Next Article in Journal
The Need for Culturally Appropriate Food Allergy Management Strategies: The Indian Milk Ladder
Next Article in Special Issue
Serum Magnesium Is Associated with Long-Term Survival of Non-ST-Elevation Myocardial Infarction Patients
Previous Article in Journal
The Circadian Nobiletin-ROR Axis Suppresses Adipogenic Differentiation and IκBα/NF-κB Signaling in Adipocytes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Opinion

Beyond Ion Homeostasis: Hypomagnesemia, Transient Receptor Potential Melastatin Channel 7, Mitochondrial Function, and Inflammation

Cardiovascular Division, Department of Medicine, The Lillehei Heart Institute, University of Minnesota at Twin Cities, Minneapolis, MN 55455, USA
*
Author to whom correspondence should be addressed.
Nutrients 2023, 15(18), 3920; https://doi.org/10.3390/nu15183920
Submission received: 7 August 2023 / Revised: 2 September 2023 / Accepted: 6 September 2023 / Published: 9 September 2023
(This article belongs to the Special Issue Magnesium Homeostasis and Magnesium Transporters in Human Health)

Abstract

:
As the second most abundant intracellular divalent cation, magnesium (Mg2+) is essential for cell functions, such as ATP production, protein/DNA synthesis, protein activity, and mitochondrial function. Mg2+ plays a critical role in heart rhythm, muscle contraction, and blood pressure. A significant decline in Mg2+ intake has been reported in developed countries because of the increased consumption of processed food and filtered/deionized water, which can lead to hypomagnesemia (HypoMg). HypoMg is commonly observed in cardiovascular diseases, such as heart failure, hypertension, arrhythmias, and diabetic cardiomyopathy, and HypoMg is a predictor for cardiovascular and all-cause mortality. On the other hand, Mg2+ supplementation has shown significant therapeutic effects in cardiovascular diseases. Some of the effects of HypoMg have been ascribed to changes in Mg2+ participation in enzyme activity, ATP stabilization, enzyme kinetics, and alterations in Ca2+, Na+, and other cations. In this manuscript, we discuss new insights into the pathogenic mechanisms of HypoMg that surpass previously described effects. HypoMg causes mitochondrial dysfunction, oxidative stress, and inflammation. Many of these effects can be attributed to the HypoMg-induced upregulation of a Mg2+ transporter transient receptor potential melastatin 7 channel (TRMP7) that is also a kinase. An increase in kinase signaling mediated by HypoMg-induced TRPM7 transcriptional upregulation, independently of any change in Mg2+ transport function, likely seems responsible for many of the effects of HypoMg. Therefore, Mg2+ supplementation and TRPM7 kinase inhibition may work to treat the sequelae of HypoMg by preventing increased TRPM7 kinase activity rather than just altering ion homeostasis. Since many diseases are characterized by oxidative stress or inflammation, Mg2+ supplementation and TRPM7 kinase inhibition may have wider implications for other diseases by acting to reduce oxidative stress and inflammation.

1. Introduction

As the second most abundant intracellular divalent cation, magnesium (Mg2+) is essential for cell functions, such as ATP production, protein and DNA synthesis, protein activity, and mitochondrial function. Mg2+ is a co-factor of hundreds of enzymes, including ATPases, endonucleases, exonucleases, polymerases, and phosphatases [1]. Mg2+ is also a calcium (Ca2+) antagonist, playing an important role in regulating Ca2+ homeostasis and function [2,3]. Traditionally, the effects of changes in Mg2+ levels have been viewed as general alterations in these Mg2+-dependent processes or through effects on the homeostasis of other ions such as Ca2+ or sodium ions (Na+).
In this review, we discuss new data that suggest Mg2+ to be in a transcriptional feedback loop with a Mg2+ transporter, transient receptor potential melastatin 7 channel (TRPM7), which also has a kinase domain. These new data raise the possibility that much of the antioxidant and anti-inflammatory effects of Mg2+ supplementation are mediated by changes in this channel/kinase level and its subsequent effects on kinase signal activity independent of any effects on Mg2+ ion distribution.

2. Hypomagnesemia and Its Association with Disease

New data on the role of TRPM7 have come to light because of experiments that study hypomagnesemia (HypoMg, serum Mg2+ < 0.8 mM) [4]. It seems increasingly clear that HypoMg is more than just a generalized dysfunction of Mg2+-dependent processes. HypoMg is a disease condition that is characterized by, among other things, increased mitochondrial dysfunction and inflammation that can lead to a type of heart failure and seizures [5,6].
HypoMg is more common than is generally known, and its incidence is increasing for several reasons. A significant decline in Mg2+ intake has been reported in developed countries because of the increased consumption of processed food and filtered/deionized water [7,8]. Studies show that ~50% of the US population, especially elderly people, consume less than the estimated average requirement of Mg2+ [9,10,11], and ~23% of US adults have hypomagnesemia (HypoMg, serum Mg < 0.8 mM) [4]. Moreover, chronic kidney, gastrointestinal diseases, and medications (e.g., diuretics, cytotoxic drugs, digoxin, amino-glycosides, and steroids) can further increase Mg2+ excretion and decrease Mg2+ absorption, causing HypoMg. A larger driver of the increase in HypoMg is likely an increase in diabetes. HypoMg and diabetes are epidemiologically associated [12,13,14,15], and with the worldwide increase in diabetes, we are likely to see a concomitant increase in HypoMg.
HypoMg is commonly observed in a variety of illnesses, including cardiovascular diseases [16,17,18,19,20], epilepsy, Alzheimer’s disease, Parkinson’s disease, osteoporosis, and diabetes (see reviews [21,22,23]). The cause-and-effect relationship between these disease states and HypoMg is an area of active investigation, and it is possible that many of these disease states actively contribute to HypoMg rather than being exacerbated by the condition.
HypoMg has been reported in heart failure [24,25], arrhythmia [26,27,28], myocardial infarction [27,28], atherosclerosis [29], diabetic cardiomyopathy [14,15], hypertension [20,30,31], and stroke [32], and Mg2+ supplementation has been used to treat some of these conditions. For example, patients with congestive heart failure often show signs of HypoMg [24,25], while Mg2+ supplementation has protective effects against heart failure [24]. HypoMg is reported in nearly 40% of patients with ventricular arrhythmias at hospital admission, and Mg2+ supplementation significantly decreases the number of episodes of non-sustained ventricular tachycardia [26]. In patients with acute myocardial ischemia, low serum Mg2+ has been observed, while Mg2+ supplementation improves heart function and mortality [27,28,33]. HypoMg and a low Mg2+ diet are associated with a higher risk of coronary artery disease, atherosclerosis, and hypertension [33,34,35], while Mg2+ supplementation improves these conditions [35,36,37].

3. Canonical Roles of Mg2+ in Cardiovascular Disease

The effects of HypoMg have generally been ascribed to alterations in Mg2+ or other cations in various subcellular compartments, and the effects of Mg2+ supplementation are thought to be mediated by reversing these generalized changes.
Mg2+ homeostasis is essential for cardiovascular function, including myocardial metabolism, myocardial contraction and relaxation, cardiac output, vascular tone, and peripheral vascular resistance [5,13,21,22,23]. In cardiomyocytes, the total intracellular Mg2+ concentration ([Mg2+]i) is ~17–20 mmol/L. Much of this Mg2+ is not free. About 50% of intracellular Mg2+ binds to ATP, phosphonucleotides, and phosphometabolites, while 25% binds to ribosomes [38]. Resting, free Mg2+ is in the range of 0.8–1.0 mmol/L both in the cytoplasm [39] and in the matrix of cardiomyocyte mitochondria [40]. Mg2+ regulates multiple cardiac ion channels, including the Na+ channel, multiple K+ channels, L- and T-type Ca2+ channels, the Na+-Ca2+ exchanger, Na+-K+-ATPase, and sarcoplasmic reticulum Ca2+ release, accounting for some of Mg2+’s effects on other cations (see review [22,41]). Mg2+ is also a critical cofactor for ATP production in cardiac mitochondria.
The disturbance of Mg2+ homeostasis can affect intracellular ion concentrations and balance, alter myocyte membrane potential, and impair heart contractility and heart rhythm. Mg2+ homeostasis is also important for vascular function and blood pressure. Decreased Mg2+ leads to vasoconstriction, while increased Mg2+ causes vasodilation and suppresses agonist-induced vasoconstriction [42,43]. Mechanistic studies show that in vascular smooth muscle cells, Mg2+ regulates excitation–contraction coupling and the myosin–actin crossbridge by modulating Ca2+ homeostasis [17,41,44].
Decreased [Mg2+]i is also associated with pulmonary hypertension, and high Mg2+ prevents the upregulated expression of NFATc3 and its nuclear translocation, which is a major Ca2+-dependent signaling pathway that regulates the proliferation and migration of pulmonary arterial smooth muscle cells in the development of pulmonary hypertension and vascular remodeling [35]. This is an important observation because it links Mg2+ and gene transcription. HypoMg also affects vascular endothelial cells’ survival, proliferation, and motility and induces higher vascular–endothelial barrier permeability [45]. Hormones and vasoactive agents such as angiotensin II, insulin, aldosterone, and vasopressin that are often altered in cardiovascular diseases also affect free [Mg2+]i in cardiomyocytes and vascular smooth muscle cells [44].

4. Mg2+ Transporters Control Mg2+ and Link Mg2+ to Other Cations

Mg2+ transporters are critical for maintaining Mg2+ homeostasis in cells and intracellular organelles. They also link Mg2+ homeostasis with other cations. Cardiac Mg2+ homeostasis is regulated and maintained by a series of sarcolemmal and organelle transporters, such as the TRPM7, solute carrier family 41 A1 (SLC41A1), Mg2+ transporter 1 (MagT1), and cyclin and CBS domain divalent metal cation transport mediator 2 (CNNM2) on the sarcolemmal membrane. In addition, there is the mitochondrial RNA splicing 2 protein (MRS2) and solute carrier family 41 A3 (SLC41A3) on mitochondrial membranes, as well as some antiporters, cotransporters, and exchangers (see review [22,46,47]). It is these antiporters, cotransporters, and exchangers that link Mg2+ homeostasis to that of other cations.
Certainly, Mg2+ transporters need more study, but some key facts are known. The SLC41A family has three members identified as Mg2+ transporters/carriers: SLC41A1, SLC41A2, and SLC41A3 (see a recent review [48]). SLC41A1 and SLC41A2 are expressed mainly on the plasma membrane [49,50], while SLC41A3 is mainly expressed on mitochondrial membranes and conducts mitochondrial Mg2+ efflux [51]. SCL41A1 conducts Mg2+ influx [52] or Na+-dependent Mg2+ efflux [49]. MRS2 regulates mitochondrial Mg2+ influx [53]. MagT1 shows strong specificity for Mg2+ with a voltage dependence [54] and localizes in the plasma and organelle membranes of endoplasmic reticulum and Golgi [55]. MagT1 regulates vascular endothelial survival, proliferation, and motility [45]. CNNM2 conducts Mg2+ uptake in a voltage-dependent manner [56] and can also regulate TRPM7-mediated Mg2+ transport [57], suggesting that these transporters can act in concert and feedback to each other.
These Mg2+ channels and transporters are not always specific to Mg2+ and can be permeable to multiple divalent cations with a selectivity that varies among the specific transporters. For example, TRPM7 also mediates Ca2+ transport, while the three SLC41A family isoforms do not transport Ca2+.

5. HypoMg Is More than Just Alterations in Ion Homeostasis

It is becoming clear that the changes with HypoMg are more specific than just the alterations in divalent cations. Recently, our group reported that Mg2+ modulates myocardial relaxation through regulating mitochondrial oxidative stress; HypoMg, in the presence or absence of diabetes mellitus, can cause mitochondrial reactive oxygen species overproduction (oxidative stress) and lead to the inability of the heart to properly relax and a type of heart failure known as diastolic heart failure, while Mg2+ supplementation or repletion can reverse these changes [5,13]. In these studies, the effects of HypoMg went beyond just having less cell energy, a loss of competition with Ca2+, and the disturbance of a divalent cation balance across the plasma membrane. HypoMg resulted in specific defects in the mitochondrial electron transport chain, increased mitochondrial oxidative stress, and mitochondrial membrane potential depolarization [5,13,41,58,59,60].
HypoMg is related to increased inflammation [61], either caused by or resulting in mitochondrial reactive oxygen species overproduction [5,62], in what likely constitutes a vicious cycle. Our recent studies show that HypoMg induces inflammation [6], which leads to mitochondrial oxidative stress and cardiac diastolic dysfunction or the failure of the heart to properly relax. This lack of relaxation can progress to a unique form of heart failure: heart failure with preserved ejection fraction (HFpEF) or diastolic heart failure, which is when the heart cannot pump enough blood because it does not fill sufficiently during the relaxation phase [63]. This type of heart failure has similar mortality and symptoms to that of heart failure, where the pump function is impaired, known as systolic heart failure. The difference in pathogenesis between these two types of heart failure is emphasized by the fact that HFpEF does not respond to therapies that are effective for systolic heart failure, and there are no known therapies for HFpEF that affect its poor mortality rate.
In an animal model of HFpEF, cardiac inflammation is manifested as increased interleukin-1β (IL-1β) levels and an altered cardiac macrophage phenotype with increased pro-inflammatory macrophages and decreased anti-inflammatory, pro-resolving macrophages [63]. IL-1β elevation triggers cardiomyocyte mitochondrial oxidative stress, causes the oxidation of cardiac myosin-binding protein C, and leads to cardiac diastolic dysfunction [63]. An IL-1 receptor antagonist, a mitochondrial-targeted antioxidant, or Mg2+ supplementation can improve mitochondrial oxidative stress and cardiac diastolic dysfunction caused either by HypoMg or diabetes [5,13,63], suggesting that HypoMg mediates this form of heart failure associated with diabetes.

6. TRPM7 May Mediate the Oxidative Stress and Inflammation of HypoMg

Among the Mg2+ transporters, TRPM7 has been extensively characterized. It regulates both whole-body Mg2+ homeostasis [64,65] and cellular Mg2+ homeostasis in many types of cells, including cardiomyocytes, vascular smooth muscle cells, and B lymphoma cells [66,67,68], but not T cells [69]. TRPM7 is highly permeable for Mg2+, Ca2+, Zn2+, Mn2+, and a few other divalent cations. TRPM7 both regulates and is regulated by Mg2+. TRPM7 currents are inhibited by the physiological concentration of Mg2+ and Mg-ATP.
In addition to its transport function, TRPM7 is a unique ion channel in that it has an α kinase domain that regulates protein transcription [70,71] and phosphorylation [72,73,74,75,76], as well as a sensor for Mg2+’s status [66]. TRPM7 is essential for early cardiogenesis [77], cardiac repolarization [77], cardiac automaticity and rhythmicity [78], vascular endothelial survival, proliferation and motility [45], and vascular smooth muscle cell growth [79].
The TRPM7 protein amount is counter-regulated to the Mg2+ concentration. That is to say, HypoMg increases TRPM7 transcription and translation. Increased TRPM7 is associated with increased kinase signaling. Under HypoMg, the upregulation of TRPM7 can be observed, including elevated TRPM7 mRNA translation [80], protein expression [6,45], and TRPM7-conduced Mg2+ currents [81,82,83]. A recent animal study on pulmonary hypertension shows an association between decreased [Mg2+]i in pulmonary arterial smooth muscle cells and the upregulation of TRPM7 expression, as well as the upregulation of SLC41A1, SCL41A2, CNNM2, MRS2, and MagT1 [35]. Furthermore, the upregulated TRPM7 function has been observed in patients with ischemia-reperfusion, arrhythmias, and hypertension [84,85,86].
Many of the effects of HypoMg can be explained by elevated TRPM7 and the activation of the TRPM7 kinase. For example, the inflammation and oxidative stress that are associated with HypoMg are correlated with TRPM7 expression and function. Our recent work shows that TRPM7 kinase activation via HypoMg contributes to the inflammation activation and oxidative stress observed in HypoMg-induced seizure activity, all of which can be significantly improved in transgenic mice with a K1646R mutation in the kinase domain that results in no kinase function while leaving transport function intact [6].
How TRPM7 kinase alters mitochondrial oxidative stress alongside inflammation is an area of active investigation. TRPM7-mediated Ca2+ signaling enhances macrophage activation in response to provocation. When selectively depleted in bone marrow macrophages, inflammatory marker IL-1β secretion was significantly reduced, and mice were resistant to peritonitis induced by a classic inflammatory mediator derived from bacteria [87]. Silencing TRPM7 improves inflammation and apoptosis in renal ischemic reperfusion injury, while the activation of TRPM7 exacerbates hypoxia/reoxygenation-induced renal injury [88].
What controls TRPM7 levels is also an area that needs more investigation. The inflammatory mediator IL-18 upregulates TRPM7 expression and currents in vascular smooth muscle cells [89], and angiotensin II and aldosterone, as mediators of hypertension and promoters of atherosclerosis, have been shown to activate TRPM7 to induce Mg2+ influx and oxidative stress production in vascular smooth muscle cells [75,79,90].
These data suggest a positive feedback loop between TRPM7 and inflammation. There is also a positive feedback loop between oxidative stress levels and TRPM7’s expression and function. Increased oxidative stress levels elevate TRPM7 expression and currents [91,92], and TRPM7 overexpression induces intracellular oxidative stress overproduction [6,93,94]. A similar positive feedback loop exists between oxidative stress and HypoMg in cardiomyocytes [5,95]. TRPM7 elevation with HypoMg could underlie this correlation between oxidative stress and HypoMg.
Other TRPM channels likely play synergistic roles in Mg2+ homeostasis and other deleterious effects of HypoMg. TRPM6, like TRPM7 with both the channel and kinase function, is predominantly expressed in the kidney, cecum, and colon. It plays an important role in Mg2+ absorption in the gastrointestinal tract and Mg2+ excretion in the kidney [96,97,98]. TRPM6 is also upregulated under HypoMg or oxidative stress [98,99,100]. This channel’s role in cardiovascular diseases is unclear, but it may amplify the effects of HypoMg or even be a primary cause of HypoMg. For example, TRPM6 upregulation may be important when understanding the association between diabetes and HypoMg [65,96,101,102]. Moreover, TRPM6 phosphorylates TRPM7 [103] and can, therefore, modulate TRPM7 activity in cardiovascular diseases. However, the full extent of the interplay between these two TRPM channels is unknown.

7. Treatment of HypoMg-Related and Other Inflammatory Disorders

The inverse relationship between TRPM7 and Mg2+ levels supports the use of Mg2+ supplementation to counteract the effect of HypoMg-related pathology. The role of TRPM7 in this pathology suggests that kinase inhibition may be another therapeutic approach. In this regard, dietary Mg2+ intake is inversely correlated to cardiovascular events, such as increased risk of heart failure, hospitalization [104], coronary heart disease [34], hypertension [20,30,31], stroke [12,32], and diabetic cardiomyopathy [12,13]. An obvious disease candidate in need of further investigation related to these proposed mechanisms is pre-eclampsia, where Mg2+ is well-established clinically as a therapy. Nevertheless, it is unclear whether Mg2+ supplementation in pre-eclampsia demonstrates some or all of its effects through the modulation of TRPM7, oxidative stress, and inflammation.
Aside from genetic approaches, the modulation of TRPM7 is a possibility in humans with small molecule therapies. For TRPM7, there are inhibitors for channel activity (e.g., NS8593, waixenicin A, and FTY720) and for the kinase function, such as TG100-115 (see review [105,106]). TG100-115 has been shown to restrict the infarct size after myocardial ischemia-reperfusion injury [107].

8. Unknowns

The discussion above suggests that there is more to the effect of Mg2+ supplementation than just generalized effects on enzyme kinetics and ion homeostasis, though there are many unknowns. Some of these are: How many diseases characterized by inflammation and oxidative stress respond to Mg2+ therapy? Do these diseases need to be associated with HypoMg or TRPM7 elevation to respond? Are specific cell types, such as white cells, mediating the effects of HypoMg, or is the response generalized to most cells? To what extent does TRPM7 kinase activity explain the many effects of HypoMg, and to what extent are cation changes important? What are the downstream effectors activated by TRPM7 kinase, and how does the kinase alter mitochondrial function? How are mitochondrial dysfunction and inflammation linked? How do reductions in Mg2+ result in TRPM7 transcriptional upregulation, and could this signaling cascade be interrupted? Can oral Mg2+ supplementation be effective for treating Mg2+-responsive diseases in humans, and can sufficient doses be achieved without inducing gastrointestinal side effects? Could TRPM7 kinase inhibition be a solution to the issues related to oral Mg2+ supplementation?

9. Conclusions

Above, we have discussed the idea that the effects of HypoMg are more than just a general loss of Mg2+ and the related shifts in other cations and their dependent processes. We have proposed that many of the effects of HypoMg could be explained by a signaling cascade involving the TRPM7 kinase, which is upregulated by HypoMg. We have argued that Mg2+ supplementation or TRPM7 kinase inhibition can prevent HypoMg-induced mitochondrial oxidative stress and inflammation and that these treatments may be useful in such conditions as diabetes, where Mg2+ deficiency is common. We have indicated that Mg2+ may be a simple therapy for a common type of heart failure: HFpEF. Moreover, we have speculated that these therapies could be helpful in diseases characterized by oxidative stress and inflammation, but not known to be associated with HypoMg. We have pointed out just some of the unknowns when following this train of thought.
What is clear is that there is more to Mg2+ than is generally appreciated, and these subtleties could lead us to new understandings of diseases and potential new therapies.

Author Contributions

Conceptualization, writing—original draft preparation, review and editing, M.L. and S.C.D.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Institutes of Health R56 HL162208 (SCD).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sharma, P.; Chung, C.; Vizcaychipi, M. Magnesium: The neglected electrolyte? A clinical review. Pharmacol. Pharm. 2014, 5, 762–772. [Google Scholar] [CrossRef]
  2. Takakuwa, Y.; Kanazawa, T. Reaction mechanism of (Ca2+, Mg2+)-ATPase of sarcoplasmic reticulum. The role of Mg2+ that activates hydrolysis of the phosphoenzyme. J. Biol. Chem. 1982, 257, 426–431. [Google Scholar] [CrossRef]
  3. Laver, D.R. Coupled calcium release channels and their regulation by luminal and cytosolic ions. Eur. Biophys. J. 2005, 34, 359–368. [Google Scholar] [CrossRef]
  4. Ford, E.S.; Mokdad, A.H. Dietary magnesium intake in a national sample of US adults. J. Nutr. 2003, 133, 2879–2882. [Google Scholar] [CrossRef]
  5. Liu, M.; Liu, H.; Feng, F.; Xie, A.; Kang, G.J.; Zhao, Y.; Hou, C.R.; Zhou, X.; Dudley, S.C., Jr. Magnesium deficiency causes a reversible, metabolic, diastolic cardiomyopathy. J. Am. Heart Assoc. 2021, 10, e020205. [Google Scholar] [CrossRef] [PubMed]
  6. Liu, M.; Liu, H.; Feng, F.; Krook-Magnuson, E.; Dudley, S.C.J. TRPM7 kinase mediates hypomagnesemia-induced seizure-related death. Sci. Rep. 2023, 13, 7855. [Google Scholar] [CrossRef]
  7. Schimatschek, H.F.; Rempis, R. Prevalence of hypomagnesemia in an unselected German population of 16,000 individuals. Magnes. Res. 2001, 14, 283–290. [Google Scholar]
  8. Guo, W.; Nazim, H.; Liang, Z.; Yang, D. Magnesium deficiency in plants: An urgent problem. Crop. J. 2016, 4, 83–91. [Google Scholar] [CrossRef]
  9. Moshfegh, A.; Goldman, J.; Ahuja, J.; Rhodes, D.; Lacomb, R. What we eat in America, NHANES 2005–2006, Usual Nutrient Intakes from Food and Water Compared to 1997 Dietary Reference Intakes for Vitamin D, Calcium, Phosphorus, and Magnesium. U.S. Department of Agriculture, Agricultural Research Service. Available online: https://www.ars.usda.gov/research/publications/publication/?seqNo115=243279 (accessed on 30 July 2020).
  10. Costello, R.B.; Elin, R.J.; Rosanoff, A.; Wallace, T.C.; Guerrero-Romero, F.; Hruby, A.; Lutsey, P.L.; Nielsen, F.H.; Rodriguez-Moran, M.; Song, Y.; et al. Perspective: The case for an evidence-based reference interval for serum magnesium: The time has come. Adv. Nutr. 2016, 7, 977–993. [Google Scholar] [CrossRef]
  11. Nielsen, F.H. Magnesium deficiency and increased inflammation: Current perspectives. J. Inflamm. Res. 2018, 11, 25–34. [Google Scholar] [CrossRef] [PubMed]
  12. Fang, X.; Wang, K.; Han, D.; He, X.; Wei, J.; Zhao, L.; Imam, M.U.; Ping, Z.; Li, Y.; Xu, Y.; et al. Dietary magnesium intake and the risk of cardiovascular disease, type 2 diabetes, and all-cause mortality: A dose-response meta-analysis of prospective cohort studies. BMC Med. 2016, 14, 210. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, M.; Jeong, E.-M.; Liu, H.; Xie, A.; So, E.Y.; Shi, G.; Jeong, G.E.; Zhou, A.; Dudley, S.C., Jr. Magnesium supplementation improves diabetic mitochondrial and cardiac diastolic function. JCI Insight 2019, 4, e123182. [Google Scholar] [CrossRef] [PubMed]
  14. Gommers, L.M.; Hoenderop, J.G.; Bindels, R.J.; de Baaij, J.H. Hypomagnesemia in type 2 diabetes: A vicious circle? Diabetes 2016, 65, 3–13. [Google Scholar] [CrossRef]
  15. Martins, I.J. Magnesium deficiency and induction of NAFLD and type 3 diabetes in Australasia. Australas. Med. J. 2017, 10, 235–237. [Google Scholar] [CrossRef]
  16. Shechter, M. Magnesium and cardiovascular system. Magnes. Res. 2010, 23, 60–72. [Google Scholar] [PubMed]
  17. Tangvoraphonkchai, K.; Davenport, A. Magnesium and cardiovascular disease. Adv. Chronic Kidney Dis. 2018, 25, 251–260. [Google Scholar] [CrossRef]
  18. Liu, M.; Dudley, S.C., Jr. Magnesium, oxidative stress, inflammation, and cardiovascular disease. Antioxidants 2020, 9, 907. [Google Scholar] [CrossRef]
  19. Reffelmann, T.; Ittermann, T.; Dörr, M.; Völzke, H.; Reinthaler, M.; Petersmann, A.; Felix, S.B. Low serum magnesium concentrations predict cardiovascular and all-cause mortality. Atherosclerosis 2011, 219, 280–284. [Google Scholar] [CrossRef]
  20. Chrysant, S.G.; Chrysant, G.S. Association of hypomagnesemia with cardiovascular diseases and hypertension. Int. J. Cardiol. Hypertens. 2019, 1, 100005. [Google Scholar] [CrossRef]
  21. Volpe, S.L. Magnesium in disease prevention and overall health. Adv. Nutr. 2013, 4, 378S–383S. [Google Scholar] [CrossRef] [PubMed]
  22. de Baaij, J.H.; Hoenderop, J.G.; Bindels, R.J. Magnesium in man: Implications for health and disease. Physiol. Rev. 2015, 95, 1–46. [Google Scholar] [CrossRef] [PubMed]
  23. Al Alawi, A.M.; Majoni, S.W.; Falhammar, H. Magnesium and human health: Perspectives and research directions. Int. J. Endocrinol. 2018, 2018, 9041694. [Google Scholar] [CrossRef]
  24. Schwinger, R.H.; Erdmann, E. Heart failure and electrolyte disturbances. Methods Find. Exp. Clin. Pharmacol. 1992, 14, 315–325. [Google Scholar]
  25. Milionis, H.J.; Alexandrides, G.E.; Liberopoulos, E.N.; Bairaktari, E.T.; Goudevenos, J.; Elisaf, M.S. Hypomagnesemia and concurrent acid–base and electrolyte abnormalities in patients with congestive heart failure. Eur. J. Heart Fail. 2002, 4, 167–173. [Google Scholar] [CrossRef] [PubMed]
  26. Ceremużyński, L.; Gębalska, J.; Wołk, R.; Makowska, E. Hypomagnesemia in heart failure with ventricular arrhythmias. Beneficial effects of magnesium supplementation. J. Intern. Med. 2000, 247, 78–86. [Google Scholar] [CrossRef]
  27. Smith, L.F.; Heagerty, A.M.; Bing, R.F.; Barnett, D.B. Intravenous infusion of magnesium sulphate after acute myocardial infarction: Effects on arrhythmias and mortality. Int. J. Cardiol. 1986, 12, 175–183. [Google Scholar] [CrossRef]
  28. Parikka, H.; Toivonen, L.; Naukkarinen, V.; Tierala, I.; Pohjola-Sintonen, S.; Heikkilä, J.; Nieminen, M.S. Decreases by magnesium of QT dispersion and ventricular arrhythmias in patients with acute myocardial infarction. Eur. Heart J. 1999, 20, 111–120. [Google Scholar] [CrossRef]
  29. Kostov, K.; Halacheva, L. Role of magnesium deficiency in promoting atherosclerosis, endothelial dysfunction, and arterial stiffening as risk factors for hypertension. Int. J. Mol. Sci. 2018, 19, 1724. [Google Scholar] [CrossRef]
  30. Ascherio, A.; Rimm, E.B.; Giovannucci, E.L.; Colditz, G.A.; Rosner, B.; Willett, W.C.; Sacks, F.; Stampfer, M.J. A prospective study of nutritional factors and hypertension among US men. Circulation 1992, 86, 1475–1484. [Google Scholar] [CrossRef] [PubMed]
  31. Ascherio, A.; Hennekens, C.; Willett, W.C.; Sacks, F.; Rosner, B.; Manson, J.; Witteman, J.; Stampfer, M.J. Prospective study of nutritional factors, blood pressure, and hypertension among US women. Hypertension 1996, 27, 1065–1072. [Google Scholar] [CrossRef]
  32. Zhao, B.; Hu, L.; Dong, Y.; Xu, J.; Wei, Y.; Yu, D.; Xu, J.; Zhang, W. The effect of magnesium intake on stroke incidence: A systematic review and meta-analysis with trial sequential analysis. Front. Neurol. 2019, 10, 852. [Google Scholar] [CrossRef]
  33. Ford, E.S. Serum magnesium and ischaemic heart disease: Findings from a national sample of US adults. Int. J. Epidemiol. 1999, 28, 645–651. [Google Scholar] [CrossRef]
  34. Abbott, R.D.; Ando, F.; Masaki, K.H.; Tung, K.H.; Rodriguez, B.L.; Petrovitch, H.; Yano, K.; Curb, J.D. Dietary magnesium intake and the future risk of coronary heart disease (the Honolulu Heart Program). Am. J. Cardiol. 2003, 92, 665–669. [Google Scholar] [CrossRef]
  35. Wang, D.; Zhu, Z.L.; Lin, D.C.; Zheng, S.Y.; Chuang, K.H.; Gui, L.X.; Yao, R.H.; Zhu, W.J.; Sham, J.S.K.; Lin, M.J. Magnesium supplementation attenuates pulmonary hypertension via regulation of magnesium transporters. Hypertension 2021, 77, 617–631. [Google Scholar] [CrossRef]
  36. Hatzistavri, L.S.; Sarafidis, P.A.; Georgianos, P.I.; Tziolas, I.M.; Aroditis, C.P.; Zebekakis, P.E.; Pikilidou, M.I.; Lasaridis, A.N. Oral magnesium supplementation reduces ambulatory blood pressure in patients with mild hypertension. Am. J. Hypertens. 2009, 22, 1070–1075. [Google Scholar] [CrossRef]
  37. Zhang, X.; Li, Y.; Del Gobbo, L.C.; Rosanoff, A.; Wang, J.; Zhang, W.; Song, Y. Effects of magnesium supplementation on blood pressure: A meta-analysis of randomized double-blind placebo-controlled trials. Hypertension 2016, 68, 324–333. [Google Scholar] [CrossRef] [PubMed]
  38. Pham, P.C.; Pham, P.A.; Pham, S.V.; Pham, P.T.; Pham, P.M.; Pham, P.T. Hypomagnesemia: A clinical perspective. Int. J. Nephrol. Renovasc. Dis. 2014, 7, 219–230. [Google Scholar] [CrossRef] [PubMed]
  39. Tashiro, M.; Inoue, H.; Konishi, M. Physiological pathway of magnesium influx in rat ventricular myocytes. Biophys. J. 2014, 107, 2049–2058. [Google Scholar] [CrossRef]
  40. Jung, D.W.; Apel, L.; Brierley, G.P. Matrix free Mg2+ changes with metabolic state in isolated heart mitochondria. Biochemistry 1990, 29, 4121–4128. [Google Scholar] [CrossRef] [PubMed]
  41. Mubagwa, K.; Gwanyanya, A.; Zakharov, S.; Macianskiene, R. Regulation of cation channels in cardiac and smooth muscle cells by intracellular magnesium. Arch. Biochem. Biophys. 2007, 458, 73–89. [Google Scholar] [CrossRef] [PubMed]
  42. Touyz, R.M.; Laurant, P.; Schiffrin, E.L. Effect of magnesium on calcium responses to vasopressin in vascular smooth muscle cells of spontaneously hypertensive rats. J. Pharmacol. Exp. Ther. 1998, 284, 998–1005. [Google Scholar] [PubMed]
  43. Yang, Z.W.; Wang, J.; Zheng, T.; Altura, B.T.; Altura, B.M. Low [Mg(2+)](o) induces contraction of cerebral arteries: Roles of tyrosine and mitogen-activated protein kinases. Am. J. Physiol. Heart Circ. Physiol. 2000, 279, H185–H194. [Google Scholar] [CrossRef] [PubMed]
  44. Laurant, P.; Touyz, R.M. Physiological and pathophysiological role of magnesium in the cardiovascular system: Implications in hypertension. J. Hypertens. 2000, 18, 1177–1191. [Google Scholar] [CrossRef]
  45. Zhu, D.; You, J.; Zhao, N.; Xu, H. Magnesium Regulates Endothelial Barrier Functions through TRPM7, MagT1, and S1P1. Adv. Sci. 2019, 6, 1901166. [Google Scholar] [CrossRef] [PubMed]
  46. Yogi, A.; Callera, G.E.; Antunes, T.T.; Tostes, R.C.; Touyz, R.M. Vascular biology of magnesium and its transporters in hypertension. Magnes. Res. 2010, 23, S207–S215. [Google Scholar] [CrossRef]
  47. Romani, A.M.P. Cellular magnesium homeostasis. Arch. Biochem. Biophys. 2011, 512, 1–23. [Google Scholar] [CrossRef] [PubMed]
  48. Nemoto, T.; Tagashira, H.; Kita, T.; Kita, S.; Iwamoto, T. Functional characteristics and therapeutic potential of SLC41 transporters. J. Pharmacol. Sci. 2023, 151, 88–92. [Google Scholar] [CrossRef]
  49. Kolisek, M.; Nestler, A.; Vormann, J.; Schweigel-Röntgen, M. Human gene SLC41A1 encodes for the Na+/Mg2+ exchanger. Am. J. Physiol. Cell Physiol. 2012, 302, C318–C326. [Google Scholar] [CrossRef]
  50. Goytain, A.; Quamme, G.A. Functional characterization of the mouse [corrected] solute carrier, SLC41A2. Biochem. Biophys. Res. Commun. 2005, 330, 701–705. [Google Scholar] [CrossRef]
  51. Mastrototaro, L.; Smorodchenko, A.; Aschenbach, J.R.; Kolisek, M.; Sponder, G. Solute carrier 41A3 encodes for a mitochondrial Mg2+ efflux system. Sci. Rep. 2016, 6, 27999. [Google Scholar] [CrossRef]
  52. Goytain, A.; Quamme, G.A. Functional characterization of human SLC41A1, a Mg2+ transporter with similarity to prokaryotic MgtE Mg2+ transporters. Physiol. Genom. 2005, 21, 337–342. [Google Scholar] [CrossRef]
  53. Kolisek, M.; Zsurka, G.; Samaj, J.; Weghuber, J.; Schweyen, R.J.; Schweigel, M. Mrs2p is an essential component of the major electrophoretic Mg2+ influx system in mitochondria. EMBO J. 2003, 22, 1235–1244. [Google Scholar] [CrossRef] [PubMed]
  54. Goytain, A.; Quamme, G.A. Identification and characterization of a novel mammalian Mg2+ transporter with channel-like properties. BMC Genom. 2005, 6, 48. [Google Scholar] [CrossRef] [PubMed]
  55. Matsuda-Lennikov, M.; Biancalana, M.; Zou, J.; Ravell, J.C.; Zheng, L.; Kanellopoulou, C.; Jiang, P.; Notarangelo, G.; Jing, H.; Masutani, E.; et al. Magnesium transporter 1 (MAGT1) deficiency causes selective defects in N-linked glycosylation and expression of immune-response genes. J. Biol. Chem. 2019, 294, 13638–13656. [Google Scholar] [CrossRef] [PubMed]
  56. Goytain, A.; Quamme, G.A. Functional characterization of ACDP2 (ancient conserved domain protein), a divalent metal transporter. Physiol. Genom. 2005, 22, 382–389. [Google Scholar] [CrossRef]
  57. Bai, Z.; Feng, J.; Franken, G.A.C.; Al’Saadi, N.; Cai, N.; Yu, A.S.; Lou, L.; Komiya, Y.; Hoenderop, J.G.J.; de Baaij, J.H.F.; et al. CNNM proteins selectively bind to the TRPM7 channel to stimulate divalent cation entry into cells. PLoS Biol. 2021, 19, e3001496. [Google Scholar] [CrossRef] [PubMed]
  58. Kumar, B.P.; Shivakumar, K. Depressed antioxidant defense in rat heart in experimental magnesium deficiency. Implications for the pathogenesis of myocardial lesions. Biol. Trace Elem. Res. 1997, 60, 139–144. [Google Scholar] [CrossRef] [PubMed]
  59. Racay, P. Effect of magnesium on calcium-induced depolarisation of mitochondrial transmembrane potential. Cell Biol. Int. 2008, 32, 136–145. [Google Scholar] [CrossRef]
  60. Shah, N.C.; Liu, J.-P.; Iqbal, J.; Hussain, M.; Jiang, X.-C.; Li, Z.; Li, Y.; Zheng, T.; Li, W.; Sica, A.C.; et al. Mg deficiency results in modulation of serum lipids, glutathione, and NO synthase isozyme activation in cardiovascular tissues: Relevance to de novo synthesis of ceramide, serum Mg and atherogenesis. Int. J. Clin. Exp. Med. 2011, 4, 103–118. [Google Scholar]
  61. Shahi, A.; Aslani, S.; Ataollahi, M.; Mahmoudi, M. The role of magnesium in different inflammatory diseases. Inflammopharmacology 2019, 27, 649–661. [Google Scholar] [CrossRef]
  62. Villa-Bellosta, R. Dietary magnesium supplementation improves lifespan in a mouse model of progeria. EMBO Mol. Med. 2020, 12, e12423. [Google Scholar] [CrossRef]
  63. Liu, H.; Huang, Y.; Zhao, Y.; Kang, G.-J.; Feng, F.; Wang, X.; Liu, M.; Shi, G.; Revelo, X.; Bernlohr, D.; et al. Inflammatory macrophage interleukin-1β mediates high fat diet induced heart failure with preserved ejection fraction. JACC. Basic Transl. Sci. 2023, 8, 174–185. [Google Scholar] [CrossRef]
  64. Ryazanova, L.V.; Rondon, L.J.; Zierler, S.; Hu, Z.; Galli, J.; Yamaguchi, T.P.; Mazur, A.; Fleig, A.; Ryazanov, A.G. TRPM7 is essential for Mg2+ homeostasis in mammals. Nat. Commun. 2010, 1, 109. [Google Scholar] [CrossRef] [PubMed]
  65. Chubanov, V.; Mittermeier, L.; Gudermann, T. Role of kinase-coupled TRP channels in mineral homeostasis. Pharmacol. Ther. 2018, 184, 159–176. [Google Scholar] [CrossRef]
  66. Schmitz, C.; Perraud, A.L.; Johnson, C.O.; Inabe, K.; Smith, M.K.; Penner, R.; Kurosaki, T.; Fleig, A.; Scharenberg, A.M. Regulation of vertebrate cellular Mg2+ homeostasis by TRPM7. Cell 2003, 114, 191–200. [Google Scholar] [CrossRef]
  67. Callera, G.E.; He, Y.; Yogi, A.; Montezano, A.C.; Paravicini, T.; Yao, G.; Touyz, R.M. Regulation of the novel Mg2+ transporter transient receptor potential melastatin 7 (TRPM7) cation channel by bradykinin in vascular smooth muscle cells. J. Hypertens. 2009, 27, 155–166. [Google Scholar] [CrossRef] [PubMed]
  68. Tashiro, M.; Inoue, H.; Konishi, M. Modulation of Mg2+ influx and cytoplasmic free Mg2+ concentration in rat ventricular myocytes. J. Physiol. Sci. 2019, 69, 97–102. [Google Scholar] [CrossRef]
  69. Jin, J.; Desai, B.N.; Navarro, B.; Donovan, A.; Andrews, N.C.; Clapham, D.E. Deletion of Trpm7 disrupts embryonic development and thymopoiesis without altering Mg2+ homeostasis. Science 2008, 322, 756–760. [Google Scholar] [CrossRef]
  70. Lee, B.C.; Hong, S.E.; Lim, H.H.; Kim, D.H.; Park, C.S. Alteration of the transcriptional profile of human embryonic kidney cells by transient overexpression of mouse TRPM7 channels. Cell Physiol. Biochem. 2011, 27, 313–326. [Google Scholar] [CrossRef] [PubMed]
  71. Krapivinsky, G.; Krapivinsky, L.; Manasian, Y.; Clapham, D.E. The TRPM7 chanzyme is cleaved to release a chromatin-modifying kinase. Cell 2014, 157, 1061–1072. [Google Scholar] [CrossRef]
  72. Qiao, W.; Wong, K.H.M.; Shen, J.; Wang, W.; Wu, J.; Li, J.; Lin, Z.; Chen, Z.; Matinlinna, J.P.; Zheng, Y.; et al. TRPM7 kinase-mediated immunomodulation in macrophage plays a central role in magnesium ion-induced bone regeneration. Nat. Commun. 2021, 12, 2885. [Google Scholar] [CrossRef] [PubMed]
  73. Ryazanova, L.V.; Dorovkov, M.V.; Ansari, A.; Ryazanov, A.G. Characterization of the protein kinase activity of TRPM7/ChaK1, a protein kinase fused to the transient receptor potential ion channel. J. Biol. Chem. 2004, 279, 3708–3716. [Google Scholar] [CrossRef]
  74. Gotru, S.K.; Chen, W.; Kraft, P.; Becker, I.C.; Wolf, K.; Stritt, S.; Zierler, S.; Hermanns, H.M.; Rao, D.; Perraud, A.L.; et al. TRPM7 kinase controls calcium responses in arterial thrombosis and stroke in mice. Arterioscler. Thromb. Vasc. Biol. 2018, 38, 344–352. [Google Scholar] [CrossRef] [PubMed]
  75. Zhang, Z.; Wang, M.; Fan, X.H.; Chen, J.H.; Guan, Y.Y.; Tang, Y.B. Upregulation of TRPM7 channels by angiotensin II triggers phenotypic switching of vascular smooth muscle cells of ascending aorta. Circ. Res. 2012, 111, 1137–1146. [Google Scholar] [CrossRef] [PubMed]
  76. Meng, X.; Cai, C.; Wu, J.; Cai, S.; Ye, C.; Chen, H.; Yang, Z.; Zeng, H.; Shen, Q.; Zou, F. TRPM7 mediates breast cancer cell migration and invasion through the MAPK pathway. Cancer Lett. 2013, 333, 96–102. [Google Scholar] [CrossRef]
  77. Sah, R.; Mesirca, P.; Mason, X.; Gibson, W.; Bates-Withers, C.; Van den Boogert, M.; Chaudhuri, D.; Pu, W.T.; Mangoni, M.E.; Clapham, D.E. Timing of myocardial trpm7 deletion during cardiogenesis variably disrupts adult ventricular function, conduction, and repolarization. Circulation 2013, 128, 101–114. [Google Scholar] [CrossRef] [PubMed]
  78. Sah, R.; Mesirca, P.; Van den Boogert, M.; Rosen, J.; Mably, J.; Mangoni, M.E.; Clapham, D.E. Ion channel-kinase TRPM7 is required for maintaining cardiac automaticity. Proc. Natl. Acad. Sci. USA 2013, 110, E3037–E3046. [Google Scholar] [CrossRef]
  79. He, Y.; Yao, G.; Savoia, C.; Touyz, R.M. Transient receptor potential melastatin 7 ion channels regulate magnesium homeostasis in vascular smooth muscle cells: Role of angiotensin II. Circ. Res. 2005, 96, 207–215. [Google Scholar] [CrossRef]
  80. Nikonorova, I.A.; Kornakov, N.V.; Dmitriev, S.E.; Vassilenko, K.S.; Ryazanov, A.G. Identification of a Mg2+-sensitive ORF in the 5’-leader of TRPM7 magnesium channel mRNA. Nucleic Acids Res. 2014, 42, 12779–12788. [Google Scholar] [CrossRef]
  81. Nadler, M.J.; Hermosura, M.C.; Inabe, K.; Perraud, A.L.; Zhu, Q.; Stokes, A.J.; Kurosaki, T.; Kinet, J.P.; Penner, R.; Scharenberg, A.M.; et al. LTRPC7 is a Mg.ATP-regulated divalent cation channel required for cell viability. Nature 2001, 411, 590–595. [Google Scholar] [CrossRef]
  82. Gwanyanya, A.; Amuzescu, B.; Zakharov, S.I.; Macianskiene, R.; Sipido, K.R.; Bolotina, V.M.; Vereecke, J.; Mubagwa, K. Magnesium-inhibited, TRPM6/7-like channel in cardiac myocytes: Permeation of divalent cations and pH-mediated regulation. J. Physiol. 2004, 559, 761–776. [Google Scholar] [CrossRef] [PubMed]
  83. Macianskiene, R.; Martisiene, I.; Zablockaite, D.; Gendviliene, V. Characterization of Mg2+-regulated TRPM7-like current in human atrial myocytes. J. Biomed. Sci. 2012, 19, 75. [Google Scholar] [CrossRef] [PubMed]
  84. Du, J.; Xie, J.; Zhang, Z.; Tsujikawa, H.; Fusco, D.; Silverman, D.; Liang, B.; Yue, L. TRPM7-mediated Ca2+ signals confer fibrogenesis in human atrial fibrillation. Circ. Res. 2010, 106, 992–1003. [Google Scholar] [CrossRef]
  85. Touyz, R.M. Transient receptor potential melastatin 6 and 7 channels, magnesium transport, and vascular biology: Implications in hypertension. Am. J. Physiol. Heart Circ. Physiol. 2008, 294, H1103–H1118. [Google Scholar] [CrossRef]
  86. Andriulė, I.; Pangonytė, D.; Almanaitytė, M.; Patamsytė, V.; Kuprytė, M.; Karčiauskas, D.; Mubagwa, K.; Mačianskienė, R. Evidence for the expression of TRPM6 and TRPM7 in cardiomyocytes from all four chamber walls of the human heart. Sci. Rep. 2021, 11, 15445. [Google Scholar] [CrossRef] [PubMed]
  87. Schappe, M.S.; Szteyn, K.; Stremska, M.E.; Mendu, S.K.; Downs, T.K.; Seegren, P.V.; Mahoney, M.A.; Dixit, S.; Krupa, J.K.; Stipes, E.J.; et al. Chanzyme TRPM7 mediates the Ca2+ influx essential for lipopolysaccharide-induced toll-like receptor 4 endocytosis and macrophage activation. Immunity 2018, 48, 59–74.e55. [Google Scholar] [CrossRef]
  88. Liu, A.; Wu, J.; Yang, C.; Wu, Y.; Zhang, Y.; Zhao, F.; Wang, H.; Yuan, L.; Song, L.; Zhu, T.; et al. TRPM7 in CHBP-induced renoprotection upon ischemia reperfusion-related injury. Sci. Rep. 2018, 8, 5510. [Google Scholar] [CrossRef]
  89. Zhang, K.; Zhang, Y.; Feng, W.; Chen, R.; Chen, J.; Touyz, R.M.; Wang, J.; Huang, H. Interleukin-18 enhances vascular calcification and osteogenic differentiation of vascular smooth muscle cells through TRPM7 activation. Arterioscler. Thromb. Vasc. Biol. 2017, 37, 1933–1943. [Google Scholar] [CrossRef]
  90. Yogi, A.; Callera, G.E.; O’Connor, S.; Antunes, T.T.; Valinsky, W.; Miquel, P.; Montezano, A.C.I.; Perraud, A.-L.; Schmitz, C.; Shrier, A.; et al. Aldosterone signaling through transient receptor potential melastatin 7 cation channel (TRPM7) and its α-kinase domain. Cell Signal. 2013, 25, 2163–2175. [Google Scholar] [CrossRef]
  91. Coombes, E.; Jiang, J.; Chu, X.P.; Inoue, K.; Seeds, J.; Branigan, D.; Simon, R.P.; Xiong, Z.G. Pathophysiologically relevant levels of hydrogen peroxide induce glutamate-independent neurodegeneration that involves activation of transient receptor potential melastatin 7 channels. Antioxid. Redox Signal. 2011, 14, 1815–1827. [Google Scholar] [CrossRef]
  92. Nunez-Villena, F.; Becerra, A.; Echeverria, C.; Briceno, N.; Porras, O.; Armisen, R.; Varela, D.; Montorfano, I.; Sarmiento, D.; Simon, F. Increased expression of the transient receptor potential melastatin 7 channel is critically involved in lipopolysaccharide-induced reactive oxygen species-mediated neuronal death. Antioxid. Redox Signal. 2011, 15, 2425–2438. [Google Scholar] [CrossRef]
  93. Su, L.-T.; Chen, H.-C.; González-Pagán, O.; Overton, J.D.; Xie, J.; Yue, L.; Runnels, L.W. TRPM7 activates m-calpain by stress-dependent stimulation of p38 MAPK and c-Jun N-terminal kinase. J. Mol. Biol. 2010, 396, 858–869. [Google Scholar] [CrossRef]
  94. Simon, F.; Varela, D.; Cabello-Verrugio, C. Oxidative stress-modulated TRPM ion channels in cell dysfunction and pathological conditions in humans. Cell Signal. 2013, 25, 1614–1624. [Google Scholar] [CrossRef]
  95. Tashiro, M.; Konishi, M.; Watanabe, M.; Yokoyama, U. Reduction of intracellular Mg2+ caused by reactive oxygen species in rat ventricular myocytes. Am. J. Physiol. Cell Physiol. 2023, 324, C963–C969. [Google Scholar] [CrossRef]
  96. Schlingmann, K.P.; Waldegger, S.; Konrad, M.; Chubanov, V.; Gudermann, T. TRPM6 and TRPM7—Gatekeepers of human magnesium metabolism. Biochim. Biophys. Acta 2007, 1772, 813–821. [Google Scholar] [CrossRef] [PubMed]
  97. Cabezas-Bratesco, D.; Brauchi, S.; González-Teuber, V.; Steinberg, X.; Valencia, I.; Colenso, C. The different roles of the channel-kinases TRPM6 and TRPM7. Curr. Med. Chem. 2015, 22, 2943–2953. [Google Scholar] [CrossRef]
  98. Tashiro, M.; Inoue, H.; Konishi, M. Magnesium homeostasis in cardiac myocytes of Mg-deficient rats. PLoS ONE 2013, 8, e73171. [Google Scholar] [CrossRef]
  99. Wolf, F.I.; Trapani, V.; Simonacci, M.; Mastrototaro, L.; Cittadini, A.; Schweigel, M. Modulation of TRPM6 and Na+/Mg2+ exchange in mammary epithelial cells in response to variations of magnesium availability. J. Cell Physiol. 2010, 222, 374–381. [Google Scholar] [CrossRef]
  100. Brandao, K.; Deason-Towne, F.; Perraud, A.L.; Schmitz, C. The role of Mg2+ in immune cells. Immunol. Res. 2013, 55, 261–269. [Google Scholar] [CrossRef] [PubMed]
  101. Kim, E.Y.; Lee, J.M. Transcriptional Control of Trpm6 by the Nuclear Receptor FXR. Int. J. Mol. Sci. 2022, 23, 1980. [Google Scholar] [CrossRef] [PubMed]
  102. Groenestege, W.M.; Hoenderop, J.G.; van den Heuvel, L.; Knoers, N.; Bindels, R.J. The epithelial Mg2+ channel transient receptor potential melastatin 6 is regulated by dietary Mg2+ content and estrogens. J. Am. Soc. Nephrol. 2006, 17, 1035–1043. [Google Scholar] [CrossRef] [PubMed]
  103. Schmitz, C.; Dorovkov, M.V.; Zhao, X.; Davenport, B.J.; Ryazanov, A.G.; Perraud, A.L. The channel kinases TRPM6 and TRPM7 are functionally nonredundant. J. Biol. Chem. 2005, 280, 37763–37771. [Google Scholar] [CrossRef]
  104. Taveira, T.H.; Ouellette, D.; Gulum, A.; Choudhary, G.; Eaton, C.B.; Liu, S.; Wu, W.C. Relation of magnesium intake with cardiac function and heart failure hospitalizations in black adults: The Jackson Heart Study. Circ. Heart Fail. 2016, 9, e002698. [Google Scholar] [CrossRef]
  105. Hofmann, T.; Schäfer, S.; Linseisen, M.; Sytik, L.; Gudermann, T.; Chubanov, V. Activation of TRPM7 channels by small molecules under physiological conditions. Pflug. Arch. 2014, 466, 2177–2189. [Google Scholar] [CrossRef] [PubMed]
  106. Chubanov, V.; Schäfer, S.; Ferioli, S.; Gudermann, T. Natural and synthetic modulators of the TRPM7 channel. Cells 2014, 3, 1089–1101. [Google Scholar] [CrossRef] [PubMed]
  107. Doukas, J.; Wrasidlo, W.; Noronha, G.; Dneprovskaia, E.; Fine, R.; Weis, S.; Hood, J.; Demaria, A.; Soll, R.; Cheresh, D. Phosphoinositide 3-kinase gamma/delta inhibition limits infarct size after myocardial ischemia/reperfusion injury. Proc. Natl. Acad. Sci. USA 2006, 103, 19866–19871. [Google Scholar] [CrossRef] [PubMed]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, M.; Dudley, S.C., Jr. Beyond Ion Homeostasis: Hypomagnesemia, Transient Receptor Potential Melastatin Channel 7, Mitochondrial Function, and Inflammation. Nutrients 2023, 15, 3920. https://doi.org/10.3390/nu15183920

AMA Style

Liu M, Dudley SC Jr. Beyond Ion Homeostasis: Hypomagnesemia, Transient Receptor Potential Melastatin Channel 7, Mitochondrial Function, and Inflammation. Nutrients. 2023; 15(18):3920. https://doi.org/10.3390/nu15183920

Chicago/Turabian Style

Liu, Man, and Samuel C. Dudley, Jr. 2023. "Beyond Ion Homeostasis: Hypomagnesemia, Transient Receptor Potential Melastatin Channel 7, Mitochondrial Function, and Inflammation" Nutrients 15, no. 18: 3920. https://doi.org/10.3390/nu15183920

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

Article Metrics

Back to TopTop