Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Metallomics Dynamic Article Links

Cite this: Metallomics, 2011, 3, 1124–1129

www.rsc.org/metallomics MINIREVIEW
The oxidative stress of zinc deficiencyw
David J. Eide
Received 13th June 2011, Accepted 7th July 2011

Downloaded from https://academic.oup.com/metallomics/article/3/11/1124/6016256 by guest on 03 November 2021


DOI: 10.1039/c1mt00064k

Zinc is an essential catalytic and structural cofactor for many enzymes and other proteins. While
Zn2+ is not redox active under physiological conditions, it has been known for many years that
zinc deficiency causes increased oxidative stress and, consequently, increased oxidative damage to
DNA, proteins, and lipids. These results have indicated that zinc plays an indirect antioxidant
role and that dietary inadequacy may contribute to human diseases such as cancer. Recent studies
are helping to identify the primary sources of oxidative stress in low zinc. In addition, through
studies of the model eukaryotic cell, Saccharomyces cerevisiae, we are now beginning to
understand the strategies cells use to limit this stress and reduce its damage.

Introduction Approximately 2 billion people on Earth consume insufficient


amounts of zinc.2 Within the US alone, it was estimated that
Zinc is required for the function of many proteins playing both almost 10% of people are at risk of inadequate zinc intake.2
catalytic and structural roles. It has been estimated that zinc is The great importance of zinc to cellular function indicates that
an essential cofactor for the function of as many as 10% of the this widespread deficiency is likely to have major health
proteins encoded by the human genome.1 These include DNA consequences. This prediction is supported by epidemiological
binding proteins such as zinc finger proteins and the p53 studies linking zinc deficiency to the increased risk of chronic
tumor suppressor protein and enzymes such as Cu/Zn super- disease including cancer.3,4
oxide dismutase. Zinc deficiency is a worldwide health problem. Given the large number of different roles that zinc plays in
biology, it is difficult to assign deficits in specific zinc-dependent
functions as contributing to the etiology of human diseases.
Department of Nutritional Sciences, University of Wisconsin-Madison, However, one obvious possible link between zinc deficiency
1415 Linden Drive, Room 340B, Madison, WI 53706-1571.
E-mail: eide@nutrisci.wisc.edu; Fax: 608-262-5860;
and disease is through zinc’s action as an antioxidant. The
Tel: 608-263-1613 antioxidant function of zinc has been a mystery for many
w This article is published as part of a themed issue on Metal Toxicity, years.5,6 Because zinc is not redox active under physiological
Guest Edited by Gregor Grass and Christopher Rensing. conditions, it is clear that the metal is not directly involved in
donating electrons to or accepting electrons from oxidant
molecules. Therefore zinc is not functioning as an antioxidant
David Eide received his BS in per se. Therefore, the antioxidant properties of zinc are a result
Microbiology from the Uni- of some indirect mechanism of action. This review explores the
versity of Minnesota in 1981 current data regarding the potential sources of oxidative stress
and his PhD in Molecular
in zinc deficiency and what cellular responses to zinc deficiency
Biology from the University
of Wisconsin-Madison in occur to limit the oxidative stress and repair the damage that
1987. He was a postdoctoral results from it.
researcher at the Massachusetts
Institute of Technology from Oxidative stress in zinc deficiency and
1987–1990 and then at the
University of Utah in its consequences
1990–1991. He began his work Several studies have demonstrated that zinc deficiency
on zinc as an Assistant
increases the production of reactive oxygen species. Many of
Professor in the Department
these studies have focused on zinc deficiency in cultured cells.
David J. Eide of Biochemistry and Molecular
Biology at the University of For example, Oteiza et al.7 showed that growth of mouse 3T3
Minnesota-Duluth. He moved to the Nutritional Science cells in low zinc results in increased reactive oxygen species
Department at the University of Missouri-Columbia in 1996 (ROS) as measured using the ROS-sensitive fluorescent probe
and to the Department of Nutritional Sciences at the University 2,7-dichlorodihydrofluorescein (DCFH). Upon reaction with
of Wisconsin-Madison in 2004. oxidants such as O2 , H2O2, and hydroxyl radical,8 DCFH is

1124 Metallomics, 2011, 3, 1124–1129 This journal is c The Royal Society of Chemistry 2011
converted to the fluorescent 2,7-dichlorofluorescein (DCF)
form. Similarly, Ho and Ames9 found that ROS levels
increased in zinc deficient rat C6 glioma cells. The increased
ROS in rat glioma cells was accompanied by an increase in
oxidative DNA damage. The effects of this damage were likely
compounded by the loss of key DNA repair mechanisms. It
was also shown that zinc deficiency in lung fibroblasts,10 liver
stellate cells,11 IMR-32 neuroblastoma cells,12 prostate epithelial
cells,13 and neuronal PC12 cells14 all show increased levels of
ROS and oxidative damage when grown in low zinc.
Increased ROS and resulting damage has also been found in
experimental animals fed low zinc diets. Perhaps the first

Downloaded from https://academic.oup.com/metallomics/article/3/11/1124/6016256 by guest on 03 November 2021


report of these effects in animals was made by Sullivan
et al.15 who showed that zinc-deficient rats had increased lipid
peroxidation in lung microsomes. These results were con-
firmed and expanded to effects in other tissues in a series of
papers by Bray and colleagues in the mid-1980s16,17 and in
subsequent papers by several other groups.18–23 Extension of
these findings to human subjects has also been made; increased
oxidative stress and resulting damage has been documented in
humans with sub-optimal zinc intakes.24,25 These findings
support the proposal that zinc deficiency may be a contributing Fig. 1 Summary of possible sources of reactive oxygen species in
factor to the progression of cancer.26,27 zinc-limited cells. Processes on the left are induced by zinc and limit
Finally, we have recently determined that the yeast reactive oxygen species (ROS) production or accumulation. Thus, low
Saccharomyces cerevisiae also shows increased levels of ROS zinc would decrease these activities and cause increased ROS.
when grown under zinc-limiting conditions.28–30 This result Conversely, the processes on the right are inhibited by zinc and increase
demonstrates that the effect of low zinc on oxidative stress ROS production. Low zinc would increase these processes and
increase ROS production.
occurs in both simple, single-cell eukaryotes and metazoans.
In addition, these results have allowed us to bring the power of
yeast genetics to bear on this fundamental issue of zinc eliminated as a possible cause in at least some cell types is loss
nutrition as described later in this review. of SOD activity. In several studies in which low zinc causes
oxidative stress, it has been noted that SOD activity actually
increased in those cells, probably as a response to the increased
Possible sources of ROS in low zinc
oxidative stress.9,10 One type of cell in which zinc limitation
A number of possible mechanisms have been proposed to explain did decrease SOD activity was yeast. Low zinc growth reduced
the increased oxidative stress in zinc-limited cells (Fig. 1). A few SOD activity by approximately 2-fold.30 However, when SOD
of these mechanisms have been discussed in prior reviews6,26,31 so was overexpressed from a plasmid vector in zinc-limited cells
those will only be briefly summarized here. First, increased and activity was restored to zinc-replete levels, oxidative stress
oxidative stress may arise from decreased activity of key anti- still occurred. Thus, loss of SOD activity is not a primary
oxidant enzymes. As one obvious example, Cu/Zn superoxide cause of oxidative stress in low zinc in yeast or mammalian
dismutase (SOD) is a zinc-dependent enzyme whose activity may cells. This is not to say, however, that SOD activity is not
be compromised under zinc deficiency. Second, Zn-metallothionein important for tolerating the oxidative stress of zinc deficiency.
has recently been shown to also have antioxidant activity To identify what antioxidant enzymes were important for low
through the oxidation of metal-bound cysteine ligands in the zinc growth, yeast mutants lacking 28 different antioxidant
protein.32,33 Expression of many metallothioneins is induced enzymes were analyzed for growth defects in low zinc.30 These
by zinc treatment possibly leading to an increased ability to mutants disrupted peroxiredoxins, glutathione peroxidases,
eliminate reactive oxygen species. Thus, in zinc-limited cells catalases, and superoxide dismutases. Of these, only two
where metallothionein expression is low, this antioxidant antioxidant enzymes were found to be critical for zinc-limited
function would be lost. As a third mechanism, it has been growth, Tsa1, the major cytosolic peroxiredoxin, and Sod1,
proposed that zinc may compete with redox active metal ions the cytosolic Cu/Zn superoxide dismutase. Superoxide dismutase
like Cu and Fe for binding to sites on proteins and other degrades superoxide to hydrogen peroxide and the Tsa1
cellular macromolecules.34,35 Such competition would inhibit peroxiredoxin degrades hydrogen peroxide to water. The fact
the site-specific production of oxygen radicals through metal- that these two enzymes work sequentially and that both are
catalyzed Fenton chemistry. Finally, zinc has been proposed required for low zinc growth suggests that the initial form of
to bind to and protect free sulfhydryl groups in proteins. This ROS generated in low zinc is superoxide.
may be another way in which zinc normally plays an anti- Additional possible sources of oxidative stress in low zinc
oxidant role and inhibits production of reactive oxygen species. include disruption of the mitochondrial electron transport
At this time, there is little evidence available to evaluate chain. Zinc is required for many functions in mitochondria
these different hypotheses in vivo. One hypothesis that can be including matrix-localized metalloproteases that proteolytically

This journal is c The Royal Society of Chemistry 2011 Metallomics, 2011, 3, 1124–1129 1125
cleave newly arrived proteins during their mitochondrial neuronal cells increased ROS production through loss of this
maturation. Mitochondrial enzymes such as the yeast Adh3 inhibitory effect.14 When activated, the NMDA receptor
alchohol dehydogenase and Leu9 a-isopropylmalate synthase triggers downstream responses by mediating calcium influx.
(for leucine synthesis) also require zinc as a cofactor. In These authors showed that zinc-limiting differentiated PC-12
addition, cytochrome c oxidase of the electron transport chain neuronal cells in culture resulted in a rapid increase in DCF
requires zinc bound to its Cox4 subunit for function. Mutation fluorescence. During that treatment, the p67phox subunit of the
of the zinc site in Cox4 disrupts the structural stability of the NADPH oxidase complex was recruited to the membrane
complex and breaks the electron transport chain.36 Such breaks fraction indicating activation of oxidase activity. These effects
result in increased generation of ROS from the release of were associated with an increase in cytosolic calcium that was
electrons to oxygen largely by coenzyme Q. It was recently blocked by the NMDA receptor antagonist MK-801. Calcium
shown that the mitochondrial matrix contains a labile cationic can activate NADPH oxidase activity through the stimulation
pool of low molecular weight zinc, primarily in the form of a of protein kinase C. Both MK-801 and the protein kinase C

Downloaded from https://academic.oup.com/metallomics/article/3/11/1124/6016256 by guest on 03 November 2021


complex with some unknown small molecule ligand, and that inhibitor Ro-320432 also blocked NADPH oxidase activation
this pool is required for respiratory function.37 In addition, zinc and the increased ROS in low zinc. Finally, direct inhibition of
deficiency was found to decrease expression of components of NADPH oxidase with apocyanin blocked ROS production.
the electron transport chain in human lung fibroblasts suggesting Thus, in neuronal cells, hyperactivation of the NMDA receptor
another mechanism through which this pathway can be appears to be the initial cause of increased oxidative stress in
disrupted.10 Thus, zinc deficiency could disrupt the function low zinc. It remains to be seen if types of cells that lack the
or expression of electron transport chain components and NMDA receptor have NADPH oxidase activity activated by
increase oxidative stress in that way. other mechanisms in low zinc. In this regard, it is intriguing to
Another candidate source of ROS in low zinc is the pathway note that the findings of Kojima-Yuasa et al.11 support the
of disulfide bond formation that occurs in the secretory path- possibility that increased NADPH oxidase activity may
way. Recent studies have shown that the ER is a potent potentially be the source of ROS in zinc-deficient hepatic stellate
generator of ROS.38,39 Newly synthesized proteins with cells. These authors found that treatment of zinc-deficient stellate
reduced thiols react with a protein disulfide isomerase (Pdi1) to cells with an oxidase inhibitor decreased ROS production.
generate disulfide bonds. Two electrons are transferred to Pdi1 The caveat with these experiments is that the inhibitor used,
during this reaction that are then passed to the Ero1 FAD- diphenyliodonium chloride, is not specific to NADPH oxidases.
dependent oxidase and then to molecular oxygen to generate Lastly, a pro-oxidant role of NADPH was also suggested by
H2O2 and/or O2 .40,41 Proteins with correct disulfide bonds Bray and colleagues.16,17 These authors found that rat lung
are stable due to their proper folding. However, incorrect microsomes isolated from zinc deficient rats generated more
disulfide bonds are sensitive to reduction by GSH (generating free radicals in the presence of NADPH than did those
GSSG) regenerating the reduced thiol state.40,42 Protein mis- isolated from zinc-replete animals. The NADPH-dependence
folding in the ER leads to more incorrect disulfide bonds and and sensitivity to inhibition by carbon monoxide suggested
cycling via the GSH-dependent pathway.43 Thus, the oxidative that the source of this oxidative stress was cytochrome P450.
protein folding system is a major generator of ROS and More recently, it was found that the level of P450 CYP4F2
GSSG under ER stress conditions.42 Several observations isozyme was elevated in liver from zinc deficient rats which
suggest that the ER is a source of ROS in low Zn. First and may explain, at least in part, this elevated ROS generation.21
foremost is the observation that loss of ER zinc transporter
function in yeast increases the sensitivity of cells to exogenous
oxidants.44 We also found that Zn deficiency increases GSSG Cellular responses to zinc deficiency to alleviate
levels28 and note that oxidative protein folding is the primary
oxidative stress
source of GSSG in cells.42 Third, Zn deficiency activates the
unfolded protein response (UPR) in both yeast and mamma- The primary mechanism cells use to prevent the oxidative
lian cells, indicating protein misfolding is occurring in the stress of zinc deficiency is to maintain zinc homeostasis. Cells
ER.45,46 As noted above, expression of misfolded proteins in of all organisms have mechanisms that provide for a constant
the ER causes oxidative stress.43 cytosolic and organellar zinc status while extracellular or
While NADPH is well recognized as an important anti- dietary levels fluctuate.50–53 In general, these mechanisms
oxidant for its role in supplying electrons to glutathione reductase include the control of zinc uptake, zinc efflux from the cell,
and thioredoxin reductase, it is also clear that NADPH can also sequestration within organelles, or zinc binding by proteins
play a pro-oxidant role at least in regard to zinc deficiency. such as metallothionein. One key player in zinc homeostasis
NADPH oxidase activity has recently been implicated as the with particular relevance to oxidative stress and zinc deficiency
source of ROS in zinc-limited neuronal cells. NADPH oxidases is the Zap1 transcription factor of yeast. Zap1 is a zinc-
generate O2 from O2 for the purpose of microbial killing in responsive transcriptional activator that induces transcription
phagocytic cells and also to provide moderate levels of ROS of its target genes in zinc-limited cells and whose activity is
for its role in cellular signaling.47 It was shown previously that shut off in zinc-replete cells.54 This regulation occurs primarily
activation of the N-methyl-D-aspartate (NMDA) receptor by zinc binding directly to the activation domains of this
caused superoxide production by activation of the NADPH transcription factor which folds those domains into conforma-
oxidase.48 Because zinc is a known inhibitor of the NMDA tions that are unable to recruit co-activator proteins that
receptor,49 Aimo et al. addressed whether zinc limitation of facilitate transcription initiation.

1126 Metallomics, 2011, 3, 1124–1129 This journal is c The Royal Society of Chemistry 2011
than a ‘‘reactive’’ mechanism because these antioxidant genes
are induced in response to the cause of the increased ROS, low
zinc, rather than in response to the ROS itself.
Another aspect of Zap1 regulation relevant to oxidative
stress in low zinc is the control of sulfate assimilation. The
assimilation of sulfate occurs through a multi-step pathway
resulting in the generation of methionine and cysteine.61,62
This pathway has a high demand for NADPH requiring 6
moles of NADPH for every one mole of methionine or
Fig. 2 Mechanism regulating sulfate assimilation in response to cysteine produced. Enzymes on this pathway include those
organic sulfur compounds and zinc. Repression of sulfate assimilation encoded by the MET3, MET14, and MET16 genes. In our
in low zinc (–Zn) conserves NADPH for its antioxidant roles. See text
recent transcriptomics analysis of zinc-responsive gene expres-

Downloaded from https://academic.oup.com/metallomics/article/3/11/1124/6016256 by guest on 03 November 2021


for further details. Names in all capital letters and italicized, e.g.
sion, we identified 36 genes that were repressed in zinc-limited
MET30, denote genes while names not italicized in which only the first
letter is capitalized, e.g. Zap1, denote proteins. Ub, ubiquitin; Met,
cells in a Zap1-dependent manner.29 These apparently Zap1-
methionine; SAM, S-adenosylmethionine; Cys, cysteine; GSH, repressed genes included MET3, MET14, and MET16. This
glutathione. observation raised the question of how a transcriptional
activator protein that induces expression could repress expres-
We and others have sought to identify the various targets of sion of these genes. The answer lies in the regulatory mechan-
Zap1 to understand the strategies cells use to thrive under ism that controls these genes in response to sulfur amino acid
zinc deficiency. Through microarray analysis and other levels. When these amino acids are low, MET3, MET14, and
approaches, it has been estimated that Zap1 induces the MET16 are all induced in expression by the Met4 transcription
expression of B80 genes in low zinc.55–59 One of these target factor (Fig. 2). Met4 activity is regulated at the level of protein
genes, TSA1 encoding the major cytosolic peroxiredoxin in the stability. When cells are replete with organic sulfur compounds
cell, is induced by Zap1. Peroxiredoxins are antioxidant (methionine, cysteine, S-adenosylmethionine, glutathione),
enzymes that degrade H2O2 to H2O through the reducing Met4 is ubiquitinated and degraded by the proteasome.62
equivalents provided by thioredoxin and NADPH-dependent Ubiquitination and targeting for degradation is mediated by
thioredoxin reductase. The TSA1 gene is expressed in zinc- the SCFMet30 ubiquitin ligase. SCFMet30 is a multisubunit
replete cell to high levels such that those cells contain ubiquitin ligase that relies on the Met30 subunit for target
B400 000 molecules of the Tsa1 protein per cell. The TSA1 protein specificity. The connection between Zap1 and MET
gene was induced 2–3-fold in zinc-limited wild type cells and gene expression is the Met30 subunit. We found that the
this induction was blocked by deletion of the Zap1 gene28 MET30 gene is a Zap1 target whose expression is upregulated
Protein levels increased to the same extent. In addition, TSA1 in low zinc.29 Increased levels of Met30 increase the activity of
expression was induced 2–3-fold in zinc-replete cells expressing SCFMet30 and cause the degradation of Met4, loss of MET
a constitutively active allele of Zap1 and the TSA1 promoter gene expression, and decreased sulfate assimilation. Consistent
contains a functional Zap1 binding site that is required for with this model, we found that zinc-limited cells had markedly
zinc-responsive regulation. Deletion of the TSA1 gene resulted reduced levels of free methionine and cysteine. When the site
in a strain that is severely defective for zinc-limited growth, of ubiquitination in the Met4 protein was mutated such that
emphasizing the importance of this antioxidant to growth this protein could no longer be down-regulated in low zinc, we
under these conditions. Zinc-limited wild type yeast show a found that these cells had decreased NADPH, elevated
marked increase in DCF fluorescence that is greatly increased GSSG : GSH ratio, increased ROS production, and increased
in a tsa1 null mutant. In addition, the GSSG : GSH ratio sensitivity to external oxidants such as H2O2. These results
increased in zinc-limited cells and this too was greatly exacerbated indicate that repression of sulfate assimilation occurs in zinc-
by loss of Tsa1 function. Regulation by Zap1 was clearly limited cells to conserve NADPH for its antioxidant functions.
important for this resistance to oxidative stress; when the
promoter of TSA1 was mutated such that the Zap1 binding
Summary and future directions
site was mutated to prevent induction in low zinc (but without
affecting basal expression), these promoter mutant cells also It has been known for many years that zinc-deficient mamma-
showed increased ROS production and a perturbed intra- lian cells in culture and whole animals on low zinc diets
cellular redox environment. Thus, the doubling of Tsa1 produce increased levels of ROS. There is also ample evidence
protein levels conferred by Zap1 induction is clearly important that zinc limitation in human subjects results in oxidative
for low zinc growth. It is also notable that the CTT1 gene, stress. This stress can lead to damage of proteins, lipids, and
which encodes the cytosolic isoform of catalase, also appears DNA. DNA damage can lead to mutation and this may
to be directly regulated by Zap1.57 UTH1, a gene of unknown explain the epidemiological link between zinc deficiency and
function that has also been implicated in oxidative stress cancer and other chronic diseases. The ultimate source of the
resistance, is a Zap1 target as well.60 Thus, through Zap1, oxidative stress of zinc deficiency remains a mystery. The
yeast cells respond to zinc deficiency by inducing expression of recent finding of elevated NADPH oxidase activity in zinc-
antioxidant enzymes that protect the cell from the oxidative limited neuronal cells suggests one possible mechanism of
stress generated under those conditions. We consider this to be ROS production. It will be interesting to determine if this
a ‘‘proactive’’ strategy of oxidative stress resistance rather mechanism also applies to cells of other types. Recent studies

This journal is c The Royal Society of Chemistry 2011 Metallomics, 2011, 3, 1124–1129 1127
of yeast have established that even single-cell eukaryotes 15 J. F. Sullivan, M. M. Jetton, H. K. Hahn and R. E. Burch,
experience increased ROS when zinc limited. This is allowing Enhanced lipid peroxidation in liver microsomes of zinc-deficient
rats, Am. J. Clin. Nutr., 1980, 33(1), 51–6.
us to use the power of yeast genetics to identify the source of 16 T. M. Bray, S. Kubow and W. J. Bettger, Effect of dietary zinc on
oxidants in yeast and this may inform us about other processes endogenous free radical production in rat lung microsomes,
that generate ROS in zinc-limited mammalian cells. Further- J. Nutr., 1986, 116(6), 1054–60.
more, we have found a direct regulatory link in yeast between 17 J. D. Hammermueller, T. M. Bray and W. J. Bettger, Effect of zinc
and copper deficiency on microsomal NADPH-dependent active
antioxidant gene expression and zinc deficiency through the oxygen generation in rat lung and liver, J. Nutr., 1987, 117(5),
Zap1 transcription factor. Zap1 mediates a proactive rather 894–901.
than a reactive response to oxidative stress by up-regulating 18 P. I. Oteiza, K. L. Olin, C. G. Fraga and C. L. Keen, Zinc
deficiency causes oxidative damage to proteins, lipids and DNA
antioxidants in response to low zinc. It will be interesting to
in rat testes, J. Nutr., 1995, 125(4), 823–9.
determine if mammalian cells also link regulation of anti- 19 A. A. Shaheen and A. A. el-Fattah, Effect of dietary zinc onlipid
oxidant enzymes directly to zinc status. peroxidation, glutathione, protein thiols levels and superoxide

Downloaded from https://academic.oup.com/metallomics/article/3/11/1124/6016256 by guest on 03 November 2021


dismutase activity in rat tissues, Int. J. Biochem. Cell Biol., 1995,
27(1), 89–95.
Acknowledgements 20 R. Canali, F. Vignolini, F. Nobili and E. Mengheri, Reduction of
oxidative stress and cytokine-induced neutrophil chemoattractant
Work in the author’s lab on zinc homeostasis and oxidative (CINC) expression by red wine polyphenols in zinc deficiency
induced intestinal damage of rat, Free Radical Biol. Med., 2000,
stress is funded by NIH grants GM056285 and GM093303. 28(11), 1661–70.
The author apologizes to the many colleagues whose work was 21 R. S. Bruno, Y. Song, S. W. Leonard, D. J. Mustacich,
not cited due the brief format of this minireview. A. W. Taylor and M. G. Traber, et al., Dietary zinc restriction
in rats alters antioxidant status and increases plasma F2 isopros-
tanes, J. Nutr. Biochem., 2007, 18(8), 509–18.
References 22 Y. Song, V. Elias, A. Loban, A. G. Scrimgeour and E. Ho, Marginal
zinc deficiency increases oxidative DNA damage in the prostate after
1 C. Andreini, L. Banci, I. Bertini and A. Rosato, Counting the zinc- chronic exercise, Free Radical Biol. Med., 2010, 48(1), 82–8.
proteins encoded in the human genome, J. Proteome Res., 2006, 23 Y. Song, S. W. Leonard, M. G. Traber and E. Ho, Zinc deficiency
5(1), 196–201. affects DNA damage, oxidative stress, antioxidant defenses, and
2 K. H. Brown, J. A. Rivera, Z. Bhutta, R. S. Gibson, J. C. King and DNA repair in rats, J. Nutr., 2009, 139(9), 1626–31.
B. Lonnerdal, et al., International Zinc Nutrition Consultative 24 A. S. Prasad, F. W. Beck, B. Bao, J. T. Fitzgerald, D. C. Snell and
Group (IZiNCG) technical document #1. Assessment of the risk of J. D. Steinberg, et al., Zinc supplementation decreases incidence of
zinc deficiency in populations and options for its control, Food infections in the elderly: effect of zinc on generation of cytokines
Nutr. Bull., 2004, 25(1 Suppl 2), S99–203. and oxidative stress, Am. J. Clin. Nutr., 2007, 85(3), 837–44.
3 N. Leone, D. Courbon, P. Ducimetiere and M. Zureik, Zinc, 25 Y. Song, C. S. Chung, R. S. Bruno, M. G. Traber, K. H. Brown
copper, and magnesium and risks for all-cause, cancer, and and J. C. King, et al., Dietary zinc restriction and repletion affects
cardiovascular mortality, Epidemiology, 2006, 17(3), 308–14. DNA integrity in healthy men, Am. J. Clin. Nutr., 2009, 90(2),
4 T. Wu, C. T. Sempos, J. L. Freudenheim, P. Muti and E. Smit, 321–8, PMCID: 2709309.
Serum iron, copper and zinc concentrations and risk of cancer 26 E. Ho, Zinc deficiency, DNA damage and cancer risk., J. Nutr.
mortality in US adults, Ann. Epidemiol., 2004, 14(3), 195–201. Biochem., 2004, 15(10), 572–8.
5 T. M. Bray and W. J. Bettger, The physiological role of zinc as an 27 B. N. Ames and P. Wakimoto, Are vitamin and mineral deficien-
antioxidant, Free Radical Biol. Med., 1990, 8(3), 281–91. cies a major cancer risk?, Nat. Rev. Cancer, 2002, 2(9), 694–704.
6 S. R. Powell, The antioxidant properties of zinc, J. Nutr., 2000, 28 C. Y. Wu, A. J. Bird, D. R. Winge and D. J. Eide, Regulation of
130(5S Suppl), 1447S–54S. the yeast TSA1 peroxiredoxin by ZAP1 is an adaptive response to the
7 P. I. Oteiza, M. S. Clegg, M. P. Zago and C. L. Keen, Zinc oxidative stress of zinc deficiency, J. Biol. Chem., 2007, 282(4), 2184–95.
deficiency induces oxidative stress and AP-1 activation in 3T3 cells, 29 C. Y. Wu, S. Roje, F. J. Sandoval, A. J. Bird, D. R. Winge and
Free Radical Biol. Med., 2000, 28(7), 1091–9. D. J. Eide, Repression of sulfate assimilation is an adaptive
8 B. Halliwell and M. Whiteman, Measuring reactive species and response of yeast to the oxidative stress of zinc deficiency,
oxidative damage in vivo and in cell culture: how should you do it J. Biol. Chem., 2009, 284(40), 27544–56, PMCID: 2785683.
and what do the results mean?, Br. J. Pharmacol., 2004, 142(2), 30 C. Y. Wu, J. Steffen and D. J. Eide, Cytosolic superoxide
231–55. dismutase (SOD1) is critical for tolerating the oxidative stress of zinc
9 E. Ho and B. N. Ames, Low intracellular zinc induces oxidative deficiency in yeast, PLoS One, 2009, 4(9), e7061, PMCID: 2737632.
DNA damage, disrupts p53, Nfkappa B, and AP1 DNA binding, 31 K. Jomova and M. Valko, Advances in metal-induced oxidative
and affects DNA repair in a rat glioma cell line, Proc. Natl. Acad. stress and human disease, Toxicology, 2011, 283(2–3), 65–87.
Sci. U. S. A., 2002, 99(26), 16770–5. 32 W. Maret, Cellular zinc and redox states converge in the metallo-
10 E. Ho, C. Courtemanche and B. N. Ames, Zinc deficiency induces thionein/thionein pair, J. Nutr., 2003, 133(5 Suppl 1), 1460S–2S.
oxidative DNA damage and increases p53 expression in human 33 W. Maret, Metallothionein redox biology in the cytoprotective and
lung fibroblasts, J. Nutr., 2003, 133(8), 2543–8. cytotoxic functions of zinc, Exp. Gerontol., 2008, 43(5), 363–9.
11 A. Kojima-Yuasa, K. Umeda, T. Ohkita, D. Opare Kennedy, 34 M. P. Zago and P. I. Oteiza, The antioxidant properties of zinc:
S. Nishiguchi and I. Matsui-Yuasa, Role of reactive oxygen species interactions with iron and antioxidants, Free Radical Biol. Med.,
in zinc deficiency-induced hepatic stellate cell activation, Free 2001, 31(2), 266–74.
Radical Biol. Med., 2005, 39(5), 631–40. 35 M. P. Zago, S. V. Verstraeten and P. I. Oteiza, Zinc in the
12 M. P. Zago, G. G. Mackenzie, A. M. Adamo, C. L. Keen and prevention of Fe2+ initiated lipid and protein oxidation, Biol.
P. I. Oteiza, Differential modulation of MAP kinases by zinc Res., 2000, 33(2), 143–50.
deficiency in IMR-32 cells: role of H(2)O(2), Antioxid. Redox 36 H. J. Coyne, 3rd, S. Ciofi-Baffoni, L. Banci, I. Bertini, L. Zhang
Signaling, 2005, 7(11–12), 1773–82. and G. N. George, et al., The characterization and role of zinc
13 M. Yan, Y. Song, C. P. Wong, K. Hardin and E. Ho, Zinc binding in yeast Cox4, J. Biol. Chem., 2007, 282(12), 8926–34.
deficiency alters DNA damage response genes in normal human 37 A. Atkinson, O. Khalimonchuk, P. Smith, H. Sabic, D. Eide and
prostate epithelial cells, J. Nutr., 2008, 138(4), 667–73. D. R. Winge, Mzm1influences a labile pool of mitochondrial zinc
14 L. Aimo, G. N. Cherr and P. I. Oteiza, Low extracellular zinc important for respiratory function, J. Biol. Chem., 2010, 285(25),
increases neuronal oxidant production through nadph oxidase and 19450–9, PMCID: 2885224.
nitric oxide synthase activation, Free Radical Biol. Med., 2010, 38 A. R. Frand, J. W. Cuozzo and C. A. Kaiser, Pathways for protein
48(12), 1577–87. disulphide bond formation, Trends Cell Biol., 2000, 10(5), 203–10.

1128 Metallomics, 2011, 3, 1124–1129 This journal is c The Royal Society of Chemistry 2011
39 B. P. Tu and J. S. Weissman, Oxidative protein folding in 52 R. A. Colvin, W. R. Holmes, C. P. Fontaine and W. Maret,
eukaryotes: mechanisms and consequences, J. Cell Biol., 2004, Cytosolic zinc buffering and muffling: their role in intracellular
164(3), 341–6. zinc homeostasis, Metallomics, 2010, 2(5), 306–17.
40 B. P. Tu and J. S. Weissman, The FAD- and O(2)-dependent 53 L. A. Lichten and R. J. Cousins, Mammalian zinc transporters:
reaction cycle of Ero1-mediated oxidative protein folding in the nutritional and physiologic regulation, Annu. Rev. Nutr., 2009, 29,
endoplasmic reticulum, Mol. Cell, 2002, 10(5), 983–94. 153–76.
41 E. Gross, C. S. Sevier, N. Heldman, E. Vitu, M. Bentzur and 54 D. J. Eide, Homeostatic and Adaptive Responses to Zinc
C. A. Kaiser, et al., Generating disulfides enzymatically: reaction Deficiency in Saccharomyces cerevisiae, J. Biol. Chem., 2009,
products and electron acceptors of the endoplasmic reticulum thiol 284(28), 18565–9, PMCID: 2707215.
oxidase Ero1p, Proc. Natl. Acad. Sci. U. S. A., 2006, 103(2), 299–304. 55 D. S. Yuan, Zinc-regulated genes in Saccharomyces cerevisiae
42 J. W. Cuozzo and C. A. Kaiser, Competition between glutathione revealed by transposon tagging, Genetics, 2000, 156(1), 45–58.
and protein thiols for disulphide-bond formation, Nat. Cell Biol., 56 T. J. Lyons, A. P. Gasch, L. A. Gaither, D. Botstein, P. O. Brown
1999, 1(3), 130–5. and D. J. Eide, Genome-wide characterization of the Zap1p zinc-
43 C. M. Haynes, E. A. Titus and A. A. Cooper, Degradation of responsive regulon in yeast, Proc. Natl. Acad. Sci. U. S. A., 2000,
misfolded proteins prevents ER-derived oxidative stress and cell 97(14), 7957–62, PMCID: 16652.

Downloaded from https://academic.oup.com/metallomics/article/3/11/1124/6016256 by guest on 03 November 2021


death, Mol. Cell, 2004, 15(5), 767–76. 57 C. Y. Wu, A. J. Bird, L. M. Chung, M. A. Newton, D. R. Winge
44 L. Li and J. Kaplan, The yeast gene MSC2, a member of the cation and D. J. Eide, Differential control of Zap1-regulated genes in
diffusion facilitator family, affects the cellular distribution of zinc, response to zinc deficiency in Saccharomyces cerevisiae, BMC
J. Biol. Chem., 2000, 276, 5036–43. Genomics, 2008, 9, 370.
45 C. D. Ellis, C. W. Macdiarmid and D. J. Eide, Heteromeric protein 58 R. De. Nicola, L. A. Hazelwood, E. A. De. Hulster, M. C. Walsh,
complexes mediate zinc transport into the secretory pathway of T. A. Knijnenburg and M. J. Reinders, et al., Physiological and
eukaryotic cells, J. Biol. Chem., 2005, 280(31), 28811–8. transcriptional responses of Saccharomyces cerevisiae to zinc
46 C. D. Ellis, F. Wang, C. W. MacDiarmid, S. Clark, T. Lyons and limitation in chemostat cultures, Appl. Environ. Microbiol., 2007,
D. J. Eide, Zinc and the Msc2 zinc transporter protein are required 73(23), 7680–92.
for endoplasmic reticulum function, J. Cell Biol., 2004, 166(3), 59 V. J. Higgins, P. J. Rogers and I. W. Dawes, Application of
325–35, PMCID: 2172251. genome-wide expression analysis to identify molecular markers
47 W. Droge, Free radicals in the physiological control of cell useful in monitoring industrial fermentations, Appl. Environ.
function, Physiol. Rev., 2002, 82(1), 47–95. Microbiol., 2003, 69(12), 7535–40.
48 A. M. Brennan, S. W. Suh, S. J. Won, P. Narasimhan, 60 P. D. Bandara, J. A. Flattery-O’Brien, C. M. Grant and
T. M. Kauppinen and H. Lee, et al., NADPH oxidase is the I. W. Dawes, Involvement of the Saccharomyces cerevisiae
primary source of superoxide induced by NMDA receptor activa- UTH1 gene in the oxidative-stress response, Curr. Genet., 1998,
tion, Nat. Neurosci., 2009, 12(7), 857–63, PMCID: 2746760. 34(4), 259–68.
49 P. Paoletti, A. M. Vergnano, B. Barbour and M. Casado, Zinc at 61 D. Thomas and Y. Surdin-Kerjan, Metabolism of sulfur amino
glutamatergic synapses, Neuroscience, 2009, 158(1), 126–36. acids in Saccharomyces cerevisiae, Microbiol. Mol. Biol. Rev., 1997,
50 D. J. Eide, Zinc transporters and the cellular trafficking of zinc, 61(4), 503–32.
Biochim. Biophys. Acta, Mol. Cell Res., 2006, 1763(7), 711–22. 62 P. Kaiser, N. Y. Su, J. L. Yen, I. Ouni and K. Flick, The yeast
51 M. R. Bleackley and R. T. Macgillivray, Transition metal homeo- ubiquitin ligase SCFMet30: connecting environmental and intra-
stasis: from yeast to human disease, Biometals, 2011. cellular conditions to cell division, Cell Div., 2006, 1, 16.

This journal is c The Royal Society of Chemistry 2011 Metallomics, 2011, 3, 1124–1129 1129

You might also like