 2009, 58, 1816–1825. [Google Scholar] [CrossRef] [Green Version]
  38. Nica, A.C.; Ongen, H.; Irminger, J.C.; Bosco, D.; Berney, T.; Antonarakis, S.E.; Halban, P.A.; Dermitzakis, E.T. Cell-type, allelic, and genetic signatures in the human pancreatic beta cell transcriptome. Genome Res. 2013, 23, 1554–1562. [Google Scholar] [CrossRef] [Green Version]
  39. Modi, H.; Jacovetti, C.; Tarussio, D.; Metref, S.; Madsen, O.D.; Zhang, F.P.; Rantakari, P.; Poutanen, M.; Nef, S.; Gorman, T.; et al. Autocrine Action of IGF2 Regulates Adult β-Cell Mass and Function. Diabetes 2015, 64, 4148–4157. [Google Scholar] [CrossRef] [Green Version]
  40. Sandovici, I.; Hammerle, C.M.; Virtue, S.; Vivas-Garcia, Y.; Izquierdo-Lahuerta, A.; Ozanne, S.E.; Vidal-Puig, A.; Medina-Gómez, G.; Constância, M. Autocrine IGF2 programmes β-cell plasticity under conditions of increased metabolic demand. Sci. Rep. 2021, 11, 7717. [Google Scholar] [CrossRef]
  41. Estil les, E.; Téllez, N.; Escoriza, J.; Montanya, E. Increased β-cell replication and β-cell mass regeneration in syngeneically transplanted rat islets overexpressing insulin-like growth factor II. Cell Transplant. 2012, 21, 2119–2129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Petrik, J.; Pell, J.M.; Arany, E.; McDonald, T.J.; Dean, W.L.; Reik, W.; Hill, D.J. Overexpression of insulin-like growth factor-II in transgenic mice is associated with pancreatic islet cell hyperplasia. Endocrinology 1999, 140, 2353–2363. [Google Scholar] [CrossRef] [PubMed]
  43. Devedjian, J.C.; George, M.; Casellas, A.; Pujol, A.; Visa, J.; Pelegrín, M.; Gros, L.; Bosch, F. Transgenic mice overexpressing insulin-like growth factor-II in beta cells develop type 2 diabetes. J. Clin. Investig. 2000, 105, 731–740. [Google Scholar] [CrossRef] [Green Version]
  44. Lu, Y.; Herrera, P.L.; Guo, Y.; Sun, D.; Tang, Z.; LeRoith, D.; Liu, J.L. Pancreatic-specific inactivation of IGF-I gene causes enlarged pancreatic islets and significant resistance to diabetes. Diabetes 2004, 53, 3131–3141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Lu, J.; Liu, K.C.; Schulz, N.; Karampelias, C.; Charbord, J.; Hilding, A.; Rautio, L.; Bertolino, P.; Östenson, C.G.; Brismar, K.; et al. IGFBP1 increases β-cell regeneration by promoting α- to β-cell transdifferentiation. EMBO J. 2016, 35, 2026–2044. [Google Scholar] [CrossRef] [PubMed]
  46. Palau, N.; Rebuffat, S.A.; Altirriba, J.; Piquer, S.; Hanzu, F.A.; Gomis, R.; Barbera, A. Role of IGFBP-3 in the regulation of β-cell mass during obesity: Adipose tissue/β-cell cross talk. Endocrinology 2012, 153, 177–187. [Google Scholar] [CrossRef] [PubMed]
  47. Kulkarni, R.N.; Brüning, J.C.; Winnay, J.N.; Postic, C.; Magnuson, M.A.; Kahn, C.R. Tissue-specific knockout of the insulin receptor in pancreatic beta cells creates an insulin secretory defect similar to that in type 2 diabetes. Cell 1999, 96, 329–339. [Google Scholar] [CrossRef] [Green Version]
  48. Okada, T.; Liew, C.W.; Hu, J.; Hinault, C.; Michael, M.D.; Krtzfeldt, J.; Yin, C.; Holzenberger, M.; Stoffel, M.; Kulkarni, R.N. Insulin receptors in beta-cells are critical for islet compensatory growth response to insulin resistance. Proc. Natl. Acad. Sci. USA 2007, 104, 8977–8982. [Google Scholar] [CrossRef] [Green Version]
  49. Otani, K.; Kulkarni, R.N.; Baldwin, A.C.; Krutzfeldt, J.; Ueki, K.; Stoffel, M.; Kahn, C.R.; Polonsky, K.S. Reduced beta-cell mass and altered glucose sensing impair insulin-secretory function in betaIRKO mice. Am. J. Physiol. Endocrinol. Metab. 2004, 286, E41–E49. [Google Scholar] [CrossRef]
  50. Shirakawa, J.; Fernandez, M.; Takatani, T.; El Ouaamari, A.; Jungtrakoon, P.; Okawa, E.R.; Zhang, W.; Yi, P.; Doria, A.; Kulkarni, R.N. Insulin Signaling Regulates the FoxM1/PLK1/CENP-A Pathway to Promote Adaptive Pancreatic β Cell Proliferation. Cell Metab. 2017, 25, 868–882.e5. [Google Scholar] [CrossRef] [Green Version]
  51. Ueki, K.; Okada, T.; Hu, J.; Liew, C.W.; Assmann, A.; Dahlgren, G.M.; Peters, J.L.; Shackman, J.G.; Zhang, M.; Artner, I.; et al. Total insulin and IGF-I resistance in pancreatic beta cells causes overt diabetes. Nat. Genet. 2006, 38, 583–588. [Google Scholar] [CrossRef]
  52. Ansarullah; Jain, C.; Far, F.F.; Homberg, S.; Wißmiller, K.; von Hahn, F.G.; Raducanu, A.; Schirge, S.; Sterr, M.; Bilekova, S.; et al. Inceptor counteracts insulin signalling in β-cells to control glycaemia. Nature 2021, 590, 326–331. [Google Scholar] [CrossRef]
  53. Folli, F.; Okada, T.; Perego, C.; Gunton, J.; Liew, C.W.; Akiyama, M.; D’Amico, A.; La Rosa, S.; Placidi, C.; Lupi, R.; et al. Altered insulin receptor signalling and β-cell cycle dynamics in type 2 diabetes mellitus. PLoS ONE 2011, 6, e28050. [Google Scholar] [CrossRef] [Green Version]
  54. Kubota, N.; Terauchi, Y.; Tobe, K.; Yano, W.; Suzuki, R.; Ueki, K.; Takamoto, I.; Satoh, H.; Maki, T.; Kubota, T.; et al. Insulin receptor substrate 2 plays a crucial role in beta cells and the hypothalamus. J. Clin. Investig. 2004, 114, 917–927. [Google Scholar] [CrossRef] [Green Version]
  55. Hashimoto, N.; Kido, Y.; Uchida, T.; Asahara, S.; Shigeyama, Y.; Matsuda, T.; Takeda, A.; Tsuchihashi, D.; Nishizawa, A.; Ogawa, W.; et al. Ablation of PDK1 in pancreatic beta cells induces diabetes as a result of loss of beta cell mass. Nat. Genet. 2006, 38, 589–593. [Google Scholar] [CrossRef]
  56. Brüning, J.C.; Gautam, D.; Burks, D.J.; Gillette, J.; Schubert, M.; Orban, P.C.; Klein, R.; Krone, W.; Müller-Wieland, D.; Kahn, C.R. Role of brain insulin receptor in control of body weight and reproduction. Science 2000, 289, 2122–2125. [Google Scholar] [CrossRef]
  57. Kappeler, L.; De Magalhaes Filho, C.; Dupont, J.; Leneuve, P.; Cervera, P.; Périn, L.; Loudes, C.; Blaise, A.; Klein, R.; Epelbaum, J.; et al. Brain IGF-1 receptors control mammalian growth and lifespan through a neuroendocrine mechanism. PLoS Biol. 2008, 6, e254. [Google Scholar] [CrossRef]
  58. Mulvihill, M.J.; Cooke, A.; Rosenfeld-Franklin, M.; Buck, E.; Foreman, K.; Landfair, D.; O’Connor, M.; Pirritt, C.; Sun, Y.; Yao, Y.; et al. Discovery of OSI-906: A selective and orally efficacious dual inhibitor of the IGF-1 receptor and insulin receptor. Future Med. Chem. 2009, 1, 1153–1171. [Google Scholar] [CrossRef]
  59. Cioce, M.; Pulito, C.; Strano, S.; Blandino, G.; Fazio, V.M. Metformin: Metabolic Rewiring Faces Tumor Heterogeneity. Cells 2020, 9, 2439. [Google Scholar] [CrossRef]
  60. Holzenberger, M.; Dupont, J.; Ducos, B.; Leneuve, P.; Géloën, A.; Even, P.C.; Cervera, P.; Le Bouc, Y. IGF-1 receptor regulates lifespan and resistance to oxidative stress in mice. Nature 2003, 421, 182–187. [Google Scholar] [CrossRef]
  61. Vellai, T.; Takacs-Vellai, K.; Zhang, Y.; Kovacs, A.L.; Orosz, L.; Müller, F. Genetics: Influence of TOR kinase on lifespan in C. elegans. Nature 2003, 426, 620. [Google Scholar] [CrossRef] [PubMed]
  62. Kapahi, P.; Zid, B.M.; Harper, T.; Koslover, D.; Sapin, V.; Benzer, S. Regulation of lifespan in Drosophila by modulation of genes in the TOR signaling pathway. Curr. Biol. 2004, 14, 885–890. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Spadaro, O.; Goldberg, E.L.; Camell, C.D.; Youm, Y.H.; Kopchick, J.J.; Nguyen, K.Y.; Bartke, A.; Sun, L.Y.; Dixit, V.D. Growth Hormone Receptor Deficiency Protects against Age-Related NLRP3 Inflammasome Activation and Immune Senescence. Cell Rep. 2016, 14, 1571–1580. [Google Scholar] [CrossRef] [Green Version]
  64. Suh, Y.; Atzmon, G.; Cho, M.O.; Hwang, D.; Liu, B.; Leahy, D.J.; Barzilai, N.; Cohen, P. Functionally significant insulin-like growth factor I receptor mutations in centenarians. Proc. Natl. Acad. Sci. USA 2008, 105, 3438–3442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Van Heemst, D.; Beekman, M.; Mooijaart, S.P.; Heijmans, B.T.; Brandt, B.W.; Zwaan, B.J.; Slagboom, P.E.; Westendorp, R.G. Reduced insulin/IGF-1 signalling and human longevity. Aging Cell 2005, 4, 79–85. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic illustration of IGF-I receptor/insulin receptor and downstream intracellular signaling. The IGF-I receptor (IGF-IR) and the insulin receptor (IR: IR-A and IR-B) possess a respective affinity for insulin, IGF-I, and IGF-II. IGF-IR and IR can interact by forming hybrid receptors or by cross-talking in downstream intracellular signaling pathways. The sizes of the arrows show the amplitudes of each signaling pathway within the multiplicity of ligand-receptor interactions.
Figure 1. Schematic illustration of IGF-I receptor/insulin receptor and downstream intracellular signaling. The IGF-I receptor (IGF-IR) and the insulin receptor (IR: IR-A and IR-B) possess a respective affinity for insulin, IGF-I, and IGF-II. IGF-IR and IR can interact by forming hybrid receptors or by cross-talking in downstream intracellular signaling pathways. The sizes of the arrows show the amplitudes of each signaling pathway within the multiplicity of ligand-receptor interactions.
Ijms 22 06817 g001
Figure 2. The effect of administration and withdrawal of OSI-906 on metabolic organs. The daily administration of OSI-906, a dual inhibitor of IGF-I receptor (IGF-IR) and insulin receptor (IR), to mice for a week provokes hepatic steatosis, induces lipoatrophy, and increases β-cell proliferation accompanied by hyperglycemia and hyperinsulinemia. The withdrawal of OSI-906 reverse hepatic steatosis, lipoatrophy, and β-cell proliferation, although the increased β-cell mass remained.
Figure 2. The effect of administration and withdrawal of OSI-906 on metabolic organs. The daily administration of OSI-906, a dual inhibitor of IGF-I receptor (IGF-IR) and insulin receptor (IR), to mice for a week provokes hepatic steatosis, induces lipoatrophy, and increases β-cell proliferation accompanied by hyperglycemia and hyperinsulinemia. The withdrawal of OSI-906 reverse hepatic steatosis, lipoatrophy, and β-cell proliferation, although the increased β-cell mass remained.
Ijms 22 06817 g002
Table 1. The phenotype of the tissue-specific IGF-I receptor, insulin receptor, or insulin receptor/IGF-I receptor double-knockout mice.
Table 1. The phenotype of the tissue-specific IGF-I receptor, insulin receptor, or insulin receptor/IGF-I receptor double-knockout mice.
TissuePromoter-Driven CreIGF-IRKOIRKODKO
LiverAlbumin-severe insulin resistance-
(Hepatocyte) overt severe diabetes
increase in β-cell mass
liver dysfunction (age-related)
reduced serum triglycerides and free fatty acids
MuscleCreatine kinasenormal body weightnormal body weightreduced body weight
normal glucose tolerancenormal glucose tolerancenormal glucose tolerance
normal cardiac performanceimpaired cardiac performancedeveloped heart failure
elevated serum triglycerides and free fatty acids
ACTA1normal body weightnormal body weightsevere muscle atrophy
no change in muscle massno change in muscle massnormal glucose tolerance
normal glucose tolerancenormal glucose tolerance
Mef2cnormal body weight--
normal serum glucose and
triglycerides
Adipose tissueAdiponectinreduced WAT and BAT massreduced body weightreduced body weight
decreased lipogenic gene
expression
reduced WAT and increased BAT massreduced WAT and BAT mass
lower plasma leptin and adiponectin levellower plasma leptin and
adiponectin level
lower plasma leptin and
adiponectin level
normal glucose toleranceovert diabetesovert diabetes
normal insulin tolerancesevere insulin resistancesevere insulin resistance
ectopic lipid accumulationectopic lipid accumulation
dyslipidemiadyslipidemia
pancreatic islet hyperplasiapancreatic islet hyperplasia
cold intolerancesevere cold intolerance
increased basal energy
expenditure
Pancreatic β-cellRat insulin 2
promoter
no change in β-cell mass and
β-cell proliferation
reduced β-cell mass and β-cell proliferationmarkedly reduced β-cell mass
impaired glucose- and arginine-induced insulin secretion and impaired glucose toleranceimpaired glucose-stimulated insulin secretion and impaired glucose toleranceovert diabetes
BrainRat nestin
promoter
reduced brain sizenormal brain size
growth retardationnormal growth
behavioral changesmild obesity
insulin resistance
IGF-IRKO, IGF-I receptor knockout; IRKO, insulin receptor knockout; DKO, IGF-I receptor/insulin receptor double knockout; BAT, brown adipose tissue; WAT, white adipose tissue.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Okuyama, T.; Kyohara, M.; Terauchi, Y.; Shirakawa, J. The Roles of the IGF Axis in the Regulation of the Metabolism: Interaction and Difference between Insulin Receptor Signaling and IGF-I Receptor Signaling. Int. J. Mol. Sci. 2021, 22, 6817. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22136817

AMA Style

Okuyama T, Kyohara M, Terauchi Y, Shirakawa J. The Roles of the IGF Axis in the Regulation of the Metabolism: Interaction and Difference between Insulin Receptor Signaling and IGF-I Receptor Signaling. International Journal of Molecular Sciences. 2021; 22(13):6817. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22136817

Chicago/Turabian Style

Okuyama, Tomoko, Mayu Kyohara, Yasuo Terauchi, and Jun Shirakawa. 2021. "The Roles of the IGF Axis in the Regulation of the Metabolism: Interaction and Difference between Insulin Receptor Signaling and IGF-I Receptor Signaling" International Journal of Molecular Sciences 22, no. 13: 6817. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22136817

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop