Please cite this paper as:

Mukhopadhyay, M., Ghosh, P., Chattopadhyay, A. and Bandyopadhyay, D. 2023. An insight into the importance of B vitamins and melatonin in the prevention of diabetes through modulation of the brain energy metabolism- a comprehensive review. Melatonin Research. 6, 3 (Sep. 2023), 397-430. DOI:https://doi.org/https://doi.org/10.32794/mr112500160.


Review

An insight into the importance of B vitamins and melatonin in the prevention of diabetes through modulation of the brain energy metabolism- a comprehensive review 

Manisha Mukhopadhyay^1, Priyanka Ghosh^1, Aindrila Chattopadhyay2, Debasish Bandyopadhyay1*

 1Oxidative Stress and Free Radical Biology Laboratory, Department of Physiology, University of Calcutta, 92, APC Road, Kolkata-700009, India

2Department of Physiology, Vidyasagar College, 39, Sankar Ghosh Lane, Kolkata-700006, India

*Correspondence: debasish63@gmail.comdbphys@caluniv.ac.inTel: +91 9433072066

^The authors have equal contribution

Running title: Melatonin and B-vitamins in brain metabolism and diabetes

ReceivedMarch 31, 2023; Accepted: September 22, 2023


ABSTRACT

     Energy metabolism is the biochemical pathway of converting macronutrients (carbohydrates, protein, and fat) to cellular energy for the maintenance of cell homeostasis. The brain is an organ that consumes unproportional energy compared to its size. Glucose (glycogen, in storage form of glucose) is the principal source of brain energy. Impairment in brain energy metabolism results in neuronal loss and subsequent neurodegenerative diseases including AD, PD, amyotrophic lateral sclerosis, Huntington’s disease, etc. However, metabolic disorders such as chronic hyperglycemia, and insulin resistance are also linked with neuronal activity. Dysregulation in neuronal transmission is associated with oxidative stress and brain insulin resistance. Diabetes mellitus jeopardizes brain function through various mechanisms including glucose toxicity, BBB damage, neuroinflammation, and gliosis. B vitamins as antioxidants and neuroprotective agents, can improve brain glucose metabolism. Melatonin is a potent free radical scavenger and it can also modulate cellular cytokine levels and prevent insulin resistance. The neuroprotective and antihyperglycemic effects of melatonin improve the brain's antioxidant defense system, decrease brain NOS activity, and prevent glucose toxicity. Hence this review suggests a therapeutic use of a combination of melatonin and  B vitamins to improve brain functioning disrupted by diabetes.

Key words: Brain, energy metabolism, glucose toxicity, diabetes, B vitamins, melatonin

___________________________________________________________________________


1. INTRODUCTION

     Metabolism is a biochemical process of converting nutrients (carbohydrates, fatty acids, and amino acids) to energy on a cellular level, involving several enzyme-dependent pathways essential for maintaining energy homeostasis and macromolecular synthesis in humans (1). The mammalian brain depends on glucose as its principal source of energy. The brain consumes 25% of the body’s energy for the restoration of the membrane gradient potential, intracellular signaling, neurotransmitter recovery, and dendritic-axonal transport (2). Adult brain neurons have the major energy requirement (3) and thus need a continuous supply of blood glucose. The production of ATP by glucose metabolism powers brain function (4). On the other hand, glycogen is the storage form of glucose in the brain, mainly located in astrocytes (5). Glucose is partially metabolized to lactate in astrocytes, which is then released into the extracellular environment and acquired by neurons (6). Other brain cells, such as glial and ependymal cells, require high energy to carry out their specific functions through glycolysis and activating glycogenolysis (5). Thus, the continuous circulation of glucose and oxygen to brain cells is necessary for brain metabolism (6). In specific conditions, such as starvation and diabetes, scarcity of carbohydrates leads  to increased ketone body (acetoacetic acid, β-hydroxybutyric acid, and acetone) production, brain can tolerate ketone body as an alternative energy source  (7). Interruption in energy metabolism will result in impaired neuronal transmission followed by neuronal loss. The dysregulation of neuronal transmission is often associated with oxidative stress (8). Neurons communicate among different regions of the brain and the peripheral nervous system via electrical impulses and chemical signals. Therefore, metabolic impairment in neurons lead to different neurodegenerative disorders, including Alzheimer’s disease (AD) (9), Huntington’s disease (10), amyotrophic lateral sclerosis, and Parkinson’s disease (11).

     Moreover, metabolic disorders such as chronic hyperglycemia, microvascular complications, and insulin resistance  also influence brain function (12–14). Numerous animal diabetic models indicate that glucose toxicity and resistant insulin signaling deteriorate brain functioning by modulating energy metabolism, synaptic plasticity, learning, and memory function resulting in poor cognitive behaviour, AD, and vascular dementia (12, 15). Oxidative stress in the peripheral tissues is responsible to induce insulin resistance (16). The brain is very susceptible to oxidative damage because of the high oxygen consumption rate, high lipid content, and a relative deficiency in antioxidant enzymes compared to other tissues. Reactive oxygen species (ROS) play a role in many neurodegenerative diseases, including diabetes, because neuronal cells are particularly vulnerable to oxidative assaults (17, 18). Diabetes affects brain function through various pathways, including glucose toxicity, BBB impairments, vascular damage, mitochondrial dysfunction, oxidative stress, brain insulin resistance, synaptic failure, neuroinflammation, and gliosis (19). The long-term effects of diabetes on the brain are manifested at the structural, neurophysiological, and neuropsychological levels (20).

     Potent antioxidants in the diet may prevent diabetes-induced brain impairment and cellular oxidative damage. Vitamins and melatonin can both prevent oxidative stress and maintain redox equilibrium (21). Tan et al. (22) referred melatonin as an antioxidant vitamin. In spite of being generated endogenously, a substantial level of melatonin is present in foodstuffs including rice, wheat, seeds, fruits and vegetables. The bioavailability of dietary melatonin is also high. Melatonin and  vitamins both are derived from plants but also frequently obtained from  consumption of animal products of meat, dairy and eggs.

     Vitamins are classic antioxidants prone to a deficit in infants and older adults. Prolonged deficiency of these micronutrients leads to malnutrition and several health complications (23). Among all, the B vitamins are a family of eight water-soluble vitamins acting as co-enzymes in diverse catabolic and anabolic reactions, that are critically related to cellular activities. Their cumulative impacts are particularly pervasive in many aspects of brain function, including energy production, DNA/RNA synthesis/repair, genomic and non-genomic methylation, and the production of several neurochemicals and signalling molecules (24). The B vitamins are synthesized in plants’ chloroplast, mitochondria, and cytosol, except for vitamin B12, which is synthesized by bacteria (25). B vitamins along with their other roles in general metabolic function have a great impact on brain function as they perform potential action in homocysteine metabolism. Increased levels of homocysteine in plasma are considered a predisposing factor to cardiovascular disease (26), AD and other dementias (27). Deficiencies in the key vitamins involved in the conversion of homocysteine to methionine including folate, vitamin B6, and vitamin B12 are implicated as the underlying cause of developing neurodegenerative diseases (28).

     Melatonin (N-acetyl-5-methoxytryptamine), an indoleamine, is synthesized in the pinealocyte from tryptophan and released into the peripheral system and cerebrospinal fluid (CSF) (29). Melatonin exerts its effect in both centrally and peripherally, and the binding sites of melatonin are present in different areas of the brain, including the hypothalamus, pars tuberalis, as well as in the immune cells, gonads, kidneys, and heart (29). Melatonin has both receptor-mediated and non-receptor-mediated actions. The non-receptor-mediated action explicates melatonin’s amphipathic properties, meaning it can easily cross the cell and nuclear membrane of the brain and other body tissues (30). The antioxidant function of melatonin against the different models of oxidative stress is also receptor-independent action (31). Therefore, the blood-brain barrier (BBB) is easily crossed by melatonin due to its amphiphilicity (32). Then, the choroid plexus enables it to permeate the central nervous system (CNS) and cerebrospinal fluid (CSF) (33). Various regions of the CNS, including the cerebellum, olfactory bulb, prefrontal cortex, hippocampus, and striatum, express the rate-limiting enzyme of melatonin biosynthesis, that is arylalkylamine N-acetyltransferase (AANAT) (34, 35). In the CNS, melatonin degrades into N1-acetyl-N2-formyl-5-methoxykynuramine (AFMK), which is then deformylated to  N1-acetyl-5-methoxykynuramine (AMK) (36).  AFMK and AMK still have cell-protective activity, including protective effects in mitochondria (37). Many studies have documented  the potent antioxidant, anti-inflammatory, and neuroprotective effects of melatonin (38), and its analgesic actions in the peripheral neurodegenerative disorders  (39). It  reduces  hyperglycemia by protecting the β cells of the pancreas (40). In this review, we summarize the action of melatonin and the B vitamins in the prevention of the impaired brain metabolism caused by hyperglycemia.


2. BRAIN AND ITS METABOLISM

     Brain uses  glucose as the main source of energy supply. In humans, brain only 2% of body weight, but consuming 20% of the total energy obtained from glucose (5.6 mg of glucose per 100 g of human brain tissue every minute) (41). The blood glucose can cross the BBB through its specific carrier and be stored as glycogen (6). Major brain operations like learning and memory consolidation depend critically on brain glycogen metabolism (42). Therefore, brain glycogen has been regarded as an emergency glucose reserve (43). In addition, during period of stress like hypoglycemia and ischemia, astrocyte  and neurons  are vitally dependent on glycogen storage with  the process of glycogenolysis to provide glucose (44–46).

2.1. The overview of glucose metabolic pathways in the brain.

     Glycogen synthesis and breakdown are intricate processes involving multiple enzymes and are precisely regulated by phosphorylation and dephosphorylation (47). Three enzymes located in the brain are critically involved in glycogen metabolisms, such as glycogen synthase (GS), glycogen phosphorylase (GP), and phosphorylase kinase (PK) (48). The brain isoform of GP is present in astrocytes and other cells, including choroid plexus and ependymal cells  (49). GS is expressed in the hippocampus, cerebellum, and olfactory bulbs (24). When the blood glucose is too low, the brain glycogen reserve is metabolised into glucose to fulfil its physiological requirement  (50). The glycogenolysis involves three specific enzymes: glucosyltransferase, glucosidase, and glycogen-debranching enzyme (GDE). These enzymes, namely PK, GP, and GDE, play a crucial role in hydrolyzing α-1,6-glycosidic branch point bonds to release glucose (51). However, the metabolic fate of brain glucose depends upon the cell type and glycolytic enzymes present in the astrocyte. In addition to generate ATP, glucose is crucial for fatty acid, amino acid, and neurotransmitter generation. In neurons, glucose is metabolized through several pathways, including glycolysis, PPP, tricarboxylic acid (TCA) cycle, and oxidative phosphorylation (52). The glycolytic pathway converts glucose to pyruvate, which is then transported into mitochondria to generate acetyl coenzyme A (acetyl-CoA). Acetyl-CoA is used to form NADH and FADH2, The energy carried by these reduced agents finally used for ATP production in the mitochondria (Figure 1).

Figure 1.png

Fig. 1. The process  of glucose metabolism in the brain.

     This graphical abstract represents that in stressed condition the brain glycogen storage has been broken down in the astrocytes and undergoes into the process of glycogenolysis to produce glucose. Whereas in normal condition, the brain glucose is used for the production of fatty acid, amino acid and neurotransmitters as well as enters into the neurons where it is used as  energy source by entering into the glycolysis, Krebs cycle and ETC chain and then stored as glycogen in astrocytes through the process of glycogenesis. The left side signifies the stressed condition whereas right side as normal condition.

     GS: glycogen synthase, GP: glycogen phosphorylase, PK: phosphorylase kinase, GDE: glycogen-debranching enzyme, G1P: glucose-1-phosphate, G6P: glucose-6-phosphate, ATP- adenosine-tri-phosphate, NAD: nicotinamide-adenine-dinucleotide (oxidized), NADH: nicotinamide-adenine-dinucleotide (reduced), FAD: flavin-adenine-dinucleotide (oxidized), FADH2- Flavin-Adenine-Dinucleotide (reduced).

2.2. Glucose transporters in the brain.

     To perform normal cellular activities brain needs a continuous supply of glucose across the  cell membrane of microvascular endothelial cells to reach neurons and glial cells. However, the cell membrane limits the  passive diffusion of glucose and other nutrients in and out of cells. Instead, the movements of glucose depend on a particular saturable transport process of two distinct classes of glucose transporters including glucose transporters (facilitated transport; GLUT) and sodium-dependent glucose transporters (secondary active transport; SGLT), each with a unique set of kinetic properties to transfer glucose across the cell membranes and tissue barriers. The majority of cells express a wide range of glucose transporters, and the pattern of expression in various organs is correlated with particular metabolic demands (Table1) (53). GLUT family proteins transfer glucose bidirectionally along a concentration gradient. The isoforms of GLUTs  have different glucose affinity, indicating adaption to the different metabolic needs of each cell (54). A high molecular weight isoform of GLUT-1 (55 kDa) is mainly responsible  to transport glucose from the blood to cells. GLUT3 and GLUT-8 are mainly found in neurons with apparently balanced distribution (55). On the other hand, GLUT-5 is highly expressed in microglial cells, and its inefficiency leads to a significantly poor supply of glucose to the brain. GLUT-4 is  insulin dependent glucose transporter (56). In hypothalamic nuclei, neuronal expression of GLUT2 and GLUT4 has been documented to implicate the central regulation of glucose homeostasis, food intake, and/or energy balance. Apart from the GLUT1-5 transporters, several other cloned isoforms are present, including GLUT-x1 (also called GLUT-8), GLUT-9 (recently called GLUT-6), GLUT-10, GLUT-11, and GLUT-12 (57). GLUT-8 is a hormonally regulated transporter susceptible to diabetes and stress (58). On the other hand, SGLTs transport glucose and galactose along a concentration gradient while also transporting Naions (59). Under hypoglycemia and hypoxemic circumstances, SGLT1, which is ubiquitously expressed in neurons, may play a crucial role in glucose metabolism (60). However, alterations in the expression of the BBB glucose transporter cause the onset of diabetes (61, 62). The GLUT isoforms and their activities in neuronal system have been summarized in the table 1.

Table 1.  Summary of GLUT and SLT isoforms in brain.

Type

Gene

Site expressed in the brain

Substrate/transporters

Facilitative/sodium independent

GLUT1

Brain endothelial and epithelial-like brain barriers, glial cells,   blood-tissue barriers, and peripheral nerves.

Glucose, galactose, mannose, glucosamine, ascorbic acid

GLUT2

Astrocytes

Mannose, galactose, fructose, glucose,   glucosamine

GLUT3

Neurons, brain endothelial cells

Glucose, galactose, mannose, xylose, dehydroascorbic acid

GLUT4

Hippocampal neurons, cerebellar neurons

Glucose, dehydroascorbic acid, glucosamine

GLUT5

Brain microglia

Fructose

GLUT6

Brain, peripheral

Glucose

GLUT8 (Insulin-responsive)

Neurons

Glucose

GLUT10

Brain

Glucose and galactose


Sodium-Glucose   Co-transporter/Sodium-dependent

SGLT1

Brain (cortical, pyramidal, and Purkinje neuronal   cells)

Glucose, galactose, water

SGLT2

SGLT3

SGLT4

brain

 

Glucose, galactose

mannose, fructose

SGLT6

Brain (neurons)

Myo-inositol, glucose

2.3. Brain glucose intolerance in hyperglycemia.

     The glucose neurotoxicity in diabetes is associated with cellular dysfunction through the following mechanisms i) increased polyol pathway flux, ii) elevation of advanced glycation end products formation, iii) activation of protein kinase C (PKC) and iv) increased hexosamine pathway flux (63). Since the glucose content in the plasma is five folds higher than that in the brain (64, 65), the endothelial cells are more prone to develop hyperglycemia than the brain parenchymal cells (66). Impairment in cerebral vasculature leads to disruption in the blood-brain barrier in diabetes, aging, and neurodegenerative diseases as well (66). The level of cognitive impairment and the progression of AD is strongly correlated with the decline in cerebral glucose metabolism, which reflects the severity of synaptic malfunction and neurodegeneration (67). It has been suggested that disturbed cerebral glucose metabolism causes oxidative stress, inflammation, mitochondrial malfunction, autophagy impairment, excitotoxicity, and apoptosis, which in turn causes Aβ deposition and tau hyperphosphorylation (68).

     Insulin resistance, seen in T2DM is associated with an increased risk of dementia and may be attributed to poor insulin signalling in neurons (12). Although brain glucose transport is not insulin-dependent, however in certain subcellular compartments it is controlled by insulin (19). Insulin signalling has been found mostly in neurons and astrocytes that may regulate glycogen metabolism (69). During low fuel supply, brain glycogen is rapidly mobilized for supporting glutamatergic neurotransmission (70, 71). Brain glucose utilization is diminished with the age or with the progression of AD (72). Willette et al. (73) reported that peripheral insulin resistance is correlated with decreased glucose metabolism in AD subjects. Insulin is  important in the specific areas of the brain including the hypothalamus which is the main regulatory part of the body’s energy homeostasis (74). While mitochondria are the energy  generating centre, their  dysfunction directly associates with neurodegenerative disorders (75). Mitochondrial dynamics, biogenesis or mitophagy in neurons and astrocytes are also regulated by insulin (76). Thus, the loss of insulin-mediated regulation of mitochondrial metabolism may lead to a reduction in the energy supply to neurons and astrocytes (19).

     Additionally, mitochondria also contribute to the breakdown of synapses through various factors including impaired ATP synthesis, Ca2+ signalling, elevated ROS, imbalanced metabolites (precursors of neurotransmitters) etc. that causes irregular mitochondrial dynamics, and mitochondria-dependent cell signalling transduction (77). Oxidative metabolism in astrocytic mitochondria and excitatory glutamatergic neurotransmission are tightly coupled (78), which is essential for memory and brain function (79). Notably, astrocytes also serve as the brain’s primary glycogen source, which is almost absent in neurons (80). Likewise, lactate produced by astrocyte glycogenosis and glycolysis has been implied for maintaining memory and brain function (79). In the hippocampus, insulin resistance is linked to abnormalities in astrocytic glycogen metabolism which may cause synaptic dysfunction associated with diabetes (81). This indicates that central insulin signalling impairment is an important factor for diabetes-induced brain injury (19). Interestingly, insulin also regulates the gene expression of memory function via the mitogen-activated protein kinase (MAPK) pathway (12, 82). Insulin also participates in regulating the key mediator of cellular activity i.e. AMP-activated protein kinase (AMPK) which might afford neuroprotection through metabolic controls (83). The brain metabolism and its relation to diabetes have illustrated in the Figure 2.

Figure 2.png

Fig. 2. Potential relationship between brain energy metabolism and diabetes.

     This graphical abstract represents that both the brain damage and diabetes are associated with each other. As brain damage causes the glucose toxicity, insulin resistance, oxidative stress, mitochondrial dysfunction, it finally results in the progression of diabetes. On the other hand, diabetes also induces brain damage by causing the BBB impairment, neuroinflammation, vascular damage and synaptic failure. BBB: blood brain barrier.


3. IMPORTANCE OF B VITAMINS IN BRAIN ENERGY METABOLISM AND DIABETES

    The B vitamins serve as cofactors are necessary for the synthesis of neurotransmitters and myelination of the spinal cord therefore, enable the CNS to function properly (84–86). B vitamins can cross the BBB and/or choroid plexus by the active transport mechanism. After cellular uptake, all B vitamins with  a turnover rate ranging from 8% to 100% per day, remain in high concentration in the brain. For example methyltetrahydrofolate (the active form of folate)  is  four times higher in the brain than that  in plasma whilst biotin and pantothenic acid levels in the brain  are 50% higher than that in  plasma (87–89).

3.1. Thiamine (Vitamin B1).

     Thiamine is converted to thiamine pyrophosphate (TPP) with the action of thiamine pyrophosphokinase. TPP acts as a cofactor of several enzymes required in brain glucose metabolism including transketolase, pyruvate dehydrogenase, and α-ketoglutarate dehydrogenase. TPP undergoes further phosphorylation to form thiamine triphosphate (TTP) or dephosphorylated to thiamine monophosphate (TMP) and thiamine diphosphate (TDP). Thiamine phosphorylation and dephosphorylation processes in the brain indicate the cellular localization of thiamine metabolizing enzymes as well as the phosphate esters themselves (90). Thiamine phosphorylated derivatives present in higher concentrations in neurons than in other brain cells (90). Previous reports have shown that clinical findings of AD are mostly related to significant loss of neurons with the concomitant decrease of thiamine metabolites (91, 92). Wernicke’s Encephalopathy (WE), a thiamine deficiency disorder is characterized by lowering the concentration of TPP and its dependent enzymes. It has been proclaimed that severe neuropathological damage is associated with WE including neuronal disintegration, mild endothelial swelling, sparing and destruction of neutrophils (93), and brain cell death (94). Potential reasons encompass impaired cerebral energy metabolism and localized lactate build-up, both of which may be triggered by diminished α-KGDH activity, and excessive release of excitotoxic amino acids (94). Thiamine levels were found to be 15% lower in red blood cells of  diabetic patients than normal subjects   (95). Alterations in erythrocyte transketolase activity have been linked to thiamine depletion in both type 1 and type 2 diabetes, however, the percentage of affected people varies among studies, ranging from 17% to 79% (96).  It has also been demonstrated that insulin-deficient rats exhibit a substantial reduction in the transfer of free thiamine and TMP but a concomitant rise in TDP levels. Insulin deficit is associated with a reduction in the rate of thiamine transport across the colon (97). Thiamine deficiency results in a substantial decrease in insulin production and secretion (98). As a result, insulin deficit may increase thiamine deficiency and vice versa (99). Studies have been proposed on the relationship between vitamin B1 and DM. Watson et al. (100) reported that 36-47% of the population suffered from thiamine-deficient hyperglycemia. Low plasma thiamine level was observed in type 1 DM. Additionally, it has been also reported that plasma thiamine levels decreased by 76% in type 1 and 75% in type 2 diabetic patients related to increased renal clearance of thiamine (101).

3.2. Riboflavin (Vitamin B2).

     The two flavoprotein-dependent co-factors flavin adenine mononucleotide (FMN) and flavin adenine dinucleotide (FAD) catalyse several important  enzymatic pathways. They primarily involve in the synthesis and conversion of other vitamins such as niacin, folate, and vitamin B6 as well as the synthesis of all heme proteins including hemoglobin, nitric oxide synthase, P450 enzymes, and proteins in the redox cycle. FAD and FMN also act as co-factors in the brain’s lipid metabolism (102). The role of riboflavin in the glutathione redox cycle is considered as its indirect antioxidative action. Alam et al. (103) have reported that riboflavin reduces  fasting blood glucose level in a dose-dependent manner while increases calcium and GLUT-4 expression as calcium is required for insulin secretion. Function as  the antioxidant, riboflavin preserves mitochondrial function, improves  many neurological conditions including AD, PD, and multiple sclerosis (MS) (104). Recent studies have elucidated that riboflavin prevents oxidative stress-mediated damage in the AD brain via modularization of the Nrf2 pathway (105). Since autoimmune responses  are directly linked to T1DM and T2DM, riboflavin as a potential antioxidant and immunomodulatory agent might be beneficial against diabetes-induced neurological disorders (106).

3.3. Niacin (Vitamin B3).

     A plethora of enzymes involved in brain metabolism is dependent on niacin derivatives such as nicotinamide adenine dinucleotide (NAD) and NAD phosphate (NADP). NAD+ is a crucial coenzyme for the processes of glycolysis, the TCA cycle, and the conversion of carbohydrates to lipids (107). Beyond energy production niacin also acts as an antioxidant vitamin required for the conversion of folate to tetrahydrofolate derivative. Niacin deficiency causes  several neurodegenerative disease-related symptoms such as dementia, and depression (108). Niacin has been widely recognized as a critical regulator of neuronal survival and development in the CNS (108). Nicotinamide can be converted to NAD by specific enzymes or methylated by nicotinamide-N-methyl transferase (NNMT) to N1-methyl nicotinamide (108). Impairment in the function of NNMT hinders methylation of NA that compromised the availability of free NAD+, which ultimately has an impact on the activities of NAD+-consuming enzymes. NNMT plays important role in brain energy metabolism and the development of many disorders, such as obesity, diabetes, aging, Parkinson’s disease, and cancer (109) because methylated nicotinamide could not be further converted into NAD+. The interaction between nicotinamide methylation and NAD+ regeneration indicates that NNMT may increase fat accumulation and limit fuel oxidation. In the event of elevated NNMT expression, the availability of nicotinamide may be compromised, hence impeding the functionality of NAD+-dependent activities (110). Moreover, NNMT gene is significantly associated with regulating glucose metabolism. The downregulation of NNMT reduces the insulin level and improves glucose tolerance. Hong et al(111) have reported  that reduced NNMT expression in primary hepatocytes lowers glucose synthesis, expression of glucose-6-phosphatase catalytic (G6pc), and phosphoenolpyruvate carboxykinase 1 cytosolic (Pck1) while  over expression of NNMT leads to increased glucose production and elevated G6pc and Pck1 expression levels. On the other hand, NNMT knockdown mice exhibited decreased fasting glucose and reduced conversion of pyruvate to glucose. All indicate NNMT involves in gluconeogenesis. In stressful conditions, nicotinamide promotes neuronal survival via multiple mechanisms including i. inhibition of cytochrome c release, ii. prevention of caspase 3 and caspase 9, iii. suppressing the degradation of FOXO3a, iv. maintenance of protein kinase B-dependent phosphorylation of FOXO3a (112). However, few studies have been done on the effect of niacin in diabetes, but niacin administration has been found to elevate high-density lipoprotein (HDL) cholesterol, lowering triglyceride and low-density lipoprotein (LDL) cholesterol (113). These modifications in lipid metabolism possess a role in diabetes-induced atherosclerosis and cell adhesion molecules (CAM) that have been observed in atherogenesis (114). Studies have depicted that niacin supplementation prevents monocyte adhesion to endothelial cells in diabetic individuals (114).

3.4. Pyridoxine (Vitamin B6).

     Apart from its role in the folate cycle vitamin B6 is essential for neurotransmitter synthesis and amino acid metabolism. The biosynthesis of dopamine, serotonin, γ aminobutyric acid (GABA), adrenaline, and melatonin requires vitamin B6 as a co-factor. The melatonin biosynthesis route involves the enzymatic activity of aromatic amino acid decarboxylase, which involves the presence of pyridoxal-5-phosphate (PLP) as a co-factor (115). A mild deficiency of vitamin B6 can negatively affect GABA and serotonin synthesis which in turn inhibits the neural activity of GABA and promote disordered sleep and behaviour, loss of cardiovascular function, and impaired hypothalamus pituitary control of hormone secretion (24). Vitamin B6 also regulates glucose metabolism in the brain (116). The association between low plasma PLP and diabetes was first reported by Lekem and Hollenbeck (117). Substantial evidences correlates low plasma PLP and the occurrence of diabetes (118, 119) and  reduced PLP level was observed in gestational diabetes mellitus (GDM) (117). Similar results were also found by Spellacy et al. who described pyridoxine therapy attenuates glucose intolerance and maintains insulin levels in gestational diabetic women (120). The underlying mechanism lies in the involvement of vitamin B6 in the kynuramine pathway. Impaired tryptophan metabolism is observed in different forms of diabetes associated with pregnancy, oral contraceptives, and emotional and metabolic stress decreasing the bioavailability of PLP (121, 122). Xanthurenic acid excretion is increased in the case of GDM which was reduced by PLP administration (121). In addition, vitamin B6 acts as a co-factor in cystathionine synthase and cystathionine lyase enzymes which are required for homocysteine metabolism. Thereby PLP deficiency leads to elevated homocysteine levels which are linked to obesity-induced insulin resistance and diabetes (123).

3.5. Biotin (Vitamin B7).

     The primary function of biotin is to act as a prosthetic group of carboxylases (124). Enzymes that require biotin as a coenzyme include acetyl-CoA carboxylase (ACC), geranyl-CoA carboxylase (GCC), 3-methylcrotonyl-CoA carboxylase (MCC), pyruvate carboxylase (PC), propionyl CoA carboxylase (PCC), and urea carboxylase (UC) (125). Biotin plays a crucial role in maintaining glucose harmony in cells through gluconeogenesis, insulin receptor transcription, and preserving β cell function (126). The physiological function of biotin is to increase the expression of hepatic glucokinase activity which was first observed in 1968 (127). Biotin controls the expression of glucokinase at both the transcription and translational levels (128).  Later, in vitro studies also reported that biotin augments the expression and activities of pancreatic glucokinase (128) hence enhancing the glycolysis pathway, and improving insulin function in muscles by promoting guanylate cyclase activity (129), to increase insulin receptor synthesis, and secretion (130–132). According to Hemmati et al., the administration of biotin as an adjuvant, in conjunction with an insulin regimen, has the potential to enhance glycaemic management and reduces plasma lipid concentrations in individuals with poorly managed type 1 diabetes (129) consequently, the levels of blood glucose, ketone bodies, triglycerides, and free fatty acids are maintained (133). Lazo et al. have reported that biotin supplementation increases β cell functioning and diminishes CAM in rodent models (134). Additionally in vivo studies also suggest a potential corelation between biotin administration and improved glucose tolerance (135, 136). The potential  mechanism behind the reduction in blood glucose in diabetic patients by biotin involves the synthesis of glycogen and reducing gluconeogenesis (137). 

3.6. Folic acid (vitamin B9) and cobalamin (vitamin B12).

     These two vitamins are interrelated due to their intrinsic role in the folate and methionine cycle. The scarcity of vitamin B12 restricts the formation of folate as it remains trapped in methyltetrahydrofolate (138). Folate deficiency usually accompanied  with reduction in purine or pyrimidine synthesis as well as genomic and non-genomic methylation which impedes the process of neuron differentiation. These cause hippocampal atrophy, progression of demyelination, and disintegration of cellular structure. Consequently, these effects have major implications on the generation of normal action potential (139). Folate and vitamin B12 deficiency downregulate of folate-dependent proteins and DNA and RNA synthesis causes foetal developmental disorder and megaloblastic anaemia (138, 139). Diabetic patients with folic acid and vitamin B12 deficient have increased level of oxidative stress due to hyper-homocysteinemia. One of the major complications related to diabetes is peripheral neuropathy which has been associated with hyper-homocysteinemia. Therefore, vitamin B12 deficiency might be considered as a predisposing factor to diabetic complications (140–142). The combined treatment of pyridoxine, folate, and vitamin B12 has shown the ability to recover retinal edema and hyperlight sensitivity in individuals with diabetic retinopathy (143) and also provide protection in conjunction with hyper-homocysteinemia (118). The actions of B vitamins on brain glucose metabolism and diabetes has been summarized in table 2 and illustrated in the figure 3.

Figure 3.png

Fig. 3. Effects  of B vitamins in brain glucose metabolism and diabetes.

     This figure illustrated  how B vitamins functions in the brain glucose metabolism and reduce the predisposition of diabetes through its different antioxidant and metabolic activities and also through the production of different neurohormones.

 PDH: pyruvate dehydrogenase, α-KGDH: α-ketogluterate dehydrogenase, FAD: flavin adenine dinucleotide, FMN: flavin mono nucleotide, GLUT-4: glucose transporter-4, NAD: nicotinamide adenine dinucleotide, TCA: tricarboxylic acid, FOXO: forkhead transcription factor, GABA: γ aminobutyric acid, ACC: acetyl- CoA carboxylase, GCC: geranyl-CoA carboxylase, MCC: 3-methylcrotonyl-CoA carboxylase, PC: pyruvate carboxylase, PCC: propionyl CoA carboxylase, UC: urea carboxylase.

Table 2  Summary of the protective action of B-vitamins against diabetes and neuro-disorders.

 

Names

Study

N

Route  and duration  

Parameters measured

Key findings

Ref.

B1,B6

Human study

30

Orally daily for 5 months

Glucose, HbA1c, Insulin, TPP, PLP, AGE DNA

Combined administration of vitamin B1 and B6 to   diabetic nephropathy patients declines DNA glycation in leucocytes.   Surprisingly B6 alone did not show any effect.

(144)

B1

Randomized control trial

12

Orally    daily for 1 month

Blood glucose

HbA1, creatinine, total cholesterol

Significant decrease in the level of   glucose and leptin in drug naïve patients with T2DM

(145)

B2

Experimental study

6

Orally  daily   for 28 days.

Glucose, MDA level, GSH, antioxidant enzymes, GLUT4   expression-comet assay

From this study, it has been concluded that   riboflavin acts as an antioxidant against lipid, protein, and DNA damage as   well as can reduce diabetic complications.

(103)

B3

Randomized controlled trials

2110

NA

Total cholesterol, LDL, HDL, plasma   glucose, HbA1c, hemoglobin

 

This study reported that niacin   supplementation can improve lipid profile without interfering with the glycaemic   level in T2DM patients.

(146)

B6

Experimental study

4

  In  drinking water for 70 days

Glucose, insulin, AGE

Administration of pyridoxamine dose-dependently can   reduce AGE formation and insulin resistance in obese diabetic KK-Ay   mice. However, fasting blood glucose level has not been altered.

(147)

B6+

B9+

B12

Human study

20 (eyes)

Orally L-methyl folate calcium, PLP, and    methyl cobalamin twice daily for 6   months.

Microperimetry measurement, mean retinal   central thickness

In this pilot study, B vitamins significantly   improve mean retinal threshold sensitivity for eyes with mild to moderate non-proliferative   diabetic retinopathy.

(143)

B7

Experimental study

15

Oral supplementation of Cr-Pic and biotin is given   for a period of 70 days.

Circulating glucose, cortisol, total cholesterol,   MDA

This study reported that chromium picolinate (Cr-Pic)   and biotin combinedly ameliorate insulin resistance, dyslipidemia, hepatic   and renal damage as well as pathophysiological alterations in the liver,   kidney, and pancreatic tissues in high-fat diet and streptozotocin-induced   T2DM.

(148)









     TPP: thiamine pyrophosphate, PLP: pyridoxal-5-phosphate, AGE: advanced glycation end product, MDA: malonaldehyde, GSH: reduced glutathione, LDL: low density lipoprotein, HDL: high density lipoprotein

.

4. IMPORTANCE OF MELATONIN IN BRAIN ENERGY METABOLISM AND DIABETES

     Melatonin is present  in almost every living organism to regulate the circadian rhythm of these organisms (149). In vertebrates, the direct release of melatonin into the blood and CSF by the pineal gland is the major source of circulating melatonin, however, it is also generated in a variety of cells, tissues, and organs primarily for local consumption (autocrine and paracrine activities) (150, 151). The suprachiasmatic nucleus (SCN) of the hypothalamus also known as the primary circadian clock, precisely regulates the activation and deactivation of this intricate neuronal network controlling pineal melatonin synthesis. Melatonin production exhibits a circadian rhythm through this pathway, which is closely synchronized with the cycle of light and dark (152). Melatonin is essential for the preservation of the internal circadian temporal rhythm, synchronizing a variety of physiological activities, such as energy metabolism, and ensuring that they occur in harmony (153). It is used to improve  clinical outcomes of several neurological diseases including dementia, AD, PD, MS, stroke, and brain ischemia/reperfusion but also in traumatic CNS injuries (traumatic brain and spinal cord injury). Additionally, experimental data support its direct and indirect antioxidant properties, including scavenging free radicals, promoting antioxidant enzymes, boosting the activities of other antioxidants, or shielding other antioxidant enzymes from oxidative damage. Melatonin is a neuroprotective substance in neurodegenerative pathologies when brain oxidative damage has been established as a common link (154).

4.1. Regulation of melatonin in the brain.

     Originally, melatonin was believed only to produce in the pineal gland but subsequent studies by  Stefulj et al. showed that gene expressions of melatonin-synthesizing enzymes AANAT and hydroxyl-indole-O-methyltransferase (HIOMT) are also  in other regions of the rat brain (155). Currently, it is well known that melatonin is exclusively synthesized in the mitochondrial matrix of the mouse brain (156), indicating every cell can synthesize melatonin in brain or in other part of the body. Liu et al. have reported that melatonin is synthesized in cultured cortical astrocytes in rats through a cascade of enzymatic events (157). For example,  serotonin (which is derived from tryptophan) is converted into N-acetyl serotonin (NAS) by the enzyme AANAT. N-acetyl serotonin is converted to melatonin utilizing the enzyme HIOMT (158).

     Melatonin acts through two G protein-coupled receptors namely MT1 and MT2 which are widely distributed in the CNS including the hippocampus, caudate putamen, suprachiasmatic nucleus, reticular thalamic nucleus, supraoptic nucleus, and inferior colliculus (159). Additionally, both receptors are also widely distributed in neurons, glial cells of the cerebral cortex, cerebellar cortex, and thalamus (160). Although several authors proposed the existence of a putative MT3 melatonin receptor subtype, recent studies have provided evidence that the MT3 melatonin receptor is a cytosolic quinine reductase 2 enzyme rather than a membrane receptor (161). Intracellular melatonin  can act  on the mitochondrial MT1 signal-transduction pathway, which prevents the release of the stress-induced cytochrome c and the activation of the caspase enzyme, therefore, reduces inflammation  and cell death. The locally produced melatonin seems to guard against neurodegeneration, known as autocrine signaling (162). However, there are numerous ways that melatonin works through its membrane receptors to prevent oligodendroglial damage, including improved membrane fluidity, decreased edema, polymorphonuclear cell infiltration, prevention of nuclear factor-kappa-B (NF-ԟB) translocation to the nucleus, and the subsequent reduction of pro-inflammatory cytokines expression, all of which have important roles in inflammation (163, 164). Melatonin may also affect astrocyte reactivity or death by increasing astrocyte anti-oxidative defense (165). Melatonin receptors and its effect on inflammatory regulation appear to involve in the neuroprotective pathway, which promotes oligodendrocyte maturation (154).

4.2. Regulation of brain energy metabolism by melatonin.

     The brain is particularly vulnerable to injury caused by free radicals because of its inherent biochemical and physiological traits, which include a high need for energy and polyunsaturated fatty acids (PUFA) (166). The accumulation of abnormal or misfolded proteins, protofibril formation, dysfunction of the ubiquitin-proteasome system, excitotoxic insult, oxidative and nitrosative stress, mitochondrial injury, and failure of axonal and dendritic transport are thought to be common unifying events in many slowly progressing neurodegenerative disorders (167). However, mitochondria play a central role in causing neural damage. They are the main source of ROS. Their dysfunction  results in an abrupt reduction in ATP production and a decrease in tricarboxylic acid (TCA) cycle activity. This dysfunction in energy production processes contributes to various brain pathologies (168). Melatonin is primarily responsible for establishing an optimal energy balance by controlling the flow of energy to and from the reserves, as well as directly regulating energy expenditure by activating brown adipose tissue and browning of white adipose tissue. Melatonin levels fall as a result of aging, shift work, or light settings at night, which also causes insulin resistance, glucose intolerance, sleep disturbances, and metabolic circadian disorganization, a condition known as chrono disruption (149). Melatonin regulates energy balance and metabolism, which have a distinct 24-hour rhythm (169). In general, the active/wakefulness phase of the day is linked to energy acquisition, eating, and subsequent energy intake, use, and storage. This is a period characterized by increased adipose tissue lipogenesis and adiponectin production, as well as elevated insulin secretion, high glucose uptake by insulin-sensitive tissues, hepatic and muscular glycogen synthesis and glycolysis, and high central and peripheral sensitivity to insulin and glucose tolerance. The normal fasting phase necessitates the utilization of stored energy for the maintenance of cellular functions, which distinguishes the rest/sleep phase of the day, in contrast with insulin resistance, enhanced hepatic gluconeogenesis and glycogenolysis, adipose tissue lipolysis, and leptin release are all present throughout this portion of the daily cycle (149) thus it also regulates brain metabolism in that way. Melatonin interacts with molecules and signalling networks such as insulin-like growth factor 1 (IGF-1), Forkhead box O (FoxOs), sirtuins, and the mammalian target of rapamycin (mTOR) signalling pathways and all of them affect energy consumption (170).

4.3. Crosstalk between brain, melatonin, and diabetes.

     The T2DM and its associated disorders, such as obesity, abnormal protein processing, oxidative stress, and proinflammatory cytokines will activate inflammatory pathways, resulting in low-grade chronic inflammation, insulin resistance, and impaired neuronal insulin signaling in the periphery (171). However, T2DM is a potential risk factor associated with the development of AD (172). The BBB disruption associated with T2DM, along with elevated levels of ROS, MMP-2, and IFNᵧ, may facilitate the entry of circulating neurotoxins into the brain due to selectivity loss on neurons and ultimately accelerate the course of AD (173). The M1 proinflammatory state that results from macrophage infiltration causes an increased release of proinflammatory cytokines and chemokines, which can cross the BBB and disrupt brain functioning (174). These alterations can be protected by melatonin through inducing the expression  of NADPH oxidase-2 and inhibition of MMP-9 in brain microvascular endothelial cells (175). Another proposed underlying mechanism is the  impairment of insulin signaling (176). AD patients even without T2DM also showed signs of insulin resistance. Through the MAPK and Akt signalling pathways as well as Nrf2 activation, and insulin receptor activation promotes glucose uptake, mitochondrial activity, anti-apoptosis, and autophagy (177). Insulin thus plays a crucial role in the development, maintenance, and function of neurons in addition to being a hormone for glucose homeostasis (178).

     Evidence has supported the circadian rhythm of melatonin in pancreatic insulin secretion (179, 180). Melatonin deems to has a negative influence on the activity of the β cells to control their  insulin decreasing as well as a reduction in glucose tolerance in rats (181, 182). High insulin and low glucose level has been found during the day when melatonin concentration is reduced. Paradoxically low insulin level was also measured in presence of high melatonin and glucose level during the night (183). In an, in vivo and in vitro study it was found that melatonin affects insulin secretion that is mediated by the MT1 and MT2 receptors. Melatonin also shows protective effects on pancreatic β-cells, which are extremely vulnerable to oxidative stress, and ROS formation. Comparing diabetic and nondiabetic rats, diabetic rats have decreased plasma melatonin levels and AANAT activity. The insulin receptor mRNA increased while AANAT mRNA dropped in the pineal gland, indicating a close relationship between insulin and melatonin (184). By these findings, melatonin concentration declines with age whereas insulin synthesis are increased which proposes the fact that melatonin inhibits age-related hyperinsulinemia (185). Corresponding to these findings the reduction of melatonin levels in diabetic hamsters has been reported (186–188). Melatonin supplementation along with exercise elevated the expression of genes involved in mitochondrial biogenesis and function in T2DM rats, including mtTFA, PGC1α, Nrf-1, and Nrf-2. Additionally, taking melatonin and working out together effectively scavenges harmful free radicals, indicating that melatonin administration exerts its anti-diabetic effects by counteracting oxidative pathways (189).

     On the other hand, melatonin has a preventive role in hyperglycemia whereas pinealectomy increases the risk (190). The gene expression of GLUT4 is reduced in pinealectomized rats resulting in glucose intolerance and insulin resistance which is ameliorated by melatonin (191). Furthermore, melatonin level is decreased in human dental pulp cells in T2DM subjects. Melatonin at pharmacological doses was found to improve SOD activity in hyperglycemic dental pulp cells (191). Another study by Oliveira et al. described that streptozotocin-induced diabetic rats with the treatment of insulin (1.5 U/100gm/day) and melatonin (0.2mg/kg/day in drinking water) improves insulin sensitivity as well as glucose metabolism of white adipose tissues (192). Decreased melatonin production has been reported in diabetic animal models which were protected by exogenous melatonin administration (100mg/kg/day in drinking water for 8 weeks) in the high-fat diet-fed mouse with insulin resistance (193). In patients with poorly controlled T2DM addition of melatonin and zinc acetate to metformin reveals better results than metformin alone (194). Thus, melatonin plays an important role in preventing the development of diabetes by modulating Aβ build-up, insulin resistance, glucose metabolism, and BBB permeability.

4.4. Role of melatonin against diabetes-induced alterations in CNS.

     Diabetes perturbs various structural and functional integrity related to the PNS and CNS (195). Cerebrovascular alterations such as neurotropic changes like decreased IGF, loss of vascular reactivity, and reduced blood flow in the brain indicate apoptosis in neuronal cells and subsequent cognitive impairment in the hippocampus (196, 197). DM is associated with poor cognitive development, AD, and dementia. Since oxidative stress is a major contributory factor in the development of diabetic complications, antioxidant administration may restore physiological functions to some extent (198–200). A plethora of studies have reported that diabetes-induced hippocampal neuronal cell damage has been prevented by antioxidant therapy (201–203). Melatonin by its antioxidant efficacy provides neuroprotective effects in different circumstances (204, 205). Melatonin administration prevents glial fibrillary acidic protein (GFAP) and S100B, prime astrogliosis indicator, and reduces lipid peroxidation in a streptozotocin-induced diabetic rat model (206,207). Melatonin exerts its function by enhancing the brain's antioxidant machinery, decreasing brain NOSs activity, and plasma cytokine levels inhibiting apoptosis, and maintaining CNS homeostasis (208–210). Another protective mechanism of melatonin lies in the inhibition of oxidative stress-mediated poly (ADP ribose) polymerases (PARPs) hyperactivation which can cause excess cellular energy depletion and cell death (211). A combined therapy of nicotinamide (300, 1000 mg/kg/day) and melatonin (3, 10 mg/kg/day) ameliorated high glucose-induced alterations in GABA neurotransmitter and glutamate levels in diabetic rats. The study revealed that suppression of the PARPs overactivation pathway can be beneficial in the treatment of DM-related CNS disorders (212).

4.5. Role of melatonin against diabetes-induced neuropathy.

     One of the most common and disabling consequences  of DM is diabetic neuropathy (DN), which harms the patient’s quality of life and imposes a considerable economic burden on the healthcare system. DN is a common adverse consequence associated with both type 1 and type 2 diabetes. Additionally, several studies have revealed that various kinds of neuropathy are also present in pre-diabetic patients (213). Worldwide, the prevalence of DN in diabetes patients ranges from 9.6 to 88.7%. Age, types of diabetes, glucose management, severity of the disease , and the accessibility of medical facilities could all contribute to this difference. Distal axon branching in particular, neurofilament polymers, which are crucial structural scaffolds of the axon, gradually cease under  chronic DM. This loss of neurofilament polymers is thought to be triggered by decreased mRNA expression of the neurofilament gene (214). Animal studies  in diabetic rats also link endoplasmic reticulum stress to the peripheral nerve damage caused by diabetes, which would impact nerve function (215). Similarly, it has been shown that hyperglycemia changes the expression patterns of heat shock proteins (HSPs) and PARP as well as the function of important plasticity molecules including growth-associated protein 43 (GAP43; also known as neuromodulin) and tubulin (216, 217). Data indicate that disruption in these networks promotes aberrant protein processing, oxidative damage, and mitochondrial dysfunction, which results in the loss of peripheral nerve function even if the exact causes of injury are yet unknown (218). Moreover, increased glucose loads cause glucose metabolism to occur through the polyol and hexosamine pathways, increasing ROS and causing inflammation, both of which are mostly caused by mitochondrial damage (218). However, melatonin has a beneficial impact on diabetic neuropathy by reducing oxidative stress and inflammatory reactions. Numerous metabolic processes, including the polyol pathway and the mitochondrial energy-production complexes, are activated by hyperglycemia, which results in the release of reactive intermediates. Additionally, a high glucose level activates the PKC pathway, which results in NF-kB activation. Melatonin reduces oxidative stress and inflammation, suppresses reactive intermediates and NF-kB, activates Nrf2, and up-regulates the production of antioxidant enzymes, all of which prevent the development of diabetic neuropathy (219). In DN patients, hyperglycemia has been linked to higher concentrations of leptin, malonaldehyde (MDA), and pro-inflammatory cytokines (TNF- and IL-6). Leptin, pro-inflammatory cytokines, and hyperglycemia were all attenuated in the melatonin-treated group, while levels of adiponectin, an anti-inflammatory adipokine, were elevated. Additionally, melatonin lowered brain MDA levels and reduced oxidative stress while raising total antioxidant capacity and GSH levels (213). The effects melatonin in brain glucose metabolism and diabetes are summarized in table 1 and illustrated in Figure 4.

Figure 4.png

Fig. 4. Effects  of melatonin in brain glucose metabolism and diabetes.

     This figure illustrated  melatonin synthesis pathways as well as its regulatory action on  brain function and  energy metabolism. Melatonin can  reduce  the CNS complications related to diabetes mellitus by upregulating, downregulating, inhibiting or activating the different signals and protein expressions and reducing oxidative stress under the condition of  diabetes mellitus.

     TPH: tryptophan hydroxylase, 5-HTTP: 5-hydroxytryptophan, AAAD: aromatic amino acid decarboxylase, 5-HT: hydroxy tryptamine, AANAT: arylalkalylamine-N-acetyl transferase, NAS: N-acetyl serotonin, HIOMT: hydroxyindole-O-methyl transferase, MT1: melatonin recpetor-1, FOXO3a: forkhead transcription factor, NF-kB: nuclear factor-kappa-B, IGF 1: insulin-like growth factor 1, MMP: matrix metalloprotenase 2, IFN γ: interferon gamma, NADPH: nicotinamide adenine dinucleotide phosphate (reduced), AMPK: activated protein kinaseMAPK: mitogen-activated protein kinas, PARP: poly ADP ribose polymerase.

4.6. Summarization  of the protective effects of melatonin on diabetes.

Type of Study

Dose and duration

Route

Parameters measured

Key findings

Ref.

In vivo

10mg/kg b.w in male Wistar rats for 2 weeks.

I.p.

Total antioxidant capacity, GSH, GPx lipid   peroxidation, IL-1β, and IL-4

Melatonin ameliorates diabetes-induced brain and   erythrocyte damage through its    antioxidant efficacy

(209)

In vivo

10 mg/kg b.w in male Wistar rats for 7   days.

I.p.

Specific neuronal protein, neural and   glial markers like neural cell adhesion molecule (NCAM) and glial fibrillary   acidic protein (GFAP), and lipid peroxidation.

Melatonin improves NCAM, GFAP, and   cognitive behaviour.

(207)

In vivo

1 mg/kg b.w in male Wistar rats for 4 weeks

I.p.

Blood glucose, hematological parameters RBC, WBC,   hemoglobin, hematocrit

Melatonin inhibits membrane protein peroxidative   damage, and non-enzymatic glycosylation of protein, and prevents diabetes-induced   hematological alterations in rats.

(208)

In vivo

10mg/kg b.w in male Wistar rats for 2   weeks.

I.p.

NOS

Melatonin administration is beneficial   in reducing oxidative stress-induced diabetes by suppressing NOS

(210)

In vivo

3 and 10mg/kg b.w in Sprague- Dawley rats for 2   weeks

I.p.

Glutamate, GABA, MDA level. Oxidative stress-PARP   pathway

Melatonin intervention inhibits oxidative   stress-PARP pathway and improves neuro behavioural pathway

(211)

In vivo

50 mg/kg b.w in male Wistar rats for 45   days.

I.p.

MDA

GPx

CAT

Immuno-

histochemistry

TEM

Melatonin significantly enhances   antioxidant enzyme activity and provides neuroprotective effects by   preserving glial cells and declining vascular damage.

(220)

In vivo

3,10 mg/kg b.w in Sprague-Dawley rats

Oral

TNF-α, IL-6, Nrf-2, COX-2

Melatonin modulates neuroinflammation in diabetic   rats by reducing inflammatory markers like TNF-α and enhancing Nrf-2   expression.

(221)

     i.p.: intraperitoneal, GPx: glutathione peroxidase, NCAM: neural cell adhesion molecule, GFAP: glial fibrillary acidic protein, RBC: red blood cells, WBC: white blood cells, NOS: nitric oxide synthase, GABA- γ: aminobutyric acid, MDA: malonaldehyde, PARP: poly ADP ribose polymerase, CAT: catalase, TEM: transmission electron microscope.


5. CONCLUSION

     In the brain, glucose metabolism seems to be the principal mechanism for producing energy supporting synaptic activity and ionic equilibrium. Memory and other cognitive and emotional functions are boosted or controlled by insulin in the brain, which also centrally modulates body metabolism. T2DM is characterized by insulin resistance, which also contributes to inflammation and oxidative stress in addition to hyperglycemia. Furthermore, uncontrolled diabetes mellitus can impact the brain’s general oxidative metabolism in a frightful condition. The biochemistry and functioning of the brain are substantially impacted by severe hyperglycemia. Thus, the development of diabetes through impaired brain metabolism and diabetes-induced disruption in brain metabolism are interrelated and vice versa process involves various underlying mechanisms. B vitamins and melatonin  play  important roles in brain metabolism under diabetic condition. The B vitamins operate as cofactors in many catalytic reactions required for synthesizing and activating neurotransmitters, allowing the CNS to function properly. Whereas, melatonin, as a broad-spectrum antioxidant, scavenges free radicals in several physiological circumstances. We can therefore conclude from this analysis that melatonin and B vitamins may be crucial for the preservation of brain glucose metabolism and the prevention of the onset of diabetes mellitus. Melatonin can therefore be an alternative when combined with B vitamins, and this pairing can result in more assertive outcomes for protecting brain metabolism-related diabetes mellitus.


ACKNOWLEDGEMENTS

     Manisha Mukhopadhyay thankfully acknowledges the receipt of UGC Senior Research Fellowship (ref no.1558/(NET-JULY  2018), Govt. of India. Priyanka Ghosh thankfully acknowledges the receipt of the UGC Junior Research Fellowship (ref no.1571/PWD (NET-NOV2017), Govt. of India. AC is supported from a research project awarded to her by the Department of Science and Technology, Govt. of West Bengal, India. DB is also supported from departmental BI grant of University of Calcutta.


AUTHORS CONTRIBUTION

     Dr. DB and Dr. AC contributed to the conception and critical revision of the manuscript and approved it. MM drafted the manuscript and prepared the table. PG prepared figures, table and edited the manuscript.

 

CONFLICT OF INTEREST

     The authors declare no conflict of interest.


ABBREVIATIONS

AANAT- Arylalkalylamine- N-Acetyl Transferase

ACC-Acetyl- Coa Carboxylase

Acetyl CoA- Acetyl Coenzyme A

AD- Alzheimer’s disease

ADP-Adenosine Diphosphate

AFMK- N1-Acetyl-N2-Formyl-5-Methoxykynuramine

AMK- N1-Acetyl-5-Methoxykynuramine

AMPK-Activated Protein Kinase

ATP-Adenosine Triphosphate

BBB-Blood-Brain Barrier

CAM- Cell Adhesion Molecules

CAMP- Adenosine Monophosphate

CNS-Central Nervous System

CSF-Cerebro Spinal Fluid

DNA-Deoxyribonucleic Acid

DN-Diabetic Neuropathy

ETC-Electron Transport Chain

FAD-Flavin Adenine Dinucleotide

FADH2- Flavin Adenine Dinucleotide (Reduced)

FMN-Flavin Adenine Mononucleotide

FOXO3a-Forkhead Transcription Factor

FoxOs -Forkhead Box O

G6P- Glucose-6-Phosphate

G6pc- Glucose-6-Phosphatase

GABA-γ Aminobutyric Acid

GAP 430- Growth-Associated Protein 43

GCC-Geranyl-Coa Carboxylase

GDE-Glycogen Debranching Enzyme

GDM- Gestational Diabetes Mellitus

GFAP- Glial Fibrillary Acidic Protein

GLUT-Glucose Transporter

GP-Glycogen Phosphorylase

GS- Glycogen Synthase

HDL-High Density Lipoprotein

HIOMT-Hydroxyindole-O-Methyl Transferase

HSPs- Heart Shock Proteins

IFN γ- Interferon Gamma

IGF 1- Insulin-Like Growth Factor 1

LDL-Low Density Lipoprotein

MAPK- Mitogen-Activated Protein Kinase

MCC-3- Methylcrotonyl-Coa Carboxylase

MDA- Malonaldehyde

MMP 2- Matrix Metalloprotenase 2

MS-Multiple Sclerosis

MT1-Melatonin Receptor1

MT2-Melatonin Receptor 2

MT3-Melatonin Receptor 3

mTOR -Mammalian Target Of Rapamycin

NA- Nicotinic Acid

NADH- Nicotinamide Adenine Dinucleotide (reduced)

NAD-Nicotinamide Adenine Dinucleotide

NADPH-Nicotinamide Adenine Dinucleotide Phosphate (reduced)

NADP-Nicotinamide Adenine Dinucleotide Phosphate

NAS- N-Acetyl serotonin

NF-kB Nuclear Factor-Kappa-B

NNMT-Nicotinamide-N-Methyl Transferase

Nrf2-Nuclear Factor 2

PARP- Poly ADP Ribose Polymerase

PC- Pyruvate Carboxylase

PCC- Propionyl CoA Carboxylase

Pck1- Phosphoenolpyruvate Carboxykinase 1

PD-Parkinson’s Disease

PKC-Protein Kinase C

PK-Phosphorylase Kinase

PLP- Pyridoxal-5-Phosphate

PPP-Pentose Phosphate Pathway

PUFA-Polyunsaturated Fatty Acids

RNA-Ribonucleic Acid

ROS-Reactive Oxygen Species

SCN- Suprachiasmatic Nucleus

SGLT- Sodium-Dependent Glucose Transporters

T2DM-Type 2 Diabetes

TCA cycle-Tri Carboxylic Acid Cycle

TDP-Thiamine Diphosphate

TMP-Thiamine Monophosphate

TPP- Thiamine Pyrophosphate

UC- Urea Carboxylase

WE-Wernicke’s Encephalopathy

α-KGDH- Alpha Ketogluterate Dehydrogenase


REFERENCES

 

  1. DeBerardinis RJ, Thompson CB (2012) Cellular metabolism and disease: what do metabolic outliers teach Us? Cell148 (6): 1132–1144. DOI: 10.1016/j.cell.2012.02.032.

  2. Attwell D, Laughlin SB (2001) An energy budget for signaling in the grey matter of the brain. J. Cereb. Blood Flow Metab. 21 (10): 1133–1145. DOI: 10.1097/00004647-200110000-00001.

  3. Howarth C, Gleeson P, Attwell D (2012) Updated energy budgets for neural computation in the neocortex and cerebellum. J. Cereb. Blood Flow Metab. 32 (7): 1222–1232. DOI: 10.1038/jcbfm.2012.35.

  4. Mergenthaler P, Lindauer U, Dienel GA, Meisel A (2013) Sugar for the brain: the role of glucose in physiological and pathological brain function. Trends Neurosci. 36 (10): 587–597. DOI: 10.1016/j.tins.2013.07.001.

  5. Brown AM, Ransom BR (2007) Astrocyte glycogen and brain energy metabolism. Glia 55 (12): 1263–1271. DOI: 10.1002/glia.20557.

  6. Zoccoli G, Silvani A, Franzini C 2017. (2017) Sleep and the peripheral vascular system. In: Reference Module in Neuroscience and Biobehavioral Psychology. Elsevier; 2017. p. 563–567. DOI: 10.1016/B978-0-12-809324-5.00964-0.

  7. FA. Beltrán, AI. Acuña MM and MC 2012. (2012) Brain energy metabolism in health and disease. In: Neuroscience - Dealing With Frontiers. In Tech; 2012. DOI: 10.5772/36092.

  8. Mazziotta JC, Phelps ME, Pahl JJ, Huang S-C, Baxter LR, Riege WH, Hoffman JM, Kuhl DE, Lanto AB, Wapenski JA, Markham CH (1987) Reduced cerebral glucose metabolism in asymptomatic subjects at risk for huntington’s disease. N. Engl. J. Med. 316 (7): 357–362. DOI: 10.1056/NEJM198702123160701.

  9. Mosconi L, Tsui WH, De Santi S, Li J, Rusinek H, Convit A, Li Y, Boppana M, de Leon MJ (2005) Reduced hippocampal metabolism in MCI and AD: Automated FDG-PET image analysis. Neurology 64 (11): 1860–1867. DOI: 10.1212/01.WNL.0000163856.13524.08.

  10. Leenders KL, Frackowiak RSJ, Quinn N, Marsden CD (1986) Brain energy metabolism and dopaminergic function in Huntington’s disease measured in vivo using positron emission tomography. Mov. Disord. 1 (1): 69–77. DOI: 10.1002/mds.870010110.

  11. Hattingen E, Magerkurth J, Pilatus U, Mozer A, Seifried C, Steinmetz H, Zanella F, Hilker R (2009) Phosphorus and proton magnetic resonance spectroscopy demonstrates mitochondrial dysfunction in early and advanced Parkinson’s disease. Brain 132 (12): 3285–3297. DOI: 10.1093/brain/awp293.

  12. Duarte J (2014) Metabolic alterations associated to brain dysfunction in diabetes. Aging Dis. 6 (5): 304–321. DOI: 10.14336/ad.2014.1104.

  13. Frisardi V, Solfrizzi V, Seripa D, Capurso C, Santamato A, Sancarlo D, Vendemiale G, Pilotto A, Panza F (2010) Metabolic-cognitive syndrome: A cross-talk between metabolic syndrome and Alzheimer’s disease. Ageing Res. Rev. 9 (4): 399–417. DOI: 10.1016/j.arr.2010.04.007.

  14. Spauwen PJJ, Köhler S, Verhey FRJ, Stehouwer CDA, van Boxtel MPJ (2013) Effects of type 2 diabetes on 12-year cognitive change. Diabetes Care36 (6): 1554–1561. DOI: 10.2337/dc12-0746.

  15. Craft S, Peskind E, Schwartz MW, Schellenberg GD, Raskind M, Porte D (1998) Cerebrospinal fluid and plasma insulin levels in Alzheimer’s disease. Neurology 50 (1): 164–168. DOI: 10.1212/WNL.50.1.164.

  16. Hurrle S, Hsu WH (2017) The etiology of oxidative stress in insulin resistance. Biomed. J. 40 (5): 257–262. DOI: 10.1016/j.bj.2017.06.007.

  17. Dugan L, Sensi S, Canzoniero L, Handran S, Rothman S, Lin T, Goldberg M, Choi D (1995) Mitochondrial production of reactive oxygen species in cortical neurons following exposure to N-methyl-D-aspartate. J. Neurosci. 15 (10): 6377–6388. DOI: 10.1523/JNEUROSCI.15-10-06377.1995.

  18. Jackson GR, Werrbach-Perez K, Pan Z, Sampath D, Perez-Polo JR (1994) Neurotrophin regulation of energy homeostasis in the central nervous system. Dev. Neurosci. 16 (5–6): 285–290. DOI: 10.1159/000112121.

  19. Garcia-Serrano AM, Duarte JMN (2020) Brain metabolism alterations in type 2 diabetes: what did we learn from diet-induced diabetes models? Front. Neurosci. 14 (229): 1–11. DOI: 10.3389/fnins.2020.00229.

  20. Muriach M, Flores-Bellver M, Romero FJ, Barcia JM (2014) Diabetes and the brain: oxidative stress, inflammation, and autophagy. Oxid. Med. Cell. Longev. 2014 (102158): 1–9. DOI: 10.1155/2014/102158.

  21. Mukhopadhyay M, Majumder R, Banerjee A, Bandyopadhyay D (2022) A comparative overview on the role of melatonin and vitamins as potential antioxidants against oxidative stress induced degenerative infirmities: An emerging concept. Melatonin Res. 5 (3): 254–277. DOI: 10.32794/mr112500131.

  22. Tan D-X, Manchester LC, Hardeland R, Lopez-Burillo S, Mayo JC, Sainz RM, Reiter RJ (2003) Melatonin: a hormone, a tissue factor, an autocoid, a paracoid, and an antioxidant vitamin. J. Pineal Res. 34 (1): 75–78. DOI: 10.1034/j.1600-079X.2003.02111.x.

  23. Darnton-Hill I (2019) Public health aspects in the prevention and control of vitamin deficiencies. Curr. Dev. Nutr. 3 (9): 1–14. DOI: 10.1093/cdn/nzz075.

  24. Kennedy D (2016) B vitamins and the Brain: mechanisms, dose and efficacy—A review. Nutrients 8 (2): 68. DOI: 10.3390/nu8020068.

  25. Smith AG, Croft MT, Moulin M, Webb ME (2007) Plants need their vitamins too. Curr. Opin. Plant Biol. 10 (3): 266–275. DOI: 10.1016/j.pbi.2007.04.009.

  26. Homocysteine Studies Collaboration (2002) Homocysteine and risk of ischemic heart disease and stroke. JAMA 288 (16): 2015. DOI: 10.1001/jama.288.16.2015.

  27. Seshadri S, Beiser A, Selhub J, Jacques PF, Rosenberg IH, D’Agostino RB, Wilson PWF, Wolf PA (2002) Plasma homocysteine as a risk factor for dementia and Alzheimer’s disease. N. Engl. J. Med. 346 (7): 476–483. DOI: 10.1056/NEJMoa011613.

  28. Smulders YM, Blom HJ (2011) The homocysteine controversy. J. Inherit. Metab. Dis. 34 (1): 93–99. DOI: 10.1007/s10545-010-9151-1.

  29. Reppart SM, Weaver DR, Godson C (1996) Melatonin receptors step into the light: cloning and classification of subtypes. Trends Pharmacol. Sci. 17 (3): 100–102. DOI: 10.1016/0165-6147(96)10005-5.

  30. Costa EJX, Lopes RH, Lamy-Freund MT (1995) Permeability of pure lipid bilayers to melatonin. J. Pineal Res. 19 (3): 123–126. DOI: 10.1111/j.1600-079X.1995.tb00180.x.

  31. Tan DX, Reiter R, Manchester L, Yan M, El-Sawi M, Sainz R, Mayo J, Kohen R, Allegra M, Hardelan R (2002) Chemical and physical properties and potential mechanisms: melatonin as a broad spectrum antioxidant and free radical scavenger. Curr. Top. Med. Chem. 2 (2): 181–197. DOI: 10.2174/1568026023394443.

  32. Reiter R, Tan DX, Manchester L, Terrón MP, Flores LJ, Koppisepi S (2007) Medical implications of melatonin: receptor-mediated and receptor-independent actions. Adv. Med. Sci. 52 : 11–28.

  33. Leston J, Harthé C, Brun J, Mottolese C, Mertens P, Sindou M, Claustrat B (2010) Melatonin is released in the third ventricle in humans. A study in movement disorders. Neurosci. Lett. 469 (3): 294–297. DOI: 10.1016/j.neulet.2009.12.008.

  34. Uz T, Ahmed R, Akhisaroglu M, Kurtuncu M, Imbesi M, Dirim Arslan A, Manev H (2005) Effect of fluoxetine and cocaine on the expression of clock genes in the mouse hippocampus and striatum. Neuroscience 134 (4): 1309–1316. DOI: 10.1016/j.neuroscience.2005.05.003.

  35. Uz T, Qu T, Sugaya K, Manev H (2002) Neuronal expression of arylalkylamine N-acetyltransferase (AANAT) mRNA in the rat brain. Neurosci. Res. 42 (4): 309–316. DOI: 10.1016/S0168-0102(02)00011-1.

  36. Tan D-X, Manchester L, Esteban-Zubero E, Zhou Z, Reiter R (2015) Melatonin as a potent and inducible endogenous antioxidant: synthesis and metabolism. Molecules 20 (10): 18886–18906. DOI: 10.3390/molecules201018886.

  37. Acuña-Castroviejo D, Escames G, LeÓn J, Carazo A, Khaldy H 2003. (2003) Mitochondrial regulation by melatonin And its metabolites. In: Advances in Experimental Medicine and Biology. 2003. p. 549–557. DOI: 10.1007/978-1-4615-0135-0_63.

  38. Rehman S, Ikram M, Ullah N, Alam S, Park H, Badshah H, Choe K, Ok Kim M (2019) Neurological enhancement effects of melatonin against brain injury-induced oxidative stress, neuroinflammation, and neurodegeneration via AMPK/CREB signaling. Cells 8 (7): 760. DOI: 10.3390/cells8070760.

  39. Turgut M, Kaplan S (2011) Effects of melatonin on peripheral nerve regeneration. Recent Pat. Endocr. Metab. Immune Drug Discov. 5 (2): 100–108. DOI: 10.2174/187221411799015336.

  40. Banerjee A, Chattopadhyay A, Bandyopadhyay D (2021) Potentially synergistic effects of melatonin and metformin in alleviating hyperglycaemia: a comprehensive review. Melatonin Res. 4 (4): 522–550. DOI: 10.32794/mr112500110.

  41. Erbsluh F, Bernsmeier A, Hillesheim HR (1958) The glucose consumption of the brain & its dependence on the liver. Arch. fur Psychiatr. und Nervenkrankheiten Ver. mit Zeitschrift fur die Gesamte Neurol. und Psychiatr. 196 (6): 611–626. DOI: 10.1007/BF00344388.

  42. Mathieu C, de la Sierra-Gallay IL, Duval R, Xu X, Cocaign A, Léger T, Woffendin G, Camadro J-M, Etchebest C, Haouz A, Dupret J-M, Rodrigues-Lima F (2016) Insights into brain glycogen metabolism. J. Biol. Chem. 291 (35): 18072–18083. DOI: 10.1074/jbc.M116.738898.

  43. Newgard CB, Hwang PK, Fletterick RJ (1989) The family of glycogen phosphorylases: structure and function. Crit. Rev. Biochem. Mol. Biol. 24 (1): 69–99. DOI: 10.3109/10409238909082552.

  44. Hossain MI, Roulston CL, Stapleton DI (2014) Molecular basis of impaired glycogen metabolism during ischemic stroke and hypoxia. Coles JA, editor. PLoS One 9 (5): e97570. DOI: 10.1371/journal.pone.0097570.

  45. Choi I-Y, Seaquist ER, Gruetter R (2003) Effect of hypoglycemia on brain glycogen metabolism in vivo. J. Neurosci. Res. 72 (1): 25–32. DOI: 10.1002/jnr.10574.

  46. Swanson RA, Sagar SM, Sharp FR (1989) Regional brain glycogen stores and metabolism during complete global ischaemia. Neurol. Res. 11 (1): 24–28. DOI: 10.1080/01616412.1989.11739856.

  47. Pfeiffer-Guglielmi B, Fleckenstein B, Jung G, Hamprecht B (2003) Immunocytochemical localization of glycogen phosphorylase isozymes in rat nervous tissues by using isozyme-specific antibodies. J. Neurochem. 85 (1): 73–81. DOI: 10.1046/j.1471-4159.2003.01644.x.

  48. Pellegri G, Rossier C, Magistretti PJ, Martin J-L (1996) Cloning, localization and induction of mouse brain glycogen synthase. Mol. Brain Res. 38 (2): 191–199. DOI: 10.1016/0169-328X(95)00305-C.

  49. Magistretti PJ, Sorg O, Martin J-L 1993. (1993) Regulation of glycogen metabolism in astrocytes: physiological, pharmacological, and pathological aspects. In: Astrocytes. Elsevier; 1993. p. 243–265. DOI: 10.1016/B978-0-12-511370-0.50015-1.

  50. Adeva-Andany MM, González-Lucán M, Donapetry-García C, Fernández-Fernández C, Ameneiros-Rodríguez E (2016) Glycogen metabolism in humans. BBA Clin. 5 : 85–100. DOI: 10.1016/j.bbacli.2016.02.001.

  51. Nadeau OW, Fontes JD, Carlson GM (2018) The regulation of glycogenolysis in the brain. J. Biol. Chem. 293 (19): 7099–7107. DOI: 10.1074/jbc.R117.803023.

  52. Camandola S, Mattson MP (2017) Brain metabolism in health, aging, and neurodegeneration. EMBO J. 36 (11): 1474–1492. DOI: 10.15252/embj.201695810.

  53. Shah K, DeSilva S, Abbruscato T (2012) The role of glucose transporters in brain disease: diabetes and alzheimer’s disease. Int. J. Mol. Sci. 13 (12): 12629–12655. DOI: 10.3390/ijms131012629.

  54. Dwyer DS, Vannucci SJ, Simpson IA (2002) Expression, regulation, and functional role of glucose transporters (GLUTs) in brain. Int. Rev. Neurobiol. 51 : 159–188. DOI: 10.1016/S0074-7742(02)51005-9.

  55. Reagan LP, Rosell DR, Alves SE, Hoskin EK, McCall AL, Charron MJ, McEwen BS (2002) GLUT8 glucose transporter is localized to excitatory and inhibitory neurons in the rat hippocampus. Brain Res. 932 (1–2): 129–134. DOI: 10.1016/S0006-8993(02)02308-9.

  56. Vannucci SJ, Koehler-Stec EM, Li K, Reynolds TH, Clark R, Simpson IA (1998) GLUT4 glucose transporter expression in rodent brain: effect of diabetes. Brain Res. 797 (1): 1–11. DOI: 10.1016/S0006-8993(98)00103-6.

  57. Joost H-G, Thorens B (2001) The extended GLUT-family of sugar/polyol transport facilitators: nomenclature, sequence characteristics, and potential function of its novel members. Mol. Membr. Biol. 18 (4): 247–256. DOI: 10.1080/09687680110090456.

  58. Reagan LP, Magariños AM, Yee DK, Swzeda LI, Van Bueren A, McCall AL, McEwen BS (2000) Oxidative stress and HNE conjugation of GLUT3 are increased in the hippocampus of diabetic rats subjected to stress. Brain Res. 862 (1–2): 292–300. DOI: 10.1016/S0006-8993(00)02212-5.

  59. Wright EM, Turk E, Hager K, Lescale-Matys L, Hirayama B, Supplisson S, Loo DD (1992) The Na+/glucose cotransporter (SGLT1). Acta Physiol. Scand. Suppl. 607 : 201–207.

  60. Koepsell H (2020) Glucose transporters in brain in health and disease. Pflügers Arch. - Eur. J. Physiol. 472 (9): 1299–1343. DOI: 10.1007/s00424-020-02441-x.

  61. Pardridge WM, Triguero D, Farrell CR (1990) Downregulation of blood-brain barrier glucose transporter in experimental diabetes. Diabetes 39 (9): 1040–1044. DOI: 10.2337/diabetes.39.9.1040.

  62. Cornford EM, Hyman S, Cornford ME, Clare-Salzler M (1995) Down-regulation of blood-brain glucose transport in the hyperglycemic nonobese diabetic mouse. Neurochem. Res. 20 (7): 869–873. DOI: 10.1007/BF00969700.

  63. Brownlee M (2001) Biochemistry and molecular cell biology of diabetic complications. Nature 414 (6865): 813–820. DOI: 10.1038/414813a.

  64. Choi I-Y, Lee S-P, Kim S-G, Gruetter R (2001) In vivo measurements of brain glucose transport using the reversible michaelis–menten model and simultaneous measurements of cerebral blood flow changes during hypoglycemia. J. Cereb. Blood Flow Metab. 21 (6): 653–663. DOI: 10.1097/00004647-200106000-00003.

  65. Gruetter R, Ugurbil K, Seaquist ER (2002) Steady-state cerebral glucose concentrations and transport in the human brain. J. Neurochem. 70 (1): 397–408. DOI: 10.1046/j.1471-4159.1998.70010397.x.

  66. Ueno M, Chiba Y, Matsumoto K, Murakami R, Fujihara R, Kawauchi M, Miyanaka H, Nakagawa T (2016) Blood-brain barrier damage in vascular dementia. Neuropathology 36 (2): 115–124. DOI: 10.1111/neup.12262.

  67. Edison P, Archer HA, Hinz R, Hammers A, Pavese N, Tai YF, Hotton G, Cutler D, Fox N, Kennedy A, Rossor M, Brooks DJ (2007) Amyloid, hypometabolism, and cognition in Alzheimer disease: An [11C]PIB and [18F]FDG PET study. Neurology 68 (7): 501–508. DOI: 10.1212/01.wnl.0000244749.20056.d4.

  68. Chen Z, Zhong C (2013) Decoding Alzheimer’s disease from perturbed cerebral glucose metabolism: Implications for diagnostic and therapeutic strategies. Prog. Neurobiol. 108 : 21–43. DOI: 10.1016/j.pneurobio.2013.06.004.

  69. Fernandez AM, Hernandez-Garzón E, Perez-Domper P, Perez-Alvarez A, Mederos S, Matsui T et al. (2017) Insulin regulates astrocytic glucose handling through cooperation with IGF-I. Diabetes 66 (1): 64–74. DOI: 10.2337/db16-0861.

  70. Gibbs ME, Lloyd HGE, Santa T, Hertz L (2007) Glycogen is a preferred glutamate precursor during learning in 1-day-old chick: Biochemical and behavioral evidence. J. Neurosci. Res. 85 (15): 3326–3333. DOI: 10.1002/jnr.21307.

  71. Duarte JMN, Morgenthaler FD, Gruetter R (2017) Glycogen supercompensation in the rat brain after acute hypoglycemia is independent of glucose levels during recovery. Neurochem. Res. 42 (6): 1629–1635. DOI: 10.1007/s11064-017-2178-z.

  72. Lee JH, Jahrling JB, Denner L, Dineley KT (2018) Targeting insulin for Alzheimer’s disease: mechanisms, status and potential directions. Perry G, Avila J, Moreira PI, Sorensen AA, Tabaton M, editors. J. Alzheimer’s Dis. 64 (s1): S427–S453. DOI: 10.3233/JAD-179923.

  73. Willette AA, Modanlo N, Kapogiannis D (2015) Insulin resistance predicts medial temporal hypermetabolism in mild cognitive impairment conversion to Alzheimer disease. Diabetes 64 (6): 1933–1940. DOI: 10.2337/db14-1507.

  74. Heni M, Kullmann S, Preissl H, Fritsche A, Häring H-U (2015) Impaired insulin action in the human brain: causes and metabolic consequences. Nat. Rev. Endocrinol. 11 (12): 701–711. DOI: 10.1038/nrendo.2015.173.

  75. Belenguer P, Duarte JMN, Schuck PF, Ferreira GC (2019) Mitochondria and the brain: bioenergetics and beyond. Neurotox. Res. 36 (2): 219–238. DOI: 10.1007/s12640-019-00061-7.

  76. Santos R, Correia S, Alves M, Oliveira PF, Cardoso S, Carvalho C, Duarte AI, Santos MS, Moreira PI (2014) Insulin therapy modulates mitochondrial dynamics and biogenesis, autophagy and tau protein phosphorylation in the brain of type 1 diabetic rats. Biochim. Biophys. Acta - Mol. Basis Dis. 1842 (7): 1154–1166. DOI: 10.1016/j.bbadis.2014.04.011.

  77. Guo L, Tian J, Du H (2017) Mitochondrial dysfunction and synaptic transmission failure in Alzheimer’s disease. J. Alzheimer’s Dis. 57 (4): 1071–1086. DOI: 10.3233/JAD-160702.

  78. Sonnay S, Gruetter R, Duarte JMN (2017) How energy metabolism supports cerebral function: insights from 13C magnetic resonance studies in vivo. Front. Neurosci. 11: 288. DOI: 10.3389/fnins.2017.00288.

  79. Alberini CM, Cruz E, Descalzi G, Bessières B, Gao V (2018) Astrocyte glycogen and lactate: New insights into learning and memory mechanisms. Glia 66 (6): 1244–1262. DOI: 10.1002/glia.23250.

  80. Duran J, Gruart A, Varea O, López-Soldado I, Delgado-García JM, Guinovart JJ (2019) Lack of neuronal glycogen impairs memory formation and learning-dependent synaptic plasticity in mice. Front. Cell. Neurosci. 13 . DOI: 10.3389/fncel.2019.00374.

  81. Soares AF, Nissen JD, Garcia‐Serrano AM, Nussbaum SS, Waagepetersen HS, Duarte JMN (2019) Glycogen metabolism is impaired in the brain of male type 2 diabetic Goto‐Kakizaki rats. J. Neurosci. Res. 97 (8): 1004–1017. DOI: 10.1002/jnr.24437.

  82. Kelly Á, Laroche S, Davis S (2003) Activation of mitogen-activated protein kinase/extracellular signal-regulated kinase in Hippocampal circuitry is required for consolidation and reconsolidation of recognition memory. J. Neurosci. 23 (12): 5354–5360. DOI: 10.1523/JNEUROSCI.23-12-05354.2003.

  83. Marinangeli C, Didier S, Vingtdeux V (2016) AMPK in neurodegenerative diseases: implications and therapeutic perspectives. Curr. Drug Targets 17 (8): 890–907. DOI: 10.2174/1389450117666160201105645.

  84. Guilarte TR (2009) Vitamin B6 and cognitive development: recent research findings from human and animal studies. Nutr. Rev. 51 (7): 193–198. DOI: 10.1111/j.1753-4887.1993.tb03102.x.

  85. Hutto BR (1997) Folate and cobalamin in psychiatric illness. Compr. Psychiatry 38 (6): 305–314. DOI: 10.1016/S0010-440X(97)90925-1.

  86. Dror DK, Allen LH (2008) Effect of vitamin B12 deficiency on neurodevelopment in infants: current knowledge and possible mechanisms. Nutr. Rev. 66 (5): 250–255. DOI: 10.1111/j.1753-4887.2008.00031.x.

  87. Spector R (2014) Vitamin transport diseases of brain: focus on folates, thiamine and riboflavin. Brain Disord. Ther. 03 (02). DOI: 10.4172/2168-975X.1000120.

  88. Spector R, Johanson CE (2007) Vitamin transport and homeostasis in mammalian brain: focus on vitamins B and E. J. Neurochem. 103 (2): 425–438. DOI: 10.1111/j.1471-4159.2007.04773.x.

  89. Uchida Y, Ito K, Ohtsuki S, Kubo Y, Suzuki T, Terasaki T (2015) Major involvement of Na + -dependent multivitamin transporter (SLC5A6/SMVT) in uptake of biotin and pantothenic acid by human brain capillary endothelial cells. J. Neurochem. 134 (1): 97–112. DOI: 10.1111/jnc.13092.

  90. Laforenza U, Patrini C, Rindi G (1988) Distribution of thiamine, thiamine phosphates, and thiamine metabolizing enzymes in neuronal and glial cell enriched fractions of rat brain. J. Neurochem. 51 (3): 730–735. DOI: 10.1111/j.1471-4159.1988.tb01805.x.

  91. Rao VLR, Richardson JS, Butterworth RF (1993) Decreased activities of thiamine diphosphatase in frontal and temporal cortex in Alzheimer’s disease. Brain Res. 631 (2): 334–336. DOI: 10.1016/0006-8993(93)91554-6.

  92. Héroux M, Raghavendra Rao VL, Lavoie J, Richardson JS, Butterworth RF (1996) Alterations of thiamine phosphorylation and of thiamine-dependent enzymes in Alzheimer’s disease. Metab. Brain Dis. 11 (1): 81–88. DOI: 10.1007/BF02080933.

  93. Torvik A (1985) Two types of brain lesions in Wernicke’s encephalopathy. Neuropathol. Appl. Neurobiol. 11 (3): 179–190. DOI: 10.1111/j.1365-2990.1985.tb00016.x.

  94. Butterworth RF (1989) Effects of thiamine deficiency on brain metabolism: implications for the pathogenesis of the wernicke-korsakoff syndrome. Alcohol Alcohol. 24 (4): 271–279. DOI: 10.1093/oxfordjournals.alcalc.a044913.

  95. Vindedzis SA, Stanton KG, Sherriff JL, Dhaliwal SS (2008) Thiamine deficiency in diabetes — is diet relevant? Diabetes Vasc. Dis. Res. 5 (3): 215–215. DOI: 10.3132/dvdr.2008.035.

  96. Jermendy G (2006) Evaluating thiamine deficiency in patients with diabetes. Diabetes Vasc. Dis. Res. 3 (2): 120–121. DOI: 10.3132/dvdr.2006.014.

  97. Kohda Y, Shirakawa H, Yamane K, Otsuka K, Kono T, Terasaki F, Tanaka T (2008) Prevention of incipient diabetic cardiomyopathy by high-dose thiamine. J. Toxicol. Sci. 33 (4): 459–472. DOI: 10.2131/jts.33.459.

  98. Mee L, Nabokina SM, Sekar VT, Subramanian VS, Maedler K, Said HM (2009) Pancreatic beta cells and islets take up thiamin by a regulated carrier-mediated process: studies using mice and human pancreatic preparations. Am. J. Physiol. Liver Physiol. 297 (1): G197–G206. DOI: 10.1152/ajpgi.00092.2009.

  99. Page GLJ, Laight D, Cummings MH (2011) Thiamine deficiency in diabetes mellitus and the impact of thiamine replacement on glucose metabolism and vascular disease. Int. J. Clin. Pract. 65 (6): 684–690. DOI: 10.1111/j.1742-1241.2011.02680.x.

  100. Watson JD, Dako DY (1975) Erythrocyte transketolase activity in adult Ghanaian subjects. Clin. Chim. Acta59 (1): 55–61. DOI: 10.1016/0009-8981(75)90218-1.

  101. Thornalley PJ, Babaei-Jadidi R, Al Ali H, Rabbani N, Antonysunil A, Larkin J, Ahmed A, Rayman G, Bodmer CW (2007) High prevalence of low plasma thiamine concentration in diabetes linked to a marker of vascular disease. Diabetologia 50 (10): 2164–2170. DOI: 10.1007/s00125-007-0771-4.

  102. Thakur K, Tomar SK, Singh AK, Mandal S, Arora S (2017) Riboflavin and health: A review of recent human research. Crit. Rev. Food Sci. Nutr. 57 (17): 3650–3660. DOI: 10.1080/10408398.2016.1145104.

  103. Alam MM, Iqbal S, Naseem I (2015) Ameliorative effect of riboflavin on hyperglycemia, oxidative stress and DNA damage in type-2 diabetic mice: Mechanistic and therapeutic strategies. Arch. Biochem. Biophys. 584 : 10–19. DOI: 10.1016/j.abb.2015.08.013.

  104. Udhayabanu T, Manole A, Rajeshwari M, Varalakshmi P, Houlden H, Ashokkumar B (2017) Riboflavin responsive mitochondrial dysfunction in neurodegenerative diseases. J. Clin. Med. 6 (5): 52. DOI: 10.3390/jcm6050052.

  105. Zhao R, Wang H, Qiao C, Zhao K (2018) Vitamin B2 blocks development of Alzheimer’s disease in APP/PS1 transgenic mice via anti-oxidative mechanism. Trop. J. Pharm. Res. 17 (6): 1049. DOI: 10.4314/tjpr.v17i6.10.

  106. King GL (2008) The role of inflammatory cytokines in diabetes and its complications. J. Periodontol. 79 (8s): 1527–1534. DOI: 10.1902/jop.2008.080246.

  107. Sauve AA (2008) NAD + and vitamin B 3 : from metabolism to therapies. J. Pharmacol. Exp. Ther. 324 (3): 883–893. DOI: 10.1124/jpet.107.120758.

  108. Gasperi V, Sibilano M, Savini I, Catani M (2019) Niacin in the central nervous system: An update of biological aspects and clinical applications. Int. J. Mol. Sci. 20 (4): 974. DOI: 10.3390/ijms20040974.

  109. Li F, Chong Z, Maiese K (2006) Cell life versus cell longevity: the mysteries surrounding the NAD+ precursor nicotinamide. Curr. Med. Chem. 13 (8): 883–895. DOI: 10.2174/092986706776361058.

  110. Trammell SAJ, Brenner C (2015) NNMT: A bad actor in fat makes good in liver. Cell Metab. 22 (2): 200–201. DOI: 10.1016/j.cmet.2015.07.017.

  111. Hong S, Moreno-Navarrete JM, Wei X, Kikukawa Y, Tzameli I, Prasad D, Lee Y, Asara JM, Fernandez-Real JM, Maratos-Flier E, Pissios P (2015) Nicotinamide N-methyltransferase regulates hepatic nutrient metabolism through Sirt1 protein stabilization. Nat. Med. 21 (8): 887–894. DOI: 10.1038/nm.3882.

  112. Chong ZZ, Lin S-H, Maiese K (2004) The NAD + precursor nicotinamide governs neuronal survival during oxidative stress through protein kinase B coupled to FOXO3a and mitochondrial membrane potential. J. Cereb. Blood Flow Metab. 24 (7): 728–743. DOI: 10.1097/01.WCB.0000122746.72175.0E.

  113. Julius U, Fischer S (2013) Nicotinic acid as a lipid-modifying drug – A review. Atheroscler. Suppl 14 (1): 7–13. DOI: 10.1016/j.atherosclerosissup.2012.10.036.

  114. Tavintharan S, Woon K, Pek LT, Jauhar N, Dong X, Lim SC, Sum CF (2011) Niacin results in reduced monocyte adhesion in patients with type 2 diabetes mellitus. Atherosclerosis 215 (1): 176–179. DOI: 10.1016/j.atherosclerosis.2010.12.020.

  115. Picone S, Ariganello P, Mondì V, Di Palma F, Martini L, Marziali S, Fariello G, Paolillo P (2019) A solution based on melatonin, tryptophan, and vitamin B6 (Melamil Tripto©) for sedation in newborns during brain MRI. Ital. J. Pediatr. 45 (1): 122. DOI: 10.1186/s13052-019-0714-y.

  116. Anitha M, Abraham PM, Paulose CS (2012) Striatal dopamine receptors modulate the expression of insulin receptor, IGF-1 and GLUT-3 in diabetic rats: Effect of pyridoxine treatment. Eur. J. Pharmacol. 696 (1–3): 54–61. DOI: 10.1016/j.ejphar.2012.09.006.

  117. Leklem JE, Hollenbeck CB (1990) Acute ingestion of glucose decreases plasma pyridoxal 5’-phosphate and total vitamin B-6 concentration. Am. J. Clin. Nutr. 51 (5): 832–836. DOI: 10.1093/ajcn/51.5.832.

  118. Satyanarayana A, Balakrishna N, Pitla S, Reddy PY, Mudili S, Lopamudra P, Suryanarayana P, Viswanath K, Ayyagari R, Reddy GB (2011) Status of B-vitamins and homocysteine in diabetic retinopathy: association with vitamin-B12 deficiency and hyperhomocysteinemia. PLoS One 6 (11): e26747. DOI: 10.1371/journal.pone.0026747.

  119. Rogers KS, Higgins ES, Kline ES (1986) Experimental diabetes causes mitochondrial loss and cytoplasmic enrichment of pyridoxal phosphate and aspartate aminotransferase activity. Biochem. Med. Metab. Biol. 36 (1): 91–97. DOI: 10.1016/0885-4505(86)90111-8.

  120. Spellacy WN, Buhi WC, Birk SA (1977) Vitamin B6 treatment of gestational diabetes mellitus. Am. J. Obstet. Gynecol. 127 (6): 599–602. DOI: 10.1016/0002-9378(77)90356-8.

  121. Bennink HJTC, Schreurs WHP (1975) Improvement of oral glucose tolerance in gestational diabetes by pyridoxine. BMJ 3 (5974): 13–15. DOI: 10.1136/bmj.3.5974.13.

  122. Connick JH, Stone TW (1985) The role of kynurenines in diabetes mellitus. Med. Hypotheses 18 (4): 371–376. DOI: 10.1016/0306-9877(85)90104-5.

  123. Liu Z, Li P, Zhao Z-H, Zhang Y, Ma Z-M, Wang S-X (2016) Vitamin B6 prevents endothelial dysfunction, insulin resistance, and hepatic lipid accumulation in Apoe −/− mice fed with high-fat diet. J. Diabetes Res. 2016 : 1–8. DOI: 10.1155/2016/1748065.

  124. Fernandez-Mejia C (2005) Pharmacological effects of biotin. J. Nutr. Biochem. 16 (7): 424–427. DOI: 10.1016/j.jnutbio.2005.03.018.

  125. Tong L (2013) Structure and function of biotin-dependent carboxylases. Cell. Mol. Life Sci. 70 (5): 863–891. DOI: 10.1007/s00018-012-1096-0.

  126. Via M (2012) The malnutrition of obesity: micronutrient deficiencies that promote diabetes. ISRN Endocrinol. 2012 (103472): 1–8. DOI: 10.5402/2012/103472.

  127. Dakshinamurti K, Cheah-Tan C (1968) Biotin-mediated synthesis of hepatic glucokinase in the rat. Arch. Biochem. Biophys. 127 (1): 17–21. DOI: 10.1016/0003-9861(68)90195-1.

  128. Romero-Navarro G (1999) Biotin regulation of pancreatic glucokinase and insulin in primary cultured rat Islets and in biotin- deficient rats. Endocrinology 140 (10): 4595–4600. DOI: 10.1210/en.140.10.4595.

  129. Hemmati M, Babaei H, Abdolsalehei M (2013) Survey of the effect of biotin on glycemic control and plasma lipid concentrations in type 1 diabetic patients in kermanshah in Iran (2008-2009). Oman Med. J. 28 (3): 195–198. DOI: 10.5001/omj.2013.53.

  130. Sone H, Ito M, Sugiyama K, Ohneda M, Maebashi M, Furukawa Y (1999) Biotin enhances glucose-stimulated insulin secretion in the isolated perfused pancreas of the rat. J. Nutr. Biochem. 10 (4): 237–243. DOI: 10.1016/S0955-2863(99)00003-0.

  131. Sone H, Ito M, Shimizu M, Sasaki Y, Komai M FY (2000) Characteristics of the biotin enhancement of glucose-induced insulin release in pancreatic islets of the rat. Biosci. Biotechnol. Biochem. 64 (3): 550–554. DOI: 10.1271/bbb.64.550.

  132. De la Vega LA, Stockert RJ (2000) Regulation of the insulin and asialoglycoprotein receptors via cGMP-dependent protein kinase. Am. J. Physiol. Physiol. 279 (6): C2037–C2042. DOI: 10.1152/ajpcell.2000.279.6.C2037.

  133. Ferre T, Pujol A, Riu E, Bosch F, Valera A (1996) Correction of diabetic alterations by glucokinase. Proc. Natl. Acad. Sci. 93 (14): 7225–7230. DOI: 10.1073/pnas.93.14.7225.

  134. Lazo de la Vega-Monroy ML, Larrieta E, German MS, Baez-Saldana A, Fernandez-Mejia C (2013) Effects of biotin supplementation in the diet on insulin secretion, islet gene expression, glucose homeostasis and beta-cell proportion. J. Nutr. Biochem. 24 (1): 169–177. DOI: 10.1016/j.jnutbio.2012.03.020.

  135. Larrieta E, de la Vega-Monroy MLL, Vital P, Aguilera A, German MS, El Hafidi M, Fernandez-Mejia C (2012) Effects of biotin deficiency on pancreatic islet morphology, insulin sensitivity and glucose homeostasis. J. Nutr. Biochem. 23 (4): 392–399. DOI: 10.1016/j.jnutbio.2011.01.003.

  136. Reddi A, DeAngelis B, Frank O, Lasker N, Baker H (1988) Biotin supplementation improves glucose and insulin tolerances in genetically diabetic KK mice. Life Sci. 42 (13): 1323–1330. DOI: 10.1016/0024-3205(88)90226-3.

  137. Xiang X, Liu Y, Zhang X, Zhang W, Wang Z (2015) [Effects of biotin on blood glucose regulation in type 2 diabetes rat model]. Wei Sheng Yan Jiu44 (2): 185–189, 195.

  138. Reynolds E (2006) Vitamin B12, folic acid, and the nervous system. Lancet Neurol. 5 (11): 949–960. DOI: 10.1016/S1474-4422(06)70598-1.

  139. Rucker RB, Zempleni J, Suttie JW MD 2007. [Internet] (2007) Handbook of Vitamins. Rucker RB, Zempleni J, Suttie JW, McCormick DB, editors. CRC Press; 2007. DOI: 10.1201/9781420005806.

  140.  Molina M, Gonzalez R, Folgado J, Real JT, Martínez-Hervás S, Priego A, Lorente R, Chaves FJ, Ascaso JF (2013) Correlation between plasma concentrations of homocysteine and diabetic polyneuropathy evaluated with the Semmes-Weinstein monofilament test in patients with type 2 diabetes mellitus. Med. Clin. (Barc). 141 (9): 382–386. DOI: 10.1016/j.medcli.2012.09.042.

  141. Ukinc K, Ersoz HO, Karahan C, Erem C, Eminagaoglu S, Hacihasanoglu AB, Yilmaz M, Kocak M (2009) Methyltetrahydrofolate reductase C677T gene mutation and hyperhomocysteinemia as a novel risk factor for diabetic nephropathy. Endocrine 36 (2): 255–261. DOI: 10.1007/s12020-009-9218-7.

  142. Jianbo L, Yuche C, Ming S, Jingrong T, Qing D, Yu Z, Jiawei C, Hongxing W (2011) Association of homocysteine with peripheral neuropathy in Chinese patients with type 2 diabetes. Diabetes Res. Clin. Pract. 93 (1): 38–42. DOI: 10.1016/j.diabres.2011.03.020.

  143. Smolek M, Notaroberto, Jaramillo, Pradillo (2013) Intervention with vitamins in patients with nonproliferative diabetic retinopathy: a pilot study. Clin. Ophthalmol. 7 : 1451. DOI: 10.2147/OPTH.S46718.

  144. Polizzi F, Andican G, Çetin E, Civelek S, Yumuk V, Burçak G (2012) Increased DNA-glycation in type 2 diabetic patients: the effect of thiamine and pyridoxine therapy. Exp. Clin. Endocrinol. Diabetes 120 (06): 329–334. DOI: 10.1055/s-0031-1298016.

  145. González-Ortiz M, Martínez-Abundis E, Robles-Cervantes JA, Ramírez-Ramírez V, Ramos-Zavala MG (2011) Effect of thiamine administration on metabolic profile, cytokines and inflammatory markers in drug-naïve patients with type 2 diabetes. Eur. J. Nutr. 50 (2): 145–149. DOI: 10.1007/s00394-010-0123-x.

  146. Xiang D, Zhang Q, Wang Y-T (2020) Effectiveness of niacin supplementation for patients with type 2 diabetes. Medicine (Baltimore). 99 (29): e21235. DOI: 10.1097/MD.0000000000021235.

  147. Unoki-Kubota H, Yamagishi S, Takeuchi M, Bujo H, Saito Y (2010) Pyridoxamine, an inhibitor of advanced glycation end product (AGE) formation ameliorates insulin resistance in obese, type 2 diabetic mice. Protein Pept. Lett. 17 (9): 1177–1181. DOI: 10.2174/092986610791760423.

  148. Sahin K, Tuzcu M, Orhan C, Sahin N, Kucuk O, Ozercan IH, Juturu V, Komorowski JR (2013) Anti-diabetic activity of chromium picolinate and biotin in rats with type 2 diabetes induced by high-fat diet and streptozotocin. Br. J. Nutr. 110 (2): 197–205. DOI: 10.1017/S0007114512004850.

  149. Cipolla-Neto J, Amaral FG, Afeche SC, Tan DX, Reiter RJ (2014) Melatonin, energy metabolism, and obesity: a review. J. Pineal Res. 56 (4): 371–381. DOI: 10.1111/jpi.12137.

  150. Simonneaux V, Ribelayga C (2003) Generation of the melatonin endocrine message in mammals: A review of the complex regulation of melatonin synthesis by norepinephrine, peptides, and other pineal transmitters. Pharmacol. Rev. 55 (2): 325–395. DOI: 10.1124/pr.55.2.2.

  151. Reiter RJ, Tan DX, Kim SJ, Cruz MHC (2014) Delivery of pineal melatonin to the brain and SCN: role of canaliculi, cerebrospinal fluid, tanycytes and Virchow–Robin perivascular spaces. Brain Struct. Funct. 219 (6): 1873–1887. DOI: 10.1007/s00429-014-0719-7.

  152. Reiter RJ (1991) Melatonin: The chemical expression of darkness. Mol. Cell. Endocrinol. 79 (1–3): C153–C158. DOI: 10.1016/0303-7207(91)90087-9.

  153. Scheer FAJL, Hilton MF, Mantzoros CS, Shea SA (2009) Adverse metabolic and cardiovascular consequences of circadian misalignment. Proc. Natl. Acad. Sci. 106 (11): 4453–4458. DOI: 10.1073/pnas.0808180106.

  154. Esposito E, Cuzzocrea S (2010) Antiinflammatory Activity of Melatonin in Central Nervous System. Curr. Neuropharmacol. 8 (3): 228–242. DOI: 10.2174/157015910792246155.

  155.  Stefulj J, Hörtner M, Ghosh M, Schauenstein K, Rinner I, Wölfler A, Semmler J, Liebmann PM (2001) Gene expression of the key enzymes of melatonin synthesis in extrapineal tissues of the rat. J. Pineal Res. 30 (4): 243–247. DOI: 10.1034/j.1600-079X.2001.300408.x.

  156. Arendt J, Aulinas A 2022. (2022) Physiology of the pineal gland and melatonin. In: Endotext. MDText.com, Inc.; 2022.

  157. Liu Y-J, Zhuang J, Zhu H-Y, Shen Y-X, Tan Z-L, Zhou J-N (2007) Cultured rat cortical astrocytes synthesize melatonin: absence of a diurnal rhythm. J. Pineal Res. 43 (3): 232–238. DOI: 10.1111/j.1600-079X.2007.00466.x.

  158. Meng X, Li Y, Li S, Zhou Y, Gan R-Y, Xu D-P, Li H-B (2017) Dietary sources and bioactivities of melatonin. Nutrients 9 (4): 367. DOI: 10.3390/nu9040367.

  159. Dubocovich ML, Markowska M (2005) Functional MT1 and MT2 melatonin receptors in mammals. Endocrine 27 (2): 101–110. DOI: 10.1385/ENDO:27:2:101.

  160.  Lacoste B, Angeloni D, Dominguez-Lopez S, Calderoni S, Mauro A, Fraschini F, Descarries L, Gobbi G (2015) Anatomical and cellular localization of melatonin MT 1 and MT 2 receptors in the adult rat brain. J. Pineal Res. 58 (4): 397–417. DOI: 10.1111/jpi.12224.

  161. Mailliet F, Ferry G, Vella F, Berger S, Cogé F, Chomarat P, Mallet C, Guénin S-P, Guillaumet G, Viaud-Massuard M-C, Yous S, Delagrange P, Boutin JA (2005) Characterization of the melatoninergic MT3 binding site on the NRH:quinone oxidoreductase 2 enzyme. Biochem. Pharmacol. 71 (1–2): 74–88. DOI: 10.1016/j.bcp.2005.09.030.

  162. Suofu Y, Li W, Jean-Alphonse FG, Jia J, Khattar NK, Li J et al. (2017) Dual role of mitochondria in producing melatonin and driving GPCR signaling to block cytochrome c release. Proc. Natl. Acad. Sci. 114 (38): E7997–E8006. DOI: 10.1073/pnas.1705768114.

  163. Mohan N, Sadeghi K, Reiter RJ, Meltz ML (1995) The neurohormone melatonin inhibits cytokine, mitogen and ionizing radiation induced NF-kappa B. Biochem. Mol. Biol. Int. 37 (6): 1063–1070.

  164. Mayo JC, Sainz RM, Tan DX, Hardeland R, Leon J, Rodriguez C, Reiter RJ (2005) Anti-inflammatory actions of melatonin and its metabolites, N1-acetyl-N2-formyl-5-methoxykynuramine (AFMK) and N1-acetyl-5-methoxykynuramine (AMK), in macrophages. J. Neuroimmunol. 165 (1–2): 139–149. DOI: 10.1016/j.jneuroim.2005.05.002.

  165. Calabrese V, Boyd-Kimball D, Scapagnini G, Butterfield DA (2004) Nitric oxide and cellular stress response in brain aging and neurodegenerative disorders: the role of vitagenes. In Vivo 18 (3): 245–267.

  166. Calabrese V, Scapagnini G, Giuffrida Stella AM, Bates TE, Clark JB (2001) Mitochondrial involvement in brain function and dysfunction: relevance to aging, neurodegenerative disorders and longevity. Neurochem. Res. 26 (6): 739–764. DOI: 10.1023/a:1010955807739.

  167. Rosenbluth J (1980) Central myelin in the mouse mutant shiverer. J. Comp. Neurol. 194 (3): 639–648. DOI: 10.1002/cne.901940310.

  168. Rubio-González A, Reiter RJ, De Luxán-Delgado B, Potes Y, Caballero B, Boga JA, Solano JJ, Vega-Naredo I, Coto-Montes A (2020) Pleiotropic role of melatonin in brain mitochondria of obese mice. Melatonin Res. 3 (4): 538–557. DOI: 10.32794/MR11250078.

  169. Laposky AD, Bass J, Kohsaka A, Turek FW (2008) Sleep and circadian rhythms: Key components in the regulation of energy metabolism. FEBS Lett. 582 (1): 142–151. DOI: 10.1016/j.febslet.2007.06.079.

  170. Jenwitheesuk A, Nopparat C, Mukda S, Wongchitrat P, Govitrapong P (2014) Melatonin regulates aging and neurodegeneration through energy metabolism, epigenetics, autophagy and circadian rhythm pathways. Int. J. Mol. Sci. 15 (9): 16848–16884. DOI: 10.3390/ijms150916848.

  171. Shen S, Liao Q, Wong YK, Chen X, Yang C, Xu C, Sun J, Wang J (2022) The role of melatonin in the treatment of type 2 diabetes mellitus and Alzheimer’s disease. Int. J. Biol. Sci. 18 (3): 983–994. DOI: 10.7150/ijbs.66871.

  172. Hölscher C (2019) Insulin signaling impairment in the brain as a risk factor in Alzheimer’s disease. Front. Aging Neurosci. 11 (APR). DOI: 10.3389/fnagi.2019.00088.

  173. Shin Y, Choi SH, Kim E, Bylykbashi E, Kim JA, Chung S, Kim DY, Kamm RD, Tanzi RE (2019) Blood–brain barrier dysfunction in a 3D in vitro model of Alzheimer’s disease. Adv. Sci. 6 (20): 1900962. DOI: 10.1002/advs.201900962.

  174. Banks W (2005) Blood-brain barrier transport of cytokines: A mechanism for neuropathology. Curr. Pharm. Des. 11 (8): 973–984. DOI: 10.2174/1381612053381684.

  175.  Alluri H, Wilson RL, Anasooya Shaji C, Wiggins-Dohlvik K, Patel S, Liu Y, Peng X, Beeram MR, Davis ML, Huang JH, Tharakan B (2016) Melatonin preserves blood-brain barrier integrity and permeability via matrix metalloproteinase-9 inhibition. Thomas B, editor. PLoS One 11 (5): e0154427. DOI: 10.1371/journal.pone.0154427.

  176. Frölich L, Blum-Degen D, Bernstein H-G, Engelsberger S, Humrich J, Laufer S, Muschner D, Thalheimer A, Türk A, Hoyer S, Zöchling R, Boissl KW, Jellinger K, Riederer P (1998) Brain insulin and insulin receptors in aging and sporadic Alzheimer’s disease. J. Neural Transm. 105 (4): 423. DOI: 10.1007/s007020050068.

  177. Heras-Sandoval D, Pérez-Rojas JM, Hernández-Damián J, Pedraza-Chaverri J (2014) The role of PI3K/AKT/mTOR pathway in the modulation of autophagy and the clearance of protein aggregates in neurodegeneration. Cell. Signal. 26 (12): 2694–2701. DOI: 10.1016/j.cellsig.2014.08.019.

  178. Ferrario CR, Reagan LP (2018) Insulin-mediated synaptic plasticity in the CNS: Anatomical, functional and temporal contexts. Neuropharmacology 136 (Pt B): 182–191. DOI: 10.1016/j.neuropharm.2017.12.001.

  179. Peschke E, Frese T, Chankiewitz E, Peschke D, Preiss U, Schneyer U, Spessert R, Muhlbauer E (2006) Diabetic Goto Kakizaki rats as well as type 2 diabetic patients show a decreased diurnal serum melatonin level and an increased pancreatic melatonin-receptor status. J. Pineal Res. 40 (2): 135–143. DOI: 10.1111/j.1600-079X.2005.00287.x.

  180. Peschke E, Stumpf I, Bazwinsky I, Litvak L, Dralle H, Mühlbauer E (2007) Melatonin and type 2 diabetes ? a possible link? J. Pineal Res. 42 (4): 350–358. DOI: 10.1111/j.1600-079X.2007.00426.x.

  181. Mirick DK, Davis S (2008) Melatonin as a biomarker of circadian dysregulation. Cancer Epidemiol. Biomarkers Prev. 17 (12): 3306–3313. DOI: 10.1158/1055-9965.EPI-08-0605.

  182.  Cagnacci A, Arangino S, Renzi A, Paoletti AM, Melis GB, Cagnacci P, Volpe A (2001) Influence of melatonin administration on glucose tolerance and insulin sensitivity of postmenopausal women. Clin. Endocrinol. (Oxf). 54 (3): 339–346. DOI: 10.1046/j.1365-2265.2001.01232.x.

  183. Boden G, Ruiz J, Urbain JL, Chen X (1996) Evidence for a circadian rhythm of insulin secretion. Am. J. Physiol. Metab. 271 (2): E246–E252. DOI: 10.1152/ajpendo.1996.271.2.E246.

  184. Peschke E (2008) Melatonin, endocrine pancreas and diabetes. J. Pineal Res. 44 (1): 26–40. DOI: 10.1111/j.1600-079X.2007.00519.x.

  185. Rasmussen DD, Mitton DR, Larsen SA, Yellon SM (2001) Aging-dependent changes in the effect of daily melatonin supplementation on rat metabolic and behavioral responses. J. Pineal Res. 31 (1): 89–94. DOI: 10.1034/j.1600-079X.2001.310113.x.

  186. Champney TH, Steger RW, Christie DS, Reiter RJ (1985) Alterations in components of the pineal melatonin synthetic pathway by acute insulin stress in the rat and Syrian hamster. Brain Res. 338 (1): 25–32. DOI: 10.1016/0006-8993(85)90244-6.

  187. Champney TH, Brainard GC, Richardson BA, Reiter RJ (1983) Experimentally-induced diabetes reduces nocturnal pineal melatonin content in the Syrian hamster. Comp. Biochem. Physiol. Part A Physiol. 76 (1): 199–201. DOI: 10.1016/0300-9629(83)90314-6.

  188.  Champney TH, Holtorf AP, Craft CM, Reiter RJ (1986) Hormonal modulation of pineal melatonin synthesis in rats and syrian hamsters: Effects of streptozotocin-induced diabetes and insulin injections. Comp. Biochem. Physiol. Part A Physiol. 83 (2): 391–395. DOI: 10.1016/0300-9629(86)90594-3.

  189. Rahman MM, Kwon H-S, Kim M-J, Go H-K, Oak M-H, Kim D-H (2017) Melatonin supplementation plus exercise behavior ameliorate insulin resistance, hypertension and fatigue in a rat model of type 2 diabetes mellitus. Biomed. Pharmacother. 92 : 606–614. DOI: 10.1016/j.biopha.2017.05.035.

  190. Conti A, Maestroni GJM (1996) Role of the pineal gland and melatonin in the development of autoimmune diabetes in non-obese diabetic mice. J. Pineal Res. 20 (3): 164–172. DOI: 10.1111/j.1600-079X.1996.tb00253.x.

  191.  Sun H, Wang X, Chen J, Gusdon AM, Song K, Li L, Qu S (2018) Melatonin treatment improves insulin resistance and pigmentation in obese patients with Acanthosis Nigricans. Int. J. Endocrinol. 2018 (2304746): 1–7. DOI: 10.1155/2018/2304746.

  192. Oliveira AC de, Andreotti S, Sertie RAL, Campana AB, de Proença ARG, Vasconcelos RP, Oliveira KA de, Coelho-de-Souza AN, Donato-Junior J, Lima FB (2018) Combined treatment with melatonin and insulin improves glycemic control, white adipose tissue metabolism and reproductive axis of diabetic male rats. Life Sci. 199 : 158–166. DOI: 10.1016/j.lfs.2018.02.040.

  193. Sartori C, Dessen P, Mathieu C, Monney A, Bloch J, Nicod P, Scherrer U, Duplain H (2009) Melatonin improves glucose homeostasis and endothelial vascular function in high-fat diet-fed insulin-resistant mice. Endocrinology 150 (12): 5311–5317. DOI: 10.1210/en.2009-0425.

  194. Kadhim HM, Ismail SH, Hussein KI, Bakir IH, Sahib AS, Khalaf BH, Hussain SA-R (2006) Effects of melatonin and zinc on lipid profile and renal function in type 2 diabetic patients poorly controlled with metformin. J. Pineal Res. 41 (2): 189–193. DOI: 10.1111/j.1600-079X.2006.00353.x.

  195. Biessels GJ, Kappelle AC, Bravenboer B, Erkelens DW, Gispen WH (1994) Cerebral function in diabetes mellitus. Diabetologia 37 (7): 643–650. DOI: 10.1007/BF00417687.

  196.  Li Z-G, Zhang W, Grunberger G, Sima AAF (2002) Hippocampal neuronal apoptosis in type 1 diabetes. Brain Res. 946 (2): 221–231. DOI: 10.1016/S0006-8993(02)02887-1.

  197. Li Z, Zhang W, Sima AAF (2005) The role of impaired insulin/IGF action in primary diabetic encephalopathy. Brain Res. 1037 (1–2): 12–24. DOI: 10.1016/j.brainres.2004.11.063.

  198. Sima AAF (2004) Diabetes underlies common neurological disorders. Ann. Neurol. 56 (4): 459–461. DOI: 10.1002/ana.20249.

  199. Ristow M (2004) Neurodegenerative disorders associated with diabetes mellitus. J. Mol. Med. 82 (8): 510–529. DOI: 10.1007/s00109-004-0552-1.

  200. Northam EA, Rankins D, Cameron FJ (2006) Therapy Insight: the impact of type 1 diabetes on brain development and function. Nat. Clin. Pract. Neurol. 2 (2): 78–86. DOI: 10.1038/ncpneuro0097.

  201. Baydas G, Canatan H, Turkoglu A (2002) Comparative analysis of the protective effects of melatonin and vitamin E on streptozocin-induced diabetes mellitus. J. Pineal Res. 32 (4): 225–230. DOI: 10.1034/j.1600-079X.2002.01856.x.

  202. Baydas G, Nedzvetskii VS, Tuzcu M, Yasar A, Kirichenko S V. (2003) Increase of glial fibrillary acidic protein and S-100B in hippocampus and cortex of diabetic rats: effects of vitamin E. Eur. J. Pharmacol. 462 (1–3): 67–71. DOI: 10.1016/S0014-2999(03)01294-9.

  203. Baydas G, Donder E, Kiliboz M, Sonkaya E, Tuzcu M, Yasar A, Nedzvetskii VS (2004) Neuroprotection by -lipoic acid in streptozotocin-induced diabetes. Biochem. 69 (9): 1001–1005. DOI: 10.1023/B:BIRY.0000043542.39691.95.

  204. Poeggeler B, Saarela S, Reiter RJ, Tan DX, Chen ID, Manchester lc BI (2006) Melatonin- a highly potent endogenous radical scavenger and electron donor: new aspects of the oxidation chemistry of this indole accessed in vitro. Ann. N. Y. Acad. Sci. 738 (1): 419–420. DOI: 10.1111/j.1749-6632.1994.tb21831.x.

  205.  Reiter RJ, Tan DX, Manchester LC, Qi W (2001) Biochemical reactivity of melatonin with reactive oxygen and nitrogen species: A review of the evidence. Cell Biochem. Biophys. 34 (2): 237–256. DOI: 10.1385/CBB:34:2:237.

  206. Baydas G, Reiter RJ, Yasar A, Tuzcu M, Akdemir I, Nedzvetskii VS (2003) Melatonin reduces glial reactivity in the hippocampus, cortex, and cerebellum of streptozotocin-induced diabetic rats. Free Radic. Biol. Med. 35 (7): 797–804. DOI: 10.1016/S0891-5849(03)00408-8.

  207. Baydas G, Nedzvetskii VS, Kirichenko S V., Nerush PA (2008) Astrogliosis in the hippocampus and cortex and cognitive deficits in rats with streptozotocin-induced diabetes: Effects of melatonin. Neurophysiology 40 (2): 91–97. DOI: 10.1007/s11062-008-9026-3.

  208. Hajam YA, Rai S, Ghosh H, Basheer M (2020) Combined administration of exogenous melatonin and insulin ameliorates streptozotocin induced toxic alteration on hematological parameters in diabetic male Wistar rats. Toxicol. Reports 7 : 353–359. DOI: 10.1016/j.toxrep.2020.01.020.

  209. Kahya MC, Naziroʇlu M, Çiʇ B (2015) Melatonin and selenium reduce plasma cytokine and brain oxidative stress levels in diabetic rats. Brain Inj. 29 (12): 1490–1496. DOI: 10.3109/02699052.2015.1053526.

  210.  Gürpınar T, Ekerbiçer N, Uysal N, Barut T, Tarakçı F, Tuglu MI (2012) The effects of the melatonin treatment on the oxidative stress and apoptosis in diabetic eye and brain. Sci. World J. 2012 : 1–5. DOI: 10.1100/2012/498489.

  211. Jangra A, Datusalia AK, Khandwe S, Sharma SS (2013) Amelioration of diabetes-induced neurobehavioral and neurochemical changes by melatonin and nicotinamide: Implication of oxidative stress–PARP pathway. Pharmacol. Biochem. Behav. 114115 : 43–51. DOI: 10.1016/j.pbb.2013.10.021.

  212. Negi G, Kumar A, Kaundal RK, Gulati A, Sharma SS (2010) Functional and biochemical evidence indicating beneficial effect of Melatonin and Nicotinamide alone and in combination in experimental diabetic neuropathy. Neuropharmacology 58 (3): 585–592. DOI: 10.1016/j.neuropharm.2009.11.018.

  213. Oliveira-Abreu K, Cipolla-Neto J, Leal-Cardoso JH (2021) Effects of melatonin on diabetic neuropathy and retinopathy. Int. J. Mol. Sci. 23 (1): 100. DOI: 10.3390/ijms23010100.

  214. Scott JN, Clark AW, Zochodne DW (1999) Neurofilament and tubulin gene expression in progressive experimental diabetes. Brain 122 (11): 2109–2118. DOI: 10.1093/brain/122.11.2109.

  215. Lupachyk S, Watcho P, Stavniichuk R, Shevalye H, Obrosova IG (2013) Endoplasmic reticulum stress plays a key role in the pathogenesis of diabetic peripheral neuropathy. Diabetes 62 (3): 944–952. DOI: 10.2337/db12-0716.

  216. Ma J, Pan P, Anyika M, Blagg BSJ, Dobrowsky RT (2015) Modulating molecular chaperones improves mitochondrial bioenergetics and decreases the inflammatory transcriptome in diabetic sensory neurons. ACS Chem. Neurosci. 6 (9): 1637–1648. DOI: 10.1021/acschemneuro.5b00165.

  217. Ilnytska O, Lyzogubov V V., Stevens MJ, Drel VR, Mashtalir N, Pacher P, Yorek MA, Obrosova IG (2006) Poly(ADP-ribose) polymerase inhibition alleviates experimental diabetic sensory neuropathy. Diabetes 55 (6): 1686–1694. DOI: 10.2337/db06-0067.

  218. Feldman EL, Nave K-A, Jensen TS, Bennett DLH (2017) New horizons in diabetic neuropathy: mechanisms, bioenergetics, and pain. Neuron 93 (6): 1296–1313. DOI: 10.1016/j.neuron.2017.02.005.

  219. Hosseini A, Samadi M, Baeeri M, Rahimifard M, Haghi-Aminjan H (2022) The neuroprotective effects of melatonin against diabetic neuropathy: A systematic review of non-clinical studies. Front. Pharmacol. 13 (984499): 1–14. DOI: 10.3389/fphar.2022.984499.

  220. Metwally MMM, Ebraheim LLM, Galal AAA (2018) Potential therapeutic role of melatonin on STZ-induced diabetic central neuropathy: A biochemical, histopathological, immunohistochemical and ultrastructural study. Acta Histochem. 120 (8): 828–836. DOI: 10.1016/j.acthis.2018.09.008.

  221. Negi G, Kumar A, Sharma SS (2010) Melatonin modulates neuroinflammation and oxidative stress in experimental diabetic neuropathy: effects on NF-κB and Nrf2 cascades. J. Pineal Res. 50 (2): no-no. DOI: 10.1111/j.1600-079X.2010.00821.x.

CCBY.png

This work is licensed under a Creative Commons Attribution 4.0 International License