Skip to main content
Erschienen in: Clinical and Experimental Medicine 1/2024

Open Access 01.12.2024 | Review

Role of zinc in health and disease

verfasst von: Lucy I. Stiles, Kevin Ferrao, Kosha J. Mehta

Erschienen in: Clinical and Experimental Medicine | Ausgabe 1/2024

Abstract

This review provides a concise overview of the cellular and clinical aspects of the role of zinc, an essential micronutrient, in human physiology and discusses zinc-related pathological states. Zinc cannot be stored in significant amounts, so regular dietary intake is essential. ZIP4 and/or ZnT5B transport dietary zinc ions from the duodenum into the enterocyte, ZnT1 transports zinc ions from the enterocyte into the circulation, and ZnT5B (bidirectional zinc transporter) facilitates endogenous zinc secretion into the intestinal lumen. Putative promoters of zinc absorption that increase its bioavailability include amino acids released from protein digestion and citrate, whereas dietary phytates, casein and calcium can reduce zinc bioavailability. In circulation, 70% of zinc is bound to albumin, and the majority in the body is found in skeletal muscle and bone. Zinc excretion is via faeces (predominantly), urine, sweat, menstrual flow and semen. Excessive zinc intake can inhibit the absorption of copper and iron, leading to copper deficiency and anaemia, respectively. Zinc toxicity can adversely affect the lipid profile and immune system, and its treatment depends on the mode of zinc acquisition. Acquired zinc deficiency usually presents later in life alongside risk factors like malabsorption syndromes, but medications like diuretics and angiotensin-receptor blockers can also cause zinc deficiency. Inherited zinc deficiency condition acrodermatitis enteropathica, which occurs due to mutation in the SLC39A4 gene (encoding ZIP4), presents from birth. Treatment involves zinc supplementation via zinc gluconate, zinc sulphate or zinc chloride. Notably, oral zinc supplementation may decrease the absorption of drugs like ciprofloxacin, doxycycline and risedronate.
Hinweise

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Introduction

Significance of zinc in human health

Zinc cannot be synthesised within the human body, so external intake of zinc is essential to maintain adequate levels in the body [1]. It is the second most abundant trace element in the body, after iron [2]. One in ten proteins found in the body is a zinc protein [3], and more than 300 enzymes and 1000 transcription factors depend on zinc for their activities [4]. Thus, zinc is an essential micronutrient involved in many cellular processes such as protein synthesis, nucleic acid metabolism including DNA synthesis, gene transcription [1], cell proliferation and differentiation, and mitosis [5].
These zinc-requiring cellular processes extend the significance of zinc to physiological level. For example, zinc is a structural component of the bone tissue and plays a role in collagen matrix synthesis, mineralisation, and bone turnover [6]. Also, zinc regulates intracellular signalling pathways of innate and adaptive immune cells [7], influences immune responses including antibody production, inflammatory signalling and lymphocyte differentiation [8], and thereby plays an essential role in the functionality of the immune system.
Zinc also plays a role in the endocrine system. For example, zinc is required in the formation and structural stability of insulin [9]. Essentially, insulin dimers form hexameric units, coordinated by two zinc ions in the central axis of the hexamer [10]. No wonder the beta cells of the pancreas contain significantly higher concentrations of zinc than other cells of the body. Furthermore, zinc ions act on the insulin signalling pathway and stimulate lipogenesis and glucose uptake into the adipocytes [9]. Zinc transporter (ZnT)-8 mediates signalling between the pancreas and liver to allow optimal insulin release, while zinc/iron-regulated-transporter-like-protein-(ZIP)7 is thought to play a role in glycaemic control within skeletal muscle. [9].
Thyroid hormones are involved in many physiological functions, such as the anabolism of proteins and increasing the basal metabolic rate and bone growth in children [11]. Zinc plays an important role in the metabolism of thyroid hormones. It regulates the synthesis of thyroid-releasing hormone (TRH) and thyroid-stimulating hormone (TSH). Zinc modulates their structure and thereby regulates the transcription factors which are essential for thyroid hormone synthesis [12]. Thus, in humans with zinc deficiency, levels of TSH, serum triiodothyronine (T3) and thyroxine (T4) also decrease [13], with several studies suggesting zinc deficiency as a cause of subclinical hypothyroidism [14]. Unsurprisingly, zinc supplementation appears to enhance thyroid hormone levels, particularly T3 [15].
Zinc is essential for male fertility. A zinc-sensing receptor, known as GPR39, has been found in the sperm tail and acrosome. When extracellular Zinc(II) binds to this receptor, it triggers an intracellular signalling pathway that ultimately results in increased sperm motility and acrosomal exocytosis [16]. Thus, zinc could have a role in the prevention, diagnosis and treatment of male infertility [17].
Additionally, zinc is important for the normal development and functioning of the central nervous system (CNS). Zinc balance is vital for neural tube formation and stem cell proliferation during development. Various zinc-dependent enzymes contribute to the function of the CNS, and ‘free’ zinc appears to modulate a variety of post-synaptic receptors. For example, zinc inhibits GABA-A receptors, which reduces their inhibitory actions. Alterations in zinc levels thereby affect the CNS and play a role in conditions such as Alzheimer’s disease and depression [18].
The human retina contains zinc in high concentrations [19]. Studies have suggested a link between higher anti-oxidant intake, including zinc (likely because zinc is a cofactor of superoxide dismutase, an anti-oxidant enzyme and also because zinc is an inhibitor of NADPH oxidase [20], which catalyses the production of reactive oxygen species [21]) and a decreased risk of age-related macular degeneration (AMD) [22], a leading cause of vision loss [23]. Studies have proposed the benefits of anti-oxidant supplementation, including zinc, in slowing the progression of AMD through the prevention of cellular damage in the retina [24].
Importantly, randomised trials in children six months to twelve years of age showed a positive effect of zinc supplementation in reducing all-cause and infectious disease mortality. It also showed a minor positive impact on linear growth [25].
Collectively, these examples highlight the significance of zinc in human health.

Micro-deficiencies and prevalence of zinc deficiency

It is estimated that 372 million (56%) preschool-aged children and 1.2 billion (69%) non-pregnant women of reproductive age across the globe have a deficiency in at least one of the micronutrients, namely zinc, folate, vitamin A and iron. Geographically, 75% of micronutrient-deficient preschool-aged children live in South Asia, sub-Saharan Africa, or East Asia and the Pacific. 57% of micronutrient-deficient non-pregnant women of reproductive age live in East Asia and the Pacific or South Asia [26]. An estimated 17% of the world’s population is at risk of insufficient zinc intake [1, 27]. In Southeast Asia and sub-Saharan Africa, zinc deficiency is endemic, affecting up to 33% of the population. Zinc deficiency is also prevalent in Turkey, Egypt, and Iran due to high phytate intake in their diets, which decreases zinc absorption and, therefore, its bioavailability. Other countries have a markedly lower prevalence of zinc deficiency, notably China, where its incidence decreased from 17 to 8%, as recorded in 2005 [1].

Zinc uptake, absorption, and regulators of its bioavailability

Zinc: location, dietary sources, and intake recommendation

Table 1 provides an overview of the level and proportion of zinc at physiological and cellular levels in the human body.
Table 1
Zinc levels in a healthy human
Compartments
Levels of zinc
Serum (Normally holds < 1% of total body zinc) [28, 29]
70–250 µg/dL [1]
109–130 µg/dL [30]
*62.13–117.72 µg/dL (conversion based on 9.5–18 µM [31])
*78.48 µg/dL–104.64 µg/dL (conversion based on 12–6 µM [29])
60–120 µg/dL, (59–125 μg/dL for male and 50–103 μg/dL for female) in Bangladesh sample population [32]
Tissues
Muscles store about 50 to 60% of the zinc found in the body [3, 29], followed by bones which have about 30 to 36.7% [3, 29], then skin (4.2%) and liver (3.4%) [29]
Prostate, pancreas, and bone, have high zinc concentration ranging from 100 to 250 µg/g [33]
Heart, brain, and plasma, have comparatively lower concentration, ranging from 1 to 23 µg/g [33]
Intracellular distribution
30–40% in nucleus, 50% in cytoplasm, and remaining 10–20% associated with membrane organelles [34, 35]
Mitochondria (0.14 pM), the mitochondrial matrix (0.2 pM), the endoplasmic reticulum (0.9 pM-5 nM), and the Golgi apparatus (0.2 pM) [28]
Total levels in an adult body
2–3 g [28, 36]
*Conversion performed using MediCalc available at https://​www.​scymed.​com/​en/​smnxtb/​tbcbpgh1.​htm
Zinc cannot be stored in substantial amounts, and so, regular dietary intake is essential to ensure sufficient zinc availability [17]. Dietary sources of zinc include fish, oysters, meat, legumes, nuts, beans, whole grains, beef, eggs, and dairy. Oysters are the richest source of zinc, while fruits and vegetables are the poorest source. Although beans, nuts, and whole grains contain zinc, the bioavailability of zinc from these is lower than food from animal sources due to the presence of phytates [37].
There are differences in the dietary recommendations of zinc. Data around this include recommendations of 7.4 mg/day (approximately) [38]. In the US, recommendations include 15 mg/day [30], 11 mg/day [39] and 11 mg/day and 8 mg/day for adult men and women, respectively, who are age 19 and above [37]. In the UK, the recommendation is 9.5 mg/day for an adult man and 7 mg/day for an adult woman. The UK Department of Health recommends that zinc intake should not exceed 25 mg/day [40].

Zinc transporters: ZIPs and ZnTs

During digestion, zinc is released from food as free Zn2+ ions. These ions need to be transported from the intestinal lumen to the enterocyte, then from here into the circulation and from circulation to the cells that require zinc for their activities. Also, within cells, there is intracellular zinc movement and compartmentalisation. These zinc transport processes are facilitated by two important types of zinc transporters: Zinc/iron-regulated-transporter-like-proteins (ZIPs) and Zinc transporters (ZnTs). ZIPs increase intracellular/cytoplasmic zinc levels by transporting zinc from the extracellular space and/or intracellular organelles into the cytoplasm. In general, ZnTs (the exception is ZnT5B, which is a bidirectional transporter) reduce intracellular/cytoplasmic zinc by transporting zinc from the cytoplasm to extracellular space (promoting zinc efflux from cells) or into an organelle for its compartmentalisation [38]. ZnTs and ZIPs are located in several different tissues/cells, on different regions of the cell surface, and on the surface of intracellular organelles (Tables 2 and 3).
Table 2
ZnT transporters: location and regulation
Transporter
Tissue and cellular distribution
Stimulus
Response
Mechanism of response
ZnT1
Ubiquitous [41], with notable abundance in the duodenum, jejunum [42] and kidney [43]
Plasma membrane (basolateral region in epithelial cells and apical region in pancreatic acinar cells) and vesicles [43, 44]
Increased cellular zinc in HepG2 cells [44, 45]
Increase in ZnT1 mRNA [44, 45]
Metal-response element-binding transcription factor-1 (MTF-1) binds to metal-response elements (MREs) in ZnT1 promoter [45]
Zinc deficiency in HepG2 cells [45]
Decreased ZnT1 protein in HepG2 cells [45]
Endocytosis of cell surface ZnT1 with subsequent degradation via proteasomal or lysosomal pathways [45]
Lipopolysaccharide in dendritic cells [44, 46]
Increase in ZnT1 mRNA [44, 46]
Process mediated via Toll/interleukin-1 receptor (TRIF) in Toll-like receptor (TLR) signalling [44, 46]
T-cell stimulation by phytohaemagglutinin (immune activation) [47]
Downregulation of ZnT1 mRNA expression in T-cells [47]
ZnT2
Vesicles, secretory granules [43]
Retina, mammary glands, small intestine, pancreas, kidney, prostate [44]
Two variants:
One variant is primarily located on the membranes of vesicles, including endosomes and lysosomes [44, 48] as well as zymogen granules in pancreatic acinar cells and the inner mitochondrial membrane in mammary cells [44]
The other variant is localised to the plasma membrane [44, 48]
High zinc levels in mammary glands, prostate, pancreas, small intestine, kidney, and retina [44]
Upregulation of ZnT2 mRNA [44]
MTF-1 binding to MRE downstream from ZnT2 transcription start site [44, 49]
Glucocorticoid hormone in pancreatic acinar cells [44]
Upregulation of ZnT2 transcription [44]
Glucocorticoid receptor and STAT5 interaction [44, 49]
Prolactin in mammary epithelial cells [44]
Upregulation of ZnT2 transcription [50]
Prolactin induced JAK2/STAT5 signalling pathway [50]
Decreased ZnT2 expression [44, 51]
Prolactin induced post-translational ZnT2 ubiquitination [44, 51]
ZnT3
Protein detected in brain, retina, and pancreas. RNA detected in testis, duodenum, airways and adipose tissue [52]
On the membranes of synaptic vesicles [44, 53]
Angiotensin II in vascular smooth muscle cells [44, 54]
Downregulation of ZnT3 mRNA expression [54]
Angiotensin II leads to reactive oxidative species which is thought to downregulate ZnT3 [54]
ZnT4
Ubiquitous, with greater abundance in the brain and digestive tract [44]
Trans-golgi network, cytoplasmic vesicles, endosomes, lysosomes, and Golgi apparatus [44]
Increased extracellular zinc [43]
Expression may not be affected but ZnT4 trafficking is induced [43]
Trafficking occurs from trans-golgi network to cytoplasmic vesicular compartment in cultured NRK cells [43]
T-cell stimulation by phytohaemagglutinin (immune activation) [47]
Downregulation of ZnT4 mRNA expression in T-cells [47]
Lipopolysaccharide in dendritic cells [46]
Upregulated expression of ZnT4 mRNA transcripts [46]
This is mediated via Toll/interleukin-1 receptor (TRIF) and myeloid differentiation primary response 88 (MyD88) protein in Toll like receptor (TLR) signalling [46]
Granulocyte–macrophage colony-stimulating factor in macrophages [44]
Upregulation of ZnT4 mRNA expression [44]
Cell differentiation in villus of small intestine [43, 44]
Increased ZnT4 expression [44]
ZnT5
ZnT5 mRNA was found in human endocrine pancreas, prostate and testis [55]. Also found in small intestine [56]
Two variants:
Variant A is located at the Golgi apparatus [57]
Variant B is a bidirectional transporter located throughout the cell, including at the plasma membrane [44, 57] and is on the apical surface of enterocytes [58]
High or low zinc levels [43]
Increased expression [43]
Decreased expression [43]
Increased mRNA stability [43]
Transcriptional repression [43], which is under control of the zinc transcriptional regulatory element (ZTRE) [44]
Lipopolysaccharide in mice liver [59]
Increased ZnT5 mRNA [59]
ZnT6
Protein detected in mouse brain, lung, small intestine, and kidney [60]
Trans-golgi network, Golgi apparatus [44, 60]
T-cell stimulation by phytohaemagglutinin (immune activation) [47]
Downregulation of ZnT6 mRNA expression in T-cells [47]
Lipopolysaccharide in dendritic cells [46]
Upregulation in ZnT6 mRNA expression [46]
Mediated through the Toll/interleukin-1 receptor (TRIF) and myeloid differentiation primary response 88 (MyD88) protein in Toll-like receptor (TLR) signalling [46]
ZnT7
In mice, protein was found in lung and small intestine. The mRNA was found in liver, kidney, spleen, heart, brain, small intestine, and lung, with abundant expression in small intestine and liver and less expression in heart [61]
Early secretory pathway including Golgi apparatus [44]
T-cell stimulation by phytohaemagglutinin (immune activation) [47]
Downregulation of ZnT7 mRNA expression in T-cells [47]
Granulocyte–macrophage colony-stimulating factor in macrophages [44]
Upregulation of ZnT7 mRNA expression [44]
ZnT8
Pancreas [62]; pancreatic β-cell-specific zinc transporter [44] on the membranes of insulin secretion granules [63]
Acute exposure to cytokines (including IL-1β, IFN-γ, IL-17, TNFα) in EndoC-βH1 cells [64]
Downregulation of ZnT8 protein [64]
ZnT10
Liver, brain [62] and intestine [65]
Early/recycling endosomes, Golgi apparatus but can localise to plasma membrane under high extracellular zinc concentrations [44]
IL-6 in human SH-SY5Y neuroblastoma cells [44]
Decrease in both ZnT10 mRNA and protein levels [66]
IL-6 may affect the transcription of the SCL30A10 (gene encoding ZnT10), possibly involving a regulation element [66] which is suggested to be the ZTRE [44]
Angiotensin II in vascular smooth muscle cells [44, 54]
Downregulation of ZnT10 mRNA expression [54]
Angiotensin II leads to reactive oxidative species which is thought to downregulate ZnT10 [54]
High manganese intake in mice [67]
Increased ZnT10 protein levels in liver and small intestine in male mice [67]
High extracellular zinc levels in human 5Y5Y neuroblastoma cells [68]
Downregulation of ZnT10 mRNA [68]
A zinc responsive element (ZRE) may be involved in ZnT10 downregulation [68]
Table 3
ZIP transporters: location and regulation
Transporter
Tissue and cellular distribution
Stimulus
Response
Putative mechanism of response
ZIP1
Ubiquitous, [69]
Plasma membrane [44]
Intracellular vesicles [69]
Zinc deficiency in vitro [70]
Increased mouse ZIP1 protein expression in transfected Human embryonic kidney cells (HEK293) [70]
(ZIP1 expression was unaffected by zinc in vivo [71])
Reduced rates of ZIP1 endocytosis due to zinc limitation [70]. Endocytosis of ZIP1 mediated through a di-leucine sorting signal [72]
Cell differentiation of pluripotent mesenchymal stem cells into osteoblast-like cells [73]
Increased ZIP1 protein expression [73]
ZIP2
Dendritic cells, ovaries, skin, liver [79]
Plasma membrane [79]
Reduced intracellular zinc in monocytes [44, 74]
Upregulation of ZIP2 mRNA in monocytes [44, 74]
Granulocyte macrophage-colony stimulating factor in macrophages [44]
Upregulation of ZIP2 mRNA in macrophages [44]
Keratinocyte differentiation [44]
Upregulation of ZIP2 mRNA in differentiating keratinocytes [44]
Macrophage polarisation to M2 [75]
Increased ZIP2 mRNA levels [75]
ZIP3
Widespread [69]
Plasma membrane but can localise to intracellular compartments after zinc treatment [44]
Zinc deficiency in zebrafish gill [76]
Increased ZIP3 mRNA [76]
Zinc deficiency in vitro [70]
Increased cell surface mouse ZIP3 expression in transfected cells [70]
Reduced rates of ZIP3 endocytosis due to zinc limitation [70]
Prolactin in secretory mammary epithelial cells [77]
Upregulation of ZIP3 mRNA and protein levels [77]
ZIP4
Small intestine and epidermis [79]
Plasma membrane [79]
Cytosolic zinc excess [28, 44]
Downregulation of ZIP4 protein [44]
Endocytosis and degradation ubiquitin-proteasomal and lysosomal degradation pathways [44]
Zinc repletion can lead to endocytosis and degradation of ZIP4 and ZIP4 mRNA destabilisation [71]
Zinc deficiency [28, 44]
Upregulation of ZIP4 [28, 44]
Non-transcriptional: ZIP4 mRNA stabilisation [44]
Transcriptional: Transcriptional upregulation mediated by Krüppel-like factor 4 (KLF4) [43, 78]
Post-translational modification: Proteolytic cleavage of extracellular amino-terminal ectodomain [28, 43, 44]
ZIP5
Intestine, kidney, liver and pancreas [69, 79]
Plasma membrane [43, 79]
Zinc availability in mice [44, 80]
Upregulation of ZIP5 translation [44, 80]
Facilitated by a conserved stem-loop and two overlapping miRNA seed sites in the 3’-untranslated region [44, 80]
Dietary zinc deficiency in mice [71]
Downregulation of ZIP5 translation [71]
ZIP5 mRNA is associated with polysomes and ZIP5 protein is endocytosed and degraded in enterocytes, acinar cells, and endoderm cells [71]
ZIP6
Widespread [69, 79]
Plasma membrane [44]
Lipopolysaccharide in dendritic cells [46]
Downregulation of ZIP6 mRNA expression [46]
Mediated through Toll/interleukin-1 receptor (TRIF) in Toll like receptor (TLR) signalling [46]
Lipopolysaccharide in mice liver [59]
Increased ZIP6 mRNA [59]
Macrophage polarisation to M2 [75]
Increased ZIP6 mRNA [75]
ZIP7
Widespread [69, 79]. Colon, skin [79]
Endoplasmic reticulum and golgi apparatus [44]
Supplemental zinc [43]
Protein abundance of ZIP7 repressed by supplemental zinc [43]
Cellular zinc levels [81]
ZIP7 expression inversely correlate with cellular zinc levels in CLN6 neurons [81]
Macrophage polarisation to M2 [75]
Increased ZIP7 mRNA levels [75]
ZIP8
Widespread [69, 79, 82], T-cells [69], highest levels in the lung [82]
Plasma membrane (apical in polarised cells) and lysosome [44]
T-cell activation in vitro [83]
Upregulation of ZIP8 expression in human T-cells [83]
Lipopolysaccharide in primary human lung epithelia, monocytes and macrophages [84]
Upregulation of ZIP8 at transcriptional level [84]
NF-κB-dependent mechanism [84]
TNF-alpha in primary human lung epithelia, monocytes and macrophages [84]
Upregulation of ZIP8 at transcriptional level [84]
NF-κB-dependent mechanism [84]
Iron loading in rat H4IIE hepatoma cells [85]
Increase in total and cell surface ZIP8 levels [85]
ZIP9
Widely distributed [79]
Plasma membrane, golgi apparatus [44]
Macrophage polarisation to M2 [75]
Increased ZIP9 mRNA levels [75]
ZIP10
Brain, liver, erythroid cell, kidney [69], renal cell, carcinoma B cell [79]
Plasma membrane [43]
Zinc deficiency in zebrafish gill [76]
Zinc excess in vitro and in vivo [76]
Upregulation of ZIP10 mRNA [76]
Downregulation of ZIP10 mRNA [76]
MTF-1 was suggested to be a negative regulator of ZIP10 expression [76]
Zinc deficiency in mice brain and liver [86]
Upregulation of ZIP10 transcription [86]
During zinc sufficient conditions, zinc-activated MTF-1 physically blocks Pol II movement through the gene, leading to ZIP10 transcription downregulation [86]
Lipopolysaccharide in dendritic cells [46]
Downregulation of ZIP10 mRNA transcript expression [46]
Mediated through Toll/interleukin-1 receptor (TRIF) in Toll-like receptor (TLR) signalling [46]
Cytokines in early B cell developmental stages [87]
Upregulated ZIP10 transcription [87]
JAK/STAT pathway involving two STAT binding sites in the promoter [87]
Thyroid hormone in intestine and kidney cells in a rat model of hypo- and hyperthyroidism [88]
Increased ZIP10 mRNA and protein levels in hyperthyroid rats and decreased ZIP10 mRNA in hypothyroid rats, when compared to euthyroid rats [88]
ZIP11
Suggested to localise to stomach and colon [82]
Nucleus, intracellular vesicles and plasma membrane of stomach and colon, golgi in mammary epithelial cells [44, 82]
Possibly zinc-dependent [89]
ZIP11 expression only modestly decreased in mouse stomach but not large or small intestine in response to dietary zinc deficiency. Upon acute zinc repletion, expression levels were not restored [89]
The presence of many MREs upstream of the first exon of the ZIP11 gene would suggest that ZIP11 expression is upregulated in response to increasing zinc levels; however, this was not seen in practice [44]
ZIP12
Brain [69, 79, 82], testis and retina [69], pulmonary vascular smooth muscle [79]
Plasma membrane [44]
Hypoxia in pulmonary vascular smooth muscle cells [90]
Upregulation of ZIP12 mRNA expression [90]
The Slc39a12 gene contains a hypoxia response element (HRE) encoding HIF-1α- and HIF-2α-binding motifs and is located 1 kb downstream of the ZIP12 transcription start site [90]
ZIP13
Widespread [69], hard and connective tissues [79], golgi apparatus, and cytoplasmic vesicles [44]
High iron levels in Drosophila [91]
Upregulation of Drosophila ZIP13 levels [91]
Iron stabilises Drosophila ZIP13 protein, protecting it from degradation [91]
ZIP14
Widespread, liver, bone, and cartilage [79]
Plasma membrane [44], endosome [79]
Zinc deficiency in mouse liver [92]
Upregulation of ZIP14 expression [92]
Mediated through the UPR [92]
IL-6 in mouse hepatocytes [59]
Increased ZIP14 mRNA and protein [59]
Inflammation induced by turpentine [59]
Increased ZIP14 mRNA [59]
Requires IL-6 [59, 93]
Lipopolysaccharide in mice liver [59]
Increased ZIP14 mRNA [59]
Partially requires IL-6 [59, 93]
Nitric oxide (induced by IL-1β) in mice liver [93]
Increased ZIP14 transcription [93]
Nitric oxide increases binding of Activator Protein-1 (AP-1) to the ZIP14 promoter [93]
High manganese intake in mice [67]
Upregulated liver ZIP14 expression in both male and female mice, but upregulated small intestine ZIP14 expression only in male mice [67]
High extracellular glucose (medium) involving INS-1E cells [94]
Upregulation of ZIP14 mRNA expression [94]
Iron loading in rat liver and pancreas, and in hypotransferrinemic mice liver [95]
Upregulated ZIP14 protein expression [95]

Process of zinc uptake, absorption, and circulation

Zinc absorption/uptake primarily occurs in the proximal part of the small intestine, in the distal duodenum and proximal jejunum [39]. Zn2+ ion entry into the enterocyte is mediated via ZIP4 and/or ZnT5B in the duodenum and jejunum (Fig. 1). Another transmembrane ion transporter potentially involved in zinc uptake into the enterocyte is Divalent Metal Transporter-1 (DMT-1) [29]. Regardless of the transporter used, once Zn+2 ions are in the enterocyte, ZnT1 transports zinc from the enterocyte into the portal blood [45]. From here, these ions travel through the hepatic portal vein, mostly likely bound to ligands such as amino acids and citrate [96]. A previous study in rats found that most zinc travels to the liver in portal blood bound to transferrin [97]. At the liver, the portal vein branches, and the blood drains through sinusoids [98]. Some zinc enters hepatocytes, most likely via ZIP14 [99]. After the blood passes through the sinusoids, it is directed to the central vein, then hepatic veins and eventually the systemic circulation [98]. From the systemic circulation, zinc ions are transported to various body tissues such as the in brain, muscle, and bone [100] (Fig. 1).
There is a wide consensus that, in the systemic circulation, the majority of zinc is bound to albumin, a lesser amount bound to α2-macroglobulin, and a fraction bound to amino acids. There is a debate over whether transferrin has a role to play as a zinc carrier in the systemic circulation. Some studies suggest it may play a role [28, 29], whilst others state that transferrin does not have a primary role in the distribution of zinc [104].

Variability in data regarding the proportion of zinc bound to albumin and α2-macroglobulin in systemic circulation

There have been various propositions regarding the proportions of these zinc carriers, particularly albumin. While some stated 80–85% of zinc is bound to albumin [105], others stated that this figure is 60% (with the remaining 30% bound to α2-macroglobulin and 10% to transferrin) [29], or 70% [38], or 80% (with the remaining 20% bound to α2-macroglobulin) [28], and yet others believe that approximately 98% zinc in the systemic circulation is bound to albumin [106].
Regardless of the exact percentage, conditions affecting albumin concentration, such as liver cirrhosis, may affect serum zinc levels [38]. For example, patients with liver cirrhosis and hepatic encephalopathy show decreased serum zinc levels [107].
Binding of Zn+2 ions to the different ligands could be the reason for the concentration of “free” Zn+2 ions in the circulation to be low (~ 0.1–1.0 nM) [96].

Regulators of zinc bioavailability

Bioavailability is the fraction of intake that can be absorbed into the blood and can be used for physiological processes in the body. Studies suggest that the typical range of zinc absorption from the intestinal lumen into the circulation is 16–50% [29], with an average of around 33% [38].
An important factor affecting zinc bioavailability is the solubility of zinc in the intestinal lumen. Previously, it was proposed that amino acids released from protein digestion enhanced zinc absorption by increasing its solubility [108]. Recent reviews have suggested that protein levels in the diet positively correlate with zinc uptake, and the presence of animal-based protein enhances zinc absorption more than the presence of plant-based protein. However, it is still uncertain whether amino acids can enhance zinc bioavailability due to a lack of consensus among studies. Citrate is a low molecular weight ligand found in milk which is thought to have a positive effect on zinc bioavailability by forming zinc-citrate complexes, thereby enhancing zinc uptake. These complexes are found in higher concentrations in human milk than in cow’s milk. Therefore, zinc absorption from human milk is higher than cow’s milk [29]. Also, food fermentation and germination enhance zinc absorption by reducing the phytate content of food [108]. Essentially, fermentation and germination both promote endogenous phytase activity [109, 110]. Germination also facilitates de novo synthesis of phytases, enzymes which hydrolyse phytic acid [110].
On the other hand, several inhibitors decrease zinc solubility and, thereby, its bioavailability, for example, phytic acid/phytates found in food such as legumes, beans, and nuts. Phytates bind to zinc in the intestine, form insoluble complexes and thereby limit zinc absorption. Thus, zinc bioavailability from plant-based foods is lower than foods from animal sources [37]. Calcium and casein in cow’s milk may also reduce the bioavailability of zinc [25].
In addition, some medications may reduce serum zinc levels. For example, long-term use of distal-tube diuretics such as chlortalidone may result in significant zinc depletion due to increased urinary zinc excretion (hyperzincuria) [111]. Angiotensin-converting enzyme inhibitors (ACEis) or angiotensin-receptor blockers (ARBs), commonly used to treat heart failure, may cause zinc deficiency [112]. Also, when consuming or prescribing oral zinc supplements for zinc deficiency, interactions of medications must be carefully monitored. For example, orally consumed zinc may decrease the absorption of many orally taken drugs, such as alendronate and risedronate, which are used to prevent and treat osteoporosis. Similarly, zinc can inhibit the absorption of many antibiotics such as ciprofloxacin and doxycycline [113].

Zinc homeostasis at physiological level

At the physiological level, zinc homeostasis is primarily maintained by controlling zinc absorption and excretion. Of the zinc that is excreted from the body, ~ 50% is lost via faeces [38] (including zinc in sloughed epithelial cells) [28], and the rest is lost through urine, sweat, menstrual flow, semen, loss of hair and nails, and shedding of skin [29]. The zinc absorption mechanism adapts more slowly, while zinc excretion mechanisms can alter quickly [29, 38]. During zinc deficiency, absorption of zinc can be increased up to 90% [28] and faecal and urinal excretion of zinc is rapidly reduced [29]. Certain tissues, such as the bone marrow, liver, and testes, secrete zinc into the circulation as a response to zinc deficiency. Other organs such as skin, skeletal muscle, heart, and kidney conserve their zinc levels even in zinc-deficient states [114].

Endogenous zinc secretion

The process of endogenous zinc secretion into the intestinal lumen may play a role in maintaining zinc homeostasis [115]. There are many ways of mediating zinc secretion into the intestinal lumen, for example, via biliary, pancreatic and gastroduodenal secretions and sloughing of mucosal cells [56, 114, 116]. Zinc transport from the portal circulation into the enterocyte is mediated by ZIP5 and ZIP14 on the basolateral membrane of enterocytes. ZnT5B transporter on the apical membrane of enterocytes is a bidirectional transporter that can transport enterocyte zinc ions into the intestinal lumen and vice versa, thereby mediating both, enterocyte uptake and endogenous secretion of zinc [29, 117].
In the context of zinc secretion into intestinal lumen from the exocrine secretions of the pancreas, there are several zinc transporters that participate in this process. For example, zinc ions are transported into the pancreas from the plasma via ZIP5. These ions are then transported into zymogen granules via ZnT2 and excreted into the digestive tract as pancreatic secretions. Interestingly, zinc concentration in pancreatic tissues and secretions is influenced by dietary zinc intake. Excess dietary zinc leads to upregulation of ZnT2 in acinar cells, and restriction of dietary zinc leads to reduced zinc concentration in both pancreatic tissue and secretions. Thus, zinc homeostasis is regulated by adjusting zinc excretion through the entero-pancreatic axis.
Regardless of the pathway, some zinc in the lumen is reabsorbed into the circulation through uptake via enterocytes. Thus, the balance between the absorption of dietary zinc, and the excretion and reabsorption of endogenous zinc collectively maintain zinc levels in the body [29].

Zinc homeostasis at the cellular level

Zinc concentration at the physiological level is determined by zinc regulation at the cellular level, which is determined by zinc transporters (Tables 2 and 3). Zinc transporters are regulated through various mechanisms, including activation of transcription, stabilisation of mRNA, modification of protein, trafficking to specific organelles, and transporter degradation. Regulatory stimuli include zinc, cytokines, hormones, endoplasmic reticulum stress, oxidative stress, and hypoxia [28].

Effect of high and low zinc on ZnTs: how zinc regulates ZnT expression

ZnT transporters are differentially regulated by zinc levels but with some similarities in mechanisms. Table 2 presents a detailed view of ZnTs, their cellular and tissue distribution, and the stimuli and mechanisms involved in the upregulation or downregulation of the ZnTs. Essentially, high zinc levels increased ZnT1 and ZnT2 mRNA expressions [44] but decreased ZnT10 mRNA levels [68]. Interestingly, ZnT5 expression is unique and complex because high or low zinc levels have been shown to increase or decrease its expression [43]. This could be due to the B variant of ZnT5, which has a bidirectional functionality in zinc transport [29].

Effect of high and low zinc on ZIPs: how zinc regulates ZIP expression

Table 3 details the ZIPs, and their cellular and tissue distribution, along with the stimuli and mechanisms involved in their upregulation or downregulation.
Most ZIP transporters were confirmed to increase their expression in response to low zinc levels including ZIP2 [74], ZIP3 [70, 76], ZIP4 [44], ZIP10 [86], and ZIP14 [92], while only ZIP5 expression was found to decrease [71]. There is some uncertainty regarding ZIP transporter regulation in response to high zinc levels. For example, ZIP7 levels inversely correlated with cellular zinc levels in CLN6 neurons [81] implying that higher cellular zinc would lead to lower ZIP7 levels. However, a causal link is yet to be confirmed because this finding could be confounded by the presence of CLN6 disease. In another example, dietary zinc restriction led to decreased ZIP11 expression in the mouse stomach. However, upon dietary zinc repletion, ZIP11 expression levels were not restored. This suggests that ZIP11 may be unaffected by zinc excess and possibly downregulated by zinc deficiency [89]. This unresponsiveness to zinc (in the form of dietary zinc repletion) [89] is unexpected, given the presence of multiple metal-response elements (MREs) upstream of the first exon of the ZIP11 gene [44].

Other regulators of ZnTs and ZIPs

Zinc transporters respond to various stimuli other than zinc (Tables 2 and 3). For example, in the immune system, T-cell stimulation by phytohaemagglutinin decreases the mRNA expressions of ZnT1, ZnT4, ZnT6 and ZnT7 [47]. These observations reiterate the importance of zinc in modulating the immune response. Moreover, in dendritic cells, lipopolysaccharide stimulation during toll-like receptor signalling increased the mRNA transcripts of ZnT1, ZnT4 and ZnT6 but decreased those of ZIP6 and ZIP10 [46].
Cytokines (namely IL-6) are known to increase ZIP14 levels [59] but decrease ZnT10 levels [44]. Hormones such as glucocorticoid and prolactin can increase ZnT2 levels [44], while thyroid hormone can increase ZIP10 levels [88]. In addition, glucose was found to increase ZIP14 levels [94].
Metals other than zinc, such as manganese [67] and iron [91, 95], can also regulate zinc transporters, which reflects their role in assisting general metal homeostasis. Interestingly, high manganese intake upregulated liver ZIP14 expression in male and female mice but upregulated ZIP14 expression in the small intestine of only male mice. ZnT10 expression was upregulated in the same regions but only in male mice [67]. These observations indicate that there might be sex-based differences in the regulation of zinc transporters.

Metallothioneins (MTs): at the interface of physiological and cellular zinc regulation

MTs are a family of proteins, ubiquitously expressed (in most cells and tissues), which have a high affinity for d10 electron configuration metals, including zinc and copper [118]. MT1 and MT2 are the main isoforms expressed in most adult mammalian tissues. MT3 has been identified in the brain, kidney, breast, pancreas, intestine, and bladder. MT4 has been reported in stratified squamous epithelium around the body and plays an important role in cell differentiation [119].
MTs are thought to play a key role in the systemic regulation of trace elements, including that of zinc [35, 118]. To execute this function, MTs within the enterocyte exhibit their regulatory effect at the absorption stage. Here, MTs can bind to zinc ions within the enterocyte cytoplasm and thereby reduce the availability of free intracellular zinc. Also, when zinc is needed by the cell, MTs can unbind zinc ions and make zinc available. So, if enterocyte zinc concentration is high, then MTs can bind to intracellular zinc and reduce free zinc ions [120, 121]. Consequently, this would reduce the amount of zinc exported into the portal blood, which, in turn, would reduce the amount of zinc distributed around the body. MTs are also thought to mediate zinc trafficking within the cell and zinc transfer to zinc transporters. Thus, through their zinc buffering and muffling properties, MTs help in maintaining zinc homeostasis [29].

Acquired zinc deficiency: diagnosis and treatment

Acquired zinc deficiency could be due to insufficient intake (seen in anorexia nervosa), increased loss (seen in chronic diarrhoea or burns patients), increased requirement (seen in pregnant and breastfeeding individuals) or malabsorption (seen in Crohn’s disease [1] and coeliac disease [122]). It shows clinical features like diarrhoea, frequent infections, and skin lesions. However, these patients usually present the symptoms later in life alongside the aforementioned factors [123]. Due to overlap of symptoms, other differentials that should be considered whilst diagnosing zinc deficiency include: depression, hypothyroidism, vitamin (A, B12 and D) deficiencies, and iron deficiency [1].
Mild zinc deficiency can manifest clinically with serum values ranging from 40 to 60 µg/dL [124]. It has been suggested that acquired zinc deficiency can be diagnosed by a simple blood test showing fasting serum zinc < 70 µg /dL. Furthermore, since low albumin levels can cause low zinc levels, serum albumin levels should also be measured [125].
Some suggest that plasma zinc as a biomarker is non-specific, and it is difficult to develop a single biomarker of zinc status due to zinc’s diverse functions [25]. However, taking a fasting sample in the morning, separating plasma or serum from cells within 45 min and using zinc-free vacuum tubes can improve accuracy [1]. In general, urinary zinc levels are not a useful diagnostic parameter for zinc deficiency, whereas hair zinc levels are useful only in the context of chronic deficiency [1, 126]. In addition to laboratory investigations, the clinical aspects comprising patient risk factors, geographical prevalence, and age of presentation, alongside physical examination and an appropriate history-taking, can help to establish the diagnosis [1].
Oral zinc supplementation, such as zinc gluconate for either short-term or long-term depending on the underlying aetiology, is usually used to cure the acquired deficiency [1]. Interestingly, zinc supplements can be formulated as zinc oxide or as salts with acetate, gluconate and sulphate [2]. A clinical trial reported that zinc oxide administered without food is less well absorbed than other zinc formulations as it is more insoluble [127]. Zinc citrate has a relatively higher zinc content, yet this is countered by the finding that zinc absorption in the form of citrate does not differ from that of zinc gluconate. However, the affordability of zinc citrate may make this an attractive alternative to zinc gluconate [127]. A potential complication of zinc deficiency treatment is overcorrection with zinc supplementation since this can cause acute zinc toxicity [1]
Other clinical conditions that may show low zinc levels are tabulated in Table 4.
Table 4
Examples of clinical conditions that show low zinc levels
Condition/disease
Possible reason for low zinc level and the clinical status
Infection with HIV
Reduced absorption of zinc from foods. These patients often have diarrhoea, which causes excess zinc loss, resulting in low serum zinc [128]
Chronic kidney disease
Serum zinc levels tend to be on the lower side due to inadequate dietary intake, malabsorption and zinc removal during haemodialysis [129]
Liver diseases
Alcoholic hepatitis patients showed lower zinc levels compared to non-alcoholic liver disease patients [130]. Patients with alcoholic liver disease often have poor diets low in zinc whilst in cirrhosis, absorption may be impaired and there usually is increased urinary zinc excretion [131]
Polycystic ovarian syndrome that increase oestrogen levels [132]
High levels of oestrogen can decrease plasma zinc levels and increase zinc in the liver [130]
Sickle cell disease or beta thalassaemia
These patients require frequent blood transfusions, which lead to iron loading. The latter is tackled via iron chelation, but this could lead to zinc deficiency, a common complication of sickle cell treatment [133]

Inherited zinc deficiency acrodermatitis enteropathica: diagnosis and treatment

Many inherited defects of zinc deficiencies are known. Most cases are associated with mutations in the SLC39A4 gene on chromosome 8. This gene encodes the zinc transporter ZIP4 [134136]. The pathological condition is referred to as acrodermatitis enteropathica, a rare autosomal recessive condition with the incidence of roughly 1 in 500,000 births [1, 137]. It affects males and females equally [137]. Because ZIP4 mediates the transport of zinc ions from the intestinal lumen into the enterocyte, a mutation in the gene encoding ZIP4 does not allow zinc ions to be transported into the enterocyte through this transporter. Consequently, insufficient zinc ions reach the systemic circulation or distant tissues (Fig. 2) [29]. Although a small amount of zinc may be taken up via the passive paracellular route [138], the result is zinc deficiency.
Notably, while ZIP4 has two zinc-binding sites and thereby can show increased efficiency in capturing and delivering zinc to the enterocytes, how ZnT5B (another zinc importer on the enterocyte) transports zinc ions into the enterocyte is not known [139]. It is conceivable that ZnT5B may have a lower affinity to zinc ions than ZIP4, and therefore, although it can allow the entry of zinc ions into the enterocyte, it cannot compensate for ZIP4 dysfunction. Left untreated, acrodermatitis enteropathica is fatal within the first few years of life [125]
Acrodermatitis enteropathica patients usually manifest symptoms early in life [1] in the phase of weaning from breastfeeding [140]. Symptoms include a triad of alopecia, diarrhoea, and dermatitis [141]. Patients may also show growth impairment, psoriasiform lesions (well-defined scaly plaques most often found on the elbows) and frequent infections [1]. Alongside the consideration of clinical symptoms, serum zinc level < 70 µg/dL in fasting and low serum alkaline phosphatase may be suggestive of acrodermatitis enteropathica [125]. Note that alkaline phosphatase is a zinc-dependent enzyme [142]. Molecular genetic testing can identify SLC39A4 mutation and confirm acrodermatitis enteropathica.
Treatment involves zinc supplementation, but the formulation of zinc depends on the route of administration. For example, zinc gluconate and sulphate [1] are commonly used orally, while zinc chloride is preferred parenterally [125]. Treatment is lifelong with patient compliance being crucial [40].
Another genetic cause of acrodermatitis enteropathica is due to a mutation in the SLC30A2 gene of the breastfeeding mother. This gene encodes for ZnT2, a zinc transporter expressed in the mammary glands [143]. In secreting mammary epithelial cells, ZnT2 imports zinc into vesicles, mediating zinc secretion into the breast milk. A mutation in this gene results in decreased zinc secretion into the breast milk. This can lead to severe zinc deficiency in exclusively breastfed infants [44]. This can be treated by supplementation of zinc at 5 mg per day whilst breastfeeding. After weaning, no further action is needed [143]. Most paediatric patients with acrodermatitis enteropathica do not present with the classic triad of periorificial and acral dermatitis, diarrhoea, and alopecia. Less than one-third of paediatric patients present in this way. Common presentations in children include recurrent infections, irritability, behavioural changes, neurological disturbances, and failure to thrive [140].
ZIP8 mutations result in cortical atrophy and, consequently, intellectual disability in the affected patient. A mutation in the SLC39A14 gene (encoding ZIP14) can lead to parkinsonism-dystonia in children [144], whilst a mutated ZIP13 protein is responsible for the spondylocheirodysplastic form of Ehlers-Danlos syndrome [145]. In mice, ZIP7 knockout was lethal, whilst a morpholino knockdown of ZIP7 caused neurodevelopmental issues in zebrafish [144].

Zinc toxicity: diagnosis and treatment

To our knowledge, there have been no reports on zinc overload/toxicity due to mutations in zinc transporters. The reported cases of zinc toxicity are due to acquired causes rather than inherited ones. Causes include pesticide exposure and exposure to compounds used to make paints, rubber and dyes [40].
The tolerable upper intake level of zinc, according to the US Institute of Medicine, is as follows: 4 mg in youngest infants, 12 mg in children 4–8 years old, 34 mg in adolescents (14–18 years), and 40 mg for persons aged 19 or older [146]. Acute and chronic zinc toxicities are defined as zinc intake of more than 200 mg/day and 50–150 mg/day, respectively [40]. Acute zinc toxicity is likely due to excessive zinc supplementation as opposed to excessive dietary zinc intake. Longer-term causes of zinc toxicity include occupational exposure to zinc [147] and iatrogenic causes such as overprescribing of zinc-containing medication, zinc present in dental fixtures (though modern preparations in the UK and US are now zinc-free) and overconsumption of over-the-counter zinc supplements [148].
At zinc doses higher than 50 mg/day, symptoms such as nausea, diarrhoea and abdominal discomfort may occur, whilst doses higher than 150 mg/day can adversely affect the body’s lipid profile and immune system. On the other hand, it has been suggested that symptoms of zinc toxicity may not manifest until intake exceeds 1–2 g [30]. The most common cause of zinc excess is taking too many zinc supplements [149]. Chronic zinc toxicity can lead to disturbances in copper metabolism causing low copper status, which affects iron distribution and causes anaemia, red blood cell microcytosis, neutropenia and reduced immune function [150, 151].
Zinc toxicity presents in different ways depending on the mode of zinc overload. For example, acute dietary ingestion presents as nausea, vomiting, diarrhoea, and muscle cramps. If toxicity is caused by inhalation of fumes, it presents with flu-like symptoms such as cough, fever, and chills. Chronic ingestion slowly leads to a syndrome of neuropathy, anaemia, fatigue, and spasticity. The 2017 Annual Report of the American Association of Poison Control Centres’ (AAPCC) National Poison Data System (NPDS) reported 1236 cases of exposure to zinc compounds, most of which were unintentional exposures in children less than five years of age. There were no deaths or major adverse health events as a consequence of this though [30].
An investigation to diagnose zinc poisoning includes several aspects like a thorough history to gain an understanding of the mode of overload, levels of serum zinc, copper and ceruloplasmin, liver function tests, platelet count, and chest X-ray. Treatment for acute ingestion involves anti-emetics, fluids and proton pump inhibitors [30]. Treatment for metal fume inhalation focuses on oral rehydration, anti-pyretics and supplemental oxygen with bronchodilators [152]. In chronic zinc toxicity, first the identification and then the removal of the source of zinc is essential. This can be followed by treatment with copper sulphate [30]. This treatment works because copper competes with zinc for absorption, so exogenous copper intake reduces zinc absorption [1]. Very severe cases may require zinc chelation with agents such as diethylenetriamine pentaacetate (DTPA) [30].

Zinc-induced copper deficiency (ZICD)

An important complication of chronic zinc toxicity is zinc-induced copper deficiency (ZICD). Excess zinc levels in the small intestine stimulate increased expression of MTs in the enterocytes. Since copper has a greater affinity for MTs compared to zinc, copper outcompetes zinc for MT binding sites, and consequently, the copper bound to MT is excreted via sloughing of enterocytes. This results in decreased absorption of copper and, therefore, copper deficiency [153]. The co-existence of hyperzincaemia (high zinc in serum) and hypocupraemia (low copper in serum) is suggestive of ZICD [154]. Measurements of urinary zinc can be useful in the diagnostics of ZICD because urinary zinc levels are usually high in this condition [155].
ZICD tends to develop slowly over many months or years, although this apparent slow onset may be due to delayed diagnosis [155]. Early manifestation includes neutropenia, leukopenia and anaemia [154]. This anaemia is likely due to the disruption of copper’s physiological role in the differentiation of haematopoietic stem cells as well as intestinal iron absorption [156].
Under physiological conditions, copper plays a vital role in the synthesis and stabilisation of myelin, and in several enzymatic pathways required for the functioning of the nervous system. Therefore, if ZICD is left undiagnosed, it can lead to severe and permanent neurological complications including gait disturbances, paraesthesia and myelopathy [156].
One study highlighted that the over-prescription of zinc was a significant cause of ZICD. This shows that zinc can have potentially serious side effects, and it is not a harmless agent that can be prescribed without a strong justification [155]. Removal of the source of excess zinc along with oral copper gluconate treatment is often sufficient to revert anaemia, neutropenia and leukopenia seen in ZICD. Neurological deficits may also improve with this treatment; however, many never fully recover and will be left with permanent neurological deficits [156].

Summary

Trace elements play an important role in human health and disease. For example, the role of iron in various diseases, including liver fibrosis, alcohol-related liver disease, and COVID-19, has been reviewed [157160].
Zinc is an essential micronutrient which cannot be stored in significant amounts, so regular dietary intake is vital. There does not seem to be a clear consensus on the recommended zinc intake, where the recommendation ranges from 7.4 to 15 mg/day. Citrate and food processing such as fermentation and germination, can enhance zinc uptake. It is unclear whether amino acids enhance zinc uptake. Phytic acid, found in cereals, legumes, and nuts, is known to decrease zinc bioavailability.
Zinc absorption occurs primarily in the proximal small intestine, where ZIP4 mediates zinc entry into enterocytes. ZnT5B and DMT-1 are also thought to play a role in this process. ZnT1 transports zinc from the enterocyte into the portal vein, via which zinc travels to the liver. Here, some zinc may enter the hepatocytes via ZIP14, and the rest may eventually drain into the systemic circulation for distribution to various bodily tissues. In the circulation, most zinc is bound to albumin, and the majority in the body can be found in skeletal muscle and bone.
Zinc deficiency can be inherited or acquired. The acquired form is due to insufficient intake, malabsorption, increased requirement, or excessive loss of zinc. Inherited zinc deficiency is mostly associated with mutations in the SLC39A4 gene (which encodes for ZIP4), resulting in a triad of alopecia, diarrhoea and dermatitis in a condition known as acrodermatitis enteropathica. In both inherited and acquired deficiencies, oral zinc supplementation is the mainstay of treatment with an excellent prognosis.
Zinc toxicity is only known to be acquired, not inherited, and may be acute or chronic. Symptoms include nausea, diarrhoea and abdominal discomfort but may vary depending on the mode of overload. Treatment involves the chelation of excess zinc using drugs.

Recommendations for future work

This review helped us identify the knowledge gaps in the literature on zinc. For example, there is no clear consensus on the proportion of zinc that is bound to albumin in the systemic circulation. Albumin levels decrease in several conditions, including liver cirrhosis, and so, knowledge of the proportion of zinc bound to albumin may inform alternative or supplementary treatments for those with albumin-depletion-induced zinc deficiency. In such patients, only oral zinc supplementation may not be enough to resolve the deficiency because there is an insufficient level of the zinc carrier albumin to distribute zinc around the body. Should albumin be the predominant zinc carrier in the circulation, albumin supplementation may play an important role in treating zinc deficiency. There are some other uncertainties in the context of zinc carriers. For example, there are contrasting suggestions on whether transferrin has a role to play as a zinc carrier in the systemic circulation or not. Thus, the knowledge on zinc carriers will aid in our understanding of the pathogenesis of zinc-related conditions.
Another knowledge gap exists in the mechanisms by which the levels of certain zinc transporters increase or decrease in cells in response to a stimulus, as reflected in Tables 2 and 3. Dedicated studies are required to elucidate these mechanisms as this might help devise ways of altering zinc levels within cells in a tissue-specific manner and thereby help ameliorate a zinc-related diseased state. Also, while our knowledge so far indicates that zinc toxicity is not caused by mutations in zinc transporters, examining genetics as a predisposing factor in the development of zinc toxicity might be helpful, as it has been in the case of zinc deficiency. Given the multi-faceted physiological role of zinc, such studies would improve the diagnostics and therapeutics of a range of conditions, positively impacting the health of the general population as well as of those with zinc-related diseases.

Acknowledgements

This article is made open access with the financial support of King’s College London, UK. Thanks to Najma Ali (GKT School of Medical Education, Faculty of Life Sciences and Medicine, King’s College London, UK) for providing some data that acted as a reference for writing a couple of paragraphs.

Declarations

Conflict of interest

Authors declare that they have no competing or financial interests to disclose.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Unsere Produktempfehlungen

e.Med Interdisziplinär

Kombi-Abonnement

Jetzt e.Med zum Sonderpreis bestellen!

Für Ihren Erfolg in Klinik und Praxis - Die beste Hilfe in Ihrem Arbeitsalltag

Mit e.Med Interdisziplinär erhalten Sie Zugang zu allen CME-Fortbildungen und Fachzeitschriften auf SpringerMedizin.de.

Jetzt bestellen und 100 € sparen!

e.Med Innere Medizin

Kombi-Abonnement

Mit e.Med Innere Medizin erhalten Sie Zugang zu CME-Fortbildungen des Fachgebietes Innere Medizin, den Premium-Inhalten der internistischen Fachzeitschriften, inklusive einer gedruckten internistischen Zeitschrift Ihrer Wahl.

Jetzt bestellen und 100 € sparen!

e.Med Allgemeinmedizin

Kombi-Abonnement

Mit e.Med Allgemeinmedizin erhalten Sie Zugang zu allen CME-Fortbildungen und Premium-Inhalten der allgemeinmedizinischen Zeitschriften, inklusive einer gedruckten Allgemeinmedizin-Zeitschrift Ihrer Wahl.

Jetzt bestellen und 100 € sparen!

Literatur
3.
Zurück zum Zitat Hernández-Camacho JD, Vicente-García C, Parsons DS, Navas-Enamorado I. Zinc at the crossroads of exercise and proteostasis. Redox Biol. 2020;35: 101529.PubMedPubMedCentralCrossRef Hernández-Camacho JD, Vicente-García C, Parsons DS, Navas-Enamorado I. Zinc at the crossroads of exercise and proteostasis. Redox Biol. 2020;35: 101529.PubMedPubMedCentralCrossRef
5.
Zurück zum Zitat Beyersmann D, Haase H. Functions of zinc in signaling, proliferation and differentiation of mammalian cells. Biometals. 2001;14:331–41.PubMedCrossRef Beyersmann D, Haase H. Functions of zinc in signaling, proliferation and differentiation of mammalian cells. Biometals. 2001;14:331–41.PubMedCrossRef
10.
Zurück zum Zitat Weiss M, Steiner DF, Philipson LH, et al. Insulin biosynthesis, secretion, structure, and structure-activity relationships. In: Feingold KR, Anawalt B, Blackman MR, Boyce A, Chrousos G, Corpas E, et al., editors. Endotext [Internet]. South Dartmouth: MDText.com Inc; 2000. Weiss M, Steiner DF, Philipson LH, et al. Insulin biosynthesis, secretion, structure, and structure-activity relationships. In: Feingold KR, Anawalt B, Blackman MR, Boyce A, Chrousos G, Corpas E, et al., editors. Endotext [Internet]. South Dartmouth: MDText.com Inc; 2000.
12.
Zurück zum Zitat Severo JS, Morais JBS, de Freitas TEC, Andrade ALP, Feitosa MM, Fontenelle LC, et al. The role of zinc in thyroid hormones metabolism. Int J Vitam Nutr Res. 2019;89:80–8.PubMedCrossRef Severo JS, Morais JBS, de Freitas TEC, Andrade ALP, Feitosa MM, Fontenelle LC, et al. The role of zinc in thyroid hormones metabolism. Int J Vitam Nutr Res. 2019;89:80–8.PubMedCrossRef
13.
14.
Zurück zum Zitat Betsy A, Binitha M, Sarita S. Zinc deficiency associated with hypothyroidism: an overlooked cause of severe alopecia. Int J Trichol. 2013;5:40–2.CrossRef Betsy A, Binitha M, Sarita S. Zinc deficiency associated with hypothyroidism: an overlooked cause of severe alopecia. Int J Trichol. 2013;5:40–2.CrossRef
15.
Zurück zum Zitat Maxwell C, Volpe SL. Effect of zinc supplementation on thyroid hormone function. ANM. 2007;51:188–94. Maxwell C, Volpe SL. Effect of zinc supplementation on thyroid hormone function. ANM. 2007;51:188–94.
17.
Zurück zum Zitat Fallah A, Mohammad-Hasani A, Colagar AH. Zinc is an essential element for male fertility: A review of Zn roles in men’s health, germination, sperm quality, and fertilization. J Reprod Infertil. 2018;19:69–81.PubMedPubMedCentral Fallah A, Mohammad-Hasani A, Colagar AH. Zinc is an essential element for male fertility: A review of Zn roles in men’s health, germination, sperm quality, and fertilization. J Reprod Infertil. 2018;19:69–81.PubMedPubMedCentral
20.
Zurück zum Zitat De Nascimento Marreiro D, Cruz KJC, Morais JBS, Beserra JB, Severo JS, de Oliveira ARS. Zinc and oxidative stress: current mechanisms. Antioxidants (Basel). 2017;6:24.CrossRef De Nascimento Marreiro D, Cruz KJC, Morais JBS, Beserra JB, Severo JS, de Oliveira ARS. Zinc and oxidative stress: current mechanisms. Antioxidants (Basel). 2017;6:24.CrossRef
21.
Zurück zum Zitat Sedeek M, Nasrallah R, Touyz RM, Hébert RL. NADPH oxidases, reactive oxygen species, and the kidney: friend and foe. J Am Soc Nephrol. 2013;24:1512–8.PubMedPubMedCentralCrossRef Sedeek M, Nasrallah R, Touyz RM, Hébert RL. NADPH oxidases, reactive oxygen species, and the kidney: friend and foe. J Am Soc Nephrol. 2013;24:1512–8.PubMedPubMedCentralCrossRef
22.
Zurück zum Zitat van Leeuwen R, Boekhoorn S, Vingerling JR, Witteman JCM, Klaver CCW, Hofman A, et al. Dietary intake of antioxidants and risk of age-related macular degeneration. JAMA. 2005;294:3101–7.PubMedCrossRef van Leeuwen R, Boekhoorn S, Vingerling JR, Witteman JCM, Klaver CCW, Hofman A, et al. Dietary intake of antioxidants and risk of age-related macular degeneration. JAMA. 2005;294:3101–7.PubMedCrossRef
23.
Zurück zum Zitat Fleckenstein M, Keenan TDL, Guymer RH, Chakravarthy U, Schmitz-Valckenberg S, Klaver CC, et al. Age-related macular degeneration. Nat Rev Dis Prim. 2021;7:31.PubMedCrossRef Fleckenstein M, Keenan TDL, Guymer RH, Chakravarthy U, Schmitz-Valckenberg S, Klaver CC, et al. Age-related macular degeneration. Nat Rev Dis Prim. 2021;7:31.PubMedCrossRef
24.
Zurück zum Zitat Evans JR, Lawrenson JG. Antioxidant vitamin and mineral supplements for slowing the progression of age-related macular degeneration. Cochrane Database Syst Rev. 2017;7:CD000254.PubMed Evans JR, Lawrenson JG. Antioxidant vitamin and mineral supplements for slowing the progression of age-related macular degeneration. Cochrane Database Syst Rev. 2017;7:CD000254.PubMed
25.
Zurück zum Zitat Bailey RL, West KP Jr, Black RE. The epidemiology of global micronutrient deficiencies. ANM. 2015;66:22–33. Bailey RL, West KP Jr, Black RE. The epidemiology of global micronutrient deficiencies. ANM. 2015;66:22–33.
26.
Zurück zum Zitat Stevens GA, Beal T, Mbuya MNN, Luo H, Neufeld LM, Addo OY, et al. Micronutrient deficiencies among preschool-aged children and women of reproductive age worldwide: a pooled analysis of individual-level data from population-representative surveys. Lancet Glob Health. 2022;10:e1590–9.PubMedCrossRef Stevens GA, Beal T, Mbuya MNN, Luo H, Neufeld LM, Addo OY, et al. Micronutrient deficiencies among preschool-aged children and women of reproductive age worldwide: a pooled analysis of individual-level data from population-representative surveys. Lancet Glob Health. 2022;10:e1590–9.PubMedCrossRef
27.
Zurück zum Zitat Wessells KR, Brown KH. Estimating the global prevalence of zinc deficiency: results based on zinc availability in national food supplies and the prevalence of stunting. PLoS ONE. 2012;7: e50568.ADSPubMedPubMedCentralCrossRef Wessells KR, Brown KH. Estimating the global prevalence of zinc deficiency: results based on zinc availability in national food supplies and the prevalence of stunting. PLoS ONE. 2012;7: e50568.ADSPubMedPubMedCentralCrossRef
28.
Zurück zum Zitat Hara T, Takeda T, Takagishi T, Fukue K, Kambe T, Fukada T. Physiological roles of zinc transporters: molecular and genetic importance in zinc homeostasis. J Physiol Sci. 2017;67:283–301.PubMedPubMedCentralCrossRef Hara T, Takeda T, Takagishi T, Fukue K, Kambe T, Fukada T. Physiological roles of zinc transporters: molecular and genetic importance in zinc homeostasis. J Physiol Sci. 2017;67:283–301.PubMedPubMedCentralCrossRef
29.
31.
Zurück zum Zitat Andrew D, Gail R, Morag B, Kishor R. Recommended reference intervals for copper and zinc in serum using the US national health and nutrition examination surveys (NHANES) data. Clin Chim Acta. 2023;546: 117397.PubMedCrossRef Andrew D, Gail R, Morag B, Kishor R. Recommended reference intervals for copper and zinc in serum using the US national health and nutrition examination surveys (NHANES) data. Clin Chim Acta. 2023;546: 117397.PubMedCrossRef
32.
Zurück zum Zitat Barman N, Salwa M, Ghosh D, Rahman MW, Uddin MN, Haque MA. Reference value for serum zinc level of adult population in Bangladesh. EJIFCC. 2020;31:117–24.PubMedPubMedCentral Barman N, Salwa M, Ghosh D, Rahman MW, Uddin MN, Haque MA. Reference value for serum zinc level of adult population in Bangladesh. EJIFCC. 2020;31:117–24.PubMedPubMedCentral
35.
Zurück zum Zitat Bafaro E, Liu Y, Xu Y, Dempski RE. The emerging role of zinc transporters in cellular homeostasis and cancer. Sig Transduct Target Ther. 2017;2:1–12.CrossRef Bafaro E, Liu Y, Xu Y, Dempski RE. The emerging role of zinc transporters in cellular homeostasis and cancer. Sig Transduct Target Ther. 2017;2:1–12.CrossRef
38.
Zurück zum Zitat Roohani N, Hurrell R, Kelishadi R, Schulin R. Zinc and its importance for human health: an integrative review. J Res Med Sci. 2013;18:144–57.PubMedPubMedCentral Roohani N, Hurrell R, Kelishadi R, Schulin R. Zinc and its importance for human health: an integrative review. J Res Med Sci. 2013;18:144–57.PubMedPubMedCentral
39.
Zurück zum Zitat Skrovanek S, DiGuilio K, Bailey R, Huntington W, Urbas R, Mayilvaganan B, et al. Zinc and gastrointestinal disease. World J Gastrointest Pathophysiol. 2014;5:496–513.PubMedPubMedCentralCrossRef Skrovanek S, DiGuilio K, Bailey R, Huntington W, Urbas R, Mayilvaganan B, et al. Zinc and gastrointestinal disease. World J Gastrointest Pathophysiol. 2014;5:496–513.PubMedPubMedCentralCrossRef
41.
Zurück zum Zitat Mor M, Beharier O, Cook DI, Campbell CR, Gheber LA, Katz A, et al. ZnT1 induces a crosstalk between T-type and L-type calcium channels through interactions with Raf-1 kinase and the calcium channel β2 subunit. Metallomics. 2023;15:mfad031.PubMedCrossRef Mor M, Beharier O, Cook DI, Campbell CR, Gheber LA, Katz A, et al. ZnT1 induces a crosstalk between T-type and L-type calcium channels through interactions with Raf-1 kinase and the calcium channel β2 subunit. Metallomics. 2023;15:mfad031.PubMedCrossRef
43.
Zurück zum Zitat Lichten LA, Cousins RJ. Mammalian zinc transporters: nutritional and physiologic regulation. Annu Rev Nutr. 2009;29:153–76.PubMedCrossRef Lichten LA, Cousins RJ. Mammalian zinc transporters: nutritional and physiologic regulation. Annu Rev Nutr. 2009;29:153–76.PubMedCrossRef
44.
Zurück zum Zitat Kambe T, Tsuji T, Hashimoto A, Itsumura N. The physiological, biochemical, and molecular roles of zinc transporters in zinc homeostasis and metabolism. Physiol Rev. 2015;95:749–84.PubMedCrossRef Kambe T, Tsuji T, Hashimoto A, Itsumura N. The physiological, biochemical, and molecular roles of zinc transporters in zinc homeostasis and metabolism. Physiol Rev. 2015;95:749–84.PubMedCrossRef
45.
Zurück zum Zitat Nishito Y, Kambe T. Zinc transporter 1 (ZNT1) expression on the cell surface is elaborately controlled by cellular zinc levels. J Biol Chem. 2019;294:15686–97.PubMedPubMedCentralCrossRef Nishito Y, Kambe T. Zinc transporter 1 (ZNT1) expression on the cell surface is elaborately controlled by cellular zinc levels. J Biol Chem. 2019;294:15686–97.PubMedPubMedCentralCrossRef
46.
Zurück zum Zitat Kitamura H, Morikawa H, Kamon H, Iguchi M, Hojyo S, Fukada T, et al. Toll-like receptor–mediated regulation of zinc homeostasis influences dendritic cell function. Nat Immunol. 2006;7:971–7.PubMedCrossRef Kitamura H, Morikawa H, Kamon H, Iguchi M, Hojyo S, Fukada T, et al. Toll-like receptor–mediated regulation of zinc homeostasis influences dendritic cell function. Nat Immunol. 2006;7:971–7.PubMedCrossRef
47.
48.
Zurück zum Zitat Lopez V, Kelleher SL. Zinc transporter-2 (ZnT2) variants are localized to distinct subcellular compartments and functionally transport zinc. Biochem J. 2009;422:43–52.PubMedCrossRef Lopez V, Kelleher SL. Zinc transporter-2 (ZnT2) variants are localized to distinct subcellular compartments and functionally transport zinc. Biochem J. 2009;422:43–52.PubMedCrossRef
49.
Zurück zum Zitat Guo L, Lichten LA, Ryu M-S, Liuzzi JP, Wang F, Cousins RJ. STAT5-glucocorticoid receptor interaction and MTF-1 regulate the expression of ZnT2 (Slc30a2) in pancreatic acinar cells. Proc Natl Acad Sci U S A. 2010;107:2818–23.ADSPubMedPubMedCentralCrossRef Guo L, Lichten LA, Ryu M-S, Liuzzi JP, Wang F, Cousins RJ. STAT5-glucocorticoid receptor interaction and MTF-1 regulate the expression of ZnT2 (Slc30a2) in pancreatic acinar cells. Proc Natl Acad Sci U S A. 2010;107:2818–23.ADSPubMedPubMedCentralCrossRef
50.
Zurück zum Zitat Qian L, Lopez V, Seo YA, Kelleher SL. Prolactin regulates ZNT2 expression through the JAK2/STAT5 signaling pathway in mammary cells. Am J Physiol Cell Physiol. 2009;297:C369-377.PubMedPubMedCentralCrossRef Qian L, Lopez V, Seo YA, Kelleher SL. Prolactin regulates ZNT2 expression through the JAK2/STAT5 signaling pathway in mammary cells. Am J Physiol Cell Physiol. 2009;297:C369-377.PubMedPubMedCentralCrossRef
51.
Zurück zum Zitat Seo YA, Lee S, Hennigar SR, Kelleher SL. Prolactin (PRL)-stimulated ubiquitination of ZnT2 mediates a transient increase in zinc secretion followed by ZnT2 degradation in mammary epithelial cells. J Biol Chem. 2014;289:23653–61.PubMedPubMedCentralCrossRef Seo YA, Lee S, Hennigar SR, Kelleher SL. Prolactin (PRL)-stimulated ubiquitination of ZnT2 mediates a transient increase in zinc secretion followed by ZnT2 degradation in mammary epithelial cells. J Biol Chem. 2014;289:23653–61.PubMedPubMedCentralCrossRef
52.
Zurück zum Zitat Smidt K, Rungby J. ZnT3: a zinc transporter active in several organs. Biometals. 2012;25:1–8.PubMedCrossRef Smidt K, Rungby J. ZnT3: a zinc transporter active in several organs. Biometals. 2012;25:1–8.PubMedCrossRef
53.
Zurück zum Zitat Niu L, Li L, Yang S, Wang W, Ye C, Li H. Disruption of zinc transporter ZnT3 transcriptional activity and synaptic vesicular zinc in the brain of Huntington’s disease transgenic mouse. Cell Biosci. 2020;10:106.PubMedPubMedCentralCrossRef Niu L, Li L, Yang S, Wang W, Ye C, Li H. Disruption of zinc transporter ZnT3 transcriptional activity and synaptic vesicular zinc in the brain of Huntington’s disease transgenic mouse. Cell Biosci. 2020;10:106.PubMedPubMedCentralCrossRef
54.
Zurück zum Zitat Patrushev N, Seidel-Rogol B, Salazar G. Angiotensin II requires zinc and downregulation of the zinc transporters ZnT3 and ZnT10 to induce senescence of vascular smooth muscle cells. PLoS ONE. 2012;7: e33211.ADSPubMedPubMedCentralCrossRef Patrushev N, Seidel-Rogol B, Salazar G. Angiotensin II requires zinc and downregulation of the zinc transporters ZnT3 and ZnT10 to induce senescence of vascular smooth muscle cells. PLoS ONE. 2012;7: e33211.ADSPubMedPubMedCentralCrossRef
55.
57.
Zurück zum Zitat Jackson KA, Helston RM, McKay JA, O’Neill ED, Mathers JC, Ford D. Splice variants of the human zinc transporter ZnT5 (SLC30A5) are differentially localized and regulated by zinc through transcription and mRNA stability. J Biol Chem. 2007;282:10423–31.PubMedCrossRef Jackson KA, Helston RM, McKay JA, O’Neill ED, Mathers JC, Ford D. Splice variants of the human zinc transporter ZnT5 (SLC30A5) are differentially localized and regulated by zinc through transcription and mRNA stability. J Biol Chem. 2007;282:10423–31.PubMedCrossRef
58.
Zurück zum Zitat Valentine RA, Jackson KA, Christie GR, Mathers JC, Taylor PM, Ford D. ZnT5 variant B is a bidirectional zinc transporter and mediates zinc uptake in human intestinal Caco-2 cells. J Biol Chem. 2007;282:14389–93.PubMedCrossRef Valentine RA, Jackson KA, Christie GR, Mathers JC, Taylor PM, Ford D. ZnT5 variant B is a bidirectional zinc transporter and mediates zinc uptake in human intestinal Caco-2 cells. J Biol Chem. 2007;282:14389–93.PubMedCrossRef
59.
Zurück zum Zitat Liuzzi JP, Lichten LA, Rivera S, Blanchard RK, Aydemir TB, Knutson MD, et al. Interleukin-6 regulates the zinc transporter Zip14 in liver and contributes to the hypozincemia of the acute-phase response. Proc Natl Acad Sci USA. 2005;102:6843–8.ADSPubMedPubMedCentralCrossRef Liuzzi JP, Lichten LA, Rivera S, Blanchard RK, Aydemir TB, Knutson MD, et al. Interleukin-6 regulates the zinc transporter Zip14 in liver and contributes to the hypozincemia of the acute-phase response. Proc Natl Acad Sci USA. 2005;102:6843–8.ADSPubMedPubMedCentralCrossRef
60.
Zurück zum Zitat Huang L, Kirschke CP, Gitschier J. Functional characterization of a novel mammalian zinc transporter, ZnT6. J Biol Chem. 2002;277:26389–95.PubMedCrossRef Huang L, Kirschke CP, Gitschier J. Functional characterization of a novel mammalian zinc transporter, ZnT6. J Biol Chem. 2002;277:26389–95.PubMedCrossRef
61.
Zurück zum Zitat Kirschke CP, Huang L. ZnT7, a novel mammalian zinc transporter, accumulates zinc in the golgi apparatus. J Biol Chem. 2003;278:4096–102.PubMedCrossRef Kirschke CP, Huang L. ZnT7, a novel mammalian zinc transporter, accumulates zinc in the golgi apparatus. J Biol Chem. 2003;278:4096–102.PubMedCrossRef
62.
Zurück zum Zitat Kambe T, Hashimoto A, Fujimoto S. Current understanding of ZIP and ZnT zinc transporters in human health and diseases. Cell Mol Life Sci. 2014;71:3281–95.PubMedCrossRef Kambe T, Hashimoto A, Fujimoto S. Current understanding of ZIP and ZnT zinc transporters in human health and diseases. Cell Mol Life Sci. 2014;71:3281–95.PubMedCrossRef
63.
Zurück zum Zitat Daniels MJ, Jagielnicki M, Yeager M. Structure/function analysis of human ZnT8 (SLC30A8): a diabetes risk factor and zinc transporter. Curr Res Struct Biol. 2020;2:144–55.PubMedPubMedCentralCrossRef Daniels MJ, Jagielnicki M, Yeager M. Structure/function analysis of human ZnT8 (SLC30A8): a diabetes risk factor and zinc transporter. Curr Res Struct Biol. 2020;2:144–55.PubMedPubMedCentralCrossRef
64.
Zurück zum Zitat Merriman C, Fu D. Down-regulation of the islet-specific zinc transporter-8 (ZnT8) protects human insulinoma cells against inflammatory stress. J Biol Chem. 2019;294:16992–7006.PubMedPubMedCentralCrossRef Merriman C, Fu D. Down-regulation of the islet-specific zinc transporter-8 (ZnT8) protects human insulinoma cells against inflammatory stress. J Biol Chem. 2019;294:16992–7006.PubMedPubMedCentralCrossRef
65.
Zurück zum Zitat Liu C, Jursa T, Aschner M, Smith DR, Mukhopadhyay S. Up-regulation of the manganese transporter SLC30A10 by hypoxia-inducible factors defines a homeostatic response to manganese toxicity. Proc Natl Acad Sci USA. 2021;118: e2107673118.PubMedPubMedCentralCrossRef Liu C, Jursa T, Aschner M, Smith DR, Mukhopadhyay S. Up-regulation of the manganese transporter SLC30A10 by hypoxia-inducible factors defines a homeostatic response to manganese toxicity. Proc Natl Acad Sci USA. 2021;118: e2107673118.PubMedPubMedCentralCrossRef
66.
Zurück zum Zitat Fujishiro H, Yoshida M, Nakano Y, Himeno S. Interleukin-6 enhances manganese accumulation in SH-SY5Y cells: implications of the up-regulation of ZIP14 and the down-regulation of ZnT10. Metallomics. 2014;6:944–9.PubMedCrossRef Fujishiro H, Yoshida M, Nakano Y, Himeno S. Interleukin-6 enhances manganese accumulation in SH-SY5Y cells: implications of the up-regulation of ZIP14 and the down-regulation of ZnT10. Metallomics. 2014;6:944–9.PubMedCrossRef
68.
Zurück zum Zitat Bosomworth HJ, Ford D, Valentine RA. ZnT10 is expressed in adult human tissues. Proc Nutr Soc. 2011;70:E127.CrossRef Bosomworth HJ, Ford D, Valentine RA. ZnT10 is expressed in adult human tissues. Proc Nutr Soc. 2011;70:E127.CrossRef
69.
Zurück zum Zitat Jeong J, Eide DJ. The SLC39 family of zinc transporters. Mol Asp Med. 2013;34:612–9.CrossRef Jeong J, Eide DJ. The SLC39 family of zinc transporters. Mol Asp Med. 2013;34:612–9.CrossRef
70.
Zurück zum Zitat Wang F, Dufner-Beattie J, Kim B-E, Petris MJ, Andrews G, Eide DJ. Zinc-stimulated endocytosis controls activity of the mouse ZIP1 and ZIP3 zinc uptake transporters. J Biol Chem. 2004;279:24631–9.PubMedCrossRef Wang F, Dufner-Beattie J, Kim B-E, Petris MJ, Andrews G, Eide DJ. Zinc-stimulated endocytosis controls activity of the mouse ZIP1 and ZIP3 zinc uptake transporters. J Biol Chem. 2004;279:24631–9.PubMedCrossRef
71.
Zurück zum Zitat Weaver BP, Dufner-Beattie J, Kambe T, Andrews GK. Novel zinc-responsive post-transcriptional mechanisms reciprocally regulate expression of the mouse Slc39a4 and Slc39a5 zinc transporters (Zip4 and Zip5). Biol Chem. 2007;388:1301–12.PubMedPubMedCentralCrossRef Weaver BP, Dufner-Beattie J, Kambe T, Andrews GK. Novel zinc-responsive post-transcriptional mechanisms reciprocally regulate expression of the mouse Slc39a4 and Slc39a5 zinc transporters (Zip4 and Zip5). Biol Chem. 2007;388:1301–12.PubMedPubMedCentralCrossRef
72.
Zurück zum Zitat Huang L, Kirschke CP. A di-leucine sorting signal in ZIP1 (SLC39A1) mediates endocytosis of the protein. FEBS J. 2007;274:3986–97.PubMedCrossRef Huang L, Kirschke CP. A di-leucine sorting signal in ZIP1 (SLC39A1) mediates endocytosis of the protein. FEBS J. 2007;274:3986–97.PubMedCrossRef
73.
Zurück zum Zitat Tang Z, Sahu SN, Khadeer MA, Bai G, Franklin RB, Gupta A. Overexpression of the ZIP1 zinc transporter induces an osteogenic phenotype in mesenchymal stem cells. Bone. 2006;38:181–98.PubMedCrossRef Tang Z, Sahu SN, Khadeer MA, Bai G, Franklin RB, Gupta A. Overexpression of the ZIP1 zinc transporter induces an osteogenic phenotype in mesenchymal stem cells. Bone. 2006;38:181–98.PubMedCrossRef
74.
Zurück zum Zitat Cousins RJ, Blanchard RK, Popp MP, Liu L, Cao J, Moore JB, et al. A global view of the selectivity of zinc deprivation and excess on genes expressed in human THP-1 mononuclear cells. Proc Natl Acad Sci USA. 2003;100:6952–7.ADSPubMedPubMedCentralCrossRef Cousins RJ, Blanchard RK, Popp MP, Liu L, Cao J, Moore JB, et al. A global view of the selectivity of zinc deprivation and excess on genes expressed in human THP-1 mononuclear cells. Proc Natl Acad Sci USA. 2003;100:6952–7.ADSPubMedPubMedCentralCrossRef
76.
Zurück zum Zitat Zheng D, Feeney GP, Kille P, Hogstrand C. Regulation of ZIP and ZnT zinc transporters in zebrafish gill: zinc repression of ZIP10 transcription by an intronic MRE cluster. Physiol Genom. 2008;34:205–14.CrossRef Zheng D, Feeney GP, Kille P, Hogstrand C. Regulation of ZIP and ZnT zinc transporters in zebrafish gill: zinc repression of ZIP10 transcription by an intronic MRE cluster. Physiol Genom. 2008;34:205–14.CrossRef
77.
Zurück zum Zitat Kelleher SL, Lönnerdal B. Zip3 plays a major role in zinc uptake into mammary epithelial cells and is regulated by prolactin. Am J Physiol Cell Physiol. 2005;288:C1042–7.PubMedCrossRef Kelleher SL, Lönnerdal B. Zip3 plays a major role in zinc uptake into mammary epithelial cells and is regulated by prolactin. Am J Physiol Cell Physiol. 2005;288:C1042–7.PubMedCrossRef
78.
Zurück zum Zitat Yin S, Duan M, Fang B, Zhao G, Leng X, Zhang T. Zinc homeostasis and regulation: Zinc transmembrane transport through transporters. Crit Rev Food Sci Nutr. 2022;63(25):7627–37.PubMedCrossRef Yin S, Duan M, Fang B, Zhao G, Leng X, Zhang T. Zinc homeostasis and regulation: Zinc transmembrane transport through transporters. Crit Rev Food Sci Nutr. 2022;63(25):7627–37.PubMedCrossRef
80.
Zurück zum Zitat Weaver BP, Andrews GK. Regulation of zinc-responsive Slc39a5 (Zip5) translation is mediated by conserved elements in the 3’-untranslated region. Biometals. 2012;25:319–35.PubMedCrossRef Weaver BP, Andrews GK. Regulation of zinc-responsive Slc39a5 (Zip5) translation is mediated by conserved elements in the 3’-untranslated region. Biometals. 2012;25:319–35.PubMedCrossRef
81.
Zurück zum Zitat Grubman A, Lidgerwood GE, Duncan C, Bica L, Tan J-L, Parker SJ, et al. Deregulation of subcellular biometal homeostasis through loss of the metal transporter, Zip7, in a childhood neurodegenerative disorder. Acta Neuropathol Commun. 2014;2:25.PubMedPubMedCentralCrossRef Grubman A, Lidgerwood GE, Duncan C, Bica L, Tan J-L, Parker SJ, et al. Deregulation of subcellular biometal homeostasis through loss of the metal transporter, Zip7, in a childhood neurodegenerative disorder. Acta Neuropathol Commun. 2014;2:25.PubMedPubMedCentralCrossRef
82.
Zurück zum Zitat Bowers K, Srai SKS. The trafficking of metal ion transporters of the Zrt- and Irt-like protein family. Traffic. 2018;19:813–22.PubMedCrossRef Bowers K, Srai SKS. The trafficking of metal ion transporters of the Zrt- and Irt-like protein family. Traffic. 2018;19:813–22.PubMedCrossRef
83.
Zurück zum Zitat Aydemir TB, Liuzzi JP, McClellan S, Cousins RJ. Zinc transporter ZIP8 (SLC39A8) and zinc influence IFN-γ expression in activated human T cells. J Leukoc Biol. 2009;86:337–48.PubMedPubMedCentralCrossRef Aydemir TB, Liuzzi JP, McClellan S, Cousins RJ. Zinc transporter ZIP8 (SLC39A8) and zinc influence IFN-γ expression in activated human T cells. J Leukoc Biol. 2009;86:337–48.PubMedPubMedCentralCrossRef
84.
Zurück zum Zitat Liu M-J, Bao S, Gálvez-Peralta M, Pyle CJ, Rudawsky AC, Pavlovicz RE, et al. The zinc transporter SLC39A8 is a negative feedback regulator of NF-κB through zinc-mediated inhibition of IKK. Cell Rep. 2013;3:386–400.PubMedPubMedCentralCrossRef Liu M-J, Bao S, Gálvez-Peralta M, Pyle CJ, Rudawsky AC, Pavlovicz RE, et al. The zinc transporter SLC39A8 is a negative feedback regulator of NF-κB through zinc-mediated inhibition of IKK. Cell Rep. 2013;3:386–400.PubMedPubMedCentralCrossRef
85.
Zurück zum Zitat Wang C-Y, Jenkitkasemwong S, Duarte S, Sparkman BK, Shawki A, Mackenzie B, et al. ZIP8 is an iron and zinc transporter whose cell-surface expression is up-regulated by cellular iron loading. J Biol Chem. 2012;287:34032–43.PubMedPubMedCentralCrossRef Wang C-Y, Jenkitkasemwong S, Duarte S, Sparkman BK, Shawki A, Mackenzie B, et al. ZIP8 is an iron and zinc transporter whose cell-surface expression is up-regulated by cellular iron loading. J Biol Chem. 2012;287:34032–43.PubMedPubMedCentralCrossRef
86.
Zurück zum Zitat Lichten LA, Ryu M-S, Guo L, Embury J, Cousins RJ. MTF-1-mediated repression of the zinc transporter Zip10 is alleviated by zinc restriction. PLoS ONE. 2011;6: e21526.ADSPubMedPubMedCentralCrossRef Lichten LA, Ryu M-S, Guo L, Embury J, Cousins RJ. MTF-1-mediated repression of the zinc transporter Zip10 is alleviated by zinc restriction. PLoS ONE. 2011;6: e21526.ADSPubMedPubMedCentralCrossRef
87.
Zurück zum Zitat Miyai T, Hojyo S, Ikawa T, Kawamura M, Irié T, Ogura H, et al. Zinc transporter SLC39A10/ZIP10 facilitates antiapoptotic signaling during early B-cell development. Proc Natl Acad Sci U S A. 2014;111:11780–5.ADSPubMedPubMedCentralCrossRef Miyai T, Hojyo S, Ikawa T, Kawamura M, Irié T, Ogura H, et al. Zinc transporter SLC39A10/ZIP10 facilitates antiapoptotic signaling during early B-cell development. Proc Natl Acad Sci U S A. 2014;111:11780–5.ADSPubMedPubMedCentralCrossRef
88.
Zurück zum Zitat Pawan K, Neeraj S, Sandeep K, Kanta Ratho R, Rajendra P. Upregulation of Slc39a10 gene expression in response to thyroid hormones in intestine and kidney. Biochim Biophys Acta. 2007;1769:117–23.PubMedCrossRef Pawan K, Neeraj S, Sandeep K, Kanta Ratho R, Rajendra P. Upregulation of Slc39a10 gene expression in response to thyroid hormones in intestine and kidney. Biochim Biophys Acta. 2007;1769:117–23.PubMedCrossRef
89.
Zurück zum Zitat Martin AB, Aydemir TB, Guthrie GJ, Samuelson DA, Chang S-M, Cousins RJ. Gastric and colonic zinc transporter ZIP11 (Slc39a11) in mice responds to dietary zinc and exhibits nuclear localization123. J Nutr. 2013;143:1882–8.PubMedPubMedCentralCrossRef Martin AB, Aydemir TB, Guthrie GJ, Samuelson DA, Chang S-M, Cousins RJ. Gastric and colonic zinc transporter ZIP11 (Slc39a11) in mice responds to dietary zinc and exhibits nuclear localization123. J Nutr. 2013;143:1882–8.PubMedPubMedCentralCrossRef
90.
Zurück zum Zitat Zhao L, Oliver E, Maratou K, Atanur SS, Dubois OD, Cotroneo E, et al. The zinc transporter, ZIP12, regulates the pulmonary vascular response to chronic hypoxia. Nature. 2015;524:356–60.ADSPubMedPubMedCentralCrossRef Zhao L, Oliver E, Maratou K, Atanur SS, Dubois OD, Cotroneo E, et al. The zinc transporter, ZIP12, regulates the pulmonary vascular response to chronic hypoxia. Nature. 2015;524:356–60.ADSPubMedPubMedCentralCrossRef
91.
Zurück zum Zitat Xu J, Wan Z, Zhou B. Drosophila ZIP13 is posttranslationally regulated by iron-mediated stabilization. Biochim Biophys Acta BBA Mol Cell Res. 2019;186:1487–97.CrossRef Xu J, Wan Z, Zhou B. Drosophila ZIP13 is posttranslationally regulated by iron-mediated stabilization. Biochim Biophys Acta BBA Mol Cell Res. 2019;186:1487–97.CrossRef
92.
Zurück zum Zitat Homma K, Fujisawa T, Tsuburaya N, Yamaguchi N, Kadowaki H, Takeda K, et al. SOD1 as a molecular switch for initiating the homeostatic ER stress response under zinc deficiency. Mol Cell. 2013;52:75–86.PubMedCrossRef Homma K, Fujisawa T, Tsuburaya N, Yamaguchi N, Kadowaki H, Takeda K, et al. SOD1 as a molecular switch for initiating the homeostatic ER stress response under zinc deficiency. Mol Cell. 2013;52:75–86.PubMedCrossRef
93.
Zurück zum Zitat Lichten LA, Liuzzi JP, Cousins RJ. Interleukin-1β contributes via nitric oxide to the upregulation and functional activity of the zinc transporter Zip14 (Slc39a14) in murine hepatocytes. Am J Physiol Gastrointest Liver Physiol. 2009;296:G860–7.PubMedPubMedCentralCrossRef Lichten LA, Liuzzi JP, Cousins RJ. Interleukin-1β contributes via nitric oxide to the upregulation and functional activity of the zinc transporter Zip14 (Slc39a14) in murine hepatocytes. Am J Physiol Gastrointest Liver Physiol. 2009;296:G860–7.PubMedPubMedCentralCrossRef
94.
Zurück zum Zitat Maxel T, Smidt K, Petersen CC, Honoré B, Christensen AK, Jeppesen PB, et al. The zinc transporter Zip14 (SLC39a14) affects beta-cell function: proteomics, gene expression, and insulin secretion studies in INS-1E cells. Sci Rep. 2019;9:8589.ADSPubMedPubMedCentralCrossRef Maxel T, Smidt K, Petersen CC, Honoré B, Christensen AK, Jeppesen PB, et al. The zinc transporter Zip14 (SLC39a14) affects beta-cell function: proteomics, gene expression, and insulin secretion studies in INS-1E cells. Sci Rep. 2019;9:8589.ADSPubMedPubMedCentralCrossRef
95.
Zurück zum Zitat Nam H, Wang C-Y, Zhang L, Zhang W, Hojyo S, Fukada T, et al. ZIP14 and DMT1 in the liver, pancreas, and heart are differentially regulated by iron deficiency and overload: implications for tissue iron uptake in iron-related disorders. Haematologica. 2013;98:1049–57.PubMedPubMedCentralCrossRef Nam H, Wang C-Y, Zhang L, Zhang W, Hojyo S, Fukada T, et al. ZIP14 and DMT1 in the liver, pancreas, and heart are differentially regulated by iron deficiency and overload: implications for tissue iron uptake in iron-related disorders. Haematologica. 2013;98:1049–57.PubMedPubMedCentralCrossRef
97.
Zurück zum Zitat Evans GW, Winter TW. Zinc transport by transferrin in rat portal blood plasma. Biochem Biophys Res Commun. 1975;66:1218–24.PubMedCrossRef Evans GW, Winter TW. Zinc transport by transferrin in rat portal blood plasma. Biochem Biophys Res Commun. 1975;66:1218–24.PubMedCrossRef
98.
Zurück zum Zitat Carneiro C, Brito J, Bilreiro C, Barros M, Bahia C, Santiago I, et al. All about portal vein: a pictorial display to anatomy, variants and physiopathology. Insights Imaging. 2019;10:38.PubMedPubMedCentralCrossRef Carneiro C, Brito J, Bilreiro C, Barros M, Bahia C, Santiago I, et al. All about portal vein: a pictorial display to anatomy, variants and physiopathology. Insights Imaging. 2019;10:38.PubMedPubMedCentralCrossRef
99.
Zurück zum Zitat Kim M-H, Aydemir TB, Kim J, Cousins RJ. Hepatic ZIP14-mediated zinc transport is required for adaptation to endoplasmic reticulum stress. Proc Natl Acad Sci. 2017;114:E5805–14.PubMedPubMedCentralCrossRef Kim M-H, Aydemir TB, Kim J, Cousins RJ. Hepatic ZIP14-mediated zinc transport is required for adaptation to endoplasmic reticulum stress. Proc Natl Acad Sci. 2017;114:E5805–14.PubMedPubMedCentralCrossRef
102.
Zurück zum Zitat Guthrie GJ, Aydemir TB, Troche C, Martin AB, Chang S-M, Cousins RJ. Influence of ZIP14 (slc39A14) on intestinal zinc processing and barrier function. Am J Physiol Gastrointest Liver Physiol. 2015;308:171–8.CrossRef Guthrie GJ, Aydemir TB, Troche C, Martin AB, Chang S-M, Cousins RJ. Influence of ZIP14 (slc39A14) on intestinal zinc processing and barrier function. Am J Physiol Gastrointest Liver Physiol. 2015;308:171–8.CrossRef
103.
Zurück zum Zitat Geiser J, De Lisle RC, Andrews GK. The zinc transporter Zip5 (Slc39a5) regulates intestinal zinc excretion and protects the pancreas against zinc toxicity. PLoS ONE. 2013;8: e82149.ADSPubMedPubMedCentralCrossRef Geiser J, De Lisle RC, Andrews GK. The zinc transporter Zip5 (Slc39a5) regulates intestinal zinc excretion and protects the pancreas against zinc toxicity. PLoS ONE. 2013;8: e82149.ADSPubMedPubMedCentralCrossRef
104.
Zurück zum Zitat Herrera C, Pettiglio MA, Bartnikas TB. Investigating the role of transferrin in the distribution of iron, manganese, copper and zinc. J Biol Inorg Chem. 2014;19:869–77.PubMedPubMedCentralCrossRef Herrera C, Pettiglio MA, Bartnikas TB. Investigating the role of transferrin in the distribution of iron, manganese, copper and zinc. J Biol Inorg Chem. 2014;19:869–77.PubMedPubMedCentralCrossRef
106.
Zurück zum Zitat Blindauer CA, Harvey I, Bunyan KE, Stewart AJ, Sleep D, Harrison DJ, et al. Structure, properties, and engineering of the major zinc binding site on human albumin. J Biol Chem. 2009;284:23116–24.PubMedPubMedCentralCrossRef Blindauer CA, Harvey I, Bunyan KE, Stewart AJ, Sleep D, Harrison DJ, et al. Structure, properties, and engineering of the major zinc binding site on human albumin. J Biol Chem. 2009;284:23116–24.PubMedPubMedCentralCrossRef
107.
108.
Zurück zum Zitat Dietary LB, Absorption FIZ. Dietary factors influencing zinc absorption. J Nutr. 2000;130:1378S-1383S.CrossRef Dietary LB, Absorption FIZ. Dietary factors influencing zinc absorption. J Nutr. 2000;130:1378S-1383S.CrossRef
109.
Zurück zum Zitat Samtiya M, Aluko RE, Puniya AK, Dhewa T. Enhancing micronutrients bioavailability through fermentation of plant-based foods: a concise review. Fermentation. 2021;7:63.CrossRef Samtiya M, Aluko RE, Puniya AK, Dhewa T. Enhancing micronutrients bioavailability through fermentation of plant-based foods: a concise review. Fermentation. 2021;7:63.CrossRef
110.
Zurück zum Zitat Gibson RS, Yeudall F, Drost N, Mtitimuni B, Cullinan T. Dietary interventions to prevent zinc deficiency123. Am J Clin Nutr. 1998;68:484S-487S.PubMedCrossRef Gibson RS, Yeudall F, Drost N, Mtitimuni B, Cullinan T. Dietary interventions to prevent zinc deficiency123. Am J Clin Nutr. 1998;68:484S-487S.PubMedCrossRef
111.
Zurück zum Zitat Reyes AJ, Leary WP, Lockett CJ, Alcocer L. Diuretics and zinc. S Afr Med J. 1982;62:373–5.PubMed Reyes AJ, Leary WP, Lockett CJ, Alcocer L. Diuretics and zinc. S Afr Med J. 1982;62:373–5.PubMed
112.
Zurück zum Zitat Cohen N, Golik A. Zinc balance and medications commonly used in the management of heart failure. Heart Fail Rev. 2006;11:19–24.PubMedCrossRef Cohen N, Golik A. Zinc balance and medications commonly used in the management of heart failure. Heart Fail Rev. 2006;11:19–24.PubMedCrossRef
114.
Zurück zum Zitat Kondaiah P, Yaduvanshi PS, Sharp PA, Pullakhandam R. Iron and zinc homeostasis and interactions: Does enteric zinc excretion cross-talk with intestinal iron absorption? Nutrients. 2019;11:1885.PubMedPubMedCentralCrossRef Kondaiah P, Yaduvanshi PS, Sharp PA, Pullakhandam R. Iron and zinc homeostasis and interactions: Does enteric zinc excretion cross-talk with intestinal iron absorption? Nutrients. 2019;11:1885.PubMedPubMedCentralCrossRef
115.
Zurück zum Zitat Gopalsamy GL, Alpers DH, Binder HJ, Tran CD, Ramakrishna BS, Brown I, et al. The relevance of the colon to zinc nutrition. Nutrients. 2015;7:572–83.PubMedPubMedCentralCrossRef Gopalsamy GL, Alpers DH, Binder HJ, Tran CD, Ramakrishna BS, Brown I, et al. The relevance of the colon to zinc nutrition. Nutrients. 2015;7:572–83.PubMedPubMedCentralCrossRef
116.
Zurück zum Zitat Krebs NF. Overview of zinc absorption and excretion in the human gastrointestinal tract. J Nutr. 2000;130:1374S-S1377.PubMedCrossRef Krebs NF. Overview of zinc absorption and excretion in the human gastrointestinal tract. J Nutr. 2000;130:1374S-S1377.PubMedCrossRef
117.
Zurück zum Zitat Ohashi W, Hara T, Takagishi T, Hase K, Fukada T. Maintenance of intestinal epithelial homeostasis by zinc transporters. Dig Dis Sci. 2019;64:2404–15.PubMedCrossRef Ohashi W, Hara T, Takagishi T, Hase K, Fukada T. Maintenance of intestinal epithelial homeostasis by zinc transporters. Dig Dis Sci. 2019;64:2404–15.PubMedCrossRef
118.
Zurück zum Zitat Isani G, Carpenè E. Metallothioneins, unconventional proteins from unconventional animals: a long journey from nematodes to mammals. Biomolecules. 2014;4:435–57.PubMedPubMedCentralCrossRef Isani G, Carpenè E. Metallothioneins, unconventional proteins from unconventional animals: a long journey from nematodes to mammals. Biomolecules. 2014;4:435–57.PubMedPubMedCentralCrossRef
119.
Zurück zum Zitat Wei H, Desouki MM, Lin S, Xiao D, Franklin RB, Feng P. Differential expression of metallothioneins (MTs) 1, 2, and 3 in response to zinc treatment in human prostate normal and malignant cells and tissues. Mol Cancer. 2008;7:7.PubMedPubMedCentralCrossRef Wei H, Desouki MM, Lin S, Xiao D, Franklin RB, Feng P. Differential expression of metallothioneins (MTs) 1, 2, and 3 in response to zinc treatment in human prostate normal and malignant cells and tissues. Mol Cancer. 2008;7:7.PubMedPubMedCentralCrossRef
120.
Zurück zum Zitat Liu S, Wang N, Long Y, Wu Z, Zhou S. Zinc homeostasis: an emerging therapeutic target for neuroinflammation related diseases. Biomolecules. 2023;13:416.PubMedPubMedCentralCrossRef Liu S, Wang N, Long Y, Wu Z, Zhou S. Zinc homeostasis: an emerging therapeutic target for neuroinflammation related diseases. Biomolecules. 2023;13:416.PubMedPubMedCentralCrossRef
121.
Zurück zum Zitat Moltedo O, Verde C, Capasso A, Parisi E, Remondelli P, Bonatti S, et al. Zinc transport and metallothionein secretion in the intestinal human cell line Caco-2. J Biol Chem. 2000;275:31819–25.PubMedCrossRef Moltedo O, Verde C, Capasso A, Parisi E, Remondelli P, Bonatti S, et al. Zinc transport and metallothionein secretion in the intestinal human cell line Caco-2. J Biol Chem. 2000;275:31819–25.PubMedCrossRef
122.
Zurück zum Zitat Fathi F, Ektefa F, Tafazzoli M, Rostami K, Nejad MR, Fathi M, et al. The concentration of serum zinc in celiac patients compared to healthy subjects in Tehran. Gastroenterol Hepatol Bed Bench. 2013;6:92.PubMedPubMedCentral Fathi F, Ektefa F, Tafazzoli M, Rostami K, Nejad MR, Fathi M, et al. The concentration of serum zinc in celiac patients compared to healthy subjects in Tehran. Gastroenterol Hepatol Bed Bench. 2013;6:92.PubMedPubMedCentral
123.
Zurück zum Zitat Human HM, Deficiency Z. Human zinc deficiency. J Nutr. 2000;130:1344S-1349S.CrossRef Human HM, Deficiency Z. Human zinc deficiency. J Nutr. 2000;130:1344S-1349S.CrossRef
126.
Zurück zum Zitat Kumar P, Lal NR, Mondal AK, Mondal A, Gharami RC, Maiti A. Zinc and skin: a brief summary. Dermatol Online J. 2012;18:1.PubMedCrossRef Kumar P, Lal NR, Mondal AK, Mondal A, Gharami RC, Maiti A. Zinc and skin: a brief summary. Dermatol Online J. 2012;18:1.PubMedCrossRef
127.
Zurück zum Zitat Wegmüller R, Tay F, Zeder C, Brnić M, Hurrell RF. Zinc absorption by young adults from supplemental zinc citrate is comparable with that from zinc gluconate and higher than from zinc oxide123. J Nutr. 2014;144:132–6.PubMedCrossRef Wegmüller R, Tay F, Zeder C, Brnić M, Hurrell RF. Zinc absorption by young adults from supplemental zinc citrate is comparable with that from zinc gluconate and higher than from zinc oxide123. J Nutr. 2014;144:132–6.PubMedCrossRef
128.
Zurück zum Zitat Baum MK, Lai S, Sales S, Page JB, Campa A. Randomized, controlled clinical trial of zinc supplementation to prevent immunological failure in HIV-infected adults. Clin Infect Dis. 2010;50:1653–60.PubMedCrossRef Baum MK, Lai S, Sales S, Page JB, Campa A. Randomized, controlled clinical trial of zinc supplementation to prevent immunological failure in HIV-infected adults. Clin Infect Dis. 2010;50:1653–60.PubMedCrossRef
129.
Zurück zum Zitat Tokuyama A, Kanda E, Itano S, Kondo M, Wada Y, Kadoya H, et al. Effect of zinc deficiency on chronic kidney disease progression and effect modification by hypoalbuminemia. PLoS ONE. 2021;16: e0251554.PubMedPubMedCentralCrossRef Tokuyama A, Kanda E, Itano S, Kondo M, Wada Y, Kadoya H, et al. Effect of zinc deficiency on chronic kidney disease progression and effect modification by hypoalbuminemia. PLoS ONE. 2021;16: e0251554.PubMedPubMedCentralCrossRef
130.
Zurück zum Zitat Squitti R, Pal A, Picozza M, Avan A, Ventriglia M, Rongioletti M, et al. Zinc therapy in early Alzheimer’s disease: safety and potential therapeutic efficacy. Biomolecules. 2020;10:1164.PubMedPubMedCentralCrossRef Squitti R, Pal A, Picozza M, Avan A, Ventriglia M, Rongioletti M, et al. Zinc therapy in early Alzheimer’s disease: safety and potential therapeutic efficacy. Biomolecules. 2020;10:1164.PubMedPubMedCentralCrossRef
131.
Zurück zum Zitat McClain C, Vatsalya V, Cave M. Role of zinc in the development/progression of alcoholic liver disease. Curr Treat Options Gastroenterol. 2017;15:285–95.PubMedPubMedCentralCrossRef McClain C, Vatsalya V, Cave M. Role of zinc in the development/progression of alcoholic liver disease. Curr Treat Options Gastroenterol. 2017;15:285–95.PubMedPubMedCentralCrossRef
132.
Zurück zum Zitat Yesiladali M, Yazici MGK, Attar E, Kelestimur F. Differentiating polycystic ovary syndrome from adrenal disorders. Diagnostics (Basel). 2022;12:2045.PubMedCrossRef Yesiladali M, Yazici MGK, Attar E, Kelestimur F. Differentiating polycystic ovary syndrome from adrenal disorders. Diagnostics (Basel). 2022;12:2045.PubMedCrossRef
136.
Zurück zum Zitat Wang K, Zhou B, Kuo Y-M, Zemansky J, Gitschier J. A novel member of a zinc transporter family is defective in acrodermatitis enteropathica. Am J Hum Genet. 2002;71:66–73.PubMedPubMedCentralCrossRef Wang K, Zhou B, Kuo Y-M, Zemansky J, Gitschier J. A novel member of a zinc transporter family is defective in acrodermatitis enteropathica. Am J Hum Genet. 2002;71:66–73.PubMedPubMedCentralCrossRef
138.
Zurück zum Zitat Goff JP. Invited review: mineral absorption mechanisms, mineral interactions that affect acid–base and antioxidant status, and diet considerations to improve mineral status. J Dairy Sci. 2018;101:2763–813.PubMedCrossRef Goff JP. Invited review: mineral absorption mechanisms, mineral interactions that affect acid–base and antioxidant status, and diet considerations to improve mineral status. J Dairy Sci. 2018;101:2763–813.PubMedCrossRef
141.
Zurück zum Zitat George AA, Mishra AK, Sahu KK, Sargent J. Acquired acrodermatitis enteropathica. Am J Med. 2021;134:e2-3.PubMedCrossRef George AA, Mishra AK, Sahu KK, Sargent J. Acquired acrodermatitis enteropathica. Am J Med. 2021;134:e2-3.PubMedCrossRef
142.
Zurück zum Zitat Cho Y-E, Lomeda R-AR, Ryu S-H, Sohn H-Y, Shin H-I, Beattie JH, et al. Zinc deficiency negatively affects alkaline phosphatase and the concentration of Ca, Mg and P in rats. Nutr Res Pract. 2007;1:113–9.PubMedPubMedCentralCrossRef Cho Y-E, Lomeda R-AR, Ryu S-H, Sohn H-Y, Shin H-I, Beattie JH, et al. Zinc deficiency negatively affects alkaline phosphatase and the concentration of Ca, Mg and P in rats. Nutr Res Pract. 2007;1:113–9.PubMedPubMedCentralCrossRef
144.
Zurück zum Zitat Woodruff G, Bouwkamp CG, De Vrij FM, Lovenberg T, Bonaventure P, Kushner SA, et al. The zinc transporter SLC39A7 (ZIP7) is essential for regulation of cytosolic zinc levels. Mol Pharmacol. 2018;94:1092–100.PubMedCrossRef Woodruff G, Bouwkamp CG, De Vrij FM, Lovenberg T, Bonaventure P, Kushner SA, et al. The zinc transporter SLC39A7 (ZIP7) is essential for regulation of cytosolic zinc levels. Mol Pharmacol. 2018;94:1092–100.PubMedCrossRef
145.
Zurück zum Zitat Bin B-H, Hojyo S, Ryong Lee T, Fukada T. Spondylocheirodysplastic Ehlers–Danlos syndrome (SCD-EDS) and the mutant zinc transporter ZIP13. Rare Dis. 2014;2: e974982.PubMedPubMedCentralCrossRef Bin B-H, Hojyo S, Ryong Lee T, Fukada T. Spondylocheirodysplastic Ehlers–Danlos syndrome (SCD-EDS) and the mutant zinc transporter ZIP13. Rare Dis. 2014;2: e974982.PubMedPubMedCentralCrossRef
146.
Zurück zum Zitat Sandstead HH. Chapter 61—Zinc. In: Nordberg GF, Fowler BA, Nordberg M, editors. Handbook on the toxicology of metals. 4th ed. San Diego: Academic Press; 2015. p. 1369–85.CrossRef Sandstead HH. Chapter 61—Zinc. In: Nordberg GF, Fowler BA, Nordberg M, editors. Handbook on the toxicology of metals. 4th ed. San Diego: Academic Press; 2015. p. 1369–85.CrossRef
147.
Zurück zum Zitat Ryu M-S, Aydemir TB. Chapter 23—Zinc. In: Marriott BP, Birt DF, Stallings VA, Yates AA, editors. Present knowledge in nutrition. 11th ed. Cambridge: Academic Press; 2020. p. 393–408.CrossRef Ryu M-S, Aydemir TB. Chapter 23—Zinc. In: Marriott BP, Birt DF, Stallings VA, Yates AA, editors. Present knowledge in nutrition. 11th ed. Cambridge: Academic Press; 2020. p. 393–408.CrossRef
148.
Zurück zum Zitat Prasad R, Hawthorne B, Durai D, McDowell I. Zinc in denture adhesive: a rare cause of copper deficiency in a patient on home parenteral nutrition. BMJ Case Rep. 2015;2015:bcr2015211390.PubMedPubMedCentralCrossRef Prasad R, Hawthorne B, Durai D, McDowell I. Zinc in denture adhesive: a rare cause of copper deficiency in a patient on home parenteral nutrition. BMJ Case Rep. 2015;2015:bcr2015211390.PubMedPubMedCentralCrossRef
150.
Zurück zum Zitat Wazir SM, Ghobrial I. Copper deficiency, a new triad: anemia, leucopenia, and myeloneuropathy. J Community Hosp Intern Med Perspect. 2017;7:265–8.PubMedPubMedCentralCrossRef Wazir SM, Ghobrial I. Copper deficiency, a new triad: anemia, leucopenia, and myeloneuropathy. J Community Hosp Intern Med Perspect. 2017;7:265–8.PubMedPubMedCentralCrossRef
151.
Zurück zum Zitat Collins JF, Prohaska JR, Knutson MD. Metabolic crossroads of iron and copper. Nutr Rev. 2010;68:133–47.PubMedCrossRef Collins JF, Prohaska JR, Knutson MD. Metabolic crossroads of iron and copper. Nutr Rev. 2010;68:133–47.PubMedCrossRef
152.
Zurück zum Zitat Greenberg MI, Vearrier D. Metal fume fever and polymer fume fever. Clin Toxicol (Phila). 2015;53:195–203.PubMedCrossRef Greenberg MI, Vearrier D. Metal fume fever and polymer fume fever. Clin Toxicol (Phila). 2015;53:195–203.PubMedCrossRef
153.
Zurück zum Zitat Munie S, Pintavorn P. Erythropoietin-resistant anemia secondary to zinc-induced hypocupremia in a hemodialysis patient. Case Rep Nephrol Dial. 2021;11:167–75.PubMedPubMedCentralCrossRef Munie S, Pintavorn P. Erythropoietin-resistant anemia secondary to zinc-induced hypocupremia in a hemodialysis patient. Case Rep Nephrol Dial. 2021;11:167–75.PubMedPubMedCentralCrossRef
154.
Zurück zum Zitat Amisha F, Saluja P, Gautam N, Biyani S, Kakadia S. Acquired hyperzincaemia due to zinc-laden denture adhesives leading to hypocupraemia as a cause of neutropenia. Eur J Case Rep Intern Med. 2021;8: 002983.PubMedPubMedCentral Amisha F, Saluja P, Gautam N, Biyani S, Kakadia S. Acquired hyperzincaemia due to zinc-laden denture adhesives leading to hypocupraemia as a cause of neutropenia. Eur J Case Rep Intern Med. 2021;8: 002983.PubMedPubMedCentral
155.
Zurück zum Zitat Duncan A, Morrison I, Bryson S. Iatrogenic copper deficiency: risks and cautions with zinc prescribing. Br J Clin Pharmacol. 2023;89:2825–9.PubMedCrossRef Duncan A, Morrison I, Bryson S. Iatrogenic copper deficiency: risks and cautions with zinc prescribing. Br J Clin Pharmacol. 2023;89:2825–9.PubMedCrossRef
156.
Zurück zum Zitat Gupta N, Carmichael MF. Zinc-induced copper deficiency as a rare cause of neurological deficit and anemia. Cureus. 2023;15:e43856.PubMedPubMedCentral Gupta N, Carmichael MF. Zinc-induced copper deficiency as a rare cause of neurological deficit and anemia. Cureus. 2023;15:e43856.PubMedPubMedCentral
157.
158.
Zurück zum Zitat Ali N, Ferrao K, Mehta KJ. Liver iron loading in alcohol-associated liver disease. Am J Pathol. 2022;193(10):1427–39.PubMedCrossRef Ali N, Ferrao K, Mehta KJ. Liver iron loading in alcohol-associated liver disease. Am J Pathol. 2022;193(10):1427–39.PubMedCrossRef
160.
Zurück zum Zitat Ferrao K, Ali N, Mehta KJ. Iron and iron-related proteins in alcohol consumers: cellular and clinical aspects. J Mol Med. 2022;100:1673–89.PubMedCrossRef Ferrao K, Ali N, Mehta KJ. Iron and iron-related proteins in alcohol consumers: cellular and clinical aspects. J Mol Med. 2022;100:1673–89.PubMedCrossRef
Metadaten
Titel
Role of zinc in health and disease
verfasst von
Lucy I. Stiles
Kevin Ferrao
Kosha J. Mehta
Publikationsdatum
01.12.2024
Verlag
Springer International Publishing
Erschienen in
Clinical and Experimental Medicine / Ausgabe 1/2024
Print ISSN: 1591-8890
Elektronische ISSN: 1591-9528
DOI
https://doi.org/10.1007/s10238-024-01302-6

Weitere Artikel der Ausgabe 1/2024

Clinical and Experimental Medicine 1/2024 Zur Ausgabe

Leitlinien kompakt für die Innere Medizin

Mit medbee Pocketcards sicher entscheiden.

Seit 2022 gehört die medbee GmbH zum Springer Medizin Verlag

Erhebliches Risiko für Kehlkopfkrebs bei mäßiger Dysplasie

29.05.2024 Larynxkarzinom Nachrichten

Fast ein Viertel der Personen mit mäßig dysplastischen Stimmlippenläsionen entwickelt einen Kehlkopftumor. Solche Personen benötigen daher eine besonders enge ärztliche Überwachung.

Nach Herzinfarkt mit Typ-1-Diabetes schlechtere Karten als mit Typ 2?

29.05.2024 Herzinfarkt Nachrichten

Bei Menschen mit Typ-2-Diabetes sind die Chancen, einen Myokardinfarkt zu überleben, in den letzten 15 Jahren deutlich gestiegen – nicht jedoch bei Betroffenen mit Typ 1.

15% bedauern gewählte Blasenkrebs-Therapie

29.05.2024 Urothelkarzinom Nachrichten

Ob Patienten und Patientinnen mit neu diagnostiziertem Blasenkrebs ein Jahr später Bedauern über die Therapieentscheidung empfinden, wird einer Studie aus England zufolge von der Radikalität und dem Erfolg des Eingriffs beeinflusst.

Costims – das nächste heiße Ding in der Krebstherapie?

28.05.2024 Onkologische Immuntherapie Nachrichten

„Kalte“ Tumoren werden heiß – CD28-kostimulatorische Antikörper sollen dies ermöglichen. Am besten könnten diese in Kombination mit BiTEs und Checkpointhemmern wirken. Erste klinische Studien laufen bereits.

Update Innere Medizin

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.