Next Article in Journal
Quantitative MRI Evaluation of Ferritin Overexpression in Non-Small-Cell Lung Cancer
Next Article in Special Issue
Metabolomics of Plasma in XLH Patients with Arterial Hypertension: New Insights into the Underlying Mechanisms
Previous Article in Journal
Pulsed Hyperoxia Acts on Plasmatic Advanced Glycation End Products and Advanced Oxidation Protein Products and Modulates Mitochondrial Biogenesis in Human Peripheral Blood Mononuclear Cells: A Pilot Study on the “Normobaric Oxygen Paradox”
Previous Article in Special Issue
Separation of Lipoproteins for Quantitative Analysis of 14C-Labeled Lipid-Soluble Compounds by Accelerator Mass Spectrometry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Changes in Cells Associated with Insulin Resistance

by
Leszek Szablewski
Chair and Department of General Biology and Parasitology, Medical University of Warsaw, Chałubińskiego Str. 5, 02-004 Warsaw, Poland
Int. J. Mol. Sci. 2024, 25(4), 2397; https://doi.org/10.3390/ijms25042397
Submission received: 6 January 2024 / Revised: 10 February 2024 / Accepted: 14 February 2024 / Published: 18 February 2024

Abstract

:
Insulin is a polypeptide hormone synthesized and secreted by pancreatic β-cells. It plays an important role as a metabolic hormone. Insulin influences the metabolism of glucose, regulating plasma glucose levels and stimulating glucose storage in organs such as the liver, muscles and adipose tissue. It is involved in fat metabolism, increasing the storage of triglycerides and decreasing lipolysis. Ketone body metabolism also depends on insulin action, as insulin reduces ketone body concentrations and influences protein metabolism. It increases nitrogen retention, facilitates the transport of amino acids into cells and increases the synthesis of proteins. Insulin also inhibits protein breakdown and is involved in cellular growth and proliferation. On the other hand, defects in the intracellular signaling pathways of insulin may cause several disturbances in human metabolism, resulting in several chronic diseases. Insulin resistance, also known as impaired insulin sensitivity, is due to the decreased reaction of insulin signaling for glucose levels, seen when glucose use in response to an adequate concentration of insulin is impaired. Insulin resistance may cause, for example, increased plasma insulin levels. That state, called hyperinsulinemia, impairs metabolic processes and is observed in patients with type 2 diabetes mellitus and obesity. Hyperinsulinemia may increase the risk of initiation, progression and metastasis of several cancers and may cause poor cancer outcomes. Insulin resistance is a health problem worldwide; therefore, mechanisms of insulin resistance, causes and types of insulin resistance and strategies against insulin resistance are described in this review. Attention is also paid to factors that are associated with the development of insulin resistance, the main and characteristic symptoms of particular syndromes, plus other aspects of severe insulin resistance. This review mainly focuses on the description and analysis of changes in cells due to insulin resistance.

1. Introduction

Insulin is a polypeptide hormone composed of 51 amino acids mainly secreted by pancreatic β-cells as a reaction to an increased concentration of glucose in serum. It is involved in the regulation of different processes in the human body and is considered a metabolic hormone that stimulates several anabolic processes, while catabolic processes are inhibited by this hormone. Insulin regulates energy storage and metabolism in organs such as the liver, kidney, brain, skeletal muscles as well as in adipose tissue. For example, in the liver, it stimulates the synthesis of glycogen and accumulation of lipids and inhibits the production of hepatic glucose in processes of gluconeogenesis and glycogenolysis. In peripheral muscles, insulin stimulates metabolism, increasing glucose uptake, glycogen synthesis and muscle mass. In the brain, this hormone is involved in the stimulation of processes associated with hunger. In adipose tissue, insulin stimulates processes such as metabolism and the uptake of glucose, fat storage in the process of lipogenesis and transport of fatty acids from the bloodstream into cells. Insulin also inhibits lipolysis. There are also several other metabolic processes which are regulated by insulin. By stimulating lipogenesis and the synthesis of proteins and inhibiting lipolysis and protein breakdown, insulin promotes the growth and proliferation of cells. The above-mentioned processes are associated with intracellular signaling pathways [1,2,3,4].
Insulin resistance (IR) is characterized by a reduced response of insulin-sensitive organs and tissues to stimulation by insulin. IR is also described as impaired insulin sensitivity due to a decreased reaction of insulin signaling for blood glucose levels. This pathophysiological reaction increases plasma insulin levels, resulting in hyperinsulinemia, which is associated with an increased risk of initiation, progression and metastasis of several cancers; poor cancer outcomes; as well as the development of metabolic diseases, such as type 2 diabetes mellitus (T2DM), metabolic syndrome, etc. [5]. Genetic and environmental factors contribute to the development of insulin resistance, specifically in mutations in genes that are associated with intracellular insulin signaling pathways. These mutations can alter the insulin protein and/or insulin receptor and can also affect insulin signaling pathways. Other mutations can affect insulin metabolism [4,5,6,7,8].

2. Characteristics of Insulin

Insulin controls a wide variety of biological processes. Activation of the insulin receptor (INSR) by hormone binding initiates the signaling pathway that leads to the activation of enzymes which control many aspects of metabolism and growth, which will be further described below. Perturbations in these signaling pathways may cause, for example, insulin resistance, hypertension and/or T2DM [1,5,9,10].

2.1. Biosynthesis and Secretion of Insulin

Insulin is an anabolic hormone comprised of 51 amino acids. It has previously been suggested that insulin is solely produced by pancreatic β-cells located in Langerhans islets; however, recent observations have revealed that low concentrations of insulin are also produced in certain neurons of the central nervous system [11]. Insulin’s biosynthesis and secretion are controlled by levels of circulating glucose and increase when glucose levels are high. The biosynthesis of insulin is stimulated when the concentration of glucose is between 2 mM and 4 mM, and secretion is inhibited when glucose levels rise above 5 mM [12,13]. After secretion, the hormone circulates systematically, reaching cells in the liver, muscle and fat, where it is taken and stored, resulting in reduced levels of glucose in the blood [2,14,15].

Biosynthesis of Insulin

As mentioned above, the main site of insulin biosynthesis is pancreatic β-cells. In humans, the hormone is encoded by the insulin gene (INS), located on the short arm of chromosome 11. The process of transcription is controlled by upstream enhancer elements, such as IDX1 (PDX1), MafA, NeuroD1 and others [2,16]. An effect of translation is synthesized preproinsulin, which, in the rough endoplasmic reticulum (RER), is processed into proinsulin upon cleavage of its signal sequence. Proinsulin is then folded and stabilized in its 3D configuration. This stage is due to linking the semihelical A and B domains by the formation of three disulfide bonds. The next stage of insulin synthesis is processed in the Golgi apparatus, where folded proinsulin is placed in still-immature secretory granules, where the convertases PC1/3 and PC2 cleave the C-peptide. The last process is the synthesis of mature insulin via the removal of C-terminal basic amino acids from the peptide chain. This process is due to carboxypeptidase E [17] (Figure 1). Most insulin granules are stored in the cytoplasm of pancreatic β-cells. To reach the plasma membrane and for secretion of mature insulin by β-cells, granules with insulin must cross a critical actin network. This process requires the reorganization of actin because the actin network plays a role as a physical barrier to the secretion of insulin [2,18,19]. This process is controlled by small G-proteins as well as their activating exchange factors, such as glucose- and Cdc42-dependent activation of Rac 1 [20]. Secretory granules must dock at the plasma membrane. The next step of insulin secretion depends on the response to the intracellular Ca2+ signal, which is associated with Ca2+ channels [21]. The secretion of insulin is biphasic. The first phase, which is rapid, is caused by fusion and secretion by docked secretory granules. In humans, this phase lasts up to 10 min. The second phase is associated with the reorganization of cytoskeletal elements, namely actin, resulting in the recruitment of granules to the plasma membrane [22].

3. Intracellular Insulin Signaling Pathways

Insulin controls many biological processes that involve the tyrosine kinase insulin receptor (INSR). Activated INSR initiates a phosphorylation of substrates associated with insulin signaling pathways, activating enzymes involved in the control of metabolic processes and growth. Disturbances in these signaling pathways may cause insulin resistance.

3.1. Insulin Receptor

Insulin action begins when insulin binds to its receptor localized on the cell membrane of the target cells. INSR is a heterotetrameric protein that consists of two extracellular α-subunits and two β-subunits which are bound together by disulfide bonds (Figure 2). Both α- and β-subunits are obtained from a single precursor by proteolytic cleavage. The INSR mRNA undergoes alternative splicing, generating two different isoforms: isoform A (INSR-A) and isoform B (INSR-B). INSR-B contains a 12-amino-acid sequence in the carboxy-terminal (C-terminal) part of α-subunit, which is missing in INSR-A. These isoforms also alter the site of expression. INSR-A is predominantly expressed in embryo and fetal tissues; the central nervous system, especially in the brain; hematopoietic cells; and tumor cells. Expression of INSR-B is highest in the liver and in insulin-targeted tissues, such as muscle, adipose tissue and the kidneys [23,24,25]. INSR-A mediates mainly mitogenic effects, whereas INSR-B is involved mainly in metabolic effects [26]. There is also the hybrid receptor. In this case, the heterotetramer is composed of an α/β dimer of INSR and an α/β dimer of insulin-like growth factor-1 receptor (IGF-1R). Insulin-like growth factor-1 (IGF-1) and insulin-like growth factor-2 (IGF-2) bind with a much higher affinity to the hybrid receptor than insulin [27].

3.2. Insulin Receptor Substrates

Insulin binds to the α-subunit of INSR, inducing a conformational change. This conformational change induces activation of tyrosine kinase in the β-subunit, causing the autophosphorylation of tyrosine residues and the subsequent activation of phosphotyrosine-binding proteins [3,4]. The substrates of INSR are, for example, members of the insulin receptor substrate (IRS) family of proteins. These proteins are IRS-1 through IRS-6. IRS-1 and IRS-2 [29,30,31] are expressed principally in human beings. There are also other insulin receptor substrates, such as growth factor receptor-bound protein-2 (GRB-2), GRB-10, Shc-transforming protein (Shc), SH2B adapter protein-2 (SH2B2), protein-tyrosine phosphatase 1B (PTP1B) and others [32]. Phosphorylation of these substrates activates two primary signaling pathways: the phosphoinositide 3-kinase (PI3K)/protein kinase B (AKT) pathway and the mitogen-activated protein kinase (MAPK) pathway. The phosphorylation of Shc activates the Ras-MAPK pathway, whereas the phosphorylation of IRS proteins mostly activates the PI3K/AKT pathway [4,28].

3.3. The PI3K/AKT Signaling Pathway

Autophosphorylation of the INSR, due to insulin binding, causes activation of the IRS protein. Activation of PI3K by IRS causes activation of downstream AKT/mTOR-network ((serine/threonine protein kinase, also known as PKB—protein kinase B) mammalian target of rapamycin complex) signaling. This process is due to phosphorylation of phosphatidylinositol 4,5-bisphosphate (PIP2) to produce phosphatidylinositol 3,4,5-triphosphate (PIP3), which phosphorylates and activates AKT and atypical protein kinase C (aPKC). AKT is involved in many cellular functions. Activated AKT phosphorylates glycogen synthase kinase, causing its deactivation, and inhibits the activity of glycogen synthase and ATP-citrate lyase, resulting in the inhibition of glycogen and fatty acids synthesis. Activated AKT by phosphorylation inactivates the mammalian target of rapamycin complex 1 to promote protein synthesis. By inhibition of the proapoptotic pathway, AKT mediates cell survival. It also activates sterol regulatory binding proteins (SREBPs), which are translocated to the nucleus, where they play a role in the transcription of genes associated with the synthesis of fatty acids and cholesterol. AKT also regulates the cell cycle and translocation of glucose transporters from the intracellular space to the plasma membrane of muscles and fat cells, increasing the uptake of glucose by these cells from the blood [3,6,33,34,35].

3.4. The MAPK Signaling Pathway

A second essential insulin signaling pathway is the MAPK signaling pathway. Its activation is independent of the PI3K/AKT signaling pathway [1]. Activated INSR and IRS proteins contain docking sites for adaptor molecules, which contain SH2 domains, such as growth factor receptor-bound protein 2 (Grb2) and Src homology (Shc). The MAPK pathway is activated when IRS-1 binds to Grb2 [4]. To Grb2 binds son-of-sevenless (SOS), which is a guanine nucleotide exchanger factor (GEF) for Ras. SOS binds to Ras, catalyzing the change of Ras from an inactive GDP-bound form (guanosine diphosphate) (Ras-GDP) to an active GTP-bound form (guanosine triphosphate) (Ras GTP). Activated Ras stimulates cRaf that phosphorylates and activates MAPK/Erk kinases, MEK1 and MEK2 (MAP kinases ERK1 and ERK2—extracellular signal-regulated kinase 1 and 2). Then, activated ERKs are translocated to the nucleus, where they cause phosphorylation and transcriptional activation by transcriptional factors, such as ELK1, promoting cell division, protein synthesis and cell growth [3,6,36].

4. Role of Insulin in Selected Cells, Tissue, Organs

As mentioned above, the main role of insulin is the regulation of the body’s energy supply. Insulin also contributes to the uptake of glucose from the blood stream into insulin-dependent cells of muscle, the liver and adipose tissue. Insulin may play a specific role in particular cells, depending on the type of cell. These specific roles depend on the intracellular insulin signaling pathway.

4.1. Insulin Signaling Pathway and Its Role in the Liver

The primary organ of insulin action is the liver. Insulin action begins after the hormone binds to INSR, which activates IRS-1 [2,37]. Activated IRS-1 recruits PI3K and then AKT. After activation, AKT suppresses the expression of the forkhead box O1 (FOXO1)-mediated gluconeogenic gene, decreasing gluconeogenesis. Insulin increases the synthesis of glycogen in the liver. This is due to the regulation of glycogen synthase 2 (GYS2) and glycogen phosphorylase. Glycogen synthase kinase 3 (GSK3) and protein phosphatase 1 (PP1) are involved in this process. Upregulating sterol regulatory element-binding protein 1c (SREBP-1c), insulin increases de novo lipogenesis in the liver [3,4,38,39].

4.2. Insulin Signaling Pathway and Its Role in Muscle

As in previous cells, the intracellular insulin signaling pathway begins when insulin binds to INSR, resulting in the activation of IRS and recruitment of PI3K, activating the AKT signaling pathway. Activated AKT stimulates the uptake of glucose due to the translocation of glucose transporter 4 (GLUT4) storage vesicles from the cytoplasmic milieu to the plasma membrane. The translocation of GLUT4 takes place within minutes of insulin binding to INSR [2]. This translocation depends on inactivation of the GTPase-activating protein (GAP) AKT substrate of 160 kDa (AS160), and the GTP-bound form of Ras-related C3 botulinum toxin substrate 1 (RAC1) must be involved. The stimulation of glycogen synthesis by insulin is associated with the inhibition of glycogen synthase kinase 3, activation of glycogen synthase as well as inactivation of glycogen phosphorylase, which depends on the dephosphorylation of phosphorylase kinase [3,4,40].

4.3. Insulin Signaling Pathway and Its Role in White Adipose Tissue

In adipocytes, insulin suppresses lipolysis, which in turn causes the suppression of hepatic glucose production (HGP) [2]. The suppression of HGP is due to the reduction of substrates which are involved in the process of gluconeogenesis. In this process, phosphodiesterase 3B (PDE3B), protein phosphatase 1 and protein phosphatase-2A (PP2A) also play an important role. In adipose tissue/adipocytes, insulin stimulates other processes, such as the transport of glucose into adipocytes, due to the translocation of GLUT4, lipogenesis and adipogenesis [2,41]. In these processes, Rab, SREBP-1c and peroxisome proliferator-activated receptor-γ (PPAR-γ) are involved, respectively. Insulin also activates several substrates, which are associated with processes taking place in adipose tissue cells. Examples of these substrates/enzymes include PIP2, PIP3, PKB, mTORC1 (also known as mechanistic target of rapamycin complex 1), glucose-6-phosphatase (G6Pase), phosphoenolopyruvate carboxykinase 1 (PEPCK; also known as PCK1), glucokinase (GCK), glycerol-3-phosphate acetyltransferase 1 (GPAT1), acetyl-CoA carboxylase (ACC), fatty acid synthase (FAS) and so on [3,42].

4.4. Insulin Signaling Pathway and Its Role in Other Organs and Cells

Insulin is also involved in regulating processes in other organs and cells. It regulates the expression and activity of different proteins, such as enzymes, transcription factors, proteins which regulate cell cycle and apoptosis as well as proteins associated with survival [43,44,45].
In the central nervous system, insulin plays an important role. It crosses the blood–brain barrier (BBB) through a receptor-mediated process [42]. In the cerebrospinal fluid, the concentration of insulin is one-third of that in circulation. Insulin action is associated with INSR localized in cell membranes of neurons and glial cells. For example, the mentioned hormone regulates appetite, reducing the expression of neuropeptide Y and agouti-related peptide (orexigenic) and increasing the expression of pro-opiomelanocortin (anorexigenic). It also regulates energy expenditure [46,47,48]. This hormone also participates in the regulation of processes associated with trophic levels and the development of neurons and glial cells. It has also been suggested that insulin contributes to the modulation of cognition, memory and mood [49]. A significantly improved process of memory after intranasal administration of single-dose (160 IU) insulin was observed in women but not in men [50]. Insulin stimulates the uptake of glucose in the spinal cord tissues and brain regions, such as the choroid plexus, the pineal gland and the pituitary [51,52,53]. Results obtained in studies showed that intranasal administration of single-dose insulin decreases food intake in healthy males, although this effect was not observed in women [50]. Precise mechanisms of many insulin actions in the nervous system are still unknown and need further investigation.
Insulin also regulates pancreatic functions. For example, it suppresses the secretion of glucagon from pancreatic α-cells [54], preventing increased glucose levels due to the process of glycogenolysis. It also stimulates δ-cells of pancreatic islets, which secrete somatostatin. As the first target cells for insulin action are the α-cells, these cells located at the periphery of pancreatic islets secrete glucagon. Insulin decreases the secretion of glucagon, which in turn increases many metabolic effects which are associated with insulin action.
In the kidneys, insulin regulates many metabolic processes, such as kidney homeostasis [55] and glucose metabolism [56]. Recently, an association of insulin with metabolic processes was suggested. Therefore, there has been an investigation, for example, into the physiological role of INSR in the kidneys [57,58]. Observations performed on animal models of diabetes and insulin resistance revealed the decreased expression of INSR and phosphorylated INSR in renal epithelial cells. The decreased expression of insulin receptors in the kidneys was detected in rat models of type 1 diabetes and patients with type 2 diabetes [57]. Insulin interacts with angiotensin type 1 receptors in the renin-angiotensin system. Angiotensin II (Ang-II) inhibits the insulin activation of PI3K signaling, causing insulin resistance [59]. Insulin is involved in regulating renal glucose metabolism, as hyperglycemia may lead to the development of diabetic kidney diseases (DKDs) [56]. Insulin is also involved in bone development [60]. This process may be regulated by insulin signaling, by which means insulin promotes osteoblast development and resorption of bones by osteoclasts [57]. It was found that these cells express INSR on their surface [61]. Based on the results, it was suggested that insulin receptors in osteoblasts are necessary for the proliferation, survival and differentiation of osteoblasts. Insulin also increases the rate of osteoblast proliferation, synthesis of collagen, production of alkaline phosphatase and uptake of glucose as well as inhibiting the activity of osteoclasts [62].
Insulin is also involved in metabolic processes of the skin and hair follicles. Hair follicles require a regular supply of oxygen and nutrients [63]. Hyperglycemia decreases oxygen and nutrient supplies, causing follicular damage and changing hair growth. Follicular damage causes thinning, fragility, sparseness of hair and a decline in hair growth [64,65]. Certain skin conditions are associated with insulin action, such as acrochordons and acne [66]. Acne is a symptom of many skin diseases associated with insulin receptors, whereas carbohydrate metabolic impairment is detected in individuals with acrochordons [67]. Deregulated insulin sensitivity is observed in individuals with psoriasis [68,69].
Insulin is also involved in ketone body metabolism. Hyperinsulinemia due to prolonged fasting or uncontrolled diabetes mellitus accelerates fat mobilization, causing an excessive influx of free fatty acids to the liver. This disturbance causes the synthesis of ketone bodies by the liver. Substrates for this synthesis are by-products of incomplete beta-oxidation of long-chain fatty acids. Products of this synthesis, ketoacids such as acetoacetate, beta-hydroxybutyrate and acetone, may be used as a source of energy by extrahepatic tissues, mainly in skeletal muscle and the heart, although under extreme conditions, these derivatives may be used by the brain as a fuel [70]. Concentrations of circulating ketone bodies are reduced by insulin via several mechanisms, for example, the inhibition of lipolysis by hormones and decreased transport of free fatty acids to the liver for ketogenesis. Insulin also inhibits hepatic ketogenesis [71]. Hyperinsulinemia is due to increased peripheral clearance of ketone bodies [72].
Insulin is also involved in protein metabolism. It increases the retention of nitrogen and protein acceleration; increases the number of ribosomes; stimulates the uptake of amino acids by hepatocytes, skeletal muscle and fibroblasts; as well as inhibits protein breakdown. These processes increase protein synthesis and protein concentration [73,74].
Insulin also interacts with the vasculature, influences endothelial cells of the vessels of systemic circulation and acts on endothelial cells, which promotes blood flow and ensures its delivery to peripheral tissues [75]. In large blood vessels, for example, the aorta and the larger arteries, insulin binds to the insulin receptors of endothelial cells. Conformational changes to INSR cause their autophosphorylation, resulting in the phosphorylation of IRS-2. The phosphorylation of IRS-2 activates the PI3K signaling pathway, which signals downstream to the serine and threonine kinase AKT/PKB. AKT activates endothelial NO synthesis, which catalyzes the conversion of L-arginine to NO [76], a potent vasodilator. The stimulation of arteries and arteriole dilation by insulin is due to the synthesis of NO [77]. These vascular effects are temporally associated with insulin action [75]. For example, the full stimulation of glucose uptake by skeletal muscle depends on prior NO-mediated vasodilation [78].

5. Insulin Resistance

Physiologically, insulin resistance (IR) is defined “as a state of reduced responsiveness in insulin-targeting tissues to physiological levels of insulin” [3]; “a state of a cell, tissue, or organism in which a greater than normal amount of insulin is required to elicit a quantitatively normal response” [79]; or “an inability of some types of tissues to respond to normal insulin levels, and thus, higher than normal levels of insulin are required to maintain the normal functions of insulin” [3]. The mechanisms of insulin resistance have not been fully established, but several suggestions have been made.

5.1. Severe Insulin Resistance Syndromes

Severe insulin resistance syndromes are rare syndromes which are characterized by profound insulin resistance. Severe insulin resistance is defined “as a severely diminished response to insulin’s biological effects and is characterized by substantial hyperinsulinemia and impaired response to endogenous and exogenous insulin” [34]. To maintain euglycemia, patients with severe insulin resistance need large amounts of exogenous insulin. These patients might present hypoglycemia, which may precede hyperglycemia [80,81]. Such syndromes are detected in 0.1–0.5% of patients in diabetes clinics [80].

5.1.1. Pathogenesis of Severe Insulin Resistance

Insulin resistance may be due to different factors, such as mutations in insulin signaling pathways, autoimmune reactions and environmental factors (Table 1).

5.1.2. Defects in the Insulin Gene

Autosomal dominant mutations in the insulin gene influence the secondary structure of the insulin protein due to the disruption of three disulfides and bonds in mature insulin. The impaired secondary structure of insulin causes endoplasmic reticulum stress and destruction of β-cell pancreatic islets of Langerhans. Autosomal recessive mutations are involved in the reduced biosynthesis of the hormone or loss of function of the insulin protein.
Insulin Wakayama, a clinical variant insulin, is an effect of the substitution in the A chain of conserved valine by leucine (ValA3→Leu). This mutation decreases the affinity of the receptor for insulin binding 500-fold [82]. Mutations in the B chain PheB24→Ser (insulin Los Angeles) and PheB25→Leu (insulin Chicago) [83] significantly reduce the bioactivity of insulin as well as decrease the binding affinity to the INSR [5]. The substitution of histidine by asparagine (HisB10→Asp) enhances the activity of the hormone caused by upregulated pathways. This mutation, described in proinsulin, is associated with hyperinsulinemia due to mis-trafficking of the protein [84].

5.1.3. Defects in the Insulin Receptor

The effects of mutations in the insulin receptor gene depend on their site. Mutations in the α-subunit of the insulin receptor may cause a decreased number of active mature insulin receptors as well as the affinity of insulin receptors for binding insulin. The affected activation of downstream signaling cascades, caused by impaired autophosphorylation of the insulin receptor, may be due to mutations in the β-subunit of INSR, the tyrosine kinase domain [25,85].
Donohue syndrome, also known as leprechaunism, is an extremely rare autosomal recessive disease due to mutations in the INSR. This syndrome affects fewer than 1 per million people worldwide, affects the ability of the hormone to bind to the receptor and is the most severe defective insulin signaling syndrome [34]. Patients with leprechaunism seldom live beyond infancy, and most affected individuals survive less than two years. Their deaths are mainly associated with intercurrent infection [80,81].
Rabson–Mendenhall syndrome (RMS) is a rare autosomal recessive disease caused by mutations in the insulin receptor gene, affecting fewer than 1 per million people worldwide. RMS is a mild form of severe insulin resistance syndrome [86]. In affected patients, extremely high levels of insulin [87], fasting hyperglycemia and failed responses to endogenous and exogenous insulin [81] may be detected. Most patients survive only up to 15 years of age, although there have been some cases of affected individuals living into their third decade of life [88].
Type A insulin resistance syndrome (TAIRS) is another rare disorder due to mutations in INSR, with an incidence of 1 in 100 000 [89]. TAIRS may be caused by autosomal dominant or recessive mutations in INSR. Heterozygous and, in some cases, homozygous mutations in the insulin receptor gene impair the function of the receptor as well as signal transduction [90,91,92,93]. TAIRS has a relatively good prognosis, and affected individuals can live beyond middle age [94]. The features of TAIRS are subtler in affected males. In some affected males, low blood sugar (hypoglycemia) can be the only indication. This syndrome is diagnosed more often in females than in males, as females have more health problems which are associated with the condition.
Type C insulin resistance syndrome is a variant of Type A insulin resistance syndrome, with less severe IR. It is also a so-called HAIR-AN syndrome (hyperandrogenic, insulin-resistant, and acanthosis nigricans). This syndrome is inherited as an autosomal dominant disease [95,96]. HAIR-AN syndrome is generally diagnosed in obese women who do not demonstrate defects of the INSR; however, these women may exhibit postreceptor defects [97]. This syndrome may affect up to 3% of women with androgen excess [98]. A significant role in determining the degree of IR may be associated with the degree of obesity [95].
Type B insulin resistance syndrome (TBIRS) is a rare autoimmune disorder. It is caused by polyclonal autoantibodies against the insulin receptor [34]. TBIRS occurs mainly between the fourth and sixth decade of life. Typically, it is diagnosed in middle-aged women in association with other autoimmune disturbances [99,100,101,102,103]. In adulthood, this syndrome is associated with a higher 10-year mortality risk [104]. The autoantibodies may act bi-phasically. In the first phase, they induce hypoglycemia, which ultimately may cause hyperglycemia. It is also suggested that high concentrations of autoantibodies antagonize INSR, causing its inhibition and leading to insulin resistance and hyperglycemia. Low levels of these autoantibodies agonize, causing hypoglycemia [105,106].

5.1.4. Lipodystrophies

The lipodystrophy syndromes are a heterogenous group of disorders. In these syndromes, severe insulin resistance and severe hypertriglyceridemia are observed, leading to pancreatitis and fatty infiltration of the liver, leading to cirrhosis [107]. Complete or partial loss of adipose tissue and depletion of capacity for the storage of lipids [108,109] is detected in affected individuals. Lipodystrophies are rare diseases, with approximately 1.3–4.7 cases per million [110]. The lipodystrophy syndromes have been classified based on the extent, the location of dystrophy and the age of onset [108,109]. According to etiology, there are congenital and acquired lipodystrophies. Based on the extent of adipose tissue deficiency, there are generalized or partial lipodystrophies, in which are included four categories: congenital generalized lipodystrophy; acquired generalized lipodystrophy; congenital partial lipodystrophy; and acquired partial lipodystrophy, also known as Barraquer–Simons syndrome [34]. These lipodystrophies may be inherited in an autosomal recessive manner, for example, congenital generalized lipodystrophy (CGL, Berardinelli–Seip syndrome) and familial partial lipodystrophies 5 and 6 (FPLD5, FPLD6); in an autosomal dominant manner, for example, FPLD1 (Köbberling type), FPLD2 (Dunnigan type), FPLD4, SHORT syndrome (short stature, hyperextensibility of joints, ocular depression, Rieger anomaly, teething delay) and mandibuloacral dysplasia (MAD); and in an X-linked manner, for example, Köbberling–Dunnigan syndrome (however, it rarely may be inherited as autosomal dominant) [34,111]. Lipodystrophies, such as congenital generalized lipodystrophies, are detected in newborns or infants, whereas in the case of acquired generalized lipodystrophy (Lawrence syndrome), patients appear normal at birth, but over days or weeks, they develop lipodystrophy [79].

5.1.5. Other Severe Insulin Resistance Syndromes

Other groups of rare genetic syndromes are described which are associated with severe insulin resistance. These syndromes include Alström syndrome, an autosomal recessive disease; myoclonic dystrophy, an autosomal dominant disorder [79]; autosomal recessive syndromes; Werner syndrome, a progeria and autosomal recessive syndrome [112]; HALS (HIV-associated lipodystrophy syndrome); Bloom syndrome; and microcephalic osteoplastic syndrome [34].

5.2. Mechanisms of Insulin Resistance

The mechanism of insulin resistance is not fully understood, although several theories have been suggested. As mentioned earlier, insulin resistance is an important pathology associated with many metabolic diseases, such as T2DM, metabolic syndrome as well as cancers. There is also an alternative theory that insulin resistance is a mechanism that protects tissue from metabolic injury due to nutrient excess [113,114]. Chronic over-nutrition causes those organs and tissues responsive to insulin glucose uptake to protect themselves from toxicity due to nutrients by becoming insulin resistant. This physiological defense may be observed, for example, in the heart and skeletal muscle [115,116]. The glucose-regulating effects of insulin, such as the suppression of hepatic glucose production and lipolysis, uptake of glucose by cells and synthesis of glycogen, in insulin-resistant tissues at normal plasma levels of insulin are not observed [117].

5.2.1. Insulin Resistance in Select Tissues and Organs

Skeletal muscle is an important tissue involved in glucose metabolism in which glucose uptake is stimulated by insulin. This tissue is responsible for up to 80% of postprandial glucose uptake from blood circulation. Therefore, muscular insulin resistance is the primary defect in T2DM and may affect whole-body metabolism [118,119,120]. Insulin resistance in skeletal muscle, due to the deletion of insulin receptor tyrosine kinase (IRTK) or GLUT4, increases adiposity in animal models [121,122]. Results obtained in studies suggest that impaired glucose transport is due to disturbances in the insulin signaling pathway [123]. There are also suggestions that IR in skeletal muscle may be due to defects at the proximal level of insulin signaling. These defects may be caused by impaired activation of insulin receptor tyrosine kinase, IRS-1, PI3K and AKT [3]. This suggestion is confirmed by observations that in skeletal muscle of insulin-resistant mice and obese individuals, the activity of tyrosine kinase of IRTK is decreased [121,124]. It was found also that in insulin-resistant skeletal muscle, the phosphorylation of IRS-1 tyrosine and IRS-1-associated activity of PI3K were damaged [125]. The increased degradation of protein in insulin-resistant muscle due to disturbances in insulin signaling pathways was confirmed also in another study [126]. Experiments performed on animals revealed an increased degradation of protein and activation of the major proteolytic systems, caspase-3 and the proteasome, in the muscle of db/db mice, a model of insulin resistance. More detailed research showed that insulin resistance causes muscle degradation. This pathology is due to a mechanism that suppresses PI3K/AKT, the signaling of which activates caspase-3 and the ubiquitin-proteasome proteolytic pathway, resulting in the degradation of protein [126]. An increased risk of insulin resistance may be due to aging skeletal muscle [127]. With increasing age, the body’s glucose regulation ability decreases and muscles atrophy [128]. Older people have decreased glucose metabolism and expression of GLUT4 in skeletal muscle [129,130], resulting in pathological changes, such as lower activity of AKT stimulated by insulin [131], disturbances in insulin signaling pathways and decreased insulin sensitivity [132]. In aging skeletal muscle, pathophysiological changes are detected, such as mitochondrial dysfunction, intramyocellular lipid accumulation, inflammation, oxidative stress, endoplasmic reticulum stress, autophagy, sarcopenia and a weakened renin-angiotensin system non-classical axis. Aging skeletal muscle is an independent risk factor for insulin resistance, but as mentioned above, pathologies may increase the risk of insulin resistance in skeletal muscle, after which insulin resistance may stimulate developing pathological changes associated with IR [127,133].
The liver is involved in the control of postprandial carbohydrate levels. This control is due to the suppression of hepatic glucose production (HGP) and the stimulation of the deposition of glucose as glycogen. During fasting, the liver plays a role as the primary source of glucose production [134]. Impaired suppression of hepatic gluconeogenesis, a pathology observed in insulin resistance, is mainly caused by defects in lipolysis in adipose tissue and the de-suppression of forkhead box 01 (FOXO1) transcription factor in the liver [134]. Insulin resistance is also associated with impaired stimulation of glycogen synthesis induced by insulin [135,136,137]. Hepatic IR also impairs the regulation of hepatic glycogen metabolism associated with fasting and feeding [136].
Endothelial dysfunction, a cardiovascular disease, is also associated with insulin resistance. As mentioned earlier, insulin stimulates synthesis of the vasodilator nitric oxide. In response to insulin stimulation, synthesized NO increases blood flow to skeletal muscle, increasing glucose uptake [138]. The PI3K-dependent insulin signaling pathway in the endothelium associated with the synthesis of NO is similar to that described in skeletal muscle. The intracellular insulin signaling pathway also regulates the secretion of the vasoconstrictor endothelin-1 (ET-1) in the endothelium [139,140,141]. Insulin resistance impairs the phosphatidylinositol 3-kinase pathway, causing disturbances in the balance between the synthesis of NO and secretion of ET-1, resulting in decreased blood flow and worsened IR. Human and animal studies revealed that endothelial dysfunction due to insulin resistance is associated with oxidative stress, advanced glycation end products (AGE), a hexoamine biosynthetic pathway, proinflammatory signaling, ceramide and inflammation [142].
Insulin resistance may disturb metabolic functions, which is commonly detected in cancer patients. Impaired metabolism causes higher recurrence in rats and decreases the overall survival of these patients [143]. It is suggested that IR may be a primary driver of cancer-associated metabolic dysfunction, which in turn increases the risk of cancer recurrence and risk of death due to cancer [33,143]. A higher risk for the development of cancers such as breast, colorectal, liver, pancreatic, endometrial, lung hepatocellular and prostate is observed in patients with diagnosed insulin resistance [144,145,146]. The progression of cancer is closely related to insulin resistance [147]. The insulin receptor is expressed at higher levels in malignant cells [148]. It is postulated that increased transcription of the genes encoding INSR may be caused by mutations in tumor suppressor genes, such as TP53, and the genes which encode BRCA1, von Hippel Lindau (VHL) and Wilms tumor protein (WT1) [149,150]. It was found that in cancer cells, their expression and signaling pathways are often dysregulated [26], and in many cancers, the INSR is overexpressed. Activation of the insulin receptor activates PI3K-AKT-mTOR and MAPK-RAS signaling pathways, which are highly controlled in normal cells, and mutations in these signaling pathways impair activation of these pathways. Results obtained in several studies showed that in solid cancers, mutations in the PI3K and/or RAS are the most common mutations, causing increased signaling through the AKT-mTOR and MAPK, respectively [151,152,153]. It was also found that the MAPK-RAS cascade plays an important role in driving tumor cell proliferation [154].
Insulin resistance in the brain impairs neuroplasticity or the release of neurotransmitters in neurons. The Tau protein participates in the formation of microtubules, which are involved in several cellular processes. The activity of this protein is regulated by phosphorylation caused by insulin. In the brains of patients with Alzheimer’s disease, the level of Tau phosphorylation is three times higher when compared to normal brains. Cognitive dysfunction may be due to brain insulin deficiency and plasma insulin resistance. Insulin resistance influences the development of cognitive dysfunction due to hyperinsulinemia and impaired insulin signaling. Brain insulin resistance that depends on IRS-1 dysfunction may promote cognitive decline, independent of the classic Alzheimer’s disease pathology. For more details, see [155,156,157].
There are also other suggested hypotheses which may explain the mechanisms of insulin resistance.

5.2.2. The Glucose-Fatty Acid Cycle (the Randle Cycle)

Insulin resistance caused by lipids in skeletal muscle, observed as a defect in the uptake of glucose, is due to limited utilization of glucose, stimulated by insulin. Limited utilization of glucose is caused by increased β-oxidation of fatty acids [158,159]. Increased oxidation of fatty acids, mainly in obese subjects, disturbs use of glucose as a main source of energy, inhibiting the activity of glycolytic enzymes. The process of fatty acid β-oxidation may increase mitochondrial acetyl-CoA, causing inactivation of pyruvate dehydrogenase. This action increases levels of citrate and inhibits phosphofructokinase, which is a key glycolytic enzyme. An effect of these biochemical processes is the accumulation of intramyocellular glucose-6-phosphate that inhibits the activity of hexokinase, resulting in the accumulation of intramyocellular glucose. Animal studies performed on rat heart and diaphragm muscles showed that the infusion of fatty acids decreases the utilization of myocellular glucose, increasing intramyocellular concentrations of glucose-6-phosphate [159,160]. Based on the proposed Randle cycle theory, the levels of glucose-6-phosphate and synthesis of glycogen should be increased by fatty acids, whereas glycolysis should be inhibited. Of note, the synthesis of muscle glycogen stimulated by insulin and the oxidation of glucose are suppressed in diabetic patients with chronic insulin resistance [161].

5.2.3. Hexosamine Biosynthesis Pathway

Elevated levels of muscular acetyl-CoA and citrate associated with increased oxidation fat suggests another pathway involved in the regulation of glucose transport. This other mechanism is associated with the hexosamine biosynthesis pathway (HBP) [3]. Fructose-6-phosphate is mainly produced from glucose-6-phosphate. During glycolysis, fructose-6-phosphate is metabolized into fructose-1,6-bisphosphate. Approximately 5% of fructose-6-phosphate is converted to glucosamine-6-phosphate. This reaction is performed by the rate-limiting enzyme of HBP, glutamine:fructose-6-phosphate amidotransferase (GFAT) [162]. Then, glucosamine-6-phosphate is converted to uridine-5′-diphosphate N-acetylglucosamine (UDP-GlcNAc), which plays a role as the donor sugar nucleotide for the glycosylation and O-GlcNAcylation substrates, such as lipids and proteins [162]. Both of these modifications may affect proteins due to the regulation of gene expression or activity of enzymes [163,164,165]. Experiments showed that O-GlcNAcylation modulates insulin sensitivity [166]. O-GlcNAcylation of proteins may compete with phosphorylation for sites involved in regulating the activity of protein and signal transduction [167]. These sites of phosphorylation include insulin signaling pathway components, such as IRS-1/2, PBK, PDK1 and Akt. These components may be modified with O-GlcNAcylation [168,169]. MO-GlcNAc modification of mammalian uncoordinated-18c (Munc18-c), a protein involved in the translocation of GLUT4 stimulated by insulin, was also found. FOXO1, an important transcriptional factor for gluconeogenic genes in the liver, may also be O-GlcNAcylated [170,171].

5.2.4. Ectopic Lipid Accumulation

A single adipose cell has limited capacity for the storage of lipids [172]. A short-term high-fat diet (HFD) causes enlarged adipocytes to stimulate insulin resistance without macrophage infiltration into adipose tissue [173]. Therefore, excess lipids in adipocytes causes insulin resistance without inflammation. This situation may be involved in ectopic lipid accumulation, resulting in IR through the synthesis of metabolically toxic products. In this case, saturated fatty acids increase the production of ceramide, which in turn stimulates insulin resistance [173,174]. The level of diacylglycerol (DAG) in the liver is strongly associated with systemic insulin resistance [172]. Numerous studies have revealed that ectopic lipid accumulation in peripheral tissues, mainly in skeletal muscle and the liver, may cause more severe insulin resistance [175,176,177]. Animal studies showed that lipid accumulation in the liver and skeletal muscle due to HFD feeding or lipid/heparin infusions causes insulin resistance in rats [178]. It was found also that deficiency or inactivation of fat transport proteins, such as CD-36 or fatty acid transport protein 1 (FATP1) stimulates the uptake of glucose in skeletal muscle [179,180]. Liver-specific knockdown FATP2 or FATP5 reduces hepatosteatosis induced by HFD, increasing glucose tolerance [181,182]. It was suggested that ectopic lipid accumulation stimulates insulin resistance and lipid metabolites, such as diacylglycerol, lipophosphatidic acid (LPA), ceramides and acylcarmitines, contributing to the development of insulin resistance in the liver and skeletal muscle.

5.2.5. Diacylglycerol

Insulin resistance induced by lipids, according to the DAG hypothesis, suggests the interference of intracellular insulin signaling associated with the activation of protein kinase C (PKC) due to the accumulation of diacylglycerol within insulin-sensitive tissues [3]. There are three groups in the PKC family: conventional, novel and atypical. Novel PKC (nPKC) has a great affinity for DAG, and results suggest its contribution to insulin resistance [183,184]. Increased levels of DAG in the liver cause the translocation of PKCε, which is the primary hepatic nPKC isoform, to the plasma membrane, inhibiting insulin receptor tyrosine kinase activity. These changes reduce insulin-stimulated activation due to the phosphorylation of IRS-2, PI3K and Akt2 [185,186]. The accumulation of intramyocellular diacylglycerol impairs intracellular insulin signaling pathways and the uptake of glucose by muscle, caused by the activation of PKCθ, a type of nPKC muscle [187,188]. Activated nPKC muscle type causes phosphorylation of IRS-1 stimulated by insulin [189,190]. Performed animal studies confirmed the hypothesis of insulin resistance due to DAG-PKC [183,191].

5.2.6. Ceramide

Ceramide is a sphingolipid. It is an essential bioactive lipid synthesized from an intracellular fatty acid and sphingosine. Ceramide is involved in insulin resistance induced by lipids [192]. It stabilizes cell membranes and regulates the distribution of signaling molecules. Results obtained in research showed a dependence between ceramide levels and insulin resistance. The molecular mechanism that describes the role of ceramide in the induction of insulin resistance has not been clearly demonstrated; there is a hypothesized association of insulin resistance with the impairment of AKT translocation due to the activation of atypical PKCζ and protein phosphatase-2A (PP2A) [193,194]. Ceramide also stimulates insulin resistance by the activation of c-Jun N-terminal kinase (JNK) [195]. Therefore, inhibition of ceramide synthesis causes amelioration of insulin resistance [196]. Unfortunately, there are several unclear and controversial results, meaning that further studies to explain the role of ceramide in insulin resistance are needed [3].

5.2.7. Endoplasmic Reticulum Stress

Obesity enhances endoplasmic reticulum stress (ER); therefore, it was suggested that ER may be associated with hepatic insulin resistance and pancreatic β-cells in obese people. An over-nutrition condition causes the liver to produce an excess of enzymes to process nutrients. After accumulating unfolded proteins in the endoplasmic reticulum, the liver should enhance the unfolded protein response (UPR). In this process, glucose-regulated protein 78 (GPR78, also known as BiP) is involved by the activation of inositol requiring enzyme-1 (IRE1α) and PKR-like ER kinase (PERK). Activating transcription factor 6 (ATF6) inhibits the translation of proteins, promotes ER chaperones and reduces unfolded protein levels [197]. Experiments revealed that induction of endoplasmic reticulum stress suppresses cellular insulin signaling caused by the phosphorylation of serine in IRS-1 by JNK. Increased demand for insulin secretion also induces ER stress in pancreatic β-cells in chronic hyperglycemic conditions [198,199].

5.2.8. Inflammation

Obesity is associated with the development of a chronic low-grade inflammatory state, influencing insulin resistance [200]. Caloric overload causes expansion of adipose tissue, increasing immune cell infiltration, resulting in a proinflammatory response [201]. Proinflammatory cytokines, which induce insulin resistance, may be secreted by adipocytes and macrophages. In adipocytes, increasing secretion of chemokine monocyte chemoattractant protein-1 (MCP-1), a chemokine ligand also known as C-C Motif Chemokine Ligand 2 (CCL2), induces the accumulation of macrophage infiltration into adipose tissue, causing IR [202,203]. Animal studies revealed that deletion of MCP-1 or its receptor (CCR2) improves insulin sensitivity [203]. Immune cells and adipocytes secrete cytokines, such as TNF-α, IL1β and IL-6, the increased secretion of which induces insulin resistance by several mechanisms. Insulin resistance may be due to the activation of Ser/Thr kinases [204,205]; decreased expression of IRS-1, GLUT4 and PPARγ [206]; or activation of suppressor of cytokine signaling 3 (SOCS3) in adipocytes [207]. Toll-like receptors (TLRs) belong to the family of pattern-recognition receptors (PRRs). They play a function in innate immunity and are involved in the identification of tissue injury by danger-associated molecular patterns. During obesity, TLR2 and TLR4 are particularly associated with inflammation-associated insulin resistance. Human and animal studies showed increased expression of TLR4 in adipocytes, hepatocytes, muscle and the hypothalamus due to obesity. Increased levels of TLR4 negatively affect insulin sensitivity [208].

5.2.9. Mitochondrial Dysfunction and ROS Formation

Insulin action is enhanced by low levels of reactive oxygen species (ROS) [209,210]. High levels of ROS may cause oxidative stress. ROS is a by-product of the electron transport chain as a consequence of mitochondrial dysfunction [211]. Changes in mitochondrial proteins due to an increased flux of metabolites into mitochondria and decreased expression of antioxidant enzymes may be due to increased levels of ROS as an effect of obesity [212,213,214,215]. Increased oxidative stress is involved in the activation of stress kinases, which phosphorylate serine in IRS proteins, inducing insulin resistance [216]. Mitochondrial fission, which causes mitochondrial dysfunction, decreases the activation of p38 MAPK, increases the activity of IRS-1 and AKT and causes insulin resistance [217]. Disturbances in mitochondrial functions may increase levels of DAG, causing activation of PKC and decreased phosphorylation of IRS-2 and activity of PI3K [218]. Mitochondrial biogenesis is controlled by nuclear genes, which encode peroxisome proliferator-activated receptor gamma (PPARγ) coactivator 1α (PGC-1α) and PGC-1β. Low expression of these genes decreases the number of mitochondria in muscle, resulting in intramyocellular fat accumulation in patients with insulin resistance [219,220]. Results obtained revealed that increased insulin resistance is associated with decreased PGC-1α beyond normal physical limits [221]. Therefore, activation of PGC-1α may be a therapeutic target for type 2 diabetes mellitus [172].

5.3. Selective Insulin Resistance

It was found that in the presence of IR, not all cellular insulin functions are less responsive. In hyperinsulinemia, there are insulin pathways which are highly responsive to insulin. This phenomenon is referred to as selective insulin resistance or pathway-selective insulin responsiveness [222]. For example, in the case of insulin resistance in the liver, insulin not does suppress hepatic glucose production; however, it does stimulate lipogenesis, causing hyperglycemia, hyperlipidemia and hepatic steatosis [222]. The mechanism associated with selective insulin resistance is not fully understood, although there are several suggestions. One hypothesis suggests that selective insulin resistance may be associated with differences in the specificities of substrates involved in AKT phosphorylation in processes of gluconeogenesis and lipogenesis [223]. According to this hypothesis, phosphorylation of AKT Ser473 may activate some AKT substrates that are associated with the process of gluconeogenesis, such as FOXO. Insulin resistance may block these activations. On the other hand, other substrates of AKT, such as glycogen synthase kinase 3 β (GSK3β) and tuberous sclerosis complex 2 (TSC2), the activation of which depends on AKT phosphorylation at Thr308, might be not disrupted, and therefore might be active [224]. Another explanation for selective insulin resistance in the liver suggests the different sensitivities of insulin-induced activation of sterol regulatory element-binding protein 1c (SREBP-1c) and suppression of gluconeogenesis. According to this hypothesis, both processes require specific insulin levels [225,226]. Selective insulin resistance in the liver, according to another hypothesis, may be due to insulin-independent lipogenesis. In the induction of lipogenesis by nutrients, carbohydrate response element-binding protein (ChREBP), the transcription factor peroxisome proliferator-activated receptor-γ (PPARγ) coactivator 1-β and liver X receptor-mediated SREBP-1c [227,228,229] are also involved. Animal studies revealed that these alternative pathways are activated by fructose and monosaccharides [230,231]. For more details, see [3].

6. Therapeutic Strategies for Insulin Resistance

Insulin resistance is associated with metabolic disturbances such as T2DM, metabolic syndrome, lipid abnormalities and hepatic derangements. There are suggested treatment strategies against insulin resistance. A sedentary lifestyle and the increased prevalence of obesity contribute to the development of insulin resistance and its consequences. Modifications to lifestyle focused on the reduction of calories and weight as well as increased physical activity are suggested as important factors against insulin resistance [3,232,233]. Suggested proposals include physical activity daily for 30 min and a suitable diet for maintaining a healthy weight, for example, the Mediterranean diet. In the case of severe hypertriglyceridemia, a very-low-fat diet may be appropriate [34]. Exercise can improve skeletal muscle insulin resistance [234]. There are several other strategies in insulin resistance prevention, such as using nature-derived compounds. Note also that there are negative results in the prevention of insulin resistance [235]: patients with diagnosed insulin resistance require insulin therapy. Patients with diagnosed insulin resistance associated with generalized lipodystrophy or mutations in insulin receptors require higher doses of insulin [236,237]. Other drugs may be used, such as insulin-like growth factor-1 in the case of patients with Donohue syndrome [238], insulin sensitizers such as metformin and thiazolidinediones in the case of patients with lipodystrophies [239] and lipid-lowering medications [240]. Etiologic therapeutic strategies, such as leptin therapy, the growth hormone releasing hormone (GHRH) analog and immunosuppressants [34] are also suggested.

7. Insulin Resistance Causes Cancer: True or False? Is Insulin Our Friend or Foe?

As mentioned earlier, insulin resistance is involved in the development of different diseases, pathologies and metabolic dysfunctions. IR is associated with hyperglycemia, metabolic syndrome, obesity, hypertriglyceridemia, nonalcoholic fatty liver disease (NAFLD), atherosclerosis, T2DM, cardiovascular disease, PCOS and so on. Insulin resistance impairs insulin secretion from pancreatic β-cells. Disturbances in insulin secretion due to insulin resistance results in compensatory insulin hypersecretion, usually at levels at which normoglycemia cannot be maintained, leading to hyperinsulinemia.
Hyperinsulinemia is a higher than usual amount of insulin in the blood. This pathology is associated with diagnoses of damaged myocardial insulin signaling, mitochondrial dysfunction, endoplasmic reticulum stress, sympathetic nervous system dysfunction, abnormalities in immune response, etc. [6]. Insulin resistance is the main cause of hyperinsulinemia. Hyperinsulinemia may be caused also by a rare tumor of pancreatic β-cells (insulinoma) as well as excessive numbers or growth of these cells (nesidioblastosis) [6]. Hyperinsulinemia is involved in increased morbidity and mortality from cardiovascular complications [241].
Insulin is proposed to be an oncogenic factor [242,243,244] and contributes to the progression of cancer. Therefore, long-lasting hyperinsulinemia due to insulin resistance may promote cancer development, stimulating different intracellular signaling pathways, enhancing cell proliferation associated with enhanced growth factors as well as affecting cell metabolism. Insulin may induce cancer growth by its mitogenic and antiapoptotic effects, as is observed in the case of breast and prostate cancers [33,245,246]. Decreasing levels of sex hormone binding globulin (SHBG) reveals a positive effect on estrogen bioavailability, causing an increased risk of breast cancer [148,247]. Physiological insulin secretion, as is observed in the normoglycemic state, and its action are critical to normal cell growth. In its physiological role in the regulation of cellular growth, insulin interacts with other mediators, such as IGF-1, IGF-2 and their receptors. Epidemiological studies revealed in patients with diagnosed insulin resistance an increased risk for several cancers, such as breast, colorectum, liver and pancreas [145,148]. It is suggested that increased levels of insulin and hyperinsulinemia are associated with the development of a number of cancers, such as colorectal, ovarian, breast and pancreatic cancer [28]. Cross-sectional and prospective studies revealed an association between hyperinsulinemia and an increased risk of breast cancer [248,249]. An association between colorectal cancer and increased serum insulin levels has been found in epidemiological studies [250,251]. Several obtained results revealed that patients with a cancer diagnosis were markedly insulin resistant [143]. In vivo and in vitro studies confirmed suggestions that insulin is involved in the growth of colon epithelial and carcinoma cells [252,253]. Epidemiological studies also revealed a link between elevated levels of insulin and an increased risk of pancreatic cancer. Results obtained in in vitro studies suggest that excessive insulin signaling may be associated with the proliferation and survival of human immortalized pancreatic ductal cells and metastatic pancreatic cancer cells [254,255,256].
As mentioned earlier, isoform B of insulin receptors is up-regulated in cancer [245]. In breast cancer epithelial cells, the insulin receptor is overexpressed up to 10 times the normal amount. This observation may indicate that overexpression of insulin receptors, mainly in the presence of hyperinsulinemia, may be associated with a selective growth advantage to breast cancer cells [28]. Disturbances in intracellular insulin signaling pathways are proposed as a mechanism of breast cancer development. It was found that insulin signaling activates PI3K/AKT/mTOR signaling pathways, which may stimulate proliferation, apoptosis resistance and invasion [257,258,259]. The PI3K/AKT/mTOR signaling pathway is involved in the regulation of central glucose metabolism and aerobic glycolysis. Aggressive cancer cells are glucose dependent, and a large amount of their energy is obtained via aerobic glycolysis, the so-called Warburg effect [260]. Aerobic glycolysis is directly associated with the aggressive biology of cancers. This effect is due to increased glycolysis and uptake of glucose, which supplies anabolic precursors for rapid growth and causes mitochondrial dysfunction, resulting in apoptosis resistance [33]. Based on results obtained, in breast cancer cells, the signaling pathway PI3K/AKT/mTOR is frequently activated [259]. Overexpression of activated AKT predicts poor prognosis in women with breast cancer [261]. There are many studies which have tested the association between the expression of phospho-AKT in breast cancer with overall survival and disease-free survival [261]. All performed studies showed that the activation of AKT signaling is an adverse prognostic factor in breast cancer [33].
Is insulin our friend or foe? The answer to this question was given to us by Paracelsus (1493–1541), Father of Medicine, who wrote: “Omnia sunt venena, nihil est sine venena. Sole dosis facit venenum”: “All substances are poisons; there is none which is not a poison. Only the dose makes the poison”. In the case of insulin, hypoinsulinemia and hyperinsulinemia are “poisons”, whereas normoinsulinemia is not “poison”. Is insulin for us friend or foe? It depends on the dose.

8. Conclusions and Perspectives

Insulin resistance, which reduces the responsiveness of cells and tissues to insulin, may be due to both genetic and environmental factors. It is associated with several pathologies and diseases, such as type 2 diabetes mellitus, polycystic ovary syndrome, glucose intolerance, hypertension, metabolic syndrome, cancers, obesity and so on. Severe insulin resistance is involved in clinical syndromes, such as Type A, Type B, Type C, HAIR-AN, Rabson–Mendenhall, leprechaunism and lipodystrophies. The prevalence of obesity and T2DM has increased dramatically worldwide, and individuals with these pathologies are at a higher risk of cancer and other diseases.
Different interactions and distinct molecular, cellular and physiological mechanisms contribute to the relationship between insulin resistance and observed metabolic dysfunctions. Unfortunately, these interactions remain unknown; however, significant progress has been made in understanding the associations between metabolic dysfunctions and severe insulin resistance. Therefore, these associations, mechanisms and interactions need further investigation. Understanding of insulin signaling and insulin resistance will stimulate new therapy designs, including gene therapies and new drugs.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Conflicts of Interest

The author declares no conflicts of interest.

References

  1. Hausler, R.A.; McGraw, T.E.; Accili, D. Biochemical and cellular properties of insulin receptor signaling. Nat. Rev. Mol. Cell Biol. 2018, 19, 31–44. [Google Scholar] [CrossRef] [PubMed]
  2. Tokarz, V.L.; MacDonald, P.E.; Klip, A. The cell biology of systemic insulin function. J. Cell Biol. 2018, 217, 2273–2289. [Google Scholar] [CrossRef] [PubMed]
  3. Lee, S.-H.; Park, S.-Y.; Choi, C.S. Insulin resistance: From mechanisms to therapeutic strategies. Diabetes Metab. J. 2022, 46, 15–37. [Google Scholar] [CrossRef] [PubMed]
  4. Boucher, J.; Kleinridders, A.; Kahn, C.R. Insulin receptor signaling in normal and insulin-resistant states. Cold Spring Harb. Perspect. Biol. 2014, 4, a009191. [Google Scholar] [CrossRef] [PubMed]
  5. Zhao, X.; Yang, C.; Sun, W.; Ji, H.; Lian, F. The crucial role of mechanism of insulin resistance in metabolic syndrome. Fron. Endocrinol. 2023, 14, 1149239. [Google Scholar] [CrossRef]
  6. Rahman, M.S.; Hossain, K.S.; Das, S.; Kundu, S.; Adegoke, E.O.; Rahman, M.A.; Hannan, M.A.; Uddin, M.J.; Pang, M.-G. Role of insulin in health and disease: An update. Int. J. Mol. Sci. 2021, 22, 6403. [Google Scholar] [CrossRef]
  7. Zimmet, P.; Albert, K.G.; Shaw, J. Global and social implications of the diabetes epidemic. Nature 2001, 414, 782–787. [Google Scholar] [CrossRef]
  8. Kahn, S.E. The relative contributions of insulin resistance and beta-cell dysfunction to the pathophysiology of type 2 diabetes. Diabetologia 2003, 46, 3–19. [Google Scholar] [CrossRef]
  9. Odegaard, J.I.; Chawla, A. Pleiotropic actions of insulin resistance and inflammation in metabolic homeostasis. Science 2013, 339, 172–177. [Google Scholar] [CrossRef]
  10. Samuel, V.T.; Shulman, G.I. Nonalcoholic fatty liver disease as a nexus of metabolic and hepatic diseases. Cell Metabol. 2018, 27, 22–41. [Google Scholar] [CrossRef]
  11. Csajbok, E.A.; Tamas, G. Cerebral cortex: A target and source of insulin? Diabetologia 2016, 59, 1609–1615. [Google Scholar] [CrossRef] [PubMed]
  12. Alarcon, C.; Lincoln, B.; Rhodes, C.I. The biosynthesis of the subtilisin-related proprotein convertase PC3, but not that of the PC2 convertase, is regulated by glucose in parallel to proinsulin biosynthesis in rat pancreatic islets. J. Biol. Chem. 1993, 268, 4276–4280. [Google Scholar] [CrossRef] [PubMed]
  13. Andrali, S.S.; Sampley, M.L.; Vanderford, N.L.; Ozcan, S. Glucose regulation of insulin gene expression in pancreatic beta-cells. Biochem. J. 2008, 415, 1–10. [Google Scholar] [CrossRef]
  14. Vasilievic, J.; Tarkko, J.M.; Knoch, K.P.; Solimena, M. The making of insulin in health and disease. Diabetologia 2020, 63, 1982–1989. [Google Scholar]
  15. Kaufman, B.A.; Li, C.; Soleimanpour, S.A. Mitochondrial regulation of beta-cell function. Maintaining the momentum for insulin resistance. Mol. Asp. Med. 2015, 42, 91–104. [Google Scholar] [CrossRef] [PubMed]
  16. Artner, I.; Stein, R. Transcriptional regulation of insulin gene expression. In Pancreatic Beta Cell in Health and Disease; Springer: Tokyo, Japan, 2008; pp. 13–30. [Google Scholar]
  17. Hutton, J.C. Insulin secretion granule biogenesis and the proinsulin-processing endopeptidases. Diabetologia 1994, 37 (Suppl. 2), S48–S56. [Google Scholar] [CrossRef] [PubMed]
  18. Li, G.; Runngerr-Brändle, E.; Just, I.; Jonas, J.C.; Aktories, K.; Wallheim, C.B. Effect of actin filaments by Clostridium botulinum C2 toxin on insulin secretion in HIT-T15 cells and pancreatic islets. Mol. Biol. Cell 1994, 5, 1199–1213. [Google Scholar] [CrossRef]
  19. Rorsman, P.; Rengström, E. Insulin granule dynamics in pancreatic beta cells. Diabetologia 2003, 46, 1029–1045. [Google Scholar] [CrossRef]
  20. Kalwat, M.A.; Thurmond, D.C. Signaling mechanisms of glucose-induced F-actin remodeling in pancreatic islet β cells. Exp. Mol. Med. 2013, 45, e37. [Google Scholar] [CrossRef]
  21. Gandasi, N.R.; Yim, P.; Riz, M.; Chibalina, M.V.; Cortese, G.; Lund, P.-E.; Matveev, V.; Rorsman, P.; Sherman, A.; Pedersen, M.G.; et al. Ca2+ channel clustering with insulin-containing granules is distributed in type 2 diabetes. J. Clin. Investig. 2017, 127, 2353–2364. [Google Scholar] [CrossRef]
  22. Wang, Z.; Thurmond, D.C. Mechanisms of biphasic insulin-granule exocytosis—Role of the cytoskeleton, small GTPases and SNARE proteins. J. Cell Sci. 2009, 122, 893–903. [Google Scholar] [CrossRef]
  23. Belfiore, A.; Malaguarnera, R.; Vella, V.; Lawrence, M.C.; Sciacca, L.; Frasca, F.; Morrione, A.; Vigneri, R. Insulin receptor isoforms in physiology and diseases: An updated view. Endocr. Rev. 2017, 38, 379–431. [Google Scholar] [CrossRef]
  24. Denley, A.; Wallace, J.C.; Cosgrove, L.J.; Forbes, B.E. The insulin receptor isoforms exon 11 (IR-A) in cancer and other diseases. A review. Horm. Metabol. Res. 2003, 35, 778–785. [Google Scholar]
  25. Escribano, O.; Beneit, N.; Rubio-Longás, C.; López-Pastor, A.R.; Gómez-Hernández, A. The role of insulin receptor isoforms in diabetes and its metabolism and vascular complications. J. Diab. Res. 2017, 2017, 1403206. [Google Scholar]
  26. Vella, V.; Milluzo, A.; Scalisi, N.M.; Vigneri, P.; Sciacca, L. Insulin receptor isoforms in cancer. Int. J. Med. 2018, 19, 3615. [Google Scholar] [CrossRef] [PubMed]
  27. Benyoucef, S.; Surinya, K.H.; Hadaschik, D.; Siddle, K. Characterization of insulin/IGF hybrid receptors: Contributions of the insulin receptor 1,2 and Fn1 domains and the alternatively spliced exon 11 sequence to ligand binding and receptor activation. Biochem. J. 2007, 403, 603–613. [Google Scholar] [CrossRef]
  28. Mantzoros, C.; Serdy, S. Insulin Action. In UpToDate Online Medical Text, Topic 1789, version 14.0; Wolters Kluwer, MEDI MEDIA: Waltham, MA, USA, 2023. [Google Scholar]
  29. Li, M.; Chi, X.; Wang, Y.; Seterrahmane, S.; Xie, W.; Xu, H. Trends in insulin resistance: Insights into mechanisms and therapeutic strategy. Signal Transduct. Target. Ther. 2022, 7, 216. [Google Scholar] [CrossRef] [PubMed]
  30. White, M.E. Regulating insulin signaling and β-cell function through IRS proteins. Can. J. Physiol. Pharmacol. 2006, 84, 725–737. [Google Scholar] [CrossRef] [PubMed]
  31. Shaw, L.M. The insulin receptor substrate (IRS) proteins: At the intersection of metabolism and cancer. Cell Cycle 2011, 10, 1750–1756. [Google Scholar] [CrossRef]
  32. Youngren, J.F. Regulation of insulin receptor function. Cell Mol. Life Sci. 2007, 64, 873–891. [Google Scholar] [CrossRef]
  33. Yee, L.D.; Mortimer, J.E.; Natarajan, R.; Dietze, E.C.; Seewaldt, V.L. Metabolic health, insulin, and breast cancer: Why oncologists should care about insulin. Front. Endocrinol. 2020, 11, 58. [Google Scholar] [CrossRef]
  34. Angeldi, A.M.; Filippaios, A.; Mantzoros, C.S. Severe insulin resistance syndromes. J. Clin. Investig. 2021, 131, e142245. [Google Scholar] [CrossRef]
  35. Vigneri, R.; Sciacca, L.; Vigneri, P. Rethinking the relationship between insulin and cancer. Trends Endocrinol. Metab. 2020, 31, 551–560. [Google Scholar] [CrossRef] [PubMed]
  36. Saltiel, A.R. Insulin signaling in health and disease. J. Clin. Investig. 2021, 131, e142241. [Google Scholar] [CrossRef] [PubMed]
  37. Sharabi, K.; Tavares, C.D.; Rines, A.G.; Puigsener, P. Molecular pathophysiology of hepatic glucose production. Mol. Asp. Med. 2015, 46, 21–33. [Google Scholar] [CrossRef] [PubMed]
  38. Tzivion, G.; Dobson, M.; Ramakrishnan, G. FoxO transcription factors: Regulation by AKT and 14-3-3 proteins. Biochim. Biophys. Acta 2011, 1813, 1938–1945. [Google Scholar] [CrossRef] [PubMed]
  39. Dong, X.C.; Coops, K.D.; Guo, S.; Li, Y.; Kollipara, R.; DePinho, R.A.; White, M.F. Inactivation of hepatic Foxo1 by insulin signaling is required for adaptative nutrient homeostasis and endocrine growth regulation. Cell Metabol. 2008, 8, 65–76. [Google Scholar] [CrossRef]
  40. Abdulla, H.; Smith, K.; Atheron, P.J.; Idris, I. Role of insulin in the regulation of human skeletal muscle protein synthesis and breakdown: A systematic review and meta-analysis. Diabetologia 2016, 59, 44–55. [Google Scholar] [CrossRef]
  41. Choi, Y.H.; Park, S.; Hockmans, S.; Zmuda-Trzebiatowska, E.; Svennelid, F.; Haluzik, M.; Gavrilova, O.; Ahmad, F.; Pepin, L.; Napolitano, M.; et al. Alterations in regulation of energy homeostasis in cyclic nucleotide phosphodiesterase 3B-null mice. J. Clin. Investig. 2006, 116, 3240–3245. [Google Scholar] [CrossRef]
  42. Woods, S.C.; Seeley, R.J.; Baskin, D.G.; Schwartz, M.W. Insulin and the blood-brain barrier. Curr. Pharm. Des. 2003, 9, 795–800. [Google Scholar] [CrossRef]
  43. Manning, B.D.; Cantley, L.C. AKT/PKB signaling: Navigating downstream. Cell 2007, 129, 1261–1274. [Google Scholar] [CrossRef] [PubMed]
  44. Avruch, J. Insulin signal transduction through protein kinase cascades. Mol. Cell Biochem. 1998, 182, 31–48. [Google Scholar] [CrossRef] [PubMed]
  45. Taniguchi, C.M.; Emanuelli, B.; Kahn, C.R. Critical nodes in signalling pathways: Insights into insulin action. Nat. Rev. Mol. Cell Biol. 2006, 7, 85–96. [Google Scholar] [CrossRef]
  46. Filippi, B.M.; Abraham, M.A.; Yue, J.T.Y.; Lam, T.K.T. Insulin and glucagon signaling in the central nervous system. Rev. Endocr. Metab. Disord. 2013, 14, 365–375. [Google Scholar] [CrossRef]
  47. Garcia-Cáceres, C.; Quarta, C.; Vareba, L.; Gao, Y.; Gruber, T.; Legutko, B.; Jastroch, M.; Johansson, P.; Ninkovic, J.; Yi, C.-X.; et al. Astrocitic insulin signaling couples brain glucose uptake with nutrient availability. Cell 2016, 166, 867–880. [Google Scholar] [CrossRef]
  48. Schwartz, M.W.; Woods, S.C.; Porte, D., Jr.; Seeley, R.J.; Baskin, D.G. Central nervous system control of food intake. Nature 2000, 404, 661–671. [Google Scholar] [CrossRef] [PubMed]
  49. Lee, S.-H.; Zabolotny, J.M.; Huang, H.; Lee, H.; Kim, Y.-B. Insulin in the nervous system in the mind: Functions in metabolism, memory, and mood. Mol. Metab. 2016, 5, 589–601. [Google Scholar] [CrossRef] [PubMed]
  50. Benedict, C.; Kern, W.; Schultes, B.; Born, J.; Hallschmidt, M. Differential sensitivity of men and women to anorexigenic and memory-improving effect of intranasal insulin. J. Clin. Endocrinol. Metab. 2008, 93, 1339–1344. [Google Scholar] [CrossRef]
  51. Havrankova, J.; Schmechel, D.; Roth, J.; Brownstein, M. Identification of insulin in rat brains. Proc. Natl. Acad. Sci. USA 1978, 75, 5737–5741. [Google Scholar] [CrossRef]
  52. Ramnanan, C.J.; Edgerton, D.S.; Cherrington, A.D. Evidence against a physiologic role for acute changes in CNS insulin action in the rapid regulation and hepatic glucose production. Cell Metab. 2012, 15, 656–664. [Google Scholar] [CrossRef]
  53. Ramnanan, C.J.; Kraft, G.; Smith, M.S.; Farmer, B.; Neal, D.; Williams, P.E.; Lautz, M.; Farmer, T.; Donohue, E.P.; Cherringtomn, A.D.; et al. Interaction between the central and peripheral effects of insulin in controlling hepatic glucose metabolism in the conscious dog. Diabetes 2013, 62, 74–84. [Google Scholar] [CrossRef] [PubMed]
  54. Asplin, C.M.; Paquette, T.L.; Palmer, J.P. In vivo inhibition of glucagon secretion by paracrine beta cell activity in man. J. Clin. Investig. 1981, 68, 314–318. [Google Scholar] [CrossRef]
  55. Singh, S.; Sharma, R.; Kumari, M.; Tiwari, S. Insulin receptors in the kidneys in health and disease. World J. Nephrol. 2019, 8, 11–22. [Google Scholar] [CrossRef]
  56. Akhtar, M.; Taha, N.M.; Nauman, A.; Mujeeb, I.B.; Al-Nabet, A. Diabetic kidney disease: Past and present. Adv. Anat. Pathol. 2020, 27, 87–97. [Google Scholar] [CrossRef] [PubMed]
  57. Gatica, R.; Bertinat, R.; Silva, P.; Carpio, D.; Ramirez, M.J.; Slebe, J.C.; San Martin, R.; Nualart, F.; Campistol, J.M.; Caelles, C.; et al. Altered expression and localization of insulin receptor in proximal tubule cells from human and rat diabetic kidneys. J. Cell Biochem. 2013, 114, 639–649. [Google Scholar] [CrossRef]
  58. Tiwari, S.; Halagappa, V.K.; Riazi, S.; Hu, X.; Ecelbarger, C.A. Reduced expression of insulin receptors in the kidneys of insulin-resistant rats. J. Am. Soc. Nephrol. 2007, 18, 2661–2671. [Google Scholar] [CrossRef]
  59. Csibi, A.; Communi, D.; Muller, N.; Bottari, S.P. Angiotensin II inhibits insulin-stimulated GLUT4 translocation and Akt activation through tyrosine nitration-dependent mechanism. PLoS ONE 2010, 5, e10070. [Google Scholar] [CrossRef] [PubMed]
  60. Malekzadeh, B.O.; Erlandsson, M.C.; Tengvall, P.; Palmquist, A.; Ransjo, M.; Bokarewa, M.I.; Westerlund, A. Effects of implant-delivered insulin on bone formation in osteoporotic rats. J. Biomed. Mater. Res. A 2018, 106, 2472–2480. [Google Scholar] [CrossRef] [PubMed]
  61. Fulzele, K.; Riddle, R.C.; DiGirolamo, D.J.; Cao, X.; Wan, C.; Chen, D.; Faugere, M.C.; Aja, S.; Hussain, M.A.; Bruning, J.C.; et al. Insulin receptor signaling in osteoblasts regulates postnatal bone acquisition and body composition. Cell 2010, 142, 309–319. [Google Scholar] [CrossRef] [PubMed]
  62. Thracilkill, K.M.; Lumpkin, C.K., Jr.; Bunn, R.C.; Kemp, S.F.; Fowlkes, J.L. Is insulin an anabolic agent in bone? Dissecting the diabetic bone for clues. Am. J. Physiol. Endocrinol. Metab. 2005, 289, E735–E745. [Google Scholar] [CrossRef]
  63. Xiao, Y.; Woo, W.M.; Nagao, K.; Li, W.; Terunuma, A.; Mukouyama, Y.S.; Oro, A.E.; Vogel, J.C.; Brownell, I. Perivascular hair follicle stem cells associated with a venule annulus. J. Investig. Dermatol. 2013, 133, 2324–2331. [Google Scholar] [CrossRef] [PubMed]
  64. Plikus, M.V.; Van Spyk, E.N.; Pham, K.; Geyfman, M.; Kumar, V.; Takahashi, J.S.; Andersen, B. The circadian clock in skin: Implications for adult stem cells, tissue regeneration, cancer, aging and immunity. J. Biol. Rhytm. 2015, 30, 163–182. [Google Scholar] [CrossRef] [PubMed]
  65. Baselga Torres, E.; Torres-Pradilla, M. Cutaneous manifestations in children with diabetes mellitus and obesity. Actas Dermosifiliogr. 2014, 105, 546–557. [Google Scholar] [CrossRef] [PubMed]
  66. Napolitano, M.; Megna, M.; Monfrecola, G. Insulin resistance and skin diseases. Sci. World J. 2015, 2015, 479354. [Google Scholar] [CrossRef] [PubMed]
  67. Kahana, M.; Grossman, E.; Feinstein, A.; Ronnen, M.; Cohen, M.; Millet, M.S. Skin tags: A cutaneous marker for diabetes mellitus. Acta Derm. Venerol. 1987, 67, 175–177. [Google Scholar] [CrossRef] [PubMed]
  68. Yadav, A.; Kataria, M.A.; Saini, V.; Yadav, A. Role of leptin and adiponectin in insulin resistance. Clin. Chim. Acta 2013, 417, 80–84. [Google Scholar] [CrossRef]
  69. Abdel Hay, R.M.; Rashed, L.A. Association between the leptin gene 2548G/A polymorphism, the plasma leptin and the metabolic syndrome with psoriasis. Exp. Dermatol. 2011, 20, 715–719. [Google Scholar] [CrossRef]
  70. Kitabachi, A.E.; Umpierrez, G.E.; Miles, J.M.; Fisher, J.N. Hyperglycemic crisis in adult patients with diabetes. Diabetes Care 2009, 32, 1335–1343. [Google Scholar] [CrossRef]
  71. Keller, U.; Gerber, P.P.; Stauffacher, W. Fatty acid-independent inhibition of hepatic ketone body production by insulin in humans. Am. J. Physiol. 1988, 254, E694–E699. [Google Scholar] [CrossRef]
  72. Keller, U.; Lustenberger, M.; Stauffacher, W. Effect of insulin on ketone body clearance studied by a ketone body “clamp” technique in normal men. Diabetologia 1988, 31, 24–29. [Google Scholar] [CrossRef]
  73. Jefferson, L.S. Lilly Lecture 1979: Role of insulin in the regulation of protein synthesis. Diabetes 1980, 29, 487–496. [Google Scholar] [CrossRef]
  74. Flakoll, P.J.; Kulaylat, M.; Frexes-Steed, M.; Hourani, H.; Brown, L.L.; Hill, J.O.; Abumrad, N.N. Amino acid augment insulin’s suppression of whole body proteolysis. Am. J. Physiol. 1989, 257, E839–E847. [Google Scholar] [CrossRef]
  75. Barrett, E.J.; Eggleston, E.M.; Inyard, A.C.; Wang, H.; Li, G.; Chai, W.; Liu, O.Z. The vascular actions of insulin control its delivery to muscle and regulate the rate-limiting step in skeletal muscle insulin action. Diabetologia 2009, 52, 752–764. [Google Scholar] [CrossRef] [PubMed]
  76. Zeng, G.; Nystrom, F.H.; Ravichandran, L.V.; Cong, L.N.; Kirby, M.; Mostowski, H.; Quon, M.J. Roles of insulin receptor, PI3-kinase, and Akt in insulin signaling pathways related to production of nitric oxide in human vascular endothelial cells. Circulation 2000, 101, 1539–1545. [Google Scholar] [CrossRef] [PubMed]
  77. Vincent, M.A.; Dawson, D.; Clark, A.D.H.; Lindner, J.R.; Rattigan, S.; Clark, M.G.; Barrett, E.J. Skeletal muscle microvascular recruitment by physiological hyperinsulinemia precedes increases in total blood flow. Diabetes 2002, 51, 42–48. [Google Scholar] [CrossRef] [PubMed]
  78. Bradley, E.A.; Richards, S.M.; Keske, M.A.; Rattigan, S. Local NOS inhibition impairs vascular and metabolic actions of insulin in rat hindleg muscle in vivo. Am. J. Physiol. Endocrinol. Metab. 2013, 305, E745–E750. [Google Scholar] [CrossRef] [PubMed]
  79. Mantzoros, C.S.; Flier, J.S. Insulin resistance: The clinical spectrum. In Advances in Endocrinology and Metabolism; Mazzaferi, E., Ed.; St. Louis Mosby-Year Book; Springer: Berlin/Heidelberg, Germany, 1995; Volume 6, pp. 193–232. [Google Scholar]
  80. Semple, R.K.; Savage, D.B.; Cohran, E.K.; Gorden, P.; O’Rahilly, S. Genetic syndromes of severe insulin resistance. Endocr. Rev. 2011, 32, 498–514. [Google Scholar] [CrossRef] [PubMed]
  81. Parker, V.E.; Semple, R.K. Genetics in endocrinology: Genetic forms of severe insulin resistance: What endocrinologists should know. Eur. J. Endocrinol. 2013, 169, R71–R80. [Google Scholar] [CrossRef] [PubMed]
  82. Wan, Z.J.; Huang, K.; Xu, B.; Hu, S.-Q.; Wang, S.; Chu, Y.-C.; Katsoyannis, P.G.; Weiss, M.A. Diabetes-associated mutation in human insulin: Crystal structure and photo-cross-linking studies of A-chain variant insulin Wakayama. Biochemistry 2005, 44, 500–516. [Google Scholar] [CrossRef] [PubMed]
  83. Ataul Islam, M.; Bhayye, S.; Adeniyi, A.A.; Soliman, M.E.S.; Pillay, T.S. Diabetes mellitus caused by mutations in human insulin: Analysis of impaired receptor binding of insulins Wakayama, Los Angeles, and Chicago using pharmacoinformatics. J. Biomol. Struct. Dyn. 2017, 35, 724–737. [Google Scholar] [CrossRef]
  84. Carroll, R.J.; Hammer, R.E.; Chan, S.J.; Swift, H.H.; Rubenstein, A.H.; Steiner, D.F. A mutant human proinsulin is secreted from islets of Langerhans in increased amounts via an upregulated pathway. Proc. Natl. Acad. Sci. USA 1988, 85, 8943–8947. [Google Scholar] [CrossRef]
  85. Hovnik, T.; Bratnič, N.; Trebusăk-Podkrajsĕk, K.; Kovač, J.; Paro, D.; Podnar, T.; Bratina, N.; Battelino, T. Severe progressive obstructive cardiomyopathy and renal tubular dysfunction in Donohue syndrome with decreased insulin receptor autophosphorylation due to a novel INSR mutation. Eur. J. Pediatr. 2013, 172, 1125–1129. [Google Scholar] [CrossRef]
  86. Zamanfa, D.; Mohamadi, F.; Moskapil, S.R. Rabson Mendenhall syndrome; a case report and review of literature. Int. J. Environ. Chem. 2020, 4, 13–19. [Google Scholar] [CrossRef]
  87. Longo, N.; Wang, Y.; Pasquali, M. Progressive decline in insulin levels in Rabson-Mendenhall syndrome. J. Clin. Endocrinol. Metab. 1999, 84, 2623–2629. [Google Scholar] [CrossRef]
  88. Gosavi, S.; Sangamesh, S.; Rao, A.S.; Patel, S.; Hodigere, V.C. Insulin, insulin everywhere: A rare case report of Rabson-Mendenhall syndrome. Cureus 2021, 13, e13126. [Google Scholar] [CrossRef]
  89. U.S. National Library of Medicine Genetics Home Reference. Type A Insulin Resistance Syndrome. 2019. Available online: https://ghr.nlm.nih.gov//condition/type-a-insulin-resistance-syndrome#statistics (accessed on 24 June 2020).
  90. Iwanishi, M.; Haruta, T.; Takata, Y.; Ishibashi, O.; Sasaoka, T.; Egawa, K.; Imamura, T.; Naitou, K.; Itazu, T.; Kobayashi, M. A mutation (Trp1193→Leu1193) in the tyrosine kinase domain of the insulin receptor associated with type A syndrome of insulin resistance. Diabetologia 1993, 36, 414–422. [Google Scholar] [CrossRef]
  91. Takasawa, K.; Tsuji-Hosokawa, A.; Takishima, S.; Wada, Y.; Nagasaki, K.; Dateki, S.; Numakura, C.; Hijikata, A.; Shirai, T.; Kashimada, K.; et al. Clinical characteristics of adolescent cases with Type A insulin resistance syndrome caused by heterozygous mutations in the β-subunit of the insulin receptor (INSR) gene. J. Diabetes 2019, 11, 46–54. [Google Scholar] [CrossRef] [PubMed]
  92. Lin, L.; Chen, C.; Fang, T.; Chen, D.; Chen, K.; Quan, H. Type A insulin resistance syndrome misdiagnosed as polycystic ovary syndrome: A case report. J. Med. Case Rep. 2019, 13, 347. [Google Scholar] [CrossRef] [PubMed]
  93. You, W.; Yang, J.; Wang, L.; Liu, Y.; Wang, W.; Zhu, L.; Wang, W.; Yang, J.; Chen, F. Case Report: A Chinese family of type A insulin resistance syndrome with diabetes mellitus, with a novel heterozygous missense mutation of the insulin receptor gene. Front. Endocrinol. 2022, 13, 895424. [Google Scholar] [CrossRef] [PubMed]
  94. Hosoe, J.; Kadowaki, H.; Miya, F.; Aizu, K.; Kawamura, T.; Miyata, I.; Satomura, K.; Ito, T.; Hara, K.; Tanaka, M.; et al. Structural basis and genotype-phenotype correlations of INSR mutations causing severe insulin resistance. Diabetes 2017, 66, 2713–2723. [Google Scholar] [CrossRef] [PubMed]
  95. Nakamura, F.; Taira, M.; Hashimoto, N.; Makino, H.; Sasaki, N. Familial type C syndrome of insulin resistance and short stature with possible autosomal dominant transmission. Endocrinology 1989, 36, 349–358. [Google Scholar] [CrossRef]
  96. Globerman, H.; Karnieli, E. Analysis of the insulin receptor gene tyrosine kinase domain in obese patients with hyperandrogenism, insulin resistance and acanthosis nigricans (type C insulin resistance). Int. J. Obes. 1998, 22, 349–353. [Google Scholar] [CrossRef]
  97. Kumakura, S.; Sakamoto, Y.; Iwamoto, Y.; Matsuda, A.; Kuzuya, T. Hyperinsulinemia, acanthosis nigricans and normal insulin binding in young women—Evidence for familial occurrence of post-binding defect in insulin action. J. Japan Diab. Soc. 1988, 31, 499–504. [Google Scholar]
  98. Moller, D.E.; Vidal-Puig, A.; Azziz, R. Severe insulin-resistance hyperandrogenic syndromes. In Contemporary Endocrinology; Azziz, R., Nestler, D., Eds.; Androgen Excess Disorders in Women; Human Press Inc.: Totova, NJ, USA, 2007; pp. 129–138. [Google Scholar]
  99. Hong, J.H.; Kim, H.J.; Park, K.S.; Ku, B.J. Paradigm shift in the management of type A insulin resistance. Ann. Transl. Med. 2018, 6 (Suppl. 2), S98. [Google Scholar] [CrossRef]
  100. Sjöholm, A.; Pereira, M.J.; Nilsson, T.; Linde, T.; Katsogiannos, P.; Saaf, J.W.; Eriksson, J.W. Type B insulin resistance syndrome in a patient with type 1 diabetes. Endocrinol. Diabetes Metab. Case Rep. 2020, 2020, 32478674. [Google Scholar] [CrossRef] [PubMed]
  101. Łebkowska, A.; Krentowska, A.; Adamska, A.; Lipińska, D.; Piasecka, B.; Kowal-Bielecka, O.; Górska, M.; Semple, R.K.; Kowalska, I. Type B insulin resistance syndrome associated with connective tissue disease and psoriasis. Endocrinol. Diabetes Metab. Case Rep. 2020, 2020, 32755965. [Google Scholar] [CrossRef] [PubMed]
  102. Martins, L.M.; Fernandes, V.O.; Dias de Carvalho, M.M.; Gadalha, D.D.; de Queiroz, P.C.; Montenegro, R.M., Jr. Type B insulin resistance syndrome: A systematic review. Arch. Endocrinol. Metab. 2020, 64, 337–348. [Google Scholar] [CrossRef] [PubMed]
  103. Caga-anan, G.; DeCastro, I.; Bates, J.T.; Sharma, M.D. Type B insulin resistance: A rare type of diabetes mellitus. AACC Clin. Case Rep. 2016, 2, e256. [Google Scholar] [CrossRef]
  104. Malek, R.; Chong, A.; Lupsa, B.C.; Lungu, A.O.; Cochran, E.K.; Soos, M.A.; Semple, R.K.; Balow, J.E.; Gorden, P. Treatment of type B insulin resistance: A novel approach to reduce insulin receptor autoantibodies. J. Clin. Endocrinol. Metab. 2010, 95, 3641–3647. [Google Scholar] [CrossRef] [PubMed]
  105. Censi, S.; Mian, C.; Betterle, C. Insulin autoimmune syndrome: From diagnosis to clinical management. Ann. Transl. Med. 2018, 6, 335. [Google Scholar] [CrossRef] [PubMed]
  106. Willard, D.L.; Stevenson, M.; Steenkamp, D. Type B insulin resistance syndrome. Curr. Opin. Endocrinol. Diabetes Obes. 2016, 4, 318–323. [Google Scholar] [CrossRef] [PubMed]
  107. Vidal-Puig, A.; Moller, D.E. Insulin resistance: Classification, prevalence, clinical manifestations, and diagnosis. In Androgen Excess Disorders in Women; Azziz, R., Nestler, J.E., Dewailly, D., Eds.; Lippincott Raven: Philadelphia, PA, USA, 1997; pp. 227–236. [Google Scholar]
  108. Brown, R.J.; Araujo-Vilar, D.; To Cheung, P.; Dunger, D.; Garg, A.; Jack, M.; Mungai, L.; Oral, E.A.; Patni, N.; Rother, K.I.; et al. The diagnosis and management of lipodystrophy syndromes: Multi-society practice guidelines. J. Clin. Endocrinol. Metab. 2016, 12, 4500–4511. [Google Scholar] [CrossRef] [PubMed]
  109. Robbins, A.L.; Savage, D.B. The genetics of lipid storage and human lipodystrophies. Trends. Mol. Med. 2015, 7, 433–438. [Google Scholar] [CrossRef] [PubMed]
  110. Chiquette, E.; Oral, E.A.; Garg, A.; Vilar, D.-A.; Dhankhar, P. Estimating the prevalence of generalized and partial lipodystrophy: Findings and challenges. Diabetes Metab. Syndr. Obes. 2017, 10, 375–383. [Google Scholar] [CrossRef]
  111. Tritos, N.A.; Mantzoros, C.S. Syndromes of severe insulin resistance. J. Clin. Endocrinol. Metab. 1998, 83, 3025–3030. [Google Scholar] [CrossRef]
  112. Uotani, S.; Yamaguchi, Y.; Yokata, A.; Yamasaki, H.; Takino, H.; Chikuba, N.; Goto, Y.; Fujishima, N.; Yano, M.; Matsumoto, K.; et al. Molecular analysis of insulin receptor gene in Werner’s syndrome. Diabetes Res. Clin. Pract. 1994, 26, 175–176. [Google Scholar] [CrossRef]
  113. Nolan, C.J.; Ruderman, N.B.; Kahn, S.E.; Pedersen, O.; Prentki, M. Insulin resistance as a physiological defense against metabolic stress: Implications for the management of subset of type 2 diabetes. Diabetes 2015, 64, 673–686. [Google Scholar] [CrossRef]
  114. Taegtmeyer, H.; Beauloye, C.; Harmancey, R.; Hue, L. Insulin resistance protects the heart from fuel overload in dysregulated metabolic states. Am. J. Physiol. Heart C 2013, 305, H1693–H1697. [Google Scholar] [CrossRef]
  115. Nolan, C.J.; Ruderman, N.B.; Prentki, M. Intensive insulin for type 2 diabetes: The risk of causing harm. Lancet Diabetes Endocrinol. 2013, 1, 9–10. [Google Scholar] [CrossRef]
  116. Nolan, C.J.; Prentki, M. Insulin resistance and insulin hypersecretion in the metabolic syndrome and type 2 diabetes: Time for a conceptual framework shift. Diabetes Vasc. Dis. Res. 2019, 16, 118–127. [Google Scholar] [CrossRef]
  117. Petersen, M.C.; Shulman, G.I. Mechanisms of insulin action and insulin resistance. Physiol. Rev. 2018, 98, 2133–2223. [Google Scholar] [CrossRef] [PubMed]
  118. DeFronzo, R.A.; Tripathy, D. Skeletal muscle insulin resistance is the primary defect in type 2 diabetes. Diabetes Care 2009, 32 (Suppl. 2), S157–S163. [Google Scholar] [CrossRef] [PubMed]
  119. Park, S.W.; Goodpaster, B.H.; Lee, J.S.; Kuller, L.H.; Boudreau, R.; de Rekeneire, N.; Harris, T.B.; Kritchevsky, S.; Tylavsky, F.A.; Nevit, M.; et al. Excessive loss of skeletal muscle mass in older adults with type 2 diabetes. Diabetes Care 2009, 32, 1993–1997. [Google Scholar] [CrossRef]
  120. Baron, A.D.; Brechtel, G.; Wallace, P.; Edelman, S.V. Rates and tissue sites of non-insulin- and insulin-mediated glucose uptake in humans. Am. J. Physiol. 1988, 255 Pt 1, E769–E774. [Google Scholar] [CrossRef]
  121. Kim, J.K.; Michael, M.D.; Previs, S.F.; Peroni, O.D.; Mauvais-Jarvis, F.; Neschen, S.; Kahn, B.B.; Kahn, C.R.; Shulman, G.I. Redistribution of substrates to adipose tissue promotes obesity in mice with selective insulin resistance in muscle. J. Clin. Investig. 2000, 105, 1791–1797. [Google Scholar] [CrossRef]
  122. Kim, J.K.; Zisman, A.; Fillmore, J.J.; Peroni, O.D.; Kotani, K.; Perret, P.; Zong, H.; Dong, J.; Kahn, C.R.; Kahn, B.B. Glucose toxicity and the development of diabetes in mice with muscle-specific inactivation of GLUT4. J. Clin. Investig. 2001, 108, 153–160. [Google Scholar] [CrossRef]
  123. O’Neill, H.M.; Maarbjerg, S.J.; Crane, J.D.; Jeppesen, J.; Jørgensen, S.B.; Schertzer, J.D.; Shyroka, O.; Kiens, B.; van Denderen, B.J.; Tarnopolsky, M.A.; et al. AMP-activated protein kinase (AMPK) β1β2 muscle null mice revealed as essential role for AMPK in maintaining mitochondrial content and glucose uptake during exercise. Proc. Natl. Acad. Sci. USA 2011, 108, 16092–16097. [Google Scholar] [CrossRef]
  124. Le Marchand-Brustel, Y.; Gremeaux, T.; Ballotti, R.; Van Obberghen, E. Insulin receptor tyrosine kinase is defective in skeletal muscle of insulin-resistant obese mice. Nature 1985, 315, 676–679. [Google Scholar] [CrossRef]
  125. Frojdo, S.; Vidal, H.; Pirola, L. Alterations of insulin signaling in type 2 diabetes: A review of the current evidence from humans. Biochim. Biophys. Acta 2009, 1792, 83–92. [Google Scholar] [CrossRef]
  126. Wang, X.; Hu, Z.; Hu, J.; Du, J.; Mitch, W.E. Insulin resistance accelerates muscle protein degradation: Activation of the ubiquitin-proteasome pathway by defects in muscle cell signaling. Endocrinology 2006, 147, 4160–4168. [Google Scholar] [CrossRef] [PubMed]
  127. Shou, J.; Chen, P.-J.; Xiao, W.-H. Mechanism of increased risk of insulin resistance in aging skeletal muscle. Diabetol. Metabol. Syndr. 2020, 12, 14. [Google Scholar] [CrossRef] [PubMed]
  128. Huffman, D.M.; Barzilai, N. Role of visceral adipose tissue in aging. BBA Gen. Subj. 2009, 1790, 1117–1123. [Google Scholar] [CrossRef] [PubMed]
  129. Meneilly, G.S.; Elliott, T.; Tessier, D.; Hards, L.; Tildesley, H. NIDDM in the elderly. Diabetes Care 1996, 19, 1320–1325. [Google Scholar] [CrossRef] [PubMed]
  130. da Costa, J.P.; Vitorino, R.; Silva, G.M.; Vogel, C.; Duarte, A.C.; Rocha-Santos, T.A. A synopsis, on aging theories, mechanisms and future prospects. Ageing Res. Rev. 2016, 29, 90–112. [Google Scholar] [CrossRef] [PubMed]
  131. Petersen, K.F.; Morino, K.; Alves, T.C.; Kibbey, R.G.; Dufour, S.; Sano, S.; Yoo, P.S.; Cline, G.E.; Shulman, G.I. Effect of aging on muscle mitochondrial substrate utilization in humans. Proc. Natl. Acad. Sci. USA 2017, 112, 11300–11334. [Google Scholar] [CrossRef] [PubMed]
  132. Dagdeviren, S.; Jung, D.Y.; Friedline, R.H.; Noh, H.L.; Kim, J.H.; Patel, P.R.; Tsitsilianos, N.; Inashima, K.; Tran, D.A.; Hu, X.; et al. IL-10 prevents aging-associated inflammation and insulin resistance in skeletal muscle. FASEB J. 2017, 31, 701–710. [Google Scholar] [CrossRef]
  133. Short, K.R.; Bigelow, M.L.; Kahl, J.; Singh, R.; Coenen-Schimke, J.; Raghavakaimal, S.; Nair, K.S. Decline in skeletal muscle mitochondrial function with aging in humans. Proc. Natl. Acad. Sci. USA 2005, 102, 5618–5623. [Google Scholar] [CrossRef]
  134. Lewis, G.F.; Carpentier, A.C.; Pereira, S.; Hahn, M.; Giacca, A. Direct and indirect control of hepatic glucose production by insulin. Cell Metab. 2021, 33, 709–720. [Google Scholar] [CrossRef]
  135. Magnusson, I.; Rothman, D.L.; Katz, L.D.; Shulman, R.G.; Shulman, G.I. Increased rate of gluconeogenesis in type II diabetes mellitus. A 13C nuclear magnetic resonance study. J. Clin. Investig. 1992, 90, 1323–1327. [Google Scholar] [CrossRef]
  136. Krssak, M.; Brehm, A.; Bernroider, E.; Anderwald, C.; Nowotny, P.; Man, C.D.; Cobelli, C.; Cline, G.W.; Shulman, G.I.; Waldhäusl, W.; et al. Alterations in postprandial hepatic glycogen metabolism in Iype 2 diabetes. Diabetes 2004, 53, 3048–3056. [Google Scholar] [CrossRef]
  137. Basu, R.; Chandramouli, V.; Dicke, B.; Landau, B.; Rizza, R. Obesity and type 2 diabetes impair insulin-induced suppression of glycogenolysis as well as gluconeogenesis. Diabetes 2005, 54, 1942–1948. [Google Scholar] [CrossRef] [PubMed]
  138. Baron, A.D.; Clark, M.G. Role of blood flow in the regulation of muscle glucose uptake. Annu. Rev. Nutr. 1997, 17, 487–499. [Google Scholar] [CrossRef]
  139. Potenza, M.A.; Marasciulo, F.L.; Chieppa, D.M.; Brigiani, G.S.; Formoso, G.; Quon, M.J.; Montagnani, M. Insulin resistance in spontaneously hypertensive rats is associated with endothelial dysfunction characterized by imbalance between NO and ET-1 production. Am. J. Physiol. 2005, 289, H813–H822. [Google Scholar] [CrossRef] [PubMed]
  140. Zeng, G.; Quon, M.J. Insulin-stimulated production of nitric oxide is inhibited by wortmannin: Direct measurement in vascular endothelial cells. J. Clin. Investig. 1996, 98, 894–898. [Google Scholar] [CrossRef] [PubMed]
  141. Montagnani, M.; Ravichandran, I.V.; Chen, H.; Esposito, D.L.; Quon, M.J. Insulin receptor substrat-1 and phosphoinositide-kinase-1 are required for insulin-stimulating productoion of nitric oxide in endothelial cells. Mol. Endocrinol. 2002, 16, 1931–1942. [Google Scholar] [CrossRef] [PubMed]
  142. Kim, J.-A.; Montagnani, M.; Koh, K.K.; Quon, M.J. Reciprocal relationship between insulin resistance and endothelial dysfunction. Molecular and pathophysiological mechanisms. Circulation 2006, 113, 1888–1904. [Google Scholar] [CrossRef] [PubMed]
  143. Mărmol, J.M.; Carlsson, M.; Raun, S.H.; Grand, M.K.; Sørensen, J.; Lang Lehrskov, L.; Richter, E.A.; Norgaard, O.; Sylow, L. Insulin resistance in patients with cancer: A systematic review and meta-analysis. Acta Oncol. 2023, 62, 364–371. [Google Scholar] [CrossRef]
  144. Esposito, K.; Chiodini, P.; Colao, A.; Lenzi, A.; Giugliano, D. Metabolic syndrome and risk of cancer: A systematic review and meta-analysis. Diabetes Care 2012, 35, 2402–2411. [Google Scholar] [CrossRef]
  145. Cowey, S.; Hardy, R.W. The metabolic syndrome: A high risk state of cancer? Am. J. Pathol. 2006, 169, 1505–1522. [Google Scholar] [CrossRef]
  146. Kundaktepe, B.P.; Durmus, S.; Cengiz, M.; Kundaktepe, F.O.; Sozer, V.; Papila, C.; Gelisgen, R.; Uzun, H. The significance of insulin resistance in nondiabetic breast cancer patients. J. Endocrinol. Metab. 2021, 11, 42–48. [Google Scholar] [CrossRef]
  147. Chen, X.; Liang, H.; Song, Q.; Xu, X.; Cao, D. Insulin promotes progression of colon cancer by upregulation of ACAT1. Lipids Health Dis. 2018, 17, 122. [Google Scholar] [CrossRef] [PubMed]
  148. Arcidiakono, B.; Iiritano, S.; Nocera, A.; Possidente, K.; Nevolo, M.T.; Ventura, V.; Foti, D.; Chiefari, E.; Brunetti, A. Insulin resistance and cancer risk: An overview of the pathogenic mechanisms. Exp. Diabetes Res. 2012, 2012, 789174. [Google Scholar] [CrossRef] [PubMed]
  149. Webster, N.J.; Resnik, J.L.; Reichart, D.B.; Struss, B.; Haas, M.; Seely, B.L. Repression of the insulin receptor promoter by the tumor suppressor gene product p53: A possible mechanism for receptor overexpression in breast cancer. Cancer Res. 1996, 56, 2781–2788. [Google Scholar]
  150. Werner, H.; Maor, S. The insulin-like growth factor-1 receptor gene: A downstream target for oncogene and tumor suppressor action. Trends Endocrinol. Metab. 2006, 17, 236–242. [Google Scholar] [CrossRef] [PubMed]
  151. Gallhager, E.J.; LeRoith, D. Hyperinsulinemia in cancer. Nat. Rev. Cancer 2020, 20, 629–644. [Google Scholar] [CrossRef]
  152. Madsen, R.R.; Vanhaese, B.; Semple, R.K. Cancer-associated PBK mutations in overgrowth disorder. Trends Mol. Med. 2018, 24, 856–870. [Google Scholar] [CrossRef]
  153. Fernandez-Medarde, A.; Santos, E. Ras in cancer and developmental diseases. Genes Cancer 2011, 2, 344–358. [Google Scholar] [CrossRef]
  154. Vatseba, T. Study of insulin resistance in patients with cancer. Arch. Clin. Med. 2020, 26, 15–19. [Google Scholar] [CrossRef]
  155. Sędzikowska, A.; Szablewski, L. Insulin and insulin resistance in Alzheimer’s disease. Int. J. Mol. Sci. 2021, 22, 9987. [Google Scholar] [CrossRef]
  156. Griffith, C.M.; Eid, T.; Rose, G.M.; Patrylo, P.R. Evidence for altered insulin receptor signaling in Alzheimer’s disease. Neuropharmacology 2018, 136, 202–215. [Google Scholar] [CrossRef]
  157. Kleinridders, A.; Ferris, H.A.; Cai, W.; Kahn, C.R. Insulin action in brain regulates systemic metabolism and brain function. Diabetes 2014, 63, 2232–2243. [Google Scholar] [CrossRef]
  158. Randle, P.J.; Garland, P.B.; Hales, C.N.; Newsholme, E.A. The glucose fatty-acid cycle. Its role in insulin sensitivity and the metabolic disturbances of diabetes mellitus. Lancet 1963, 1, 785–789. [Google Scholar] [CrossRef] [PubMed]
  159. Schwartzman, L.I.; Brown, J. Glucose inhibition of fatty acid oxidation by rat diaphragm. Am. J. Physiol. 1960, 199, 235–237. [Google Scholar] [CrossRef]
  160. Christe, M.; Rodgers, R.L. Cardiac glucose and fatty acid oxidation in the streptozotocin-induced diabetic spontaneously hypersensitive rat. Hypertension 1995, 25, 235–241. [Google Scholar] [CrossRef]
  161. Cline, G.W.; Petersen, K.F.; Krssak, M.; Shen, J.; Hundal, R.S.; Trajanoski, Z.; Inzucchi, S.; Dresner, A.; Rothman, D.L.; Shulman, G.I. Impaired glucose transport as a cause of decreased insulin-stimulated muscle glycogen synthesis in type 2 diabetes. N. Engl. J. Med. 1999, 341, 240–246. [Google Scholar] [CrossRef]
  162. Marshall, S.; Bacote, V.; Traxinger, R.R. Discovery of a metabolic pathway mediating glucose-induced desensitization of the glucose transport system. Role of hexosamine biosynthesis in the induction of insulin resistance. J. Biol. Chem. 1991, 266, 4706–4712. [Google Scholar] [CrossRef] [PubMed]
  163. Hawkins, M.; Barzilai, N.; Liu, R.; Hu, M.; Chen, W.; Rossetti, L. Role of the glucosamine pathway in fat-induced insulin resistance. J. Clin. Investig. 1997, 99, 2173–2182. [Google Scholar] [CrossRef] [PubMed]
  164. Akimoto, Y.; Hart, G.W.; Wells, L.; Vosseller, K.; Yamamoto, K.; Munetomo, E.; Ohara-Imaizumi, M.; Nishiwaki, C.; Nagamatsu, S.; Hirano, H.; et al. Elevation of the post-translational modification of the glucose-stimulated insulin secretion in the pancreas of diabetic Goto-Kakizaki rats. Glycobiology 2007, 17, 127–140. [Google Scholar] [CrossRef]
  165. Akimoto, Y.; Kawakami, H.; Yamamoto, K.; Munetomo, E.; Hida, T.; Hirano, H. Elevated expression of O-GlcNAc-modified proteins and O-GlcNAc transferase in corneas of diabetic Goto-Kakizaki rats. Investig. Ophtalmol. Vis. Sci. 2003, 44, 3802–3809. [Google Scholar] [CrossRef]
  166. Marshall, S.; Bacote, V.; Traxinger, R.R. Complete inhibition of glucose-induced desensitization of the glucose transport system by inhibitors of mRNA synthesis. Evidence for rapid turnover of glutamine:fructose-6-phosphate aminotransferase. J. Biol. Chem. 1991, 266, 10155–10161. [Google Scholar] [CrossRef]
  167. Copeland, R.J.; Bullen, J.; Hart, G.W. Cross-talk between GlcNAcylation and phosphorylation: Roles of insulin resistance and glucose toxicity. Am. J. Physiol. Endocrinol. Metab. 2008, 295, E17–E28. [Google Scholar] [CrossRef] [PubMed]
  168. Ball, L.E.; Berkaw, M.N.; Buse, M.G. Identification of the major site of O-linked β-N-acetylglucosamine modification in the C terminus of insulin receptor substrate-1. Mol. Cell Proteom. 2006, 5, 313–323. [Google Scholar] [CrossRef] [PubMed]
  169. Kang, E.S.; Han, D.; Park, J.; Kwak, T.K.; Oh, M.A.; Lee, S.A.; Choi, S.; Park, Z.Y.; Kim, Y.; Lee, J. O-GlcNAc modulation at Akt1 Ser473 correlates with apoptosis of murine pancreatic beta cells. Exp. Cell Res. 2008, 314, 2238–2248. [Google Scholar] [CrossRef] [PubMed]
  170. Housley, M.P.; Rodgers, J.T.; Udeshi, N.D.; Kelly, T.J.; Shabanowitz, J.; Hunt, D.F.; Puigserver, P.; Hart, G.W. O-GlcNAc regulates FoxO activation in response to glucose. J. Biol. Chem. 2008, 283, 16283–16292. [Google Scholar] [CrossRef]
  171. Chen, G.; Liu, P.; Thurmond, D.C.; Elmendorf, J.S. Glucosamine-induced insulin resistance is coupled to O-linked glycosylation of Munc 18c. FEBS Lett. 2003, 534, 54–60. [Google Scholar] [CrossRef]
  172. Wondmkun, Y.T. Obesity, insulin resistance, and type 2 diabetes: Associations and therapeutic implications. Diabet. Metab. Syndr. Obes. Targets Ther. 2020, 13, 3611–3616. [Google Scholar] [CrossRef]
  173. Eshima, H.; Tamura, Y.; Kakehi, S.; Kurebayashi, N.; Murayama, T.; Nakamura, K.; Kakigi, R.; Okada, T.; Sakurai, T.; Kawamori, R.; et al. Long-term, but not short-term high-fat diet induces fiber composition changes and impaired contractile force in mouse fast-twitch skeletal muscle. Physiol. Rep. 2017, 5, e13250. [Google Scholar] [CrossRef]
  174. Badin, M.; Louche, K.; Mairal, A.; Liebisch, G.; Schmitz, G.; Rustan, A.C.; Smith, S.R.; Langin, D.; Moro, C. Altered skeletal muscle lipase expression and activity contribute to insulin resistance in humans. Diabetes 2011, 60, 1734–1742. [Google Scholar] [CrossRef]
  175. Mann, J.P.; Savage, D.B. What lipodystrophies teach us about the metabolic syndrome. J. Clin. Investig. 2019, 129, 4009–4021. [Google Scholar] [CrossRef]
  176. Fabbrini, E.; Magkos, F.; Mohammed, B.S.; Pietka, T.; Abumrad, N.A.; Patterson, B.W.; Okunade, A.; Klein, S. Intrahepatic fat, not visceral fat, is linked with metabolic complications of obesity. Proc. Natl. Acad. Sci. USA 2009, 106, 15430–15435. [Google Scholar] [CrossRef]
  177. Krssak, M.; Falk Petersen, K.; Dresner, A.; DiPietro, L.; Vogel, S.M.; Rothman, D.L.; Roden, M.; Shulman, G.I. Intramyocellular lipid concentrations are correlated with insulin sensitivity in humans: A 1H NMR spectroscopy study. Diabetologia 1999, 42, 113–116. [Google Scholar] [CrossRef]
  178. Samuel, V.T.; Liu, Z.X.; Qu, X.; Elder, B.D.; Bilz, S.; Befroy, D.; Romanelli, A.J.; Shulman, G.I. Mechanism of hepatic insulin resistance in non-alcoholic fatty liver disease. J. Biol. Chem. 2004, 279, 32345–32353. [Google Scholar] [CrossRef]
  179. Goudriaan, J.R.; Dahlmans, C.E.H.; Teusink, B.; Ouwers, M.; Fabbriao, M.; Maassen, J.A.; Romijn, J.A.; Havekes, L.M.; Voshol, P.J. CD36 deficiency increases insulin sensitivity in muscle, but induces insulin resistance in the liver in mice. J. Lipid Res. 2003, 44, P2270–P2277. [Google Scholar] [CrossRef] [PubMed]
  180. Kim, J.K.; Gimeno, R.E.; Higoshimori, T.; Kim, H.-J.; Choi, H.; Punreddy, S.; Mozell, R.L.; Tan, C.; Stricker-Krongrad, A.; Hirsch, D.J.; et al. Inactivation of fatty acid transport protein 1 prevents fat-induced insulin resistance in skeletal muscle. J. Clin. Investig. 2004, 113, 756–763. [Google Scholar] [CrossRef] [PubMed]
  181. Doege, H.; Grimm, D.; Falcon, A.; Tsang, B.; Storm, T.A.; Xu, H.; Ortegon, A.M.; Kazantzis, M.; Kay, M.A.; Stahl, A. Silencing of hepatic fatty acid transporter protein 5 in vivo reverses diet-induced non-alcoholic fatty liver disease and improves hyperglycemia. J. Biol. Chem. 2008, 283, 22186–22192. [Google Scholar] [CrossRef] [PubMed]
  182. Falcon, A.; Doege, H.; Fliitt, A.; Tsang, B.; Watson, N.; Kay, M.A.; Stahl, A. FATP2 is a hepatic fatty acid transporter and proximal very long-chain acyl-CoA synthase. Am. J. Physiol. Endocrinol. Metab. 2010, 299, E384–E393. [Google Scholar] [CrossRef] [PubMed]
  183. Considine, R.V.; Nyce, M.R.; Allen, L.E.; Morales, L.M.; Triester, S.; Serrano, J.; Colberg, J.; Lanza-Jacoby, S.; Caro, J.F. Protein kinase C is increased in the liver of humans and rats with non-insulin-dependent diabetes mellitus: An alteration not due to hyperglycemia. J. Clin. Investig. 1995, 95, 2938–2944. [Google Scholar] [CrossRef] [PubMed]
  184. Karasik, A.; Rothenberg, P.L.; Yamada, K.; White, M.F.; Kahn, C.R. Increased protein kinase C activity is linked to reduced insulin receptor autophosphorylation in liver of starved rats. J. Biol. Chem. 1990, 265, 10226–10231. [Google Scholar] [CrossRef] [PubMed]
  185. Petersen, M.C.; Madiraju, A.K.; Gassaway, B.M.; Marcel, M.; Nasiri, A.R.; Butrico, G.; Marcucci, M.J.; Zhang, D.; Abulizi, A.; Zhang, X.-M.; et al. Insulin receptor thr1160 phosphorylation mediates lipid-induced hepatic insulin resistance. J. Clin. Investig. 2016, 126, 4361–4371. [Google Scholar] [CrossRef] [PubMed]
  186. Samuel, V.T.; Liu, Z.X.; Wang, A.; Beddow, S.A.; Geisler, J.G.; Kahn, M.; Zhang, X.-m.; Monia, B.P.; Bhanot, S.; Shulman, G.I. Inhibition of protein kinase Cepsilon prevents hepatic insulin resistance in nonalcoholic fatty liver disease. J. Clin. Investig. 2007, 117, 739–745. [Google Scholar] [CrossRef]
  187. Schmitz-Peiffer, C.; Browne, C.L.; Oakes, N.D.; Watkinson, A.; Chisholm, D.J.; Kraegen, E.W.; Biden, T.J. Alterations in the expression and cellular localization of proteins kinase C isoenzymes epsilon and theta are associated with insulin resistance in skeletal muscle of the high-fat-fed rat. Diabetes 1997, 46, 169–178. [Google Scholar] [CrossRef] [PubMed]
  188. Yu, C.; Chen, Y.; Cline, G.W.; Zhang, D.; Zong, H.; Wang, Y.; Bergeron, R.; Kim, J.K.; Cushman, S.W.; Cooney, G.J.; et al. Mechanism by which fatty acids inhibit insulin activation of insulin receptor substrate-1 (IRS-1)-associated phosphatidylinositol 3-kinase activity in muscle. J. Biol. Chem. 2002, 277, 50230–50236. [Google Scholar] [CrossRef] [PubMed]
  189. Li, Y.; Soos, T.J.; Li, X.; Wu, J.; Degennaro, M.; Sun, X.; Littman, D.R.; Birnbaum, M.J.; Polakiewicz, R.D. Protein kinase C Theta inhibits insulin signaling by phosphorylating IRS1 at Ser(1101). J. Biol. Chem. 2004, 279, 45304–45307. [Google Scholar] [CrossRef] [PubMed]
  190. Szendroedi, J.; Yoshimura, T.; Phielix, E.; Koliaki, C.; Marcucci, M.; Zhang, D.; Jelenik, T.; Müller, J.; Herder, C.; Nowotny, P.; et al. Role of diacylglycerol activation of PKCθ in lipid-induced muscle insulin resistance in humans. Proc. Natl. Acad. Sci. USA 2014, 111, 9597–9602. [Google Scholar] [CrossRef] [PubMed]
  191. Avignon, A.; Yamada, K.; Zhou, X.; Spencer, B.; Cardona, O.; Saba-Sidolique, S.; Galloway, M.l.; Standaert, M.L.; Farese, R.V. Chronic activation of protein kinase C in soleus muscles and other tissues of insulin-resistant type II diabetic Goto-Kakizaki (GK), obese/aged, and obese/Zucker rats: A mechanism for inhibiting glycogen synthesis. Diabetes 1996, 45, 1396–1404. [Google Scholar] [CrossRef]
  192. Chavez, J.A.; Summers, S.A. A ceramide—Centric view of insulin resistance. Cell Metab. 2012, 15, 585–594. [Google Scholar] [CrossRef]
  193. Powell, D.J.; Hajduch, E.; Kular, G.; Hundal, H.S. Ceramide disables 3-phosphoinositide binding to the pleckstrin homology domain of protein kinase B (PKB)/Aky by a PKCζ-dependent mechanism. Mol. Cell Biol. 2003, 23, 7794–7808. [Google Scholar] [CrossRef]
  194. Stratford, S.; Hoehn, K.L.; Liu, F.; Summers, S.A. Regulation of insulin action by ceramide: Dual mechanisms linking ceramide accumulation to the inhibition of Akt/protein kinase B. J. Biol. Chem. 2004, 279, 36608–36615. [Google Scholar] [CrossRef]
  195. Westwick, J.K.; Bielawska, A.E.; Dbaibo, G.; Hannun, Y.A.; Brenner, D.A. Ceramide activates the stress-activated protein kinases. J. Biol. Chem. 1995, 270, 22689–22692. [Google Scholar] [CrossRef]
  196. Holland, W.L.; Brozinick, J.T.; Wang, L.P.; Hawkins, E.D.; Sargent, K.M.; Liu, Y.; Narra, K.; Hoehn, K.L.; Knotts, T.A.; Siesky, A.; et al. Inhibition of ceramide synthesis ameliorates glucocorticoid-, saturated-fat, and obesity-induced insulin resistance. Cell Metab. 2007, 5, 167–179. [Google Scholar] [CrossRef] [PubMed]
  197. Okada, T.; Yoshida, H.; Akazawa, R.; Negishi, M.; Mori, K. Distinct role of activating transcription factor 6 (ATF6) and double-stranded RNA-activated protein kinase-like endoplasmic reticulum kinase (PERK) in transcription during the mammalian unfolded protein response. Biochem. J. 2002, 366 Pt 2, 585–594. [Google Scholar] [CrossRef]
  198. Laybutt, D.R.; Preston, A.M.; Akerfeldt, M.C.; Kench, J.G.; Busch, A.K.; Biankin, A.V.; Biden, T.J. Endoplasmic reticulum stress contributes to beta cell apoptosis in type 2 diabetes. Diabetologia 2007, 50, 752–763. [Google Scholar] [CrossRef]
  199. Marchetti, P.; Bugliani, M.; Lupi, R.; Marselli, L.; Masini, M.; Boggi, U.; Filipponi, F.; Weir, G.C.; Eizirik, D.L.; Cnop, M. The endoplasmic reticulum in pancreatic beta cells of type 2 diabetes patients. Diabetologia 2007, 50, 2486–2494. [Google Scholar] [CrossRef]
  200. Osborn, O.; Olefsky, J.M. The cellular and signaling networks linking the immune system and metabolism in disease. Nat. Med. 2012, 18, 363–374. [Google Scholar] [CrossRef] [PubMed]
  201. Sun, K.; Kusminsky, C.M.; Scherer, P.E. Adipose tissue remodeling and obesity. J. Clin. Investig. 2011, 121, 2094–2102. [Google Scholar] [CrossRef] [PubMed]
  202. Kamei, N.; Tobe, K.; Suzuki, R.; Ohsugi, M.; Watanabe, T.; Kubota, N.; Ohtsuka-Kowatari, N.; Kumagai, K.; Sakamoto, K.; Kobayashi, M.; et al. Overexpression of monocyte chemoattractant protein-1 in adipose tissues causes macrophage recruitment and insulin resistance. J. Biol. Chem. 2006, 281, 26602–26614. [Google Scholar] [CrossRef] [PubMed]
  203. Kanda, H.; Tateya, S.; Tamori, Y.; Kotani, K.; Hiasa, K.; Kitazawa, R.; Miyachi, H.; Maeda, S.; Egashira, K.; Kasuga, M. MCP-1 contributes to macrophage infiltration into adipose tissue, insulin resistance, and hepatic steatosis in obesity. J. Clin. Investig. 2006, 116, 1494–1505. [Google Scholar] [CrossRef] [PubMed]
  204. Fan, Y.; Yu, Y.; Shi, Y.; Sun, W.; Xie, M.; Ge, N.; Mao, R.; Chang, A.; Chang, A.; Xu, G.; et al. Lysine 63-linked polyubiquination of TAK1 at lysine 158 is required for tumor necrosis factor-α- and interleukin-1/3-induced IKK/NF-κB and JNK/AP-1 activation. J. Biol. Chem. 2010, 285, 5347–5360. [Google Scholar] [CrossRef] [PubMed]
  205. Zhang, J.; Gao, Z.; Yin, J.; Quon, M.J.; Ye, J. S6K directly phosphorylates IRS-1 on Ser-270 to promote insulin resistance in response to TNF-α signaling through IKK2. J. Biol. Chem. 2008, 283, 35375–35382. [Google Scholar] [CrossRef]
  206. Jager, J.; Gremeaux, T.; Cormont, M.; Le Marchand-Brustel, Y.; Tanti, J.E. Interleukin-1β-induced insulin resistance in adipocytes through down-regulation of insulin receptor substrate-1 expression. Endocrinology 2007, 148, 241–251. [Google Scholar] [CrossRef]
  207. Steppan, C.M.; Wang, J.; Whiteman, E.L.; Birnbaum, M.J.; Lazar, M.A. Activation of SOCS-3 by resistin. Mol. Cell Biol. 2005, 25, 1569–1575. [Google Scholar] [CrossRef] [PubMed]
  208. Morgensen, T.H. Pathogen recognition and inflammation signaling in innate immune defects. Clin. Microbiol. Rev. 2009, 22, 240–273. [Google Scholar] [CrossRef] [PubMed]
  209. Mahadev, K.; Zilbering, A.; Zhu, L.; Goldstein, B.J. Insulin-stimulated hydrogen peroxide reversibly inhibits protein-tyrosine phosphate 1b in vivo and enhances the early insulin action cascade. J. Biol. Chem. 2001, 276, 21938–21942. [Google Scholar] [CrossRef]
  210. Krieger-Bauer, H.I.; Kather, H. Human fat cells possess a plasma membrane-bound H2O2-generating system that is activated by insulin via a mechanism bypassing the receptor kinase. J. Clin. Investig. 1992, 89, 1006–1013. [Google Scholar] [CrossRef] [PubMed]
  211. Chang, Y.C.; Chuang, L.M. The role of oxidative stress in the pathogenesis of type 2 diabetes: From molecular mechanism to clinical implication. Am. J. Transl. Res. 2010, 2, 316–331. [Google Scholar]
  212. Fridlyand, L.E.; Philipson, L.H. Reactive species and early manifestation of insulin resistance in type 2 diabetes. Diabetes Obes. Metab. 2006, 8, 136–145. [Google Scholar] [CrossRef]
  213. Evans, J.L.; Maddeux, B.A.; Goldfine, I.D. The molecular basis for oxidative stress-induced insulin resistance. Antioxid. Redox Signal. 2005, 7, 1040–1052. [Google Scholar] [CrossRef]
  214. West, I.C. Radicals and oxidative stress in diabetes. Diabet. Med. 2000, 17, 171–180. [Google Scholar] [CrossRef]
  215. Rosen, P.; Nawroth, P.P.; King, G.; Moller, W.; Tritschler, H.J.; Packer, I. The role of oxidative stress in the onset and progression of diabetes and complications: A summary of a Congress Series sponsored by UNESCO-MCBN, the American Diabetes Association and the German Diabetes Society. Diabetes Metab. Res. Rev. 2001, 17, 189–212. [Google Scholar] [CrossRef]
  216. Dokhen, D.B.; Saengsirisuwan, V.; Kim, J.S.; Teachy, M.K.; Henriksen, E.J. Oxidative stress-induced insulin resistance in rat skeletal muscle: Role of glycogen synthase kinase-3. Am. J. Physiol. Endocrinol. Metab. 2008, 294, E615–E621. [Google Scholar] [CrossRef]
  217. Jheng, H.F.; Tsai, P.J.; Guo, S.M.; Kuo, L.H.; Chang, C.S.; Su, I.J.; Chang, C.R.; Tsai, Y.S. Mitochondrial fission contributes to mitochondrial dysfunction and insulin resistance in skeletal muscle. Mol. Cell Biol. 2012, 32, 309–319. [Google Scholar] [CrossRef]
  218. Zhang, D.; Liu, Z.X.; Choi, C.S.; Tian, I.; Kibbey, R.; Dong, J.; Cline, G.W.; Wood, P.A.; Shulman, G.J. Mitochondrial dysfunction due to long-chain Acyl-CoA dehydrogenase deficiency causes hepatic steatosis and hepatic insulin resistance. Proc. Natl. Acad. Sci. USA 2007, 104, 17075–17080. [Google Scholar] [CrossRef]
  219. Petersen, K.F.; Shulman, G.I. Etiology of insulin resistance. Am. J. Med. 2006, 119, S10–S16. [Google Scholar] [CrossRef]
  220. St-Pierre, J.; Lin, J.; Krauss, S.; Tarr, P.T.; Yang, R.; Newgard, C.B.; Spigelman, B.M. Bioenergetic analysis of peroxisome proliferator-activated receptor γ coactivators 1α and 1β (PGC-1α and PGC-1β) in muscle cells. J. Biol. Chem. 2003, 278, 26597–26603. [Google Scholar] [CrossRef]
  221. Yuan, D.; Xiao, D.; Gao, Q.; Zeng, L. PGC-1α activation: A therapeutic target for type 2 diabetes. Eat. Weight Disord. 2019, 24, 385–395. [Google Scholar] [CrossRef]
  222. Wu, X.; Chen, K.; Williams, K.J. The role of pathway-selective insulin resistance and responsiveness, in diabetic dyslipoproteinemia. Curr. Opin. Lipidol. 2012, 23, 334–344. [Google Scholar] [CrossRef]
  223. Li, S.; Brown, M.S.; Goldstein, J.L. Bifurcation of insulin signaling pathway in rat liver: mTORC1 required for stimulation of lipogenesis, but not inhibition of gluconeogenesis. Proc. Natl. Acad. Sci. USA 2010, 107, 3441–3446. [Google Scholar] [CrossRef]
  224. Guertin, D.A.; Stevens, D.M.; Thorens, C.C.; Burds, A.A.; Kalaany, N.Y.; Moffat, J.; Brown, M.; Fitzgerald, K.J.; Sabatini, D.M. Ablation in mice of the mTOR complement raptor, rictor, or mLST8 reveals that mTOR2 is required for signaling to Akt-FOXO and PKCalpha, but not S6K1. Dev. Cell 2006, 11, 859–871. [Google Scholar] [CrossRef]
  225. Kim, K.; Qiang, L.; Hayden, M.S.; Sparling, D.P.; Purcell, N.H.; Pajvani, U.B. mTORC1-independent raptor prevents hepatic steatosis by stabilizing PHLPP2. Nat. Commun. 2016, 7, 10255. [Google Scholar] [CrossRef]
  226. Kim, K.; Ryu, D.; Dongiovanni, P.; Ozcan, L.; Nayak, S.; Ueberheide, B.; Valenti, L.; Auwerx, J.; Pajvani, U.B. Degradation of PHLPP2 by KCTD17, via a glucagon-dependent pathway promotes hepatic steatosis. Gastroenterology 2017, 153, 1568–1580. [Google Scholar] [CrossRef]
  227. Nagai, Y.; Yonemitsu, S.; Erion, D.M.; Iwasaki, T.; Stark, R.; Weissmann, D.; Dong, J.; Zhang, D.; Jurczak, M.J.; Löffer, M.G.; et al. The role of peroxisome proliferator-activated receptor γ coactivator-1 β in the pathogenesis of fructose-induced insulin resistance. Cell Metab. 2009, 9, 252–264. [Google Scholar] [CrossRef]
  228. Bindesboll, C.; Fan, Q.; Norgaard, R.C.; MacPherson, L.; Ruan, H.B.; Wu, J.; Pedersen, T.; Steffensen, K.R.; Yang, X.; Mathews, J.; et al. Liver X receptor regulates hepatic nuclear O-GlcNAc signaling and carbohydrate responsive element-binding protein activity. J. Lipid Res. 2015, 56, 771–785. [Google Scholar] [CrossRef]
  229. Postic, C.; Dentin, R.; Denechaud, P.D.; Girard, J. ChREBP, a transcriptional regulator of glucose and lipid metabolism. Annu. Rev. Nutr. 2007, 27, 179–192. [Google Scholar] [CrossRef]
  230. Erion, D.M.; Popov, V.; Hsiao, J.J.; Vatner, D.; Mitchell, K.; Yonemitsu, S.; Nagai, Y.; Kahn, M.; Gillum, M.P.; Dong, J.; et al. The role of the carbohydrate response element-binding protein in male fructose-fed rats. Endocrinology 2013, 154, 36–44. [Google Scholar] [CrossRef]
  231. Uyeda, K.; Repa, J.J. Carbohydrate response element binding protein, ChREBP, a transcription factor coupling hepatic glucose utilization and lipid synthesis. Cell Metab. 2006, 4, 107–110. [Google Scholar] [CrossRef]
  232. Magkos, F.; Yannakoulia, M.; Chan, J.L.; Mantzoros, C.S. Management of the metabolic syndrome and type 2 diabetes through life style modification. Annu. Rev. Nutr. 2009, 29, 223–256. [Google Scholar] [CrossRef]
  233. Lloyd-Jons, D.M.; Liu, K.; Colangelo, L.A.; Yan, L.L.; Klein, L.; Loria, C.M.; Lewis, C.E.; Savage, P. Consistently stable or decreased body mass index in young adulhood and longitudal changes in metabolic syndrome components: The Coronary Artery Risk Development in Young Adults Study. Circulation 2007, 115, 1004–1011. [Google Scholar] [CrossRef]
  234. Nieman, D.C.; Wentz, L.M. The compelling link between physical activity and the body’s defense system. J. Sport Health Sci. 2019, 8, 201–217. [Google Scholar] [CrossRef]
  235. Wolosowicz, M.; Prokopiuk, S.; Kaminski, T.W. Recent advances in the treatment of insulin resistance targeting molecular and metabolic pathways: Fighting a losing battle? Medicina 2022, 58, 472. [Google Scholar] [CrossRef]
  236. Moore, M.M.; Bailey, A.M.; Flannery, A.H.; Baum, R.A. Treatment of diabetic ketoacidosis with intravenous U-500 insulin in patient with Rabson-Mendenhall syndrome: A case report. J. Pharm. Pract. 2017, 30, 468–475. [Google Scholar] [CrossRef]
  237. Semple, R.K.; Williams, R.M.; Dunger, D.B. What is the best management strategy for patients with severe insulin resistance? Clin. Endocrinol. 2010, 73, 286–290. [Google Scholar] [CrossRef] [PubMed]
  238. Kawashima, Y.; Nishimura, R.; Utsunomiya, A.; Kagawa, R.; Funata, H.; Fujimoto, M.; Hanaki, K.; Kanzaki, S. Leprechaunism (Donohue syndrome): A case bearing novel compound heterozygous mutations in the insulin receptor gene. Endocr. J. 2013, 60, 107–112. [Google Scholar] [CrossRef]
  239. Caldwell, S.H. Efficacy and safety of troglitazone for lipodystrophy syndrome. Ann. Intern. Med. 2001, 134, 1008. [Google Scholar] [CrossRef]
  240. Ray, K.K.; Bays, H.E.; Catapano, A.L.; Lalwani, N.D.; Bloedon, L.T.; Sterling, L.R.; Robinson, P.L.; Ballantyne, C.M. Safety and efficacy of bempedoic acid to reduce LDL cholesterol. N. Engl. J. Med. 2019, 380, 1022–1032. [Google Scholar] [CrossRef]
  241. Thomas, N.J.; Lynam, A.L.; Hill, A.V.; Weedon, M.N.; Shields, B.M.; Oram, R.A.; McDonald, T.J.; Hattersley, A.T.; Jones, A.G. Type 1 diabetes defined by severe insulin deficiency occurs after 30 years of age and is commonly treated as type 2 diabetes. Diabetologia 2019, 62, 1167–1172. [Google Scholar] [CrossRef]
  242. Zhang, A.M.Y.; Magrill, J.; de Winter, T.J.J.; Hu, X.; Skovsø, S.; Schaeffer, D.E.; Kopp, J.L.; Johnson, J.D. Endogenous hyperinsulinemia contributes to pancreatic cancer development. Cell Metab. 2019, 30, 403–404. [Google Scholar] [CrossRef]
  243. Zhong, W.; Mao, Y. Daily insulin dose and cancer risk among patients with type 1 diabetes. JAMA Oncol. 2022, 8, 1356. [Google Scholar] [CrossRef]
  244. Iqbal, M.A.; Siddiqui, F.A.; Gupta, K.; Chattopadhyay, S.; Gopinath, P.; Kumar, B.; Manvati, S.; Chaman, N.; Bamezai, R.N.K. Insulin enhances metabolic capacities of cancer cells by dual regulation of glycolytic enzyme pyruvate kinase M2. Mol. Cancer 2013, 12, 72. [Google Scholar] [CrossRef]
  245. Frasca, F.; Pandini, P.; Scalia, P.; Sciacca, L.; Mineo, R.; Constantino, A.; Goldfine, I.D.; Belfore, A.; Vigneri, R. Insulin receptor isoforms A, a newly recognized high-affinity insulin-like growth factor II receptor in fetal and cancer cells. Mol. Cell Biol. 1999, 19, 3278–3288. [Google Scholar] [CrossRef]
  246. Vigneri, P.; Frasca, F.; Sciacca, L.; Pandini, G.; Vigneri, R. Diabetes and cancer. Endocr.-Relat. Cancer 2009, 16, 1103–1123. [Google Scholar] [CrossRef]
  247. Goodwin, P.J.; Stambolic, V. Obesity and insulin resistance in breast cancer—Chemoprevention strategies with a focus on metformin. Breast 2011, 20, S31–S35. [Google Scholar] [CrossRef] [PubMed]
  248. Del Giudice, M.E.; Fantus, I.G.; Ezzat, S.; McKeown-Eyssen, G.; Page, D.; Goodwin, P.J. Insulin and related factors in premenopausal breast cancer risk. Breast Cancer Res. Treat. 1998, 47, 111–120. [Google Scholar] [CrossRef] [PubMed]
  249. Gunter, M.J.; Hoover, D.R.; Yu, H.; Wassertheil-Smoller, S.; Rohan, T.E.; Manson, J.E.; Li, J.; Ho, G.Y.F.; Xue, X.; Anderson, G.I.; et al. Insulin, insulin-like growth factor-1 and risk of breast cancer in postmenopausal women. J. Natl. Cancer Inst. 2009, 101, 48–60. [Google Scholar] [CrossRef] [PubMed]
  250. Kaaks, R.; Toniolo, P.; Akmedkhanov, A.; Lukanova, A.; Biessy, C.; Dechaud, H.; Rinaldi, S.; Zeleniuch-Jacquotte, A.; Shore, P.E.; Riboli, E. Serum C-peptide, insulin-like growth factor (IGF)-I, IGF-binding protein, and colorectal cancer risk in women. J. Natl. Cancer Inst. 2000, 92, 1592–1600. [Google Scholar] [CrossRef] [PubMed]
  251. Schoen, R.E.; Tangen, C.M.; Kuller, L.H.; Burke, G.L.; Cushman, M.; Tracy, R.P.; Dobs, A.; Savage, P.J. Increased blood glucose and insulin, body size, and incident colorectal cancer. J. Natl. Cancer Inst. 1999, 91, 1147–1154. [Google Scholar] [CrossRef] [PubMed]
  252. Koohestani, N.; Tran, T.T.; Lee, W.; Wolever, T.M.; Bruce, W.R. Insulin resistance and promotion of aberrant crypt foci in the colons of rats on a high-fat diet. Nutr. Cancer 1997, 29, 69–76. [Google Scholar] [CrossRef]
  253. Watkins, L.F.; Lewis, L.R.; Lewine, A.E. Characterization of the synergistic effect of insulin and transferin and the regulation of their receptors on a human colon carcinoma cell line. Int. J. Cancer 1990, 45, 372–375. [Google Scholar] [CrossRef]
  254. Moschos, S.J.; Mantzoros, C.S. The role of the IGF system in cancer: From basic to clinical studies and clinical applications. Oncology 2002, 63, 317–332. [Google Scholar] [CrossRef]
  255. Pisani, P. Hyper-insulinaemia and cancer, meta-analyses of epidemiological studies. Arch. Physiol. Biochem. 2008, 114, 63–70. [Google Scholar] [CrossRef]
  256. Chan, M.T.; Lim, G.E.; Skovsø, S.; Yang, Y.H.C.; Albrecht, T.; Alejandro, E.U.; Hoesli, C.A.; Piret, J.M.; Warnock, G.L.; Johnson, J.D. Effects of insulin on human pancreatic cancer progression modeled in vitro. BMC Cancer 2014, 14, 814. [Google Scholar] [CrossRef]
  257. Nicholson, K.M.; Anderson, N.G. The protein kinase B/Akt signalling pathway in human malignancy. Cell Signal. 2002, 14, 381–395. [Google Scholar] [CrossRef]
  258. Altomare, D.A.; Testa, J.R. Perturbations of the AKT signaling pathway in human cancer. Oncogene 2005, 24, 7455–7464. [Google Scholar] [CrossRef] [PubMed]
  259. Cancer Genome Atlas Network. Comprehensive molecular partaits of human breast tumours. Nature 2012, 490, 61–70. [Google Scholar] [CrossRef] [PubMed]
  260. Sundaram, S.; Johnson, A.R.; Makowski, L. Obesity, metabolism and the microenvironment links to cancer. J. Carcinog. 2013, 12, 19. [Google Scholar] [PubMed]
  261. Yang, Z.Y.; Di, M.Y.; Yuan, J.Q.; Shen, W.X.; Zheng, D.Y.; Chen, J.Z.; Mao, C.; Tang, J.L. The prognostic value of phosphorylated Akt in breast cancer: A systematic review. Sci. Rep. 2015, 5, 7758. [Google Scholar] [CrossRef]
Figure 1. Process of biosynthesis and secretion of insulin. After transcription of preproinsulin mRNA from the INS gene, it is translated into preproinsulin peptides. Then, preproinsulin transits through the rough endoplasmic reticulum and trans-Golgi network, causing the preproinsulin to change to its mature form and be stored in mature secretory granules in pancreatic β-cells [2,17,22].
Figure 1. Process of biosynthesis and secretion of insulin. After transcription of preproinsulin mRNA from the INS gene, it is translated into preproinsulin peptides. Then, preproinsulin transits through the rough endoplasmic reticulum and trans-Golgi network, causing the preproinsulin to change to its mature form and be stored in mature secretory granules in pancreatic β-cells [2,17,22].
Ijms 25 02397 g001
Figure 2. Structure of insulin receptor [4,28]. The insulin receptor is composed of two extracellular insulin-binding α-subunits and two signal transduction β-subunits, which are bound together by disulfide bonds. Conformational change of α-subunits due to binding of insulin to the receptor enables binding of adenosine triphosphate (ATP) to the β-subunit. Binding of ATP activates a tyrosine kinase in the β-subunits, causing autophosphorylation of the receptor. The phosphorylated insulin receptor in turn phosphorylates other protein substrates, such as IRS-1 and IRS-2.
Figure 2. Structure of insulin receptor [4,28]. The insulin receptor is composed of two extracellular insulin-binding α-subunits and two signal transduction β-subunits, which are bound together by disulfide bonds. Conformational change of α-subunits due to binding of insulin to the receptor enables binding of adenosine triphosphate (ATP) to the β-subunit. Binding of ATP activates a tyrosine kinase in the β-subunits, causing autophosphorylation of the receptor. The phosphorylated insulin receptor in turn phosphorylates other protein substrates, such as IRS-1 and IRS-2.
Ijms 25 02397 g002
Table 1. Characteristics of severe insulin resistance syndromes.
Table 1. Characteristics of severe insulin resistance syndromes.
Site of DefectsCauseSyndromeFeatures
Insulin geneMutation ValA3→LeuInsulin WakayamaDecreased binding of insulin to the INSR.
Mutation PheB24→SerInsulin Los AngelesSignificantly reduced bioactivity of insulin.
Mutation PheB25→LeuInsulin ChicagoDecreased binding of insulin to the INSR.
Insulin receptorAutosomal recessive mutationsDonohue syndromeAffected patients seldom live beyond infancy. Most affected patients survive fewer than 2 years.
Rabson–Mendenhall syndromeHyperinsulinemia, fasting hyperglycemia, failed reaction on insulin. Most affected patients survive up to 15 years.
Autosomal dominant or recessive mutationsType A insulin resistance syndrome (TAIRS)Syndrome has a relatively good prognosis. Affected patients can live beyond middle age. Hypoglycemia is observed.
Autosomal dominant mutationsType C insulin resistanceIt is a variant of Type A insulin resistance with less severe insulin resistance. It is so-called HAIR-AN (hyperandrogenic, insulin resistance, acanthosis nigricans).
An autoimmune disorder caused by polyclonal autoantibodies acting against the insulin receptor.Type B insulin resistance (TBIRS)Depending on levels of autoantibodies, hypoglycemia or hyperglycemia may be observed.
Depending on the type of lipodystrophy, these syndromes may be due to autosomal dominant, autosomal recessive or X-linked mutations.LipodystrophiesThe lipodystrophy syndromes are a heterogenous group of disorders. Depending on the disorder, different pathologies, such as severe insulin resistance, severe hypertriglyceridemia, pancreatitis, complete or partial loss of adipose tissue and depletion of lipid storage capacity, are observed.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Szablewski, L. Changes in Cells Associated with Insulin Resistance. Int. J. Mol. Sci. 2024, 25, 2397. https://doi.org/10.3390/ijms25042397

AMA Style

Szablewski L. Changes in Cells Associated with Insulin Resistance. International Journal of Molecular Sciences. 2024; 25(4):2397. https://doi.org/10.3390/ijms25042397

Chicago/Turabian Style

Szablewski, Leszek. 2024. "Changes in Cells Associated with Insulin Resistance" International Journal of Molecular Sciences 25, no. 4: 2397. https://doi.org/10.3390/ijms25042397

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