UTILIZATION OF SELENIUM IN THE MOUSE BRAIN ......various brain diseases. Mouse models have been...

146
UTILIZATION OF SELENIUM IN THE MOUSE BRAIN: IMPLICATIONS FOR NEUROLOGICAL DISEASE A DISSERTATION SUBMITTED TO THE GRADUATE DIVISION OF THE UNIVERSITY OF HAWAI‘I AT MĀNOA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY IN CELL AND MOLECULAR BIOLOGY (NEUROSCIENCES) AUGUST 2012 By Arjun Venkat Raman Dissertation Committee: Marla Berry, Chairperson Frederick Bellinger Scott Lozanoff Robert Nichols Bruce Shiramizu

Transcript of UTILIZATION OF SELENIUM IN THE MOUSE BRAIN ......various brain diseases. Mouse models have been...

UTILIZATION OF SELENIUM IN THE MOUSE BRAIN:

IMPLICATIONS FOR NEUROLOGICAL DISEASE

A DISSERTATION SUBMITTED TO

THE GRADUATE DIVISION OF THE UNIVERSITY OF HAWAI‘I AT MĀNOA

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

IN

CELL AND MOLECULAR BIOLOGY (NEUROSCIENCES)

AUGUST 2012

By

Arjun Venkat Raman

Dissertation Committee:

Marla Berry, Chairperson

Frederick Bellinger

Scott Lozanoff

Robert Nichols

Bruce Shiramizu

ii

ABSTRACT

Selenium is a chemical element that is an essential micronutrient associated with various

aspects of human health. The biochemical activity of selenium is mediated by proteins,

which incorporate it into the amino acid selenocysteine (Sec). There are 25 genes in

humans encoding Sec-containing selenoproteins. The functionally characterized

selenoproteins are oxidoreductase enzymes involved in cellular oxidation-reduction

reactions. Most selenoproteins are expressed in the mammalian brain, and dietary

selenium deficiency causes preferential retention in brain relative to other body organs.

Further, dietary selenium deficiency and specific selenoproteins are associated with

various brain diseases. Mouse models have been extensively used to study the function

and handling of selenium in mammals. Genetic deletion of a Sec-rich protein in mice

causes brain selenium deficiency, neurodegeneration and neurological impairment, and

disruption of a phospholipid hydroperoxidase selenoenzyme causes rapid

neurodegeneration. Therefore, selenoprotein expression and function promotes a healthy

nervous system. However, selenium metabolism and the function of several

selenoproteins in brain are not clearly defined.

The overall purpose of this work is to clarify the function and utilization of selenium in

the mammalian brain, to reveal implications for developmental and neurological diseases.

The goal of these studies is to investigate changes in brain function and selenoprotein

expression under conditions of altered selenium metabolism in mice. The research

presented in this dissertation covers three major topics that are separated into chapters. To

investigate selenium distribution in the brain, the neurological consequences of disrupting

selenium transport and recycling in mice are assessed and compared. Disruption of

selenium transport caused more profound neurological consequences than disruption of

selenium recycling. To investigate potential functions of selenium in the nervous system,

select selenoproteins were examined for cellular and subcellular expression in cells and

brain tissue from transgenic and control mice. Select selenoproteins and synthesis factors

were observed at synapses, suggesting localized expression and physiological relevance.

To investigate a potential interaction between selenium and methamphetamine,

iii

expression profiling of selenoproteins in mouse brain after exposure to methamphetamine

is described. Selenoprotein synthesis was adversely affected by methamphetamine

administration in mice. These results confirm the importance of selenium in the

mammalian brain.

iv

PAPERS ARISING FROM THIS DISSERTATION

Published, in press:

Raman, A. V., Pitts, M. W., Seyedali, A., Hashimoto, A. C., Seale, L. A., Bellinger, F. P.

and Berry, M. J. (2012), Absence of selenoprotein P but not selenocysteine lyase results

in severe neurological dysfunction. Genes, Brain and Behavior. doi: 10.1111/j.1601-

183X.2012.00794.x

In preparation for publication:

Raman, A. V., Pitts, M. W., Hashimoto, A. C., Nichols, R.A., Bellinger, F. P. and Berry,

M. J. (2012), Expression of selenoprotein W in neurons extends into processes and is

highly dependent on selenoprotein P. In preparation.

v

DEDICATION

I would like to dedicate this dissertation directly to my parents, Neerja and Vasan Raman,

and indirectly to all of my ancestors, family, and friends. Needless to say, nothing about

me would have been possible without generations of love, dedication, and perseverance.

vi

TABLE OF CONTENTS

ABSTRACT........................................................................................................... ii

PAPERS ARISING FROM THIS DISSERTATION............................................ iv

DEDICATION....................................................................................................... v

LIST OF FIGURES................................................................................................ viii

CHAPTER 1: INTRODUCTION

1.1 Selenium in Chemistry and Biology.................................................... 1

1.2 Synthesis of Selenoproteins................................................................. 3

1.3 Selenoproteins as Oxidoreductase Enzymes........................................ 6

1.4 Selenoproteins and Metabolism........................................................... 9

1.5 The Selenoprotein Family.................................................................... 12

1.6 Selenoproteins in the Nervous System................................................. 18

1.7 References............................................................................................ 26

CHAPTER 2: ABSENCE OF SELENOPROTEIN P BUT NOT

SELENOCYSTEINE LYASE RESULTS IN SEVERE NEURLOGICAL

DYSFUNCTION

2.1 Abstract................................................................................................ 38

2.2 Introduction.......................................................................................... 39

2.3 Methods................................................................................................ 40

2.4 Results.................................................................................................. 46

2.5 Discussion............................................................................................ 51

2.6 References............................................................................................ 56

CHAPTER 3: EXPRESSION OF SELENOPROTEIN W IN NEURONS

EXTENDS INTO PROCESSES AND IS HIGHLY DEPENDENT ON

SELENOPROTEIN P

3.1 Abstract................................................................................................ 75

3.2 Introduction.......................................................................................... 76

3.3 Methods................................................................................................ 78

3.4 Results.................................................................................................. 82

3.5 Discussion............................................................................................ 85

3.6 References............................................................................................ 89

vii

CHAPTER 4: METHAMPHETAMINE-INDUCED ALTERATIONS IN

SELENOPROTEIN EXPRESSION IN MICE

4.1 Abstract................................................................................................ 101

4.2 Introduction.......................................................................................... 102

4.3 Methods................................................................................................ 104

4.4 Results.................................................................................................. 108

4.5 Discussion............................................................................................ 110

4.6 References............................................................................................ 112

CHAPTER 5: CONCLUSION

5.1 Summary and Discussion..................................................................... 123

5.2 Oxidative Stress and Selenoproteins.................................................... 128

5.3 Redox Systems and Selenoproteins in the Brain.................................. 130

5.4 References............................................................................................ 134

viii

LIST OF FIGURES

1.1 Assimilation of Se into selenoproteins 24

1.2 Selenoprotein localization to subcellular domains 25

2.1 Spontaneous activity and motor coordination is more reduced in male

than female Sepp1-/- mice compared to control mice fed a standard diet

65

2.2 High Se diet improves spontaneous activity and motor coordination

more in male than female Sepp1-/- mice compared to control mice

66

2.3 Generation of Scly-knockout mice 67

2.4 Spontaneous activity and motor coordination is normal in male and

female Scly-/- mice fed a low Se diet

68

2.5 Spatial learning and memory is disrupted in Sepp1-/- mice fed a

standard diet

69

2.6 Spatial learning and memory is not disrupted in Scly-/- mice fed a

standard diet

70

2.7 Spatial learning is mildly impaired in Scly-/- mice fed a low Se diet 71

2.8 Expression of selenoprotein transcripts is increased in Se-deficient

Scly-/- mice brains

72

2.9 Expression of selenoproteins is decreased in Se-deficient Scly-/- mice

brains

73

2.10 Glutathione peroxidase activity is decreased in Se-deficient Scly-/- mice

brains

74

3.1 Sepw1 is expressed in cell bodies and processes of cultured neurons 95

3.2 Sepw1 is expressed in cell bodies and processes of pyramidal neurons in

cortex and hippocampus

96

3.3 Regional expression of Sepw1 in neurons of mouse brain 97

3.4 Sepw1 is present in isolated nerve terminals 98

3.5 Sepw1 expression in isolated nerve terminals is greatly reduced in mice

lacking Sepp1

99

3.6 Several selenoprotein synthesis factors are present in isolated nerve

terminals

100

ix

4.1 Expression of selenoprotein mRNAs is altered by methamphetamine 116

4.2 Selenoprotein W and selenophosphate synthetase 2 mRNA are

upregulated in midbrain after two weeks of methamphetamine

administration

117

4.3 Selenoprotein expression following two weeks of methamphetamine

administration is not changed in mice brains

118

4.4 Long term methamphetamine administration dramatically reduces

selenoprotein mRNAs in striatum

119

4.5 Sepp1 impacts brain expression of selenoproteins more than

methamphetamine

120

4.6 Methamphetamine reduces tyrosine hydroxylase expression in

substantia nigra independently of Sepp1

121

4.7 Apoptotic cell death is not increased in methamphetamine-treated

selenoprotein P knockout mice

122

1

CHAPTER 1

INTRODUCTION

SELENIUM IN CHEMISTRY AND BIOLOGY

In 1817 the Swedish chemist Jöns Jacob Berzelius discovered the element selenium (Se)

after analyzing an impurity that was contaminating the sulfuric acid being produced at a

nearby factory. Since the factory workers were suffering an illness caused by Se, it was

initially considered to be a poisonous chemical. Nearly 140 years later, Se was identified

as an essential micronutrient in humans and livestock (Hatfield, 2006). Se is a trace

mineral present in the Earth’s crust and ocean water at wide ranging abundances,

averaging between 0.05 and 2.0 ppm. Perhaps due to low environmental quantities, Se

has a narrow concentration range from deficiency to toxicity in animals. The profound

biological effects of Se suggested that it was involved in enzymatic activity. This

speculation was confirmed when Se was discovered to be present as selenocysteine in the

active site of two unrelated enzymes (Behne et al., 1990, Berry et al., 1991b, Forstrom et

al., 1978, Rotruck et al., 1973b).

Like other group 16 elements oxygen and sulfur, Se has two unpaired electrons,

producing highly reactive atoms. Se is classified a nonmetal but has some metalloid

properties. For instance, it is a photoconductive semiconductor, meaning the electrical

conductivity is enhanced by light and somewhere in between that of a true conductor and

an insulator (Liao et al., 2010). This unique reactivity proved instrumental in production

of the first commercial photocopiers made by Xerox, and has been exploited for use in

solid-state electronics and photoelectric cells. Se and sulfur share some similar chemical

and physical properties, such as electronegativity and atomic radius, however selenium is

much heavier. Thus Se and sulfur compounds exist naturally as structurally analogous

molecules typically complexed with a metal. These inorganic and organic molecules exist

in a variety of oxidation states ranging from -2 to +6, and include selenides (Se2−

),

2

selenites (SeO32−

), and selenates (SeO42–

), as well as the amino acids selenocysteine and

selenomethionine, and methylated forms (Johansson et al., 2005).

In biological systems, however, the substantial differences between Se and sulfur are

highlighted by their non-interchangeable nature. For example, sulfur represents about one

quarter of one percent of human body weight and possesses almost no toxicity risk even

at very high levels. In contrast, Se is required in nanomolar concentrations, and the

median lethal dose of in humans is 1.5 - 6.0 mg/kg body weight with inorganic salts

being more acutely toxic than organic forms (Koller & Exon, 1986). In vertebrates, Se is

obtained almost exclusively by ingestion of plants and animals. The amount of Se in

dietary sources depends on the concentration of environmental Se in the soil or water.

Currently, the recommended daily allowance of dietary Se for healthy adults is 55 µg/day

and the upper tolerable intake level is 400 µg/day (Monsen, 2000).

Most cellular biological functions of Se are attributed to selenoproteins, which

incorporate Se as the amino acid selenocysteine (Sec). Sec is structurally analogous to

cysteine with the exception of Se replacing sulfur in the side chain. However, the

similarity ends there. Se and sulfur have similar properties, but their differences generate

divergent character between Sec and cysteine. Having Sec in the active site of an enzyme

generally produces more efficient catalysis (Berry et al., 1992). The cause of the higher

catalytic capacity of Sec is still under investigation, and several ideas have been

proposed. For example, Sec is described as having stronger nucleophilic and electrophilic

character than cysteine, potentially aiding in electron and proton transfer during catalysis.

The acid dissociation constant for the selenol (pKa = 5.2) of Sec is lower than the thiol

(pKa = 8.3) of cysteine. This means that at physiological pH, Sec tends towards

ionization whereas cysteine tends towards protonation, and at low pH Sec retains much

higher reactivity than cysteine. The selenol is further thought to be a superior leaving

group due to higher acidity (Johansson et al., 2005).

The biological rationale for Sec utilization has been enigmatic. Due to the relative

stabilities of each oxidation state of Sec and cysteine near physiological pH, Sec has two

3

main advantages over cysteine in terms of reaction with oxygen. Kinetic calculations

suggest that oxidized Sec, both selenenic and seleninic acid, will cycle more rapidly to

the reduced form than oxidized forms of cysteine, sulfenic and sulfinic acid. Second,

further oxidation to selenonic acid is extremely unfavorable, unlike irreversible oxidation

of cysteine to sulfonic acid. The actual reactivity of the residue will ultimately depend on

the context of the Sec within the local and global structure of the peptide and enzyme.

However, these unique reactive properties may potentially explain the biological pressure

to use Sec in proteins during evolution (Hondal & Ruggles, 2011).

SYNTHESIS OF SELENOPROTEINS

Beyond the chemistry, Sec is an extremely unusual amino acid, and can be considered an

anomaly in the genetic code. Roughly defined, the genetic code is the system in which

genetic information is first transcribed from deoxyribonucleic acid (DNA) to messenger

ribonucleic acid (mRNA), and subsequently interpreted by ribosomes and several transfer

RNAs (tRNAs) to produce polypeptides. Three-nucleotide sequences in mRNA termed

codons specify binding to various tRNAs, each carrying single amino acids to be

positioned during protein synthesis. This system of mRNA translation also uses initiation

and termination codons to determine the precise peptide sequence. Termination codons

typically recruit proteins called release factors rather than a tRNA, thus dissociating the

ribosomal subunits and releasing the peptide chain. In discordance with the standard rules

of the system, the opal termination codon specified by UGA can be recognized by a

specific tRNA (tRNASec) (Diamond et al., 1981). However, in order for tRNASec to

efficiently recognize the in-frame UGA codon, the mRNA must additionally contain a

Sec insertion sequence (SECIS). The SECIS element is a stem-loop structure downstream

of the UGA codon in selenoprotein mRNAs that recruits proteins, which facilitate

recoding of the UGA and incorporation of Sec (Fig. 1) (Berry et al., 1991a).

Although selenoproteins are present in eukarya, archaea, and bacteria, the mechanism of

Sec incorporation differs between domains in several respects. For example, the bacterial

SECIS element is located immediately following the Sec UGA codon, whereas the

4

eukaryotic and archaeal SECIS elements are typically located much farther downstream

in the 3’ untranslated region (UTR) of the transcript. Additionally, the archaeal genus

Methanococcus has a 5’ UTR SECIS element, while eukaryotes have two distinct forms

of 3’ UTR SECIS elements (Kryukov et al., 2003, Kryukov & Gladyshev, 2004, Wilting

et al., 1997). An elongation factor specific to tRNASec has been identified as selB in

bacteria. In eukaryotes, the coordinated function of a SECIS binding protein (SBP2) and

a Sec-specific elongation factor (EFSec) allow the co-translational incorporation of Sec

instead of termination (Papp et al., 2007). Biosynthesis of Sec in eukarya and archaea is

accomplished by the sequential actions of O-phosphoseryl-tRNA kinase (PSTK) and O-

phosphoserine-tRNA:Sec-tRNA synthase (SEPSECS), which convert seryl-tRNASec to

phosphoseryl-tRNASec and ultimately to selenocysteinyl-tRNASec. Thus Sec formation

in archaea and eukarya is a two-step process, whereas bacterial selenocysteine synthase

(selA) can synthesize Sec directly from seryl-tRNASec. The Se donor utilized by both

SEPSECS and selA is selenophosphate, which is generated from selenide (H2Se) by

selenophosphate synthetase (SPS) enzymes. Selenophosphate is first transferred to O-

phosphoseryl-tRNASec by SEPSECS while displacing the phosphoseryl moiety, and

subsequent hydrolysis of the phosphate group yields Sec charged to its cognate tRNA

(Palioura et al., 2009). Bacteria have one SPS enzyme termed selD, while two (SPS1,

SPS2) have been identified in eukarya. Selenophosphate-synthetase 2 (SPS2) is a

eukaryotic selenoprotein that is required for the synthesis of all selenoproteins including

itself, and may provide feedback for global selenoprotein synthesis. Selenophosphate is

generated by SPS2 in the presence of selenide and ATP (Guimaraes et al., 1996). A

related protein called SPS1 contains a cysteine residue in place of Sec, but its

involvement in Sec and selenoprotein biosynthesis is uncertain (Low et al., 1995).

Intriguingly SPS2, but not SPS1, is required for selenoprotein synthesis in NIH3T3

mouse fibroblasts (Xu et al., 2007). Studies in Drosophila melanogaster, indicate that the

function of SPS1 is primarily in metabolism of vitamin B6, but whether this holds true

for vertebrates is unknown. It has alternatively been suggested that SPS2 assimilates

selenite, whereas SPS1 recycles Sec in a Se-salvage pathway (Tamura et al., 2004).

However an in vivo requirement of SPS2 and specificity of SPS1 and SPS2 with different

Se substrates have not been reported. Selenocysteine lyase (SCLY) is a putative Se

5

recycling enzyme, which is able to catalyze the hydrolysis of Sec into selenide and

alanine (Esaki et al., 1982). It facilitates Se incorporation into selenoproteins when Sec is

the Se source, and may deliver selenide to SPS enzymes for phosphorylation and

subsequent insertion into nascent selenoproteins (Kurokawa et al., 2011, Tobe et al.,

2009). Pyridoxal phosphate is a prosthetic group derived from vitamin B6 and is required

by enzymes involved in metabolism of amino acids, glucose, lipids and

neurotransmitters. The enzymatic activity of both SCLY and SEPSECS are pyridoxal

phosphate-dependent, implying that Sec metabolism shares regulatory elements with

standard amino acid metabolic pathways (Lacourciere et al., 2000, Mihara et al., 2000,

Palioura et al., 2009).

Several steps in selenoprotein synthesis are regulated by Se availability (dietary or

environmental depending on the organism), as well as by the local oxidation state. For

example in addition to being a limiting substrate for SPS2, Se levels affect different

SECIS elements differentially, potentially regulating selenoprotein synthesis efficiency at

the level of translation (Papp et al., 2007). Oxidation of the cysteine-rich, redox-sensitive

domain of SBP2 masks the nuclear-export signal (NES) causing importation into the

nucleus. Subsequently it is either sequestered there during high oxidative burden, or else

reduced by nuclear-specific isoforms of selenoproteins, unmasking the NES for binding

and exportation by the nuclear export receptor CRM-1 (Papp et al., 2006). Individual

selenoprotein expression also responds differentially to Se-availability and oxidative

stress, providing another level of regulatory control for selenoprotein synthesis. Thus

there is a complex, nonlinear interaction between Se-status and oxidative burden that

coordinates the synthesis of the numerous selenoproteins. Despite a low environmental

availability of Se and the various elaborate mechanisms for Sec incorporation,

selenoproteins are widespread in organisms, with certain plants and fungi being the only

major exceptions that lack selenoproteins and an essential biological function for Se

(Lobanov et al., 2009).

6

SELENOPROTEINS AS OXIDOREDUCTASE ENZYMES

The human genome codes for 25 selenoproteins, most of which have been identified

recently by bioinformatics approaches looking for SECIS elements downstream of in-

frame UGA codons (Kryukov et al., 2003). The various members display wide

subcellular and tissue distribution, and several are known to have multiple transcript

variants and protein isoforms. Functionally, selenoproteins are closely linked with the

cellular thiol-disulfide couples, particularly the glutathione (GSH) and thioredoxin (TXN)

couples. GSH is a tripeptide made of glycine, cysteine, and glutamate and is the most

abundant thiol in cells, present at millimolar concentrations. Oxidation of the cysteine

thiol links two molecules of GSH to form glutathione disulfide (GSSG). This reaction can

be spontaneous in the presence of electrophiles or alternatively can be catalyzed by a

number of enzymes that utilize GSH as an electron donor including glutathione

peroxidase (GPX), glutaredoxin, and glutathione S-transferase enzymes. Reduction of

GSSG, producing two molecules of GSH, is performed by the homodimeric flavoenzyme

glutathione reductase.

All of the characterized selenoproteins that function as enzymes are oxidoreductases that

catalyze thiol-disulfide oxidation-reduction (redox) reactions and contain Sec in the

active site. The first Se-dependent enzyme discovered was glutathione peroxidase 1

(GPX1), which catalyzes the reduction of hydrogen peroxide (H2O2) by oxidation of two

molecules of GSH to GSSG (Rotruck et al., 1973a). Cytosolic and mitochondrial forms

of the GPX1 enzyme are transcribed from the same gene containing one Sec-encoding

TGA. The enzyme functions as a homotetramer utilizing four Se atoms per active enzyme

(Esworthy et al., 1997). GPX1 is the most abundant glutathione peroxidase and the most

abundant selenoprotein in rats, representing a significant fraction of the total circulating

Se pool (Hawkes et al., 1985).

Four homologous GPX selenoproteins have subsequently been identified in humans.

GPX2, also known as gastrointestinal GPX, is a cytosolic enzyme that is specific to

epithelial cells and is abundant in the gut (Chu et al., 1997). The extracellular GPX3 has

7

broad substrate specificity and is found in most extracellular compartments but is

abundant in kidney and blood plasma (Olson et al., 2009). GPX6 is closely related to

GPX3 but is only expressed in the olfactory system, and exists as a cysteine homolog in

rodents (Kryukov et al., 2003). GPX4 is structurally and functionally different from the

other GPX enzymes because it is active as a monomer rather than a tetramer, and can

directly reduce membrane lipid hydroperoxides and free fatty acid hydroperoxides.

Alternative splicing and transcription initiation generates three distinct isoforms of GPX4

that localize to the cytosol, mitochondria, and nucleus (Maiorino et al., 2003).

Additionally, GPX4 translation is regulated by the mRNA binding proteins guanine-rich

sequence-binding factor 1 (Ufer et al., 2008) and PARK7/DJ-1 (Blackinton et al., 2009,

Van Der Brug et al., 2008). Genetic deletion of GPX4 in mice causes embryonic lethality

and knockdown of GPX4 in cells leads to rapid lipoxygenase-mediated lipid peroxidation

and subsequent apoptosis, suggesting that removal of lipid hydroperoxides by GPX4 is

essential for cell viability (Seiler et al., 2008, Yant et al., 2003). There are three

additional GPX enzymes (GPX5, GPX7, and GPX8) that are not selenoproteins in

humans.

TXN is a small protein of ~12 kDa and is present at concentrations several orders of

magnitude below GSH. It contains an active site dithiol that is highly conserved in

evolution and widely distributed among the TXN superfamily of proteins, which includes

protein disulfide isomerase enzymes. Through oxidation of the dithiol to a disulfide, TXN

can directly reduce cysteine sulfenic acids and control the state of dithiol-disulfide motifs

in target proteins, and can also serve as an electron donating cofactor for enzymes such as

ribonucleotide reductase, peroxiredoxins and methionine sulfoxide reductases (Arner &

Holmgren, 2000). In turn, reducing oxidized TXN is mediated exclusively by the

thioredoxin reductase (TXNRD) family of selenoproteins (Zhong et al., 2000). At least

four selenoproteins (GPX3, GPX4, SEPP1, SEPX1) can utilize TXN as a cofactor for

enzymatic reduction, and it is possible that others do as well (Bjornstedt et al., 1994,

Takebe et al., 2002). There are numerous thioredoxin-like proteins that may depend on

TXN or act in parallel to provide additional substrate specificity beyond that provided by

TXN. Several selenoproteins contain a TXN-like fold, which is a well described

8

secondary/tertiary structure pattern with a conserved Cys-X-X-Cys or Cys-X-X-Ser/Thr

active-site motif characteristic of oxidoreductases (Dikiy et al., 2007), where X is any

amino acid. It is tempting to speculate that these selenoproteins operate similarly to TXN,

namely by controlling the redox state of cysteine residues and dithiol motifs, but there is

little evidence to support or deny this notion at present.

Three mammalian thioredoxin reductases are selenoenzymes encoded by individual

genes. TXNRD1, TXNRD2, and TXNRD3 encode homodimeric flavoproteins that

localize to the cytosol, mitochondria, and testes respectively. They are members of the

pyridine nucleotide-disulfide oxidoreductase family and contain two redox-sensitive sites

in the N- and C-terminus that interact in the head to tail dimer conformation of the active

enzyme. These enzymes are capable of reducing a number of substrates, but depend on

NADPH for donating electrons, which are first transferred to the FAD group, then passed

to the N-terminal dithiol of one subunit and subsequently to the C-terminal selenenyl-

sulfide of the other subunit. The highly conserved Sec-containing C-terminal motif is

absolutely critical for catalytic function of TXNRD enzymes (Arner & Holmgren, 2000,

Hatfield, 2006). The main substrate for TXNRDs is the small redox-sensitive protein

TXN, which is integral to physiological processes such as cell communication,

metabolism, proliferation, and apoptosis. In general, the reactive dithiol of TXN will

become oxidized to a disulfide during reduction of an oxidized target protein.

Regeneration of reduced TXN proteins requires TXNRD, and thus the TXN/TXNRD

system is completely dependent on Se in mammals. The importance of this system is

highlighted by the fact that knockout of either TXNRD1 or TXNRD2 is embryonic lethal

in mice (Conrad et al., 2004, Jakupoglu et al., 2005). It is worth noting that TXNRDs

from mammals differ from the Se-independent enzymes of archaea, bacteria, yeast, and

plants.

Reactions between target proteins and TXN can be spontaneous, but several enzymes can

catalyze the reduction of target proteins using TXN as an electron donating cofactor. The

human genome codes for four Methionine Sulfoxide Reductase (MSR) enzymes that

reduce oxidized methionine residues in proteins utilizing TXN as a cofactor. There is

9

now considerable evidence that like cysteine, reversible methionine oxidation can

regulate protein function (Levine et al., 2000). For example, calcium/calmodulin-

dependent protein kinase II and the phosphatase calcineurin, among many other proteins,

are regulated by methionine redox status (Agbas & Moskovitz, 2009, Erickson et al.,

2008). A single MSRA and three MSRB enzymes stereo-specifically reduce S- and R-

sulfoxidated methionines respectively. MSRB1 is a selenoprotein also known as

Selenoprotein R and SEPX1, while MSRA as well as MSRB2 and MSRB3 are Se-

independent enzymes. SEPX1 is a zinc-containing protein present in the cytosol and

nucleus and exhibits the highest methionine-R-sulfoxide reductase activity because of the

presence of Sec in its active site (Kim & Gladyshev, 2004). Interestingly, redox status of

the cysteine-rich metallothionein/thionein couple dictates zinc loading in that reduced

thionein binds zinc and oxidation of metallothionein releases it. Moreover, reduction of

non-selenoprotein MSRB3 by TXN, TXNRD, and NADPH is more efficient in the

presence of thionein (Sagher et al., 2006). Therefore regulation of specific kinases,

phosphatases, and other proteins by methionine-R-sulfoxide reduction is mediated by two

selenoproteins (MSRB1, TXNRD1) and NADPH.

SELENOPROTEINS AND METABOLISM

Specific selenoproteins function at the intersection of cellular and organism metabolism

by modulating insulin and thyroid hormone signaling. The iodothyronine deiodinases

(DIO) function in activation and deactivation of thyroid hormone and were the second

family of enzymes determined to be Sec-containing selenoenzymes (Berry et al., 1991b).

Thyroid hormone metabolism both at the level of production in the thyroid and local

hormone activity in the periphery is reliant on the DIO family of selenoenzymes. Most

vertebrates have three DIO enzymes that can deiodinate thyroid hormones to control local

availability. These integral membrane protein enzymes are thiol-requiring

oxidoreductases that remove iodine atoms from the aromatic rings of thyroxine (T4),

triiodothyronine (T3), and reverse triiodothyronine (rT3) (Bianco & Kim, 2006). DIO1 is

a plasma membrane protein found mainly in cells of the liver and kidney, is capable of

deiodinating both the inner and outer rings, and produces most of the circulating T3.

10

DIO2 is found in the endoplasmic reticulum of cells in several tissues including the

thyroid, heart, skeletal muscle, fat, and the central nervous system and selectively

removes the outer ring iodine, making it the primary tissue activator of thyroid hormone

by converting T4 to T3. DIO3 is also a plasma membrane protein, however it is mainly

found in fetal tissue and the placenta, selectively removes the inner ring iodine, and thus

contributes to thyroid hormone inactivation.

A specific role for a selenoprotein in redox regulation of insulin signaling was established

when it was found that overexpression of GPX1 causes hyperinsulinemia and insulin

resistance in mice (Mcclung et al., 2004). Moreover genetic deletion of GPX1 promotes

glucose tolerance and insulin sensitivity in mice on a high-fat diet by enhancement of

insulin-induced PI3K/Akt signaling (Loh et al., 2009). Dietary studies in humans have

further suggested that supranutritional levels of Se are associated with type II diabetes,

while animal studies confirm that both excessive dietary Se and GPX1 overexpression

lead to hyperinsulinemia and insulin resistance (Labunskyy et al., 2011, Lippman et al.,

2009, Stranges et al., 2007, Wang et al., 2008). Peroxide-induced oxidation of PTEN and

S-glutathionylation of protein tyrosine phosphatase 1B are affected by GPX1 activity,

which thereby modulates insulin receptor activation and insulin resistance (Mueller et al.,

2009).

Selenoprotein P (SEPP1) is a unique selenoprotein that contains multiple Sec residues

and is also implicated in insulin resistance. Diabetic patients display an increase in

hepatic SEPP1 mRNA and serum SEPP1 protein, and purified SEPP1 administered to

mice is able to induce insulin resistance and glucose intolerance. Furthermore,

knockdown or knockout of SEPP1 in mice improves glucose tolerance and insulin

sensitivity, and SEPP1 knockout mice are protected against glucose intolerance and

insulin resistance even when on an obesity-inducing diet (Misu et al., 2010). Since

SEPP1 expression can dictate expression of other selenoproteins including GPX1, its

effect on insulin resistance may be direct or indirect. Primate and rodent SEPP1 contains

up to ten Sec residues while as many as 17 Sec residues are present in zebrafish. SEPP1

is the most abundant selenoprotein in blood and accounts for as much as 65% of plasma

11

Se in rats (Burk & Hill, 1994). The abundant Sec residues of SEPP1 are divided into two

regions with the bulk being located in the C-terminal domain that is required for the Se

transport function. SEPP1 also contains an N-terminus Cys-X-X-Sec motif and catalyzes

the reduction of lipid hydroperoxides in vitro utilizing TXN as a cofactor (Saito et al.,

1999, Takebe et al., 2002). Aside from the high Se content in the form of Sec residues,

SEPP1 is distinct in that it is one of two extracellular selenoproteins (the other being

GPX3), and the carboxy-terminal Se transport domain appears to be a metazoan

adaptation (Lobanov et al., 2008). SEPP1 is abundantly produced by the liver and

secreted into blood, however local production and secretion in nearly all tissue systems

has been described (Burk & Hill, 2005). Bodily transport of Se to extrahepatic tissues,

particularly the brain, testes and kidneys, is facilitated by receptor-mediated uptake of

SEPP1 by the low-density lipoprotein receptor family members ApoER2 (LRP8) and

Megalin (LRP2) (Burk & Hill, 2009). In addition to a Se transport function and

peroxidase activity, SEPP1 exhibits pH-dependent heparin binding and heavy metal

binding that likely also function in redox-dependent processes.

In our lab, selenocysteine lyase (SCLY) has also shown to be involved in metabolic

syndrome in an in vivo mouse model. Although SCLY is not itself a selenoprotein, it is an

enzyme with high specific activity for Sec, and liberates selenide (Mihara et al., 2000).

Since selenoproteins are a reservoir of Sec and SEPP1 supplies Se in the form of Sec,

SCLY is a potentially relevant factor in the global distribution of Se and the expression of

all selenoproteins. Transgenic mice lacking SCLY were characterized and found to have

altered energy metabolism in the liver. When given adequate Se, the animals exhibit

hyperinsulinemia and mild hepatic steatosis, along with an increase in blood SEPP1.

Upon dietary Se restriction, the animals develop obesity, fatty liver, hypercholesteremia,

and insulin resistance (Seale, et al. 2012, unpublished). The results of studies on mice

with disrupted GPX1, SEPP1, and SCLY do not reconcile with a simple explanation.

Instead, they suggest a nonlinear, dynamic explanation. Regardless, these data

demonstrate the importance of proper Se handling in vertebrate carbohydrate and lipid

metabolism. Se utilization is tightly integrated with insulin and thyroid hormone

signaling, thus disruption of selenoproteins and Sec-related proteins alters vertebrate

12

metabolic systems. A connection between Se and metabolism in multicellular organisms

suggests that cell-autonomous regulation of redox systems and signaling may similarly

depend on one or more selenoproteins.

THE SELENOPROTEIN FAMILY

Several selenoproteins with uncertain functions could have a role in regulating target

protein oxidation state. The regulatory control of cellular redox signaling by

Selenoprotein W (SEPW1) will be discussed next as an example, bearing in mind that

selenoproteins with unknown roles can impact the cellular response to environmental

changes, particularly in relation to growth and stress. Following the section on SEPW1, a

brief summary of what is known of the remaining selenoproteins is presented, with the

biological functions described possibly owing to the activity of the selenoproteins in

undefined redox circuits.

SEPW1 was purified in the early 1990s but putatively identified much earlier due to its

absence in Se-deficient lambs suffering a myopathy called White Muscle Disease

(Vendeland et al., 1993). Mammalian SEPW1 is a highly conserved cytosolic protein of

just 87 amino acids, and SEPW1 orthologs are among the most widely distributed

selenoproteins in all species including prokaryotes (Kryukov & Gladyshev, 2004, Zhang

et al., 2005). The expression level of SEPW1 in vertebrates is very sensitive to dietary Se

intake as well as the expression level of SEPP1 (Hoffmann et al., 2007, Vendeland et al.,

1995, Yeh et al., 1995). Abundant SEPW1 expression is observed in muscle, and SEPW1

transcription during myocyte differentiation is maintained by binding of the myogenic

transcription factor MyoD to the SEPW1 promoter (Noh et al., 2010). A putative metal-

response element in the promoter of the SEPW1 gene was probed in vitro using a

luciferase reporter fusion construct, and luciferase specific activity was found to be

stimulated by copper and zinc, but not cadmium (Amantana et al., 2002). Although a

bona fide enzymatic activity has not been attributed to SEPW1, the presence of a Cys-X-

X-Sec motif in a thioredoxin-like fold may indicate thioredoxin-like redox activity (Dikiy

et al., 2007).

13

Recently, SEPW1 was shown to pull-down and co-immunoprecipitate with the beta and

gamma isoforms of 14-3-3 protein. This interaction was further confirmed by NMR

spectroscopy, and extended to identify three loops of SEPW1 that interact with 14-3-3

proteins (Aachmann et al., 2007). 14-3-3 beta and gamma proteins are scaffolding

proteins derived from the YWHAB and YWHAG genes respectively, and bind a diverse

assortment of proteins including kinases, phosphatases, and receptors. In this way 14-3-3

proteins coordinate molecular interactions and participate in cell cycle regulation,

metabolism, apoptosis, protein trafficking and gene transcription (Fu et al., 2000). A

computational study of SEPW1/14-3-3 interaction suggests that a conserved cysteine of

14-3-3 beta and gamma (Cys191 and Cys195 respectively) can be reversibly oxidized,

with SEPW1 acting as a reducing agent (Musiani et al., 2010). The oxidized cysteine

sulfenic acid of 14-3-3 can putatively react with Sec of SEPW1, producing a mixed

complex. Subsequently the formation of an intramolecular selenenyl-sulfide within

SEPW1 would result in 14-3-3 being fully reduced. Oxidized SEPW1 can then migrate

away and likely be reduced to its parent state by GSH. This speculated reduction is

supported by evidence that a SEPW1 cysteine residue conserved from rodents to primates

can be S-glutathionylated (Beilstein et al., 1996, Gu et al., 1999).

Redox regulation of 14-3-3 proteins by SEPW1 could serve several cellular functions, but

an intriguing possibility is presented by the in vitro finding that SEPW1 expression is

regulated by the cell-cycle, and knockdown of SEPW1 induces p53-dependent cell-cycle

arrest (Hawkes et al., 2009). Recently, it has been demonstrated that growth factor-

induced receptor tyrosine kinase phosphorylation, and downstream JNK and p38 MAPK

signaling, leading to cell proliferation requires SEPW1 (Hawkes, WC; personal

communication). Therefore SEPW1, through redox regulation of 14-3-3 proteins, may

coordinate Se availability and oxidative burden with cellular proliferation, differentiation

and death via the p53 pathway. Deficiency of SEPW1-mediated redox functionality may

serve as the basis for the myopathies in livestock (Whanger, 2009). Interestingly SEPW1

is also associated with multiple myeloma in humans, where SEPW1 overactivity may

promote malignancy and the overangiogenic phenotype of endothelial cells in active

14

disease (Ria et al., 2009). These studies further suggest a pivotal role for SEPW1 and Se

in regulatory control of the cell cycle.

A subset of selenoproteins is observed in mitochondrion to combat against electron leak

during oxidative respiration and phosphorylation. Mitochondria-specific isoforms of

GPX1, GPX4, and TXNRD2 regulate peroxide metabolism and oxidative tone within this

organelle. The selenoproteins GPX4, TXNRD1, MSRB1, and SELH have been shown to

exhibit varying degrees of nuclear localization (Fig. 2). Apart from SELH, the other three

selenoproteins are presumably involved in reduction of lipid peroxides, oxidized TXN,

and sulfoxidized methionine residues within the nuclear envelope.

SELH is the only DNA-binding selenoprotein described and has a role in regulation of

gene expression. Similar to SEPW1, Selenoprotein H (SELH) is a small selenoprotein

that is highly expressed during development and is sensitive to dietary Se intake (Kipp et

al., 2009, Novoselov et al., 2007). Like several selenoproteins it contains a Cys-X-X-Sec

sequence within a thioredoxin-like fold, but unlike any other selenoprotein described to

date, it is a DNA-binding protein of the AT-hook family. SELH is primarily located in

the nucleus and is implicated in redox-sensitive transcription of genes whose products are

involved in de novo glutathione synthesis and phase II detoxification (Panee et al., 2007).

Multiple metal-response elements are present in the SELH gene (Stoytcheva et al., 2010),

and one group has confirmed in vivo that SELH mRNA and protein are upregulated under

conditions of elevated copper in mouse liver (Burkhead et al., 2010). Although it is a

nuclear protein, mitochondrial biogenesis and function are also linked with SELH.

Overexpression of SELH in a transformed neuronal cell line attenuates the UVB-induced

increase of p53 protein and caspase-mediated apoptosis (Mendelev et al., 2009).

Additionally SELH overexpression increases mitochondrial size, cytochrome c content,

and expression of mitochondrial biogenesis proteins while boosting respiration

(Mendelev et al., 2011). Collectively these findings suggest that SELH is a Se- and

metal-regulated selenoprotein that is able to transduce oxidant signals by modulating

gene expression in conjunction with other redox-sensitive transcription factors. Further

15

investigation is warranted to determine if SELH modifies cysteine S-glutathionylation or

disulfide formation in target proteins such as p53 to regulate gene expression.

The endoplasmic reticulum (ER) regulates the synthesis, folding, and transport of

proteins, and additionally constitutes the main intracellular store for calcium ions, which

are integral in cell signaling. Seven selenoproteins are enriched in the ER (Fig. 2) and

some are postulated to have a role in protein folding and ER calcium handling, since

oxidative mechanisms within the ER are known to regulate these processes (Shchedrina

et al., 2010). The ER is a relatively oxidizing environment compared to other intracellular

organelles and contains oxidase enzymes to facilitate the formation of disulfide bonds in

proteins destined for export. Simultaneously GSH and TXN system components are

transported into the ER, providing both oxidation and reduction mechanisms for dynamic

redox regulation associated with protein processing and secretion. Redox state affects

calcium homeostasis by modulating ER calcium channels and chaperones, and oxidative

stress and ER-stress are intimately related in signaling for apoptosis (Gorlach et al.,

2006). Similar to the ER, the secretory and endosomal/lysosomal pathways are also more

oxidized than other subcellular compartments (Austin et al., 2005, Hwang et al., 1992).

Ligand binding to various receptors stimulates endocytosis of redox-active endosomes

whose luminal redox activity directs spatiotemporally-regulated signaling and prevents

nonspecific redox reactions (Li et al., 2006, Oakley et al., 2009). Therefore redox-

mediated processes are vital for secretory and endocytic function, and the ability of

selenoproteins to transmit oxidative signals from reactive intermediates to disulfide bonds

or exposed thiols of target proteins may help to explain the enrichment of selenoproteins

in the ER.

Selenoprotein T (SELT) is an ER- and Golgi-localized selenoprotein that is ubiquitously

expressed from development through adulthood, and shares some sequence similarity

with SEPW1 and SELH including the thioredoxin-like fold containing a Cys-X-X-Sec

motif (Dikiy et al., 2007). Deficiency of SELT in murine fibroblasts causes an

upregulation of SEPW1, in addition to altering cell adhesion and redox regulation

(Sengupta et al., 2009). A biological role for SELT in neuroendocrine secretion and

16

calcium mobilization in vitro has also been presented. SELT was identified as a target

gene of the neuropeptide pituitary adenylate cyclase-activating polypeptide (PACAP),

and manipulation of SELT expression altered PACAP-induced intracellular Ca2+

changes

and growth hormone secretion (Grumolato et al., 2008). Furthermore, PACAP regulation

of SELT is associated with ontogenesis, tissue maturation and regeneration of nervous,

endocrine, and metabolic tissues (Tanguy et al., 2011).

Two of the more abundant and ubiquitous ER selenoproteins, Selenoprotein M (SELM)

and 15-kDa selenoprotein (SEP15), are both 15-kDa proteins that share 31% sequence

homology and have Cys-X-X-Sec and Cys-X-Sec redox motifs respectively. SELM and

SEP15 have thioredoxin-like tertiary structure and homology to protein disulfide

isomerases that suggest oxidoreductase activity, but direct evidence in support of this

notion is lacking (Ferguson et al., 2006). SEP15 associates with UDP-

glucose:glycoprotein glucosyltransferase (UGTR), a protein involved in protein

conformation quality control, and has been suggested to facilitate proper protein folding

(Labunskyy et al., 2005). SEP15 knockout mice have been generated, and the only major

phenotype found was a predispostion to cataract formation that was suspected to be due

to improper folding of lens proteins. A role for SELM in protein folding has been

proposed, and recent evidence suggests that SELM may also be involved in regulating the

flux of calcium ions. Overexpression of SELM in a neuronal cell line in vitro reduced

peroxide-induced calcium influx, whereas knockdown of SELM increased the baseline

intracellular calcium concentration (Reeves et al., 2010).

Selenoprotein N (SEPN1) is a large (70 kDa) single-spanning transmembrane protein

localized to the ER membrane with two known isoforms generated by alternative splicing

of exon 3 (Moghadaszadeh et al., 2001). Several congenital muscular dystrophy

syndromes such as multiminicore disease, rigid spine muscular dystrophy, and desmin-

related myopathy with Mallory body-like inclusions are directly associated with

mutations in the SEPN1 gene and have been classified as SEPN1-related myopathies

(Ferreiro et al., 2004). Interestingly, mutations in SEPN1 leading to congenital fiber-type

disproportion are associated with insulin resistance (Clarke et al., 2006). To date, SEPN1

17

is the only selenoprotein gene in which mutations are directly and causatively linked to

human disease. An in vivo study using zebrafish determined that SEPN1 associates with

ER/SR ryanodine receptors, and that this interaction is necessary for the calcium-induced

release of calcium from intracellular stores (Jurynec et al., 2008). Ryanodine receptor

channels are homotetramers with several redox-regulated cysteine residues and SEPN1

contains a Cys-Sec-Gly-Ser motif, suggesting that SEPN1 regulates ryanodine receptor-

mediated calcium flux in muscle by redox-dependent signaling. Finally, SEPN1 is

integral to the generation and/or maintenance of skeletal muscle satellite cells, which are

an adult stem cell population involved in muscle growth and regeneration (Castets et al.,

2011).

Selenoprotein K (SELK) and Selenoprotein S (SELS) are also predominantly ER-

localized single-spanning transmembrane proteins, however they are much smaller than

SEPN1, and also show some localization to the plasma membrane (Chen et al., 2006,

Kryukov et al., 2003, Ye et al., 2004). Both are widely expressed in a variety of tissues

and have been implicated in the cellular response to ER-stress. Specifically, ER-stress

agents regulated the expression of SELK in HepG2 hepatoma cells, and knockdown of

SELK exacerbated cell death when challenged with ER-stress (Du et al., 2010). Genetic

deletion of SELK in mice decreases receptor-mediated calcium flux in immune cells,

impairs calcium-dependent immune cell function, and increases West Nile virus-induced

lethality (Verma et al., 2011). An interesting link between metabolism and inflammation

is presented in the case of SELS, which was originally identified as a glucose-regulated

protein in a rodent model of diabetes (Walder et al., 2002). The relationship between

SELS and type II diabetes was confirmed in humans, and there is evidence that SELS can

be secreted from liver and identified in blood sera where it associates with LDL (Gao et

al., 2007, Karlsson et al., 2004). SELS is now known to also be regulated by

inflammatory cytokines (Gao et al., 2006) and reciprocally, reduced expression of SELS,

due to polymorphisms in the gene promoter, influences the levels of IL-1, TNFα, and IL-

6 (Curran et al., 2005). SELS participates in removal of misfolded proteins from the ER

lumen (Ye et al., 2005, Ye et al., 2004) and was demonstrated to prevent ER-stress and

have anti-apoptotic function in macrophages and astrocytes (Fradejas et al., 2008, Kim et

18

al., 2007). Further work in mice indicates that in brain, SELS is mainly expressed

neuronally under basal conditions, but is intensely upregulated in reactive astrocytes

following brain injury (Fradejas et al., 2011).

Three selenoproteins remain largely unexplored with very little published data currently

available. The sequences of selenoprotein I (SELI), selenoprotein O (SELO), and

selenoprotein V (SELV) were identified in the human genome several years ago,

however almost no information is available on the cellular localizations or physiological

functions of these selenoproteins. SELI mRNA is known to be expressed in several

tissues, and is a putative transmembrane protein hypothesized to function in phospholipid

biosynthesis based on the presence of a CDP-alcohol phosphatidyltransferase motif that

is conserved in phospholipid synthases (Horibata & Hirabayashi, 2007, Kryukov et al.,

2003). SELO is predicted to be a 669 amino-acid selenoprotein containing a Cys-X-X-

Sec motif, but experimental data demonstrating a redox function is unavailable (Kryukov

et al., 2003). SELV appears to be a testes-restricted protein with a predicted thioredoxin-

like fold housing a Cys-X-X-Sec motif, and also has some sequence homology with

SEPW1, SELH, and SELT (Kryukov et al., 2003).

SELENOPROTEINS IN THE NERVOUS SYSTEM

The nervous system has a unique relationship with Se. This is most clearly demonstrated

by the fact that dietary Se deprivation causes a much greater drop in peripheral tissue Se

concentration than in brain Se concentration. In other words, the brain preferentially

retains Se compared to other organs during times of deficiency, suggesting that it has

some essential function in brain (Chen & Berry, 2003). Additionally, nearly all

selenoproteins are expressed in the brain, and neurons appear to be the major functional

sites of selenoprotein expression (Zhang et al., 2008).

This organ-specific prioritization of Se to the brain is at least partially explained by

receptor-mediated uptake of SEPP1 in a tissue specific manner. This unique

selenoprotein is only found in metazoans, and has high Se content with as many as 28

19

Sec residues in the sea urchin Strongylocentrotus purpuratus (Lobanov et al., 2008). The

C-terminal region of SEPP1, containing nine of the ten Sec residues in primates and

rodents, is involved in maintaining stable brain Se concentration during dietary variation.

SEPP1 is thought to be primarily produced in the liver and secreted to blood to transport

Se, however hepatic SEPP1 deficiency does not alter brain Se levels (Schweizer et al.,

2005). Rather, Se levels in brain, particularly hippocampus, are lowered by genetic

deletion of full-length SEPP1 or the C-terminus (Hill et al., 2007, Nakayama et al.,

2007). These mutations cause brain Se concentration to drop by a greater extent than can

be achieved by dietary Se deficiency, providing further evidence that SEPP1 facilitates a

steady supply of Se to the brain.

Mice lacking SEPP1 develop sensory, motor, and cognitive neurological impairment

including spasticity, hyperreflexia, gait disruption, and a spatial learning deficit. These

animals also display widespread neurodegeneration which, along with the behavioral

phenotype, is modulated by dietary Se status (Burk & Hill, 2009). However, Se-

supplemented SEPP1-deficient mice still exhibit a profound deficit in synaptic long term

potentiation (Peters et al., 2006). This suggests that beyond Se delivery, SEPP1 may

function in cell signaling, which is established for the SEPP1 receptor ApoER2.

ApoER2-deficient mice develop a phenotype similar to but less severe than SEPP1

knockout mice in terms of brain Se concentration and neurological impairment (Burk et

al., 2007). Alternatively, SEPP1 could be a preferred Se source within brain that is not

readily compensated for by dietary sources. Indeed Se supply to cultured Jurkat cells,

assessed by stimulation of GPX activity, is 5-100 times more efficient with SEPP1 than

other Se-containing proteins and compounds (Saito & Takahashi, 2002).

ApoER2 is a member of the low-density lipoprotein receptor-related protein family that is

highly expressed in brain. It exists in multiple isoforms, and has been observed in all

brain cell types (Fan et al., 2001, Korschineck et al., 2001). Expression at some

excitatory synapses, where it is associated with the postsynaptic densities, can cause

formation of a functional complex with N-methyl-D-aspartate (NMDA) receptors

(Beffert et al., 2005). This association allows ApoER2 to modulate NMDA receptor

20

activity, synaptic neurotransmission, and memory in mice. Although these effects have

been demonstrated using a different ApoER2 ligand, Reelin, alterations in synaptic

plasticity and memory in SEPP1-deficient mice suggest that both ligands can actively

modulate synaptic function.

The central nervous system is protected from drugs, peptides, and other substances in the

peripheral circulation by the blood-brain-barrier (BBB). It is chiefly composed of tight

junctions between endothelial cells of the cerebral vasculature. Additionally, the vast

majority of cerebral vessels are lined by astrocytic processes termed end-feet. This

anatomical microarchitecture provides a high trans-endothelial electrical resistance, and

prohibits passage of the vast majority of circulating blood components, except for small,

lipophilic molecules. Since blood constituents, chiefly oxygen and glucose, are absolutely

essential for nervous system function, the BBB contains an array of transmembrane

proteins that transport required substrates (Zlokovic, 2008).

Before the knowledge of particular receptors for SEPP1, endothelial cells of the cerebral

vasculature were observed to bind high quantities of this protein (Burk et al., 1997).

Recently, ApoER2 was identified as a specific component of the BBB based on its highly

enriched mRNA expression in mouse brain microvascular endothelial cells compared to

liver and lung endothelial cells (Daneman et al., 2010). However at the BBB, ApoER2 is

apparently present on the abluminal or basolateral side of endothelial cells, in between

the capillary lumen and the astrocytic endfeet (Elali & Hermann, 2010). In other words

ApoER2 is on the brain side of the cerebral microvascular endothelium. Thus, uptake of

SEPP1 within brain by endothelial ApoER2 may retain Se in brain. However, transfer of

SEPP1 across the BBB may involve additional receptors, such as Megalin. Megalin is the

primary SEPP1 receptor in kidney and is responsible for reabsorption of SEPP1 from the

glomerular filtrate by proximal convoluted tubule epithelial cells (Olson et al., 2008).

Interestingly, Megalin facilitates transcytosis of apolipoprotein J from blood to brain at

the cerebral vascular endothelium and the choroid plexus epithelium (Zlokovic et al.,

1996). Furthermore apolipoprotein J, a.k.a. Clusterin, associates with SEPP1 in high

molecular weight complexes in plasma, and Megalin-mutant mice have reduced Se in

21

brain and the periphery (Cheung et al., 2010, Chiu-Ugalde et al., 2010). Therefore, the

two receptors may work in tandem to take up blood SEPP1 and/or retain central SEPP1

depending on local demand, physiological activity, and developmental stage. Since

transgenic mice lacking SEPP1 or its receptors have some Se in brain, complementary

mechanisms affecting brain and whole body Se distribution are additionally inferred to

exist.

The choroid plexus is a heavily vascularized cuboidal epithelium protruding from the

ependymal walls of the cerebral ventricles. This structure produces cerebrospinal fluid

(CSF), the extracellular fluid bathing the nervous system, essentially by filtering blood.

SEPP1 is found in CSF, and choroid plexus highly expresses the mRNA, implying that it

is secreted (Scharpf et al., 2007). The ventricular regions are known to have high Se by

autoradiography, positioning the choroid plexus and CSF as primary sites of brain Se

homeostasis by virtue of SEPP1 expression and secretion (Kuhbacher et al., 2009).

Choroid plexus also expresses abundant mRNA for nearly all other selenoproteins, the

synthesis factors, the SEPP1 receptors ApoER2 and Megalin, and SCLY, further

suggesting a role in transfer of Se from peripheral blood to the central nervous system

(Lein et al., 2007).

The role of SCLY in the brain has received little attention. Although SEPP1 is

established in promoting brain Se concentration and selenoprotein expression, a complete

mechanism is lacking. This is because SEPP1 supplies Se in the form of Sec residues,

while selenoprotein synthesis is thought to require selenide to generate selenophosphate

and Sec co-translationally. The enzymatic activity of SCLY generates selenide from Sec,

and the protein has been suggested to associate with SPS enzymes to recycle Se from Sec

in support of selenoprotein synthesis (Tobe et al., 2009). SCLY protein and activity is

found in mouse brain, albeit at much lower levels than in kidney and liver (Mihara et al.,

2000). Curiously, the testes, which also has a blood-tissue barrier and relies on SEPP1 for

Se, show a pattern whereby cells expressing ApoER2 do not express SCLY and vice

versa (Kurokawa et al., 2011). Specifically, cells exposed to circulating SEPP1 express

ApoER2, whereas germ cells expressing other selenoproteins express SCLY. This

22

suggests that uptake of SEPP1 and utilization of the Sec residues may depend on multiple

interacting cell types, rather than cell autonomous utilization. In brain, where neurons,

astrocytes, oligodendrocytes, microglia and endothelial cells interact in a complex

system, Se distribution between cells and extracellular compartments may similarly occur

by cell-type and -context dependent expression of SEPP1, ApoER2, and SCLY. However

the exact nature of this system, both in body and brain, has not been fully elucidated.

23

FIGURE LEGENDS

Figure 1: Assimilation of Se into selenoproteins. Dietary forms of Se in vertebrates

include selenate, selenite, selenocysteine, and selenomethionine and are highlighted in

green. Dietary forms are converted to the intermediate metabolite selenide for

selenoprotein synthesis or Se excretion. Several factors are required to synthesize Sec on

its tRNA, and position it at UGA codons during translation.

Figure 2: Selenoprotein localization to subcellular domains. A partial list of mammalian

selenoproteins is depicted schematically in a prototypical cell. GPX isoforms are

distributed in the cytosolic, mitochondrial, nuclear, and extracellular compartments.

Seven selenoproteins display localization to the endoplasmic reticulum, and other

selenoproteins are also shown. Several steps involved in cellular signal transduction are

known to be redox-sensitive, and may be functionally affected by selenoproteins.

24

FIGURE 1

25

FIGURE 2

26

REFERENCES

Aachmann, F.L., Fomenko, D.E., Soragni, A., Gladyshev, V.N. & Dikiy, A. (2007)

Solution structure of selenoprotein W and NMR analysis of its interaction with

14-3-3 proteins. J Biol Chem, 282, 37036-37044.

Agbas, A. & Moskovitz, J. (2009) The Role of Methionine Oxidation/Reduction in the

Regulation of Immune Response. Curr Signal Transduct Ther, 4, 46-50.

Amantana, A., Vorachek, W.R., Butler, J.A., Costa, N.D. & Whanger, P.D. (2002) Effect

of copper, zinc and cadmium on the promoter of selenoprotein W in glial and

myoblast cells. J Inorg Biochem, 91, 356-362.

Arner, E.S. & Holmgren, A. (2000) Physiological functions of thioredoxin and

thioredoxin reductase. Eur J Biochem, 267, 6102-6109.

Austin, C.D., Wen, X., Gazzard, L., Nelson, C., Scheller, R.H. & Scales, S.J. (2005)

Oxidizing potential of endosomes and lysosomes limits intracellular cleavage of

disulfide-based antibody-drug conjugates. Proceedings of the National Academy

of Sciences of the United States of America, 102, 17987-17992.

Beffert, U., Weeber, E.J., Durudas, A., Qiu, S., Masiulis, I., Sweatt, J.D., Li, W.P.,

Adelmann, G., Frotscher, M., Hammer, R.E. & Herz, J. (2005) Modulation of

synaptic plasticity and memory by Reelin involves differential splicing of the

lipoprotein receptor Apoer2. Neuron, 47, 567-579.

Behne, D., Kyriakopoulos, A., Meinhold, H. & Kohrle, J. (1990) Identification of type I

iodothyronine 5'-deiodinase as a selenoenzyme. Biochemical and biophysical

research communications, 173, 1143-1149.

Beilstein, M.A., Vendeland, S.C., Barofsky, E., Jensen, O.N. & Whanger, P.D. (1996)

Selenoprotein W of rat muscle binds glutathione and an unknown small molecular

weight moiety. J Inorg Biochem, 61, 117-124.

Berry, M.J., Banu, L., Chen, Y.Y., Mandel, S.J., Kieffer, J.D., Harney, J.W. & Larsen,

P.R. (1991a) Recognition of UGA as a selenocysteine codon in type I deiodinase

requires sequences in the 3' untranslated region. Nature, 353, 273-276.

Berry, M.J., Banu, L. & Larsen, P.R. (1991b) Type I iodothyronine deiodinase is a

selenocysteine-containing enzyme. Nature, 349, 438-440.

Berry, M.J., Maia, A.L., Kieffer, J.D., Harney, J.W. & Larsen, P.R. (1992) Substitution

of cysteine for selenocysteine in type I iodothyronine deiodinase reduces the

catalytic efficiency of the protein but enhances its translation. Endocrinology,

131, 1848-1852.

Bianco, A.C. & Kim, B.W. (2006) Deiodinases: implications of the local control of

thyroid hormone action. J Clin Invest, 116, 2571-2579.

Bjornstedt, M., Xue, J., Huang, W., Akesson, B. & Holmgren, A. (1994) The thioredoxin

and glutaredoxin systems are efficient electron donors to human plasma

glutathione peroxidase. J Biol Chem, 269, 29382-29384.

Blackinton, J., Kumaran, R., van der Brug, M.P., Ahmad, R., Olson, L., Galter, D., Lees,

A., Bandopadhyay, R. & Cookson, M.R. (2009) Post-transcriptional regulation of

mRNA associated with DJ-1 in sporadic Parkinson disease. Neuroscience letters,

452, 8-11.

27

Burk, R.F. & Hill, K.E. (1994) Selenoprotein P. A selenium-rich extracellular

glycoprotein. J Nutr, 124, 1891-1897.

Burk, R.F. & Hill, K.E. (2005) Selenoprotein P: an extracellular protein with unique

physical characteristics and a role in selenium homeostasis. Annu Rev Nutr, 25,

215-235.

Burk, R.F. & Hill, K.E. (2009) Selenoprotein P-expression, functions, and roles in

mammals. Biochim Biophys Acta, 1790, 1441-1447.

Burk, R.F., Hill, K.E., Boeglin, M.E., Ebner, F.F. & Chittum, H.S. (1997) Selenoprotein

P associates with endothelial cells in rat tissues. Histochem Cell Biol, 108, 11-15.

Burk, R.F., Hill, K.E., Olson, G.E., Weeber, E.J., Motley, A.K., Winfrey, V.P. & Austin,

L.M. (2007) Deletion of apolipoprotein E receptor-2 in mice lowers brain

selenium and causes severe neurological dysfunction and death when a low-

selenium diet is fed. The Journal of neuroscience : the official journal of the

Society for Neuroscience, 27, 6207-6211.

Burkhead, J.L., Ralle, M., Wilmarth, P., David, L. & Lutsenko, S. (2010) Elevated

Copper Remodels Hepatic RNA Processing Machinery in the Mouse Model of

Wilson's Disease. J Mol Biol.

Castets, P., Bertrand, A.T., Beuvin, M., Ferry, A., Le Grand, F., Castets, M., Chazot, G.,

Rederstorff, M., Krol, A., Lescure, A., Romero, N.B., Guicheney, P. & Allamand,

V. (2011) Satellite cell loss and impaired muscle regeneration in selenoprotein N

deficiency. Hum Mol Genet, 20, 694-704.

Chen, C.L., Shim, M.S., Chung, J., Yoo, H.S., Ha, J.M., Kim, J.Y., Choi, J., Zang, S.L.,

Hou, X., Carlson, B.A., Hatfield, D.L. & Lee, B.J. (2006) G-rich, a Drosophila

selenoprotein, is a Golgi-resident type III membrane protein. Biochemical and

biophysical research communications, 348, 1296-1301.

Chen, J. & Berry, M.J. (2003) Selenium and selenoproteins in the brain and brain

diseases. Journal of neurochemistry, 86, 1-12.

Cheung, M.C., Vaisar, T., Han, X., Heinecke, J.W. & Albers, J.J. (2010) Phospholipid

transfer protein in human plasma associates with proteins linked to immunity and

inflammation. Biochemistry, 49, 7314-7322.

Chiu-Ugalde, J., Theilig, F., Behrends, T., Drebes, J., Sieland, C., Subbarayal, P., Kohrle,

J., Hammes, A., Schomburg, L. & Schweizer, U. (2010) Mutation of megalin

leads to urinary loss of selenoprotein P and selenium deficiency in serum, liver,

kidneys and brain. Biochem J, 431, 103-111.

Chu, F.F., Esworthy, R.S., Ho, Y.S., Bermeister, M., Swiderek, K. & Elliott, R.W. (1997)

Expression and chromosomal mapping of mouse Gpx2 gene encoding the

gastrointestinal form of glutathione peroxidase, GPX-GI. Biomed Environ Sci, 10,

156-162.

Clarke, N.F., Kidson, W., Quijano-Roy, S., Estournet, B., Ferreiro, A., Guicheney, P.,

Manson, J.I., Kornberg, A.J., Shield, L.K. & North, K.N. (2006) SEPN1:

associated with congenital fiber-type disproportion and insulin resistance. Ann

Neurol, 59, 546-552.

Conrad, M., Jakupoglu, C., Moreno, S.G., Lippl, S., Banjac, A., Schneider, M., Beck, H.,

Hatzopoulos, A.K., Just, U., Sinowatz, F., Schmahl, W., Chien, K.R., Wurst, W.,

Bornkamm, G.W. & Brielmeier, M. (2004) Essential role for mitochondrial

28

thioredoxin reductase in hematopoiesis, heart development, and heart function.

Mol Cell Biol, 24, 9414-9423.

Curran, J.E., Jowett, J.B., Elliott, K.S., Gao, Y., Gluschenko, K., Wang, J., Abel Azim,

D.M., Cai, G., Mahaney, M.C., Comuzzie, A.G., Dyer, T.D., Walder, K.R.,

Zimmet, P., MacCluer, J.W., Collier, G.R., Kissebah, A.H. & Blangero, J. (2005)

Genetic variation in selenoprotein S influences inflammatory response. Nat

Genet, 37, 1234-1241.

Daneman, R., Zhou, L., Agalliu, D., Cahoy, J.D., Kaushal, A. & Barres, B.A. (2010) The

mouse blood-brain barrier transcriptome: a new resource for understanding the

development and function of brain endothelial cells. PLoS One, 5, e13741.

Diamond, A., Dudock, B. & Hatfield, D. (1981) Structure and properties of a bovine liver

UGA suppressor serine tRNA with a tryptophan anticodon. Cell, 25, 497-506.

Dikiy, A., Novoselov, S.V., Fomenko, D.E., Sengupta, A., Carlson, B.A., Cerny, R.L.,

Ginalski, K., Grishin, N.V., Hatfield, D.L. & Gladyshev, V.N. (2007) SelT, SelW,

SelH, and Rdx12: genomics and molecular insights into the functions of

selenoproteins of a novel thioredoxin-like family. Biochemistry, 46, 6871-6882.

Du, S., Zhou, J., Jia, Y. & Huang, K. (2010) SelK is a novel ER stress-regulated protein

and protects HepG2 cells from ER stress agent-induced apoptosis. Arch Biochem

Biophys, 502, 137-143.

ElAli, A. & Hermann, D.M. (2010) Apolipoprotein E controls ATP-binding cassette

transporters in the ischemic brain. Sci Signal, 3, ra72.

Erickson, J.R., Joiner, M.L., Guan, X., Kutschke, W., Yang, J., Oddis, C.V., Bartlett,

R.K., Lowe, J.S., O'Donnell, S.E., Aykin-Burns, N., Zimmerman, M.C.,

Zimmerman, K., Ham, A.J., Weiss, R.M., Spitz, D.R., Shea, M.A., Colbran, R.J.,

Mohler, P.J. & Anderson, M.E. (2008) A dynamic pathway for calcium-

independent activation of CaMKII by methionine oxidation. Cell, 133, 462-474.

Esaki, N., Nakamura, T., Tanaka, H. & Soda, K. (1982) Selenocysteine lyase, a novel

enzyme that specifically acts on selenocysteine. Mammalian distribution and

purification and properties of pig liver enzyme. J Biol Chem, 257, 4386-4391.

Esworthy, R.S., Ho, Y.S. & Chu, F.F. (1997) The Gpx1 gene encodes mitochondrial

glutathione peroxidase in the mouse liver. Arch Biochem Biophys, 340, 59-63.

Fan, Q.-W., Iosbe, I., Asou, H., Yanagisawa, K. & Michikawa, M. (2001) Expression and

regulation of apolipoprotein E receptors in the cells of the central nervous system

in culture: A review. Age (Dordr), 24, 1-10.

Ferguson, A.D., Labunskyy, V.M., Fomenko, D.E., Arac, D., Chelliah, Y., Amezcua,

C.A., Rizo, J., Gladyshev, V.N. & Deisenhofer, J. (2006) NMR structures of the

selenoproteins Sep15 and SelM reveal redox activity of a new thioredoxin-like

family. J Biol Chem, 281, 3536-3543.

Ferreiro, A., Ceuterick-de Groote, C., Marks, J.J., Goemans, N., Schreiber, G., Hanefeld,

F., Fardeau, M., Martin, J.J., Goebel, H.H., Richard, P., Guicheney, P. &

Bonnemann, C.G. (2004) Desmin-related myopathy with Mallory body-like

inclusions is caused by mutations of the selenoprotein N gene. Ann Neurol, 55,

676-686.

Forstrom, J.W., Zakowski, J.J. & Tappel, A.L. (1978) Identification of the catalytic site

of rat liver glutathione peroxidase as selenocysteine. Biochemistry, 17, 2639-

2644.

29

Fradejas, N., Pastor, M.D., Mora-Lee, S., Tranque, P. & Calvo, S. (2008) SEPS1 gene is

activated during astrocyte ischemia and shows prominent antiapoptotic effects. J

Mol Neurosci, 35, 259-265.

Fradejas, N., Serrano-Perez Mdel, C., Tranque, P. & Calvo, S. (2011) Selenoprotein S

expression in reactive astrocytes following brain injury. Glia, 59, 959-972.

Fu, H., Subramanian, R.R. & Masters, S.C. (2000) 14-3-3 proteins: structure, function,

and regulation. Annu Rev Pharmacol Toxicol, 40, 617-647.

Gao, Y., Hannan, N.R., Wanyonyi, S., Konstantopolous, N., Pagnon, J., Feng, H.C.,

Jowett, J.B., Kim, K.H., Walder, K. & Collier, G.R. (2006) Activation of the

selenoprotein SEPS1 gene expression by pro-inflammatory cytokines in HepG2

cells. Cytokine, 33, 246-251.

Gao, Y., Pagnon, J., Feng, H.C., Konstantopolous, N., Jowett, J.B., Walder, K. & Collier,

G.R. (2007) Secretion of the glucose-regulated selenoprotein SEPS1 from

hepatoma cells. Biochemical and biophysical research communications, 356, 636-

641.

Gorlach, A., Klappa, P. & Kietzmann, T. (2006) The endoplasmic reticulum: folding,

calcium homeostasis, signaling, and redox control. Antioxid Redox Signal, 8,

1391-1418.

Grumolato, L., Ghzili, H., Montero-Hadjadje, M., Gasman, S., Lesage, J., Tanguy, Y.,

Galas, L., Ait-Ali, D., Leprince, J., Guerineau, N.C., Elkahloun, A.G., Fournier,

A., Vieau, D., Vaudry, H. & Anouar, Y. (2008) Selenoprotein T is a PACAP-

regulated gene involved in intracellular Ca2+ mobilization and neuroendocrine

secretion. FASEB J, 22, 1756-1768.

Gu, Q.P., Beilstein, M.A., Barofsky, E., Ream, W. & Whanger, P.D. (1999) Purification,

characterization, and glutathione binding to selenoprotein W from monkey

muscle. Arch Biochem Biophys, 361, 25-33.

Guimaraes, M.J., Peterson, D., Vicari, A., Cocks, B.G., Copeland, N.G., Gilbert, D.J.,

Jenkins, N.A., Ferrick, D.A., Kastelein, R.A., Bazan, J.F. & Zlotnik, A. (1996)

Identification of a novel selD homolog from eukaryotes, bacteria, and archaea: is

there an autoregulatory mechanism in selenocysteine metabolism? Proceedings of

the National Academy of Sciences of the United States of America, 93, 15086-

15091.

Hatfield, D.L. (2006) Selenium : its molecular biology and role in human health,

Springer, New York, NY.

Hawkes, W.C., Wang, T.T., Alkan, Z., Richter, B.D. & Dawson, K. (2009) Selenoprotein

W modulates control of cell cycle entry. Biol Trace Elem Res, 131, 229-244.

Hawkes, W.C., Wilhelmsen, E.C. & Tappel, A.L. (1985) Abundance and tissue

distribution of selenocysteine-containing proteins in the rat. J Inorg Biochem, 23,

77-92.

Hill, K.E., Zhou, J., Austin, L.M., Motley, A.K., Ham, A.J., Olson, G.E., Atkins, J.F.,

Gesteland, R.F. & Burk, R.F. (2007) The selenium-rich C-terminal domain of

mouse selenoprotein P is necessary for the supply of selenium to brain and testis

but not for the maintenance of whole body selenium. J Biol Chem, 282, 10972-

10980.

30

Hoffmann, P.R., Hoge, S.C., Li, P.A., Hoffmann, F.W., Hashimoto, A.C. & Berry, M.J.

(2007) The selenoproteome exhibits widely varying, tissue-specific dependence

on selenoprotein P for selenium supply. Nucleic Acids Res, 35, 3963-3973.

Hondal, R.J. & Ruggles, E.L. (2011) Differing views of the role of selenium in

thioredoxin reductase. Amino acids, 41, 73-89.

Horibata, Y. & Hirabayashi, Y. (2007) Identification and characterization of human

ethanolaminephosphotransferase1. J Lipid Res, 48, 503-508.

Hwang, C., Sinskey, A.J. & Lodish, H.F. (1992) Oxidized redox state of glutathione in

the endoplasmic reticulum. Science, 257, 1496-1502.

Jakupoglu, C., Przemeck, G.K., Schneider, M., Moreno, S.G., Mayr, N., Hatzopoulos,

A.K., de Angelis, M.H., Wurst, W., Bornkamm, G.W., Brielmeier, M. & Conrad,

M. (2005) Cytoplasmic thioredoxin reductase is essential for embryogenesis but

dispensable for cardiac development. Mol Cell Biol, 25, 1980-1988.

Johansson, L., Gafvelin, G. & Arner, E.S. (2005) Selenocysteine in proteins-properties

and biotechnological use. Biochim Biophys Acta, 1726, 1-13.

Jurynec, M.J., Xia, R., Mackrill, J.J., Gunther, D., Crawford, T., Flanigan, K.M.,

Abramson, J.J., Howard, M.T. & Grunwald, D.J. (2008) Selenoprotein N is

required for ryanodine receptor calcium release channel activity in human and

zebrafish muscle. Proceedings of the National Academy of Sciences of the United

States of America, 105, 12485-12490.

Karlsson, H.K., Tsuchida, H., Lake, S., Koistinen, H.A. & Krook, A. (2004) Relationship

between serum amyloid A level and Tanis/SelS mRNA expression in skeletal

muscle and adipose tissue from healthy and type 2 diabetic subjects. Diabetes, 53,

1424-1428.

Kim, H.Y. & Gladyshev, V.N. (2004) Methionine sulfoxide reduction in mammals:

characterization of methionine-R-sulfoxide reductases. Mol Biol Cell, 15, 1055-

1064.

Kim, K.H., Gao, Y., Walder, K., Collier, G.R., Skelton, J. & Kissebah, A.H. (2007)

SEPS1 protects RAW264.7 cells from pharmacological ER stress agent-induced

apoptosis. Biochemical and biophysical research communications, 354, 127-132.

Kipp, A., Banning, A., van Schothorst, E.M., Meplan, C., Schomburg, L., Evelo, C.,

Coort, S., Gaj, S., Keijer, J., Hesketh, J. & Brigelius-Flohe, R. (2009) Four

selenoproteins, protein biosynthesis, and Wnt signalling are particularly sensitive

to limited selenium intake in mouse colon. Mol Nutr Food Res, 53, 1561-1572.

Koller, L.D. & Exon, J.H. (1986) The two faces of selenium-deficiency and toxicity--are

similar in animals and man. Canadian journal of veterinary research = Revue

canadienne de recherche veterinaire, 50, 297-306.

Korschineck, I., Ziegler, S., Breuss, J., Lang, I., Lorenz, M., Kaun, C., Ambros, P.F. &

Binder, B.R. (2001) Identification of a novel exon in apolipoprotein E receptor 2

leading to alternatively spliced mRNAs found in cells of the vascular wall but not

in neuronal tissue. J Biol Chem, 276, 13192-13197.

Kryukov, G.V., Castellano, S., Novoselov, S.V., Lobanov, A.V., Zehtab, O., Guigo, R. &

Gladyshev, V.N. (2003) Characterization of mammalian selenoproteomes.

Science, 300, 1439-1443.

Kryukov, G.V. & Gladyshev, V.N. (2004) The prokaryotic selenoproteome. EMBO Rep,

5, 538-543.

31

Kuhbacher, M., Bartel, J., Hoppe, B., Alber, D., Bukalis, G., Brauer, A.U., Behne, D. &

Kyriakopoulos, A. (2009) The brain selenoproteome: priorities in the hierarchy

and different levels of selenium homeostasis in the brain of selenium-deficient

rats. J Neurochem, 110, 133-142.

Kurokawa, S., Takehashi, M., Tanaka, H., Mihara, H., Kurihara, T., Tanaka, S., Hill, K.,

Burk, R. & Esaki, N. (2011) Mammalian selenocysteine lyase is involved in

selenoprotein biosynthesis. J Nutr Sci Vitaminol (Tokyo), 57, 298-305.

Labunskyy, V.M., Ferguson, A.D., Fomenko, D.E., Chelliah, Y., Hatfield, D.L. &

Gladyshev, V.N. (2005) A novel cysteine-rich domain of Sep15 mediates the

interaction with UDP-glucose:glycoprotein glucosyltransferase. J Biol Chem, 280,

37839-37845.

Labunskyy, V.M., Lee, B.C., Handy, D.E., Loscalzo, J., Hatfield, D.L. & Gladyshev,

V.N. (2011) Both maximal expression of selenoproteins and selenoprotein

deficiency can promote development of type 2 diabetes-like phenotype in mice.

Antioxid Redox Signal.

Lacourciere, G.M., Mihara, H., Kurihara, T., Esaki, N. & Stadtman, T.C. (2000)

Escherichia coli NifS-like proteins provide selenium in the pathway for the

biosynthesis of selenophosphate. J Biol Chem, 275, 23769-23773.

Lein, E.S. & Hawrylycz, M.J. & Ao, N. & Ayres, M. & Bensinger, A. & Bernard, A. &

Boe, A.F. & Boguski, M.S. & Brockway, K.S. & Byrnes, E.J. & Chen, L. &

Chen, L. & Chen, T.M. & Chin, M.C. & Chong, J. & Crook, B.E. & Czaplinska,

A. & Dang, C.N. & Datta, S. & Dee, N.R. & Desaki, A.L. & Desta, T. & Diep, E.

& Dolbeare, T.A. & Donelan, M.J. & Dong, H.W. & Dougherty, J.G. & Duncan,

B.J. & Ebbert, A.J. & Eichele, G. & Estin, L.K. & Faber, C. & Facer, B.A. &

Fields, R. & Fischer, S.R. & Fliss, T.P. & Frensley, C. & Gates, S.N. &

Glattfelder, K.J. & Halverson, K.R. & Hart, M.R. & Hohmann, J.G. & Howell,

M.P. & Jeung, D.P. & Johnson, R.A. & Karr, P.T. & Kawal, R. & Kidney, J.M. &

Knapik, R.H. & Kuan, C.L. & Lake, J.H. & Laramee, A.R. & Larsen, K.D. &

Lau, C. & Lemon, T.A. & Liang, A.J. & Liu, Y. & Luong, L.T. & Michaels, J. &

Morgan, J.J. & Morgan, R.J. & Mortrud, M.T. & Mosqueda, N.F. & Ng, L.L. &

Ng, R. & Orta, G.J. & Overly, C.C. & Pak, T.H. & Parry, S.E. & Pathak, S.D. &

Pearson, O.C. & Puchalski, R.B. & Riley, Z.L. & Rockett, H.R. & Rowland, S.A.

& Royall, J.J. & Ruiz, M.J. & Sarno, N.R. & Schaffnit, K. & Shapovalova, N.V.

& Sivisay, T. & Slaughterbeck, C.R. & Smith, S.C. & Smith, K.A. & Smith, B.I.

& Sodt, A.J. & Stewart, N.N. & Stumpf, K.R. & Sunkin, S.M. & Sutram, M. &

Tam, A. & Teemer, C.D. & Thaller, C. & Thompson, C.L. & Varnam, L.R. &

Visel, A. & Whitlock, R.M. & Wohnoutka, P.E. & Wolkey, C.K. & Wong, V.Y.,

et al. (2007) Genome-wide atlas of gene expression in the adult mouse brain.

Nature, 445, 168-176.

Levine, R.L., Moskovitz, J. & Stadtman, E.R. (2000) Oxidation of methionine in

proteins: roles in antioxidant defense and cellular regulation. IUBMB Life, 50,

301-307.

Li, Q., Harraz, M.M., Zhou, W., Zhang, L.N., Ding, W., Zhang, Y., Eggleston, T.,

Yeaman, C., Banfi, B. & Engelhardt, J.F. (2006) Nox2 and Rac1 regulate H2O2-

dependent recruitment of TRAF6 to endosomal interleukin-1 receptor complexes.

Mol Cell Biol, 26, 140-154.

32

Liao, Z.M., Hou, C., Liu, L.P. & Yu, D.P. (2010) Temperature dependence of

photoelectrical properties of single selenium nanowires. Nanoscale research

letters, 5, 926-929.

Lippman, S.M., Klein, E.A., Goodman, P.J., Lucia, M.S., Thompson, I.M., Ford, L.G.,

Parnes, H.L., Minasian, L.M., Gaziano, J.M., Hartline, J.A., Parsons, J.K.,

Bearden, J.D., 3rd, Crawford, E.D., Goodman, G.E., Claudio, J., Winquist, E.,

Cook, E.D., Karp, D.D., Walther, P., Lieber, M.M., Kristal, A.R., Darke, A.K.,

Arnold, K.B., Ganz, P.A., Santella, R.M., Albanes, D., Taylor, P.R., Probstfield,

J.L., Jagpal, T.J., Crowley, J.J., Meyskens, F.L., Jr., Baker, L.H. & Coltman,

C.A., Jr. (2009) Effect of selenium and vitamin E on risk of prostate cancer and

other cancers: the Selenium and Vitamin E Cancer Prevention Trial (SELECT).

JAMA, 301, 39-51.

Lobanov, A.V., Hatfield, D.L. & Gladyshev, V.N. (2008) Reduced reliance on the trace

element selenium during evolution of mammals. Genome Biol, 9, R62.

Lobanov, A.V., Hatfield, D.L. & Gladyshev, V.N. (2009) Eukaryotic selenoproteins and

selenoproteomes. Biochim Biophys Acta, 1790, 1424-1428.

Loh, K., Deng, H., Fukushima, A., Cai, X., Boivin, B., Galic, S., Bruce, C., Shields, B.J.,

Skiba, B., Ooms, L.M., Stepto, N., Wu, B., Mitchell, C.A., Tonks, N.K., Watt,

M.J., Febbraio, M.A., Crack, P.J., Andrikopoulos, S. & Tiganis, T. (2009)

Reactive oxygen species enhance insulin sensitivity. Cell Metab, 10, 260-272.

Low, S.C., Harney, J.W. & Berry, M.J. (1995) Cloning and functional characterization of

human selenophosphate synthetase, an essential component of selenoprotein

synthesis. J Biol Chem, 270, 21659-21664.

Maiorino, M., Scapin, M., Ursini, F., Biasolo, M., Bosello, V. & Flohe, L. (2003) Distinct

promoters determine alternative transcription of gpx-4 into phospholipid-

hydroperoxide glutathione peroxidase variants. J Biol Chem, 278, 34286-34290.

McClung, J.P., Roneker, C.A., Mu, W., Lisk, D.J., Langlais, P., Liu, F. & Lei, X.G.

(2004) Development of insulin resistance and obesity in mice overexpressing

cellular glutathione peroxidase. Proceedings of the National Academy of Sciences

of the United States of America, 101, 8852-8857.

Mendelev, N., Mehta, S.L., Witherspoon, S., He, Q., Sexton, J.Z. & Li, P.A. (2011)

Upregulation of human selenoprotein H in murine hippocampal neuronal cells

promotes mitochondrial biogenesis and functional performance. Mitochondrion,

11, 76-82.

Mendelev, N., Witherspoon, S. & Li, P.A. (2009) Overexpression of human

selenoprotein H in neuronal cells ameliorates ultraviolet irradiation-induced

damage by modulating cell signaling pathways. Exp Neurol, 220, 328-334.

Mihara, H., Kurihara, T., Watanabe, T., Yoshimura, T. & Esaki, N. (2000) cDNA

cloning, purification, and characterization of mouse liver selenocysteine lyase.

Candidate for selenium delivery protein in selenoprotein synthesis. J Biol Chem,

275, 6195-6200.

Misu, H., Takamura, T., Takayama, H., Hayashi, H., Matsuzawa-Nagata, N., Kurita, S.,

Ishikura, K., Ando, H., Takeshita, Y., Ota, T., Sakurai, M., Yamashita, T.,

Mizukoshi, E., Honda, M., Miyamoto, K., Kubota, T., Kubota, N., Kadowaki, T.,

Kim, H.J., Lee, I.K., Minokoshi, Y., Saito, Y., Takahashi, K., Yamada, Y.,

33

Takakura, N. & Kaneko, S. (2010) A liver-derived secretory protein,

selenoprotein P, causes insulin resistance. Cell Metab, 12, 483-495.

Moghadaszadeh, B., Petit, N., Jaillard, C., Brockington, M., Roy, S.Q., Merlini, L.,

Romero, N., Estournet, B., Desguerre, I., Chaigne, D., Muntoni, F., Topaloglu, H.

& Guicheney, P. (2001) Mutations in SEPN1 cause congenital muscular

dystrophy with spinal rigidity and restrictive respiratory syndrome. Nat Genet, 29,

17-18.

Monsen, E.R. (2000) Dietary reference intakes for the antioxidant nutrients: vitamin C,

vitamin E, selenium, and carotenoids. Journal of the American Dietetic

Association, 100, 637-640.

Mueller, A.S., Bosse, A.C., Most, E., Klomann, S.D., Schneider, S. & Pallauf, J. (2009)

Regulation of the insulin antagonistic protein tyrosine phosphatase 1B by dietary

Se studied in growing rats. J Nutr Biochem, 20, 235-247.

Musiani, F., Ciurli, S. & Dikiy, A. (2010) The interaction of selenoprotein W with 14-3-3

proteins: a computational approach. J Proteome Res.

Nakayama, A., Hill, K.E., Austin, L.M., Motley, A.K. & Burk, R.F. (2007) All regions of

mouse brain are dependent on selenoprotein P for maintenance of selenium. J

Nutr, 137, 690-693.

Noh, O.J., Park, Y.H., Chung, Y.W. & Kim, I.Y. (2010) Transcriptional regulation of

selenoprotein W by MyoD during early skeletal muscle differentiation. J Biol

Chem, 285, 40496-40507.

Novoselov, S.V., Kryukov, G.V., Xu, X.M., Carlson, B.A., Hatfield, D.L. & Gladyshev,

V.N. (2007) Selenoprotein H is a nucleolar thioredoxin-like protein with a unique

expression pattern. J Biol Chem, 282, 11960-11968.

Oakley, F.D., Abbott, D., Li, Q. & Engelhardt, J.F. (2009) Signaling components of

redox active endosomes: the redoxosomes. Antioxid Redox Signal, 11, 1313-1333.

Olson, G.E., Whitin, J.C., Hill, K.E., Winfrey, V.P., Motley, A.K., Austin, L.M., Deal, J.,

Cohen, H.J. & Burk, R.F. (2009) EXTRACELLULAR GLUTATHIONE

PEROXIDASE (Gpx3) BINDS SPECIFICALLY TO BASEMENT

MEMBRANES OF MOUSE RENAL CORTEX TUBULE CELLS. Am J Physiol

Renal Physiol.

Olson, G.E., Winfrey, V.P., Hill, K.E. & Burk, R.F. (2008) Megalin mediates

selenoprotein P uptake by kidney proximal tubule epithelial cells. J Biol Chem,

283, 6854-6860.

Palioura, S., Sherrer, R.L., Steitz, T.A., Soll, D. & Simonovic, M. (2009) The human

SepSecS-tRNASec complex reveals the mechanism of selenocysteine formation.

Science, 325, 321-325.

Panee, J., Stoytcheva, Z.R., Liu, W. & Berry, M.J. (2007) Selenoprotein H is a redox-

sensing high mobility group family DNA-binding protein that up-regulates genes

involved in glutathione synthesis and phase II detoxification. J Biol Chem, 282,

23759-23765.

Papp, L.V., Lu, J., Holmgren, A. & Khanna, K.K. (2007) From selenium to

selenoproteins: synthesis, identity, and their role in human health. Antioxid Redox

Signal, 9, 775-806.

34

Papp, L.V., Lu, J., Striebel, F., Kennedy, D., Holmgren, A. & Khanna, K.K. (2006) The

redox state of SECIS binding protein 2 controls its localization and selenocysteine

incorporation function. Mol Cell Biol, 26, 4895-4910.

Peters, M.M., Hill, K.E., Burk, R.F. & Weeber, E.J. (2006) Altered hippocampus

synaptic function in selenoprotein P deficient mice. Molecular

neurodegeneration, 1, 12.

Reeves, M.A., Bellinger, F.P. & Berry, M.J. (2010) The neuroprotective functions of

selenoprotein M and its role in cytosolic calcium regulation. Antioxid Redox

Signal, 12, 809-818.

Ria, R., Todoerti, K., Berardi, S., Coluccia, A.M., De Luisi, A., Mattioli, M., Ronchetti,

D., Morabito, F., Guarini, A., Petrucci, M.T., Dammacco, F., Ribatti, D., Neri, A.

& Vacca, A. (2009) Gene expression profiling of bone marrow endothelial cells in

patients with multiple myeloma. Clin Cancer Res, 15, 5369-5378.

Rotruck, J.T., Pope, A.L., Ganther, H.E., Swanson, A.B., Hafeman, D.G. & Hoekstra,

W.G. (1973a) Selenium: biochemical role as a component of glutathione

peroxidase. Science, 179, 588-590.

Rotruck, J.T., Pope, A.L., Ganther, H.E., Swanson, A.B., Hafeman, D.G. & Hoekstra,

W.G. (1973b) Selenium: biochemical role as a component of glutathione

peroxidase. Science, 179, 588-590.

Sagher, D., Brunell, D., Hejtmancik, J.F., Kantorow, M., Brot, N. & Weissbach, H.

(2006) Thionein can serve as a reducing agent for the methionine sulfoxide

reductases. Proceedings of the National Academy of Sciences of the United States

of America, 103, 8656-8661.

Saito, Y., Hayashi, T., Tanaka, A., Watanabe, Y., Suzuki, M., Saito, E. & Takahashi, K.

(1999) Selenoprotein P in human plasma as an extracellular phospholipid

hydroperoxide glutathione peroxidase. Isolation and enzymatic characterization of

human selenoprotein p. J Biol Chem, 274, 2866-2871.

Saito, Y. & Takahashi, K. (2002) Characterization of selenoprotein P as a selenium

supply protein. Eur J Biochem, 269, 5746-5751.

Scharpf, M., Schweizer, U., Arzberger, T., Roggendorf, W., Schomburg, L. & Kohrle, J.

(2007) Neuronal and ependymal expression of selenoprotein P in the human

brain. J Neural Transm, 114, 877-884.

Schweizer, U., Streckfuss, F., Pelt, P., Carlson, B.A., Hatfield, D.L., Kohrle, J. &

Schomburg, L. (2005) Hepatically derived selenoprotein P is a key factor for

kidney but not for brain selenium supply. Biochem J, 386, 221-226.

Seiler, A., Schneider, M., Forster, H., Roth, S., Wirth, E.K., Culmsee, C., Plesnila, N.,

Kremmer, E., Radmark, O., Wurst, W., Bornkamm, G.W., Schweizer, U. &

Conrad, M. (2008) Glutathione peroxidase 4 senses and translates oxidative stress

into 12/15-lipoxygenase dependent- and AIF-mediated cell death. Cell Metab, 8,

237-248.

Sengupta, A., Carlson, B.A., Labunskyy, V.M., Gladyshev, V.N. & Hatfield, D.L. (2009)

Selenoprotein T deficiency alters cell adhesion and elevates selenoprotein W

expression in murine fibroblast cells. Biochem Cell Biol, 87, 953-961.

Shchedrina, V.A., Zhang, Y., Labunskyy, V.M., Hatfield, D.L. & Gladyshev, V.N.

(2010) Structure-function relations, physiological roles, and evolution of

mammalian ER-resident selenoproteins. Antioxid Redox Signal, 12, 839-849.

35

Stoytcheva, Z.R., Vladimirov, V., Douet, V., Stoychev, I. & Berry, M.J. (2010) Metal

transcription factor-1 regulation via MREs in the transcribed regions of

selenoprotein H and other metal-responsive genes. Biochim Biophys Acta, 1800,

416-424.

Stranges, S., Marshall, J.R., Natarajan, R., Donahue, R.P., Trevisan, M., Combs, G.F.,

Cappuccio, F.P., Ceriello, A. & Reid, M.E. (2007) Effects of long-term selenium

supplementation on the incidence of type 2 diabetes: a randomized trial. Ann

Intern Med, 147, 217-223.

Takebe, G., Yarimizu, J., Saito, Y., Hayashi, T., Nakamura, H., Yodoi, J., Nagasawa, S.

& Takahashi, K. (2002) A comparative study on the hydroperoxide and thiol

specificity of the glutathione peroxidase family and selenoprotein P. J Biol Chem,

277, 41254-41258.

Tamura, T., Yamamoto, S., Takahata, M., Sakaguchi, H., Tanaka, H., Stadtman, T.C. &

Inagaki, K. (2004) Selenophosphate synthetase genes from lung adenocarcinoma

cells: Sps1 for recycling L-selenocysteine and Sps2 for selenite assimilation.

Proceedings of the National Academy of Sciences of the United States of America,

101, 16162-16167.

Tanguy, Y., Falluel-Morel, A., Arthaud, S., Boukhzar, L., Manecka, D.L., Chagraoui, A.,

Prevost, G., Elias, S., Dorval-Coiffec, I., Lesage, J., Vieau, D., Lihrmann, I.,

Jegou, B. & Anouar, Y. (2011) The PACAP-regulated gene selenoprotein T is

highly induced in nervous, endocrine, and metabolic tissues during ontogenetic

and regenerative processes. Endocrinology, 152, 4322-4335.

Tobe, R., Mihara, H., Kurihara, T. & Esaki, N. (2009) Identification of proteins

interacting with selenocysteine lyase. Bioscience, biotechnology, and

biochemistry, 73, 1230-1232.

Ufer, C., Wang, C.C., Fahling, M., Schiebel, H., Thiele, B.J., Billett, E.E., Kuhn, H. &

Borchert, A. (2008) Translational regulation of glutathione peroxidase 4

expression through guanine-rich sequence-binding factor 1 is essential for

embryonic brain development. Genes Dev, 22, 1838-1850.

van der Brug, M.P., Blackinton, J., Chandran, J., Hao, L.Y., Lal, A., Mazan-Mamczarz,

K., Martindale, J., Xie, C., Ahmad, R., Thomas, K.J., Beilina, A., Gibbs, J.R.,

Ding, J., Myers, A.J., Zhan, M., Cai, H., Bonini, N.M., Gorospe, M. & Cookson,

M.R. (2008) RNA binding activity of the recessive parkinsonism protein DJ-1

supports involvement in multiple cellular pathways. Proceedings of the National

Academy of Sciences of the United States of America, 105, 10244-10249.

Vendeland, S.C., Beilstein, M.A., Chen, C.L., Jensen, O.N., Barofsky, E. & Whanger,

P.D. (1993) Purification and properties of selenoprotein W from rat muscle. J Biol

Chem, 268, 17103-17107.

Vendeland, S.C., Beilstein, M.A., Yeh, J.Y., Ream, W. & Whanger, P.D. (1995) Rat

skeletal muscle selenoprotein W: cDNA clone and mRNA modulation by dietary

selenium. Proceedings of the National Academy of Sciences of the United States

of America, 92, 8749-8753.

Verma, S., Hoffmann, F.W., Kumar, M., Huang, Z., Roe, K., Nguyen-Wu, E.,

Hashimoto, A.S. & Hoffmann, P.R. (2011) Selenoprotein k knockout mice exhibit

deficient calcium flux in immune cells and impaired immune responses. J

Immunol, 186, 2127-2137.

36

Walder, K., Kantham, L., McMillan, J.S., Trevaskis, J., Kerr, L., De Silva, A.,

Sunderland, T., Godde, N., Gao, Y., Bishara, N., Windmill, K., Tenne-Brown, J.,

Augert, G., Zimmet, P.Z. & Collier, G.R. (2002) Tanis: a link between type 2

diabetes and inflammation? Diabetes, 51, 1859-1866.

Wang, X.D., Vatamaniuk, M.Z., Wang, S.K., Roneker, C.A., Simmons, R.A. & Lei, X.G.

(2008) Molecular mechanisms for hyperinsulinaemia induced by overproduction

of selenium-dependent glutathione peroxidase-1 in mice. Diabetologia, 51, 1515-

1524.

Whanger, P.D. (2009) Selenoprotein expression and function-selenoprotein W. Biochim

Biophys Acta, 1790, 1448-1452.

Wilting, R., Schorling, S., Persson, B.C. & Bock, A. (1997) Selenoprotein synthesis in

archaea: identification of an mRNA element of Methanococcus jannaschii

probably directing selenocysteine insertion. J Mol Biol, 266, 637-641.

Xu, X.M., Carlson, B.A., Irons, R., Mix, H., Zhong, N., Gladyshev, V.N. & Hatfield,

D.L. (2007) Selenophosphate synthetase 2 is essential for selenoprotein

biosynthesis. Biochem J, 404, 115-120.

Yant, L.J., Ran, Q., Rao, L., Van Remmen, H., Shibatani, T., Belter, J.G., Motta, L.,

Richardson, A. & Prolla, T.A. (2003) The selenoprotein GPX4 is essential for

mouse development and protects from radiation and oxidative damage insults.

Free Radic Biol Med, 34, 496-502.

Ye, Y., Shibata, Y., Kikkert, M., van Voorden, S., Wiertz, E. & Rapoport, T.A. (2005)

Recruitment of the p97 ATPase and ubiquitin ligases to the site of

retrotranslocation at the endoplasmic reticulum membrane. Proceedings of the

National Academy of Sciences of the United States of America, 102, 14132-14138.

Ye, Y., Shibata, Y., Yun, C., Ron, D. & Rapoport, T.A. (2004) A membrane protein

complex mediates retro-translocation from the ER lumen into the cytosol. Nature,

429, 841-847.

Yeh, J.Y., Beilstein, M.A., Andrews, J.S. & Whanger, P.D. (1995) Tissue distribution

and influence of selenium status on levels of selenoprotein W. FASEB J, 9, 392-

396.

Zhang, Y., Fomenko, D.E. & Gladyshev, V.N. (2005) The microbial selenoproteome of

the Sargasso Sea. Genome Biol, 6, R37.

Zhang, Y., Zhou, Y., Schweizer, U., Savaskan, N.E., Hua, D., Kipnis, J., Hatfield, D.L. &

Gladyshev, V.N. (2008) Comparative analysis of selenocysteine machinery and

selenoproteome gene expression in mouse brain identifies neurons as key

functional sites of selenium in mammals. J Biol Chem, 283, 2427-2438.

Zhong, L., Arner, E.S. & Holmgren, A. (2000) Structure and mechanism of mammalian

thioredoxin reductase: the active site is a redox-active

selenolthiol/selenenylsulfide formed from the conserved cysteine-selenocysteine

sequence. Proceedings of the National Academy of Sciences of the United States

of America, 97, 5854-5859.

Zlokovic, B.V. (2008) The blood-brain barrier in health and chronic neurodegenerative

disorders. Neuron, 57, 178-201.

Zlokovic, B.V., Martel, C.L., Matsubara, E., McComb, J.G., Zheng, G., McCluskey,

R.T., Frangione, B. & Ghiso, J. (1996) Glycoprotein 330/megalin: probable role

in receptor-mediated transport of apolipoprotein J alone and in a complex with

37

Alzheimer disease amyloid beta at the blood-brain and blood-cerebrospinal fluid

barriers. Proc Natl Acad Sci U S A, 93, 4229-4234.

38

CHAPTER 2

ABSENCE OF SELENOPROTEIN P BUT NOT SELENOCYSTEINE LYASE

RESULTS IN SEVERE NEUROLOGICAL DYSFUNCTION

ABSTRACT

Dietary selenium restriction in mammals causes bodily selenium to be preferentially

retained in the brain relative to other organs. Almost all of the known selenoproteins are

found in brain, where expression is facilitated by selenocysteine-laden selenoprotein P.

The brain also expresses selenocysteine lyase, an enzyme that putatively salvages

selenocysteine and recycles the selenium for selenoprotein translation. We compared

mice with a genetic deletion of selenocysteine lyase to selenoprotein P knockout mice for

similarity of neurological impairments, and whether dietary selenium modulates these

parameters. We report that selenocysteine lyase knockout mice do not display

neurological dysfunction comparable to selenoprotein P knockout mice. Feeding a low-

selenium diet to selenocysteine lyase knockout mice revealed a mild spatial learning

deficit without disrupting motor coordination. Additionally, we report that the

neurological phenotype caused by the absence of selenoprotein P is exacerbated in male

versus female mice. These findings indicate that selenocysteine recycling via

selenocysteine lyase becomes limiting under selenium deficiency, and suggest the

presence of a complementary mechanism for processing selenocysteine. Our studies

illuminate the interaction between selenoprotein P and selenocysteine lyase in the

distribution and turnover of body and brain selenium, and emphasize the consideration of

sex differences when studying selenium and selenoproteins in vertebrate biology.

39

INTRODUCTION

Selenium (Se) is an essential dietary micronutrient with antioxidant properties, and the

human health consequences of Se-deficiency are under extensive study (Rayman, 2000).

Sex-specific differences are observed in Se status and metabolism, which has

complicated this research (Combs et al., 2011, Galan et al., 2005). The functions of Se in

biochemical reactions and cellular processes of organisms are principally mediated by

selenoproteins that incorporate Se into the amino acid selenocysteine (Sec). Biosynthesis

of Sec occurs on its UGA-recognizing tRNA, and is catalyzed by Sec synthetase

(SepSecS) in the presence of selenophosphate [reviewed in (Bellinger et al., 2009)].

Recoding the UGA stop codon for Sec incorporation requires a Sec-specific elongation

factor (EFSec), an mRNA stem-loop termed a Sec insertion sequence (SECIS), and a

SECIS-binding protein (SBP2). Among 25 human selenoproteins, the glutathione

peroxidase, thioredoxin reductase, and iodothyronine deiodinase families of enzymatic

selenoproteins are relatively well characterized, and crucial for the health of mammals.

Selenoprotein P (Sepp1) uniquely contains up to 10 Sec residues in primates and rodents.

During Se-deficiency, brain Se is maintained compared to other organs by tissue-specific

uptake of Sepp1 by ApoER2 and Megalin (Burk et al., 2007, Chiu-Ugalde et al., 2010).

Mutant mice lacking full-length Sepp1 or the Sec-rich C-terminus show a greater

depletion of brain Se than can be achieved through dietary Se deprivation (Hill et al.,

2007). Sepp1 gene disruption in mice additionally causes cognitive, motor and sensory

symptoms that can be exacerbated by dietary Se restriction and diminished by Se

supplementation. These mice present spatial learning deficits, spasticity, and

hyperreflexia that coincide with deficient synaptic plasticity and widespread

neurodegeneration (Caito et al., 2011, Peters et al., 2006, Valentine et al., 2008,

Valentine et al., 2005).

If endocytosis of Sepp1 delivers Se to cells, the Sec residues from Sepp1 must be

processed for incorporation into selenoproteins. Sec lyase (Scly) catalyzes the

decomposition of Sec into alanine and hydrogen selenide (Esaki et al., 1982), and

40

promotes the production of selenophosphate in the presence of Sec and selenophosphate

synthetase (SPS) (Tobe et al., 2009). Scly mRNA and protein are expressed in mouse

brain (Mihara et al., 2000), where it is posited to recycle Sec from Sepp1 for

selenoprotein synthesis (Schweizer et al., 2005).

We hypothesized that Scly liberates Se from Sepp1 in brain, and that deletion of Scly in

mice would cause similar neurological deficits as observed in Sepp1-/- mice. To test this

hypothesis, we assessed whether a novel transgenic mouse strain lacking functional Scly

develops a phenotype similar to Sepp1-deficient mice. Here we report that, in contrast to

Sepp1-/- mice, Scly-/- mice display few neurological abnormalities. However, spatial

learning and selenoprotein expression are sensitive to Scly disruption when the mice are

challenged with a low-Se diet. In addition, we extensively characterized sex differences

in the behavioral phenotype of Sepp1-/- mice, and report that male mice are more

dependent on Sepp1 and Se than females for normal brain function.

MATERIALS AND METHODS

Animals: Genetically modified male and female mice on a C57BL/6 background lacking

Sepp1 or Scly were bred on commercially available diets containing adequate Se (~0.25

ppm). Animals were given food and water ad libitum on a 12-hour light-cycle and group

housed until behavioral experimentation. All behavioral experiments were conducted on

single-housed adult mice aged 4-6 months during the light cycle. Male and female mice

of all genotypes were used in approximately equal numbers to examine sex differences

present in the animals. All animal procedures and experimental protocols were approved

by the University of Hawaii Institutional Animal Care and Use Committee.

Generation of Sepp1-/- and Scly-/- mice: Sepp1-/- mice were generated by

electroporating a construct into 129S9/SvEvH-derived embryonic stem (ES) cells that

were subsequently injected into C57BL/6 blastocysts. The resulting chimeric males were

bred with C57BL/6J females (Hill et al., 2003). Mutant mice were backcrossed to

C57BL/6J for at least 10 generations before arriving in our lab, and were bred with our

41

C57BL/6J colony to ensure congenic strains (Hoffmann et al., 2007). Sepp1+/- mice

were bred to generate littermate pups of Sepp1-/- knockout and Sepp1+/+ control mice.

Genotyping of the mice was carried out using methods previously described (Hoffmann

et al., 2007).

A targeting vector was generated by the NCRR-NIH supported KnockOut Mouse Project

(KOMP) Repository, and included an ~11-kb region of the wild type Scly locus

subcloned from a positively identified C57BL/6 BAC clone. The vector was designed

with one homology arm extending 5.5 kb 5’ to exon 4, and the other 5.5 kb homology

arm terminating 3’ to exon 7. A promoterless trapping cassette (L1L2_gt0) with flanking

Flp-recombinase target (FRT) sites was inserted in an intron 5’ of exon 4. Efficient

splicing to the reporter cassette results in truncation of the endogenous transcript, causing

a constitutive null mutation in the Scly gene. Cre-recombinase target loxP sites were

inserted 5’ and 3’ of critical coding exon 4. The total size of the targeting construct,

including vector backbone (L3L4_pZero_DTA_kan) and Neo cassette, was 21.642 kb.

The targeting vector was transfected into C57BL/6 embryonic stem cells by

electroporation. After selection with antibiotic, surviving clones were expanded and

analyzed by PCR to identify recombinant ES cell clones. ES cell clones were

microinjected into C57BL/6 blastocysts to produce chimeras with one wild-type and one

mutant Scly allele, which were then mated to generate Scly-/- mice on a pure C57BL/6

background. Mutant mice were backcrossed to C57BL/6J to ensure genetic comparability

with wild-type control C57BL/6J mice. The latter were maintained not more than five

generations after arrival from The Jackson Laboratory. Deletion of Scly was confirmed in

all offspring using PCR that amplified a 1.2-kb product in the targeted region present in

the wild-type allele (forward, 5’-CAC AGG TGC GGC CAT GAG GG-3’; reverse, 5’-

CTG GCT GTC CCT GAA CTA GCT TCA TA-3’) and a 233-bp product in the mutant

allele (forward, 5’-GAG ATG GCG CAA CGA AAT TAA T-3’; reverse, 5’- CTG GCT

GTC CCT GAA CTA GCT TCA TA -3’).

Diets: For dietary experiments, animals were switched from standard laboratory diets

containing ~0.25 mg/kg Se to defined diets at the time of weaning (3–4 wk of age). Mice

were fed Open Source Diets purchased from Research Diets containing either 0.08 mg/kg

42

(cat.#D19101) or 1.0 mg/kg (cat.#D05050403) Se. The diets were formulated with

purified ingredients and contained 20.3% protein, 66% carbohydrate, and 5% fat. The

protein source was casein, which was also the source of Se in the low Se diet. For high Se

diet, sodium selenite was added to achieve final Se levels. Multiple lots were

independently tested to confirm the Se concentration by inductively coupled plasma-MS

(Bodycote) with lot-to-lot variation at or below the detection limit of inductively coupled

plasma-MS testing (0.02 mg/kg). Mice were fed the defined diets from weaning until the

time of sacrifice. Se-adequate standard lab diets containing ~0.25 mg/kg Se are not

completely defined, and may be considered slightly supplemented compared to the rodent

RDA of 0.15 mg/kg Se. The low-Se diet containing 0.08 mg/kg Se is marginally

deficient, and causes moderate selenium deficiency in mice (Hoffmann et al., 2010). The

high-Se diet containing 1.0 mg/kg Se has been shown to prevent many of the

neurological symptoms in Sepp1-/- mice (Hill et al., 2003). This range of dietary Se

concentration reflects the global spectrum of human Se intake, from deficiency to

therapeutic supplementation.

Animal Behavior:

Spontaneous activity. Animals were placed in a transparent cylinder (20 cm diameter, 20

cm height) and activity was videotaped for three minutes. The cylinder was situated on

clear plexiglass with a mirror placed at an angle underneath for clear view of movement

along the ground as well as along the walls of the cylinder. The number of rears, forelimb

and hindlimb steps, and time spent grooming were measured. Videotapes were scored in

slow motion by an experimenter blind to the mouse genotype. A rear was counted when

an animal made a vertical movement with both forelimbs removed from the ground.

Forelimb and hindlimb steps were counted when an animal moved both forelimbs or both

hindlimbs across the floor of the cylinder. Number of steps, rears, and time spent

grooming were compared for wild-type and knockout mice (Fleming et al., 2004).

Open field. Animals were placed in a square box (50 cm sides, 40 cm walls) and

monitored by overhead camera linked to computer-assisted tracking software. During the

test, the mice were allowed to move freely around the open field and to explore the

environment for five minutes. The path of each mouse was automatically recorded, and

43

recordings were then analyzed. Total distance traveled, number of rears, time spent

grooming, and center time was compared between groups.

Vertical pole. The vertical pole descent test has been used to assess coordination and

basal ganglia related movement disorders in transgenic mice (Fleming et al., 2004).

Animals were placed head-up on top of a vertical wooden pole 50 cm long (1.2 cm in

diameter). The base of the pole was fixed in plexiglass and put in the home cage. When

placed on the pole, animals orient themselves downward and descend the length of the

pole back into their home cage. Knockout and wild-type mice received two days of

training consisting of five trials per day. On the third day, animals received five trials,

and time to orient downward (turn) and total time to descend (total) were measured with

a stopwatch. The best performance over the five trials was used for both wild-type and

knockout mice.

Inverted grid. The ability to hang upside down is a test of neuromuscular strength

(Crawley, 1999). Each mouse was placed on a wire grid (mesh, 12 cm2 with 0.5 cm

2

squares) 20 cm above a table top for 120 sec and videotaped. The lid was gently turned

upside down, 60 cm above a soft surface to avoid injuries. The latency to fall was timed.

Each mouse was given up to two attempts to hold on to the inverted grid for a maximum

of 120 seconds and the longest period was recorded.

Morris water maze. A circular pool (2 m in diameter, 1 m deep) surrounded by constant

external cues was located in an observation room and filled with 24°C water. White non-

toxic paint was added to make the water translucent. Tests were performed under dim

light conditions. Lights below the height of the tank, necessary for video capture, also

provided light for the swimming mouse. A small circular escape platform (7 cm diameter,

located just below the water surface, or protruding just above the water) was placed in a

constant location in the center of quadrant 1. Four equally spaced points around the wall

of the pool were used as starting points. The mice were given one block of four trials

each day with an inter-trial interval of 5 to 10 min. Each trial started from one of four

different points, in a semi-random order. The mouse was allowed to swim around until it

located the platform or 60 sec, at which point the mouse was placed on the platform by

the experimenter and allowed to stay on the platform for 15 sec. The time to locate the

platform was recorded as escape latency during training days. After sufficient training a

44

one-minute probe trial was performed, in which the platform was removed and the path

of each mouse on each trial was automatically recorded and then analyzed. Time

investigating the target, opposite, and adjacent quadrants, platform latency and platform

crossings were measured. Average swim speed was calculated from total distance

traveled per 60 sec trial. Animals that spent significant time floating, which occurred

sporadically in very few animals independent of genotype, were excluded from this

analysis.

Quantitative RT-PCR: Animals were sacrificed by CO2 asphyxiation, or deeply

anesthetized with tribromoethanol and decapitated, and the brains rapidly excised,

washed in PBS and snap-frozen in liquid nitrogen. Tissue was ground into powder, using

a mortar and pestle on dry ice, and collected into pre-chilled tubes. Total RNA from

tissue was prepared by Trizol extraction (Invitrogen, Carlsbad, CA, USA) followed by

purification using the RNeasy kit (QIAgen, Valencia, CA, USA). Concentration and

purity of extracted RNA and synthesized cDNA was determined using A260/A280 ratio

measured on an ND1000 Spectrophotometer (NanoDrop Technologies, Wilmington, DE,

USA). Synthesis of cDNA was carried out using High Capacity cDNA Reverse

Transcription Kit (Applied Biosystems, Foster City, CA, USA), with 1 g RNA per 20 l

reaction. For real-time PCR, 100 ng of the cDNA was used in 5 l reactions with

Platinum SYBR Green qPCR SuperMix-UDG (Invitrogen). Reactions were carried out in

triplicate or quadruplicate in a LightCycler 480 II thermal cycler (Roche, Indianapolis,

IN, USA). Cycling conditions followed the manufacturers suggestions in the SYBR

Green kit instructions. All qPCR results were normalized to 18S rRNA expression as a

housekeeping gene and analyzed using Absolute Quantification Software (Roche).

SDS-PAGE and Western blot: Total protein was extracted from powdered mouse tissues

by light sonication in CelLytic MT buffer (Sigma, St. Louis, MO, USA), followed by

centrifugation according to the manufacturers’ protocol. Protein was added to reduced

Laemmli buffer, boiled for 10 minutes, and loaded into 4-20% gradient polyacrylamide

gels (Bio-Rad, Hercules, CA, USA). Following electrophoresis, gel contents were

transferred to PVDF membranes, which were blocked with undiluted Odyssey Blocking

45

Buffer (Li-Cor Biosciences, Lincoln, NE, USA) for one hour. Membranes were then

probed for 90 minutes with the following primary antibodies: Goat-anti-GPX1 (R&D

Systems, Minneapolis, MN, USA), Rabbit-anti-GPX4 (AbFrontier, Seoul, Korea),

Rabbit-anti-SEPW1 (Rockland, Gilbertsville, PA, USA), and Mouse-anti-alpha Tubulin

(Novus, Littleton, CO, USA). After washing with PBS containing 0.05% Tween-20

(PBST), membranes were incubated in the dark in secondary antibodies labeled with

infrared fluorophores (Li-Cor Biosciences). After further washes in PBST, blots were

imaged and quantified with the Odyssey infrared imaging system (Li-Cor Biosciences).

Glutathione peroxidase activity assay: Total glutathione peroxidase activity was

measured using the Bioxytech GPx-340 Assay kit (Oxis International, Foster City, CA,

USA). Mouse tissues for the assay were collected in the same manner as above.

Powdered tissue was homogenized in lysis buffer by sonication and centrifuged at 15,000

x g. The resulting supernatant was serially diluted to determine the linear range of the

assay. 5 l of diluted sample (1:10 for brain, 1:200 for liver) was added to 25 l of assay

buffer and 25 l of NADPH reagent in a 96-well plate. 25 l of tert-butyl hydroperoxide

was added to initiate the reaction, which was monitored for 10 minutes at room

temperature by measuring kinetic absorbance at 340 nm on a SpectraMax M3 microplate

reader (Molecular Devices, Sunnyvale, CA, USA).

Statistical analysis: Data were analyzed using Microsoft Excel (Redmond, WA, USA),

and plotted using GraphPad Prism software (San Diego, CA, USA). Two-way ANOVA

was used to determine an interaction between genotype and sex for all experiments. If no

interaction effect was observed, male and female groups were sometimes combined.

Repeated measures ANOVA was used to assess training in the water maze, and genotype

and quadrant were the two factors for analyzing quadrant investigation on the probe trial.

Post hoc test using Bonferroni correction for multiple comparisons was used to determine

significance between individual groups. When male and female groups were combined,

unpaired t-tests comparing genotypes were performed for end-point measures in the

water maze and for biochemical experiments. The significance criteria were set at p <

0.05 for all statistical measures.

46

RESULTS

Neurological motor phenotype in Sepp1-/- mice is more pronounced in males

Genetic deletion of Sepp1 in mice results in neuromotor impairments (Hill et al., 2004,

Schomburg et al., 2003). We characterized the phenotype of Sepp1-deficient mice raised

on a standard lab diet using a battery of behavioral tests to assess motor function in male

and female mice. We found that motor coordination was worse in male knockout animals

compared to females. Spontaneous locomotor activity, as assessed using the cylindrical

chamber, was greatly decreased in the male Sepp1-deficient mice while being slightly

decreased in the female knockout animals, when compared to wild-type mice [Fig. 1A].

Two-way ANOVA revealed an effect of genotype on spontaneous rearing

(F(1,32)=13.06, p=0.0010) [Fig. 1A, left] and grooming (F(1,32)=16.93, p=0.0003) [Fig.

1A, center]. Post hoc analysis indicated a statistically significant decrease in rearing

(t(16)=3.464, p<0.01) and grooming (t(16)=2.532, p<0.05) in male Sepp1-/- mice,

whereas only grooming (t(16)=3.287, p<0.01) was decreased in female Sepp1-/- mice,

when compared to wild-type. Despite these differences, we did not observe any genotype

or sex interaction effects on total distance traveled in the open field [Fig. 1A, right]. To

assess motor coordination and strength, animals were subjected to the pole descent and

inverted hang tests. Two-way ANOVA revealed an interaction between genotype and sex

on time to turn (F(1,42)=16.57, p=0.0002) [Fig. 1B, left] and descend (F(1,42)=5.222,

p=0.0274) [Fig. 1B, center] a vertically oriented pole. Genotype strongly affected ability

to hang upside down for 120 seconds (F(1,32)=16.10, p=0.0003), however only male

Sepp1-/- mice were significantly different from wild-type control mice (t(16)=4.045,

p<0.001) [Fig. 1B, right].

Dietary Se supplementation alleviates motor deficits in Sepp1-/- mice

Previous studies have shown that Se supplementation can attenuate neuromotor

impairments in Sepp1-/- mice (Schweizer et al., 2004). We determined if gender could

influence this attenuation. Animals were supplemented with 1 mg/kg Se in the diet

starting at weaning, and assayed in the same manner as those fed a standard diet. We

found that Se-supplemented Sepp1-/- mice do not differ from wild-type mice in

47

spontaneous rearing [Fig. 2A, left], grooming [Fig. 2A, center] and distance traveled [Fig.

2A, right]. Performance on the pole descent test also improved in Se-supplemented mice,

however a main effect of genotype remained for the time to invert (F(1,33)=12.68,

p=0.0011) [Fig. 2B, left] and descend (F(1,33)=4.989, p=0.0324) [Fig. 2B, center] the

vertical pole, and post hoc test indicated male Sepp1-/- mice were slower than wild-type

littermates in turn time only (t(19)=4.004, p<0.001). Performance on the inverted hang

test also improved in the Se-supplemented Sepp1-/- mice, and an effect of genotype did

not reach statistical significance (F(1,31)=3.257, p=0.081) [Fig. 2B, right]. In all motor

tests, Se-supplemented Sepp1-/- mice showed subtle qualitative behavioral deficits in

both sexes, but the male bias was largely eliminated.

Generation and characterization of Scly-/- animals

Scly knockout mice were generated by a conditional knockout approach in the event that

deletion of Scly caused embryonic lethality. Mice were generated with a trapping cassette

inserted upstream of exon 4 of the Scly gene, resulting in a constitutive null mutation in

the whole animal [Fig 3A]. Mutation of the targeted region of the Scly gene was

confirmed by PCR of mouse tail DNA [Fig. 3B], and Scly mRNA was undetectable by

qPCR in all tissues examined from Scly-/- mice, including brain [Fig. 3C]. Scly-/- mice

were born at the expected Mendelian ratio, appeared generally healthy, and developed to

adulthood. Unlike Sepp1-/- mice, Scly-/- mice were fertile, producing viable offspring

that did not differ in size at birth compared to wild-type control mice.

Neurological motor phenotype is largely absent in Scly-/- mice

Scly is a relatively uncharacterized enzyme thought to be involved in recycling Se from

Sec in support of selenoprotein synthesis. Sepp1 contains up to 10 Sec residues and is

proposed to be a source of Se for the brain. We hypothesized that mice lacking Scly

would be unable to efficiently catalyze Sec degradation, and therefore would manifest a

phenotype similar to the Sepp1-/- mice. Contrary to our prediction, we found that Scly-/-

animals raised on a Se-adequate diet displayed no neurological phenotype. Spontaneous

rearing and movement were unaffected by Scly genotype, and performance on the

vertical pole descent was similar between Scly+/+ and Scly-/- mice [data not shown].

48

As no apparent behavioral phenotype presented in the Scly-/- mice raised on a standard

lab diet, we subjected the animals to a marginally low Se diet, containing 0.08 mg/kg Se,

to assess whether the animals display enhanced sensitivity to Se deficiency as measured

by locomotor behavior. Total distance traveled in the open field was unaffected by Scly

genotype under low-Se conditions, however two-way ANOVA revealed an interaction

between Scly genotype and sex on rearing activity (F(1,22)=4.937, p=0.0369) [Fig. 4A].

Motor coordination assayed by turn and total time in the vertical pole descent test

indicated no difference due to genotype in Se-deficient Scly-/- animals [Fig. 4B].

Spatial learning is sensitive to disruptions in Se availability

The Morris water maze is a paradigm for assessing spatial learning and memory in

rodents (Morris, 1984). Sepp1 knockout mice raised on a Se-supplemented diet have mild

impairments in learning measured with this paradigm, despite having a large deficit in

long-term potentiation, a cellular model for learning and memory (Peters et al., 2006). As

Se supplementation attenuates neurological impairments in Sepp1 knockout mice, we

questioned if learning deficits would be greater for Sepp1-/- mice raised on a normal Se

diet. Mice were initially trained over several days to locate a hidden platform in a large

tank of water. Subsequently the platform was removed for one final trial, in which the

amount of time the animal investigated the area where the platform used to be was

measured. We compared male and female Sepp1-/- and wild-type animals raised on

standard laboratory chow with adequate Se for learning deficits.

In contrast to our findings of gender differences in neuromotor function, we did not

observe significant learning differences between Sepp1-/- male and female mice.

Therefore male and female groups were combined after eliminating sex as an interacting

variable. We found that learning was impaired in Sepp1-/- mice fed a standard diet [Fig.

5A]. Two-way ANOVA revealed an interaction between genotype and training

(F(7,112)=2.361, p=0.0275). Escape latency over time was not substantially reduced in

Sepp1-/- as compared to control mice, suggesting the mice did not learn the spatial

location of the platform. The probe trial results are ambiguous since the training was

ineffective in Sepp1-/- mice; however a genotype x quadrant interaction effect was

49

observed (F(3,64)=3.277, p=0.0266) during quadrant investigation [Fig. 5B]. Sepp1-/-

mice spent significantly more time than controls investigating the opposite quadrant,

which was not due to uncoordinated swimming, and likely indicates failure to learn the

platform location. There was a strong trend towards reduced number of platform

crossings (t(16)=2.064, p=0.0557) [Fig. 5C], but swim speed (t(14)=0.7262, p=0.48)

[Fig. 5D] was not significantly different between genotypes by t-test. Similar to published

work on Se-supplemented Sepp1-/- mice in the water maze, we found a genotype

difference during training, but not on the probe trial in Sepp1-/- mice fed a high Se diet

[data not shown].

Although we found little difference between Scly-/- and control mice in locomotor

activity, we assayed the Scly-/- animals using the Morris water maze for comparison with

Sepp1-/- mice. Scly-/- mice fed standard chow performed like wild-type mice during

training [Fig. 6A], as the only effect observed by two-way ANOVA was for training day

(F(7,154)=24.07, p<0.0001). Similarly, in the probe trial we found an effect of quadrant

(F(3,88)=5.78, p=0.0012), but no interaction with genotype [Fig. 6B]. Platform crossings

(t(22)=0.5415, p=0.59) [Fig. 6C], latency to platform location [data not shown], and

swim speed (t(22)=0.4955, p=0.63) [Fig. 6D] were not significantly different between

genotypes when assessed by t-test. Scly-/- mice showed a trend towards delayed learning

on days three and four. The lack of spatial learning impairment in Se-adequate Scly-/-

mice starkly contrasted with Sepp1-/- mice fed the same diet.

The trend toward mild learning deficits in Scly-/- mice led us to question if a restricted Se

diet would result in greater learning impairments in these animals. When Scly-/- animals

fed a Se-deficient diet were assessed, two-way ANOVA revealed an interaction between

genotype and training (F(5,120)=2.496, p=0.0345), while post hoc tests indicated longer

latency times on days 2 (t(24)=2.942, p<0.05), 3 (t(24)=3.467, p<0.01), 5 (t(24)=2.895,

p<0.05), and 6 (t(24)=3.037, p<0.05) in the knockout mice [Fig. 7A], which suggests

reduced learning. However the mice learned the platform location and the deficit was not

as severe as in non-supplemented Sepp1-/- animals. Scly-/- animals performed as well as

wild-type animals in the probe trial. We found a main effect on quadrant investigation

50

(F(3,96)=11.19, p<0.0001) but no Scly genotype interaction [Fig. 7B]. We found no

difference in platform crossings (t(24)=1.089, p=0.29) [Fig. 7C], latency [data not

shown], or average swim speed (t(22)=1.60, p=0.124) [Fig. 7D] in Scly-/- mice on a low

Se diet compared to control mice.

Selenoprotein expression and Glutathione Peroxidase activity

When dietary Se is restricted in mammals, Sec-enriched Sepp1 helps maintain

selenoprotein expression in the brain (Hill et al., 2007). We therefore investigated the

expression of selenoproteins in brains of Scly-/- animals by quantitative RT-PCR,

western blot, and glutathione peroxidase (GPX) activity. We did not find a significant

change in the mRNA for Sepp1 (t(9)=1.536, p=0.1589, n=5-6), GPX1 (t(10)=1.056,

p=0.3157, n=6), GPX4 (t(10)=0.6899, p=0.5060, n=6), or selenoprotein W (Sepw1)

(t(8)=2.149, p=0.0639, n=5) in brains of Scly-/- mice fed normal chow [data not shown].

Thus Scly-/- mice on a Se-adequate diet show neither a behavioral phenotype, nor any

major changes in selenoprotein mRNA expression in brain. Therefore western blotting

and GPX activity assays to assess the severity of Se-deficiency were performed in mice

fed a low-Se diet only.

Scly-/- animals on a low-Se diet displayed significantly elevated Sepp1 mRNA

expression in brain compared to control mice (t(14)=2.686, p=0.0177) [Fig. 8A]. GPX1

(t(12)=3.40, p=0.0053) [Fig. 8B] and GPX4 (t(14)=3.093, p=0.0079) [Fig. 8C] mRNA

were also significantly increased in brain, while Sepw1 was not changed (t(14)=0.1380,

p=0.89) [Fig. 8D]. In the same mice, we did not detect an increase in mRNA expression

of Nfs1 [(t(14)=0.7154, p=0.4861, not pictured], a cysteine desulfurase enzyme with Sec

lyase activity.

In contrast to the selenoprotein mRNA expression data, the corresponding protein levels

assessed by western blot are consistently decreased in the brains of Se-deficient Scly-/-

mice. Both GPX1 (t(14)=12.39, p<0.0001) [Fig. 9A] and Sepw1 (t(14)=8.294,

p<0.0001)) [Fig. 9C] are expressed at ~37% of the wild-type level, while GPX4

expression (t(14)=5.365, p<0.0001) [Fig. 9B] is reduced to ~60% compared to control

51

brains. These results confirm that Scly contributes to selenoprotein expression in brain

during Se-deficiency.

GPX activity in brain was reduced to 43% of the wild-type control level (t(14)=4.495,

p=0.0005) [Fig. 10A], and was reduced to 54% in liver (t(14)=9.9297, p<0.0001) [Fig.

10B] of Scly-/- mice fed a low-Se diet. However, GPX activity in serum was not

significantly reduced [Fig. 10C]. These data indicate that Scly helps maintain

selenoprotein expression and activity when dietary Se availability is limiting. They

further suggest that Se status in tissues may be more affected than plasma in the absence

of Scly. The contrast of increased selenoprotein mRNA expression in the Scly-/- mouse

brain despite reduced selenoproteins and enzyme activity suggests that Scly is a

significant contributor to the Se pool for selenoprotein translation.

DISCUSSION

The results reported herein describe the first characterization of a novel mouse strain

deficient in Scly. These mice exhibit minimal neurological deficits, an unexpected

finding given the phenotypic effects of Sepp1 knockout and the proposed role of Scly in

recycling the essential trace element Se from Sec. Unlike Sepp1-/- animals, deletion of

Scly does not result in neuromotor impairments or spatial learning deficits, except under

low-Se conditions. We also report sex differences in the motor phenotype of mice with

genetic deletion of Sepp1, highlighting the importance of considering gender on studies

addressing the biological functions of Se or selenoproteins in mammals.

Sepp1 is a plasma protein that is considered to be the physiological transporter of Se from

liver to brain. Sepp1 is additionally found in grey and white matter and cerebrospinal

fluid, and may store Se within brain (Scharpf et al., 2007). Sepp1 and ApoER2 maintain

a high Se concentration in the testes, and prioritize Se to brain albeit at a lower

concentration (Burk & Hill, 2009). Most tissues produce Sepp1 and Scly, and the whole

body turnover rate of Sepp1 is high, fluxing a significant proportion of bodily Se even

when dietary availability is limiting (Burk & Hill, 2005). Brain Se level is not dependent

52

on hepatic Sepp1 in Se-adequate adult animals (Schweizer et al., 2005). However Se-

deficiency directs Se from liver-derived Sepp1 to the brain (Nakayama et al., 2007,

Renko et al., 2008). The turnover rate of Sepp1 within the nervous system and the

interaction with the circulation are uncertain.

Unlike most trace element transporters, Sepp1 cannot rapidly load and unload cargo

because Se is covalently incorporated as the amino acid Sec, which must be degraded to

supply Se. Biosynthesis and incorporation of Sec is a protracted, energy intensive process

that requires organized interaction of specific proteins (SPS, SepSecS, EFSec, SBP2) and

nucleic acids (tRNA(Sec)

, SECIS-containing mRNA), in addition to ATP, pyridoxal

phosphate, and the translation machinery. Receptor-mediated uptake by ApoER2

facilitates entry of Sepp1 into cells, but recycling of the Sec residues would depend on

the lysosome or proteasome to hydrolyze peptide bonds followed by liberation of Se from

Sec, presumably by Scly.

To test the hypothesis that Scly recycles Sec from Sepp1 in the brain, we investigated

whether Scly-/- mice manifest a phenotype similar to Sepp1-/- mice. Surprisingly, very

little neurological dysfunction was present in the Scly-/- mice, even when fed a diet low

in Se. The lack of behavioral changes in Scly-/- mice, compared to Sepp1-/- mice, could

be due to Nfs1 catalyzing Sec to selenide conversion for selenoprotein synthesis

(Lacourciere et al., 2000). Although we did not detect increased Nfs1 mRNA in Scly-/-

mice brains, the normal activity of the enzyme might be compensating for the absence of

Scly. Alternatively, Sepp1 may have an acute function in brain not strictly related to Se

delivery. Disrupted synaptic plasticity in Se-supplemented Sepp1-/- mice (Peters et al.,

2006) supports the notion that Sepp1 has a role in cell signaling via its receptor, ApoER2.

ApoER2 is found at synaptic sites (Beffert et al., 2005), in cultured astrocytes,

oligodendrocytes, and microglia (Fan et al., 2001), and in the brain vasculature

(Korschineck et al., 2001), suggesting that all brain cells can take up Sepp1. In adult

brain, mRNA for Sepp1 is apparent in glia (Lein et al., 2007) but protein expression is

dominant in neurons (Bellinger et al., 2008, Scharpf et al., 2007). Sepp1 mRNA and

53

protein are abundant in choroid plexus epithelium (Bellinger et al., 2008, Steinert et al.,

1998, Zhang et al., 2008). Scly mRNA appears enriched in grey matter and neurons of

mouse brain (Lein et al., 2007). Further, a mouse proteomics study identified Scly in

synaptoneurosomes (Filiou et al., 2010).

We found that mRNA for Scly was not significantly changed in brain of Sepp1-/- mice

fed a standard diet. Since Sepp1-/- mice have depressed Se in brain, this finding suggests

that Scly is not Se-regulated in brain, and the enzyme is minimally affected by dietary Se

or tissue Se levels (Deagen et al., 1987). The mRNA expression of Sepp1 trended up in

Scly-/- mice fed normal chow, and was significantly increased in brain of Scly-/- mice

fed low-Se chow. Se-deficient Scly-/- mice also had increased GPX1 and GPX4 mRNA,

while that of Sepw1 was unchanged. However, GPX protein and activity in brain of the

low-Se Scly-/- mice were dramatically reduced compared to wild-type animals. Liver

GPX activity was similarly reduced, while serum activity was less affected. Therefore

Scly supports selenoprotein expression and function under conditions of dietary Se

deficiency. Increased GPX mRNA despite reduced protein and activity in the Se-deficient

Scly-/- brain could be a compensatory mechanism to boost inefficient selenoprotein

translation. These findings extend a recent study, which demonstrated reduced GPX1

expression and reduced incorporation of Se derived from radiolabeled Sepp1, in cells

with Scly knocked down by siRNA (Kurokawa et al., 2011). In addition, we observe that

tissues are more reliant on Scly than blood.

Our finding that Se-deficient Scly-/- mice manifest a subtle spatial learning deficit in the

water maze corresponds with results on Se-supplemented Sepp1-/- mice (Peters et al.,

2006). Spatial learning requires the hippocampus, which is more dependent on Sepp1 for

optimal Se concentration than other brain regions (Nakayama et al., 2007). Therefore

Scly, Sepp1, and probably other selenoproteins in the hippocampus support spatial

learning. It is likely that the kinetics of selenoprotein degradation and synthesis are even

more disrupted in Scly-/- mice than the steady-state mRNA and protein expression levels.

Brain regions with high metabolism or cellular turnover could be more dependent on a

54

putative Sepp1-Scly recycling mechanism, while other cell populations might efficiently

utilize an alternate Se source.

These results are the most extensive characterization of behavioral sex differences in

Sepp1-/- mice to date. It has been suggested that the phenotype of Sepp1-/- mice is sex-

dependent (Riese et al., 2006), however studies on Sepp1-/- mice have focused on males

and data regarding behavioral sex differences are limited. Despite variations in behavioral

testing paradigms, our results showing impaired motor performance in male Sepp1-/-

mice are in agreement with previous reports (Hill et al., 2004, Renko et al., 2008,

Schweizer et al., 2004). We additionally assessed motor impairment in female Sepp1-/-

mice, and found it to be minimal compared to males. We also report that the spatial

learning deficit in Sepp1-/- mice on a Se-adequate diet is worse than in Se-supplemented

mice, building on a previous study that used only male mice on a high-Se diet (Peters et

al., 2006).

Selenoprotein expression is modulated by sex in mammals (Meplan et al., 2007, Riese et

al., 2006, Stoedter et al., 2010), and Sepp1 is an androgen responsive gene (Takahashi et

al., 2006). Additionally, testosterone secretion declines during Se-deficiency in male rats

(Behne et al., 1996). Male mice displayed increased sensitivity to Sepp1 deletion, and

male Sepp1-/- mice greatly improved when given supranutritional dietary Se, indicating

that males have a higher demand than females for Sepp1 and Se in the nervous system.

However the phenotype is not exclusive to males, suggesting that some aspect of

metabolism or development that is more prominent in males is dependent on Se.

Male gender is a risk factor for poor neurodevelopmental outcome after premature birth.

Cerebral palsy and related developmental disorders are more common in males than

females (Johnston & Hagberg, 2007). The phenotype of Sepp1-/- mice resembles cerebral

palsy in that the developmental onset of spasticity and ataxia often presents with

intellectual impairment and seizures. Moreover, this phenotype is not rapidly progressive

and remains stable in adulthood when given adequate Se. Perinatal infection and

hypoxia-ischemia are synergistic risk factors for cerebral palsy (Johnston & Hagberg,

55

2007, Mayoral et al., 2009), while Sepp1 is known to modulate immunity (Bosschaerts et

al., 2008) and metabolism (Misu et al., 2010). Metabolically demanding brain regions

and cells, with a presumably higher rate of selenoprotein synthesis, are susceptible to

neurodegeneration in Sepp1-/- mice (Valentine et al., 2008). Moreover, an autosomal-

recessive human disease termed progressive cerebellocerebral atrophy has been linked to

mutations in SepSecS that globally disrupt selenoprotein synthesis (Agamy et al., 2010).

The sequelae of these patients, including mental retardation, spasticity and seizures, are

similar to those found in Sepp1-/- mice and emphasize the importance of selenoproteins

in the function and health of the nervous system.

In conclusion, these results indicate that a novel mouse strain lacking Scly does not

develop a neurological phenotype similar to Sepp1-/- mice. A subtle learning deficit is

observed when Scly-/- animals are fed a low-Se diet, and these animals also have reduced

expression of selenoproteins in brain. We further report a male bias in the neurological

motor phenotype of Sepp1-/- mice. The disparity of neurological problems in Scly-/- and

Sepp1-/- mice suggests that Sepp1 is more critical than Scly for maintenance of brain Se,

but that recycling Se from Sec via Scly is physiologically important during dietary Se

deficiency. Altogether these findings highlight that Se is critically important for the

nervous system, and that Se metabolism through Sepp1 and Scly impacts spatial learning.

56

REFERENCES

Agamy, O., Ben Zeev, B., Lev, D., Marcus, B., Fine, D., Su, D., Narkis, G., Ofir, R.,

Hoffmann, C., Leshinsky-Silver, E., Flusser, H., Sivan, S., Soll, D., Lerman-

Sagie, T. & Birk, O.S. (2010) Mutations disrupting selenocysteine formation

cause progressive cerebello-cerebral atrophy. Am J Hum Genet, 87, 538-544.

Beffert, U., Weeber, E.J., Durudas, A., Qiu, S., Masiulis, I., Sweatt, J.D., Li, W.P.,

Adelmann, G., Frotscher, M., Hammer, R.E. & Herz, J. (2005) Modulation of

synaptic plasticity and memory by Reelin involves differential splicing of the

lipoprotein receptor Apoer2. Neuron, 47, 567-579.

Behne, D., Weiler, H. & Kyriakopoulos, A. (1996) Effects of selenium deficiency on

testicular morphology and function in rats. J Reprod Fertil, 106, 291-297.

Bellinger, F.P., He, Q.P., Bellinger, M.T., Lin, Y., Raman, A.V., White, L.R. & Berry,

M.J. (2008) Association of selenoprotein p with Alzheimer's pathology in human

cortex. J Alzheimers Dis, 15, 465-472.

Bellinger, F.P., Raman, A.V., Reeves, M.A. & Berry, M.J. (2009) Regulation and

function of selenoproteins in human disease. Biochem J, 422, 11-22.

Bosschaerts, T., Guilliams, M., Noel, W., Herin, M., Burk, R.F., Hill, K.E., Brys, L.,

Raes, G., Ghassabeh, G.H., De Baetselier, P. & Beschin, A. (2008) Alternatively

activated myeloid cells limit pathogenicity associated with African

trypanosomiasis through the IL-10 inducible gene selenoprotein P. J Immunol,

180, 6168-6175.

Burk, R.F. & Hill, K.E. (2005) Selenoprotein P: an extracellular protein with unique

physical characteristics and a role in selenium homeostasis. Annu Rev Nutr, 25,

215-235.

Burk, R.F. & Hill, K.E. (2009) Selenoprotein P-expression, functions, and roles in

mammals. Biochim Biophys Acta, 1790, 1441-1447.

Burk, R.F., Hill, K.E., Olson, G.E., Weeber, E.J., Motley, A.K., Winfrey, V.P. & Austin,

L.M. (2007) Deletion of apolipoprotein E receptor-2 in mice lowers brain

selenium and causes severe neurological dysfunction and death when a low-

selenium diet is fed. J Neurosci, 27, 6207-6211.

Caito, S.W., Milatovic, D., Hill, K.E., Aschner, M., Burk, R.F. & Valentine, W.M.

(2011) Progression of neurodegeneration and morphologic changes in the brains

of juvenile mice with selenoprotein P deleted. Brain Res, 1398, 1-12.

Chiu-Ugalde, J., Theilig, F., Behrends, T., Drebes, J., Sieland, C., Subbarayal, P., Kohrle,

J., Hammes, A., Schomburg, L. & Schweizer, U. (2010) Mutation of megalin

leads to urinary loss of selenoprotein P and selenium deficiency in serum, liver,

kidneys and brain. Biochem J, 431, 103-111.

Combs, G.F., Jackson, M.I., Watts, J.C., Johnson, L.K., Zeng, H., Idso, J., Schomburg,

L., Hoeg, A., Hoefig, C.S., Chiang, E.C., Waters, D.J., Davis, C.D. & Milner, J.A.

(2011) Differential responses to selenomethionine supplementation by sex and

genotype in healthy adults. Br J Nutr, 1-12.

Crawley, J.N. (1999) Behavioral phenotyping of transgenic and knockout mice:

experimental design and evaluation of general health, sensory functions, motor

abilities, and specific behavioral tests. Brain Res, 835, 18-26.

57

Deagen, J.T., Butler, J.A., Beilstein, M.A. & Whanger, P.D. (1987) Effects of dietary

selenite, selenocystine and selenomethionine on selenocysteine lyase and

glutathione peroxidase activities and on selenium levels in rat tissues. J Nutr, 117,

91-98.

Esaki, N., Nakamura, T., Tanaka, H. & Soda, K. (1982) Selenocysteine lyase, a novel

enzyme that specifically acts on selenocysteine. Mammalian distribution and

purification and properties of pig liver enzyme. J Biol Chem, 257, 4386-4391.

Fan, Q.-W., Iosbe, I., Asou, H., Yanagisawa, K. & Michikawa, M. (2001) Expression and

regulation of apolipoprotein E receptors in the cells of the central nervous system

in culture: A review. Age (Dordr), 24, 1-10.

Filiou, M.D., Bisle, B., Reckow, S., Teplytska, L., Maccarrone, G. & Turck, C.W. (2010)

Profiling of mouse synaptosome proteome and phosphoproteome by IEF.

Electrophoresis, 31, 1294-1301.

Fleming, S.M., Salcedo, J., Fernagut, P.O., Rockenstein, E., Masliah, E., Levine, M.S. &

Chesselet, M.F. (2004) Early and progressive sensorimotor anomalies in mice

overexpressing wild-type human alpha-synuclein. J Neurosci, 24, 9434-9440.

Galan, P., Viteri, F.E., Bertrais, S., Czernichow, S., Faure, H., Arnaud, J., Ruffieux, D.,

Chenal, S., Arnault, N., Favier, A., Roussel, A.M. & Hercberg, S. (2005) Serum

concentrations of beta-carotene, vitamins C and E, zinc and selenium are

influenced by sex, age, diet, smoking status, alcohol consumption and corpulence

in a general French adult population. Eur J Clin Nutr, 59, 1181-1190.

Hill, K.E., Zhou, J., Austin, L.M., Motley, A.K., Ham, A.J., Olson, G.E., Atkins, J.F.,

Gesteland, R.F. & Burk, R.F. (2007) The selenium-rich C-terminal domain of

mouse selenoprotein P is necessary for the supply of selenium to brain and testis

but not for the maintenance of whole body selenium. J Biol Chem, 282, 10972-

10980.

Hill, K.E., Zhou, J., McMahan, W.J., Motley, A.K., Atkins, J.F., Gesteland, R.F. & Burk,

R.F. (2003) Deletion of selenoprotein P alters distribution of selenium in the

mouse. J Biol Chem, 278, 13640-13646.

Hill, K.E., Zhou, J., McMahan, W.J., Motley, A.K. & Burk, R.F. (2004) Neurological

dysfunction occurs in mice with targeted deletion of the selenoprotein P gene. J

Nutr, 134, 157-161.

Hoffmann, F.W., Hashimoto, A.C., Shafer, L.A., Dow, S., Berry, M.J. & Hoffmann, P.R.

(2010) Dietary selenium modulates activation and differentiation of CD4+ T cells

in mice through a mechanism involving cellular free thiols. J Nutr, 140, 1155-

1161.

Hoffmann, P.R., Hoge, S.C., Li, P.A., Hoffmann, F.W., Hashimoto, A.C. & Berry, M.J.

(2007) The selenoproteome exhibits widely varying, tissue-specific dependence

on selenoprotein P for selenium supply. Nucleic Acids Res, 35, 3963-3973.

Johnston, M.V. & Hagberg, H. (2007) Sex and the pathogenesis of cerebral palsy. Dev

Med Child Neurol, 49, 74-78.

Korschineck, I., Ziegler, S., Breuss, J., Lang, I., Lorenz, M., Kaun, C., Ambros, P.F. &

Binder, B.R. (2001) Identification of a novel exon in apolipoprotein E receptor 2

leading to alternatively spliced mRNAs found in cells of the vascular wall but not

in neuronal tissue. J Biol Chem, 276, 13192-13197.

58

Kurokawa, S., Takehashi, M., Tanaka, H., Mihara, H., Kurihara, T., Tanaka, S., Hill, K.,

Burk, R. & Esaki, N. (2011) Mammalian selenocysteine lyase is involved in

selenoprotein biosynthesis. J Nutr Sci Vitaminol (Tokyo), 57, 298-305.

Lacourciere, G.M., Mihara, H., Kurihara, T., Esaki, N. & Stadtman, T.C. (2000)

Escherichia coli NifS-like proteins provide selenium in the pathway for the

biosynthesis of selenophosphate. J Biol Chem, 275, 23769-23773.

Lein, E.S. & Hawrylycz, M.J. & Ao, N. & Ayres, M. & Bensinger, A. & Bernard, A. &

Boe, A.F. & Boguski, M.S. & Brockway, K.S. & Byrnes, E.J. & Chen, L. &

Chen, L. & Chen, T.M. & Chin, M.C. & Chong, J. & Crook, B.E. & Czaplinska,

A. & Dang, C.N. & Datta, S. & Dee, N.R. & Desaki, A.L. & Desta, T. & Diep, E.

& Dolbeare, T.A. & Donelan, M.J. & Dong, H.W. & Dougherty, J.G. & Duncan,

B.J. & Ebbert, A.J. & Eichele, G. & Estin, L.K. & Faber, C. & Facer, B.A. &

Fields, R. & Fischer, S.R. & Fliss, T.P. & Frensley, C. & Gates, S.N. &

Glattfelder, K.J. & Halverson, K.R. & Hart, M.R. & Hohmann, J.G. & Howell,

M.P. & Jeung, D.P. & Johnson, R.A. & Karr, P.T. & Kawal, R. & Kidney, J.M. &

Knapik, R.H. & Kuan, C.L. & Lake, J.H. & Laramee, A.R. & Larsen, K.D. &

Lau, C. & Lemon, T.A. & Liang, A.J. & Liu, Y. & Luong, L.T. & Michaels, J. &

Morgan, J.J. & Morgan, R.J. & Mortrud, M.T. & Mosqueda, N.F. & Ng, L.L. &

Ng, R. & Orta, G.J. & Overly, C.C. & Pak, T.H. & Parry, S.E. & Pathak, S.D. &

Pearson, O.C. & Puchalski, R.B. & Riley, Z.L. & Rockett, H.R. & Rowland, S.A.

& Royall, J.J. & Ruiz, M.J. & Sarno, N.R. & Schaffnit, K. & Shapovalova, N.V.

& Sivisay, T. & Slaughterbeck, C.R. & Smith, S.C. & Smith, K.A. & Smith, B.I.

& Sodt, A.J. & Stewart, N.N. & Stumpf, K.R. & Sunkin, S.M. & Sutram, M. &

Tam, A. & Teemer, C.D. & Thaller, C. & Thompson, C.L. & Varnam, L.R. &

Visel, A. & Whitlock, R.M. & Wohnoutka, P.E. & Wolkey, C.K. & Wong, V.Y.,

et al. (2007) Genome-wide atlas of gene expression in the adult mouse brain.

Nature, 445, 168-176.

Mayoral, S.R., Omar, G. & Penn, A.A. (2009) Sex differences in a hypoxia model of

preterm brain damage. Pediatr Res, 66, 248-253.

Meplan, C., Crosley, L.K., Nicol, F., Beckett, G.J., Howie, A.F., Hill, K.E., Horgan, G.,

Mathers, J.C., Arthur, J.R. & Hesketh, J.E. (2007) Genetic polymorphisms in the

human selenoprotein P gene determine the response of selenoprotein markers to

selenium supplementation in a gender-specific manner (the SELGEN study).

FASEB J, 21, 3063-3074.

Mihara, H., Kurihara, T., Watanabe, T., Yoshimura, T. & Esaki, N. (2000) cDNA

cloning, purification, and characterization of mouse liver selenocysteine lyase.

Candidate for selenium delivery protein in selenoprotein synthesis. J Biol Chem,

275, 6195-6200.

Misu, H., Takamura, T., Takayama, H., Hayashi, H., Matsuzawa-Nagata, N., Kurita, S.,

Ishikura, K., Ando, H., Takeshita, Y., Ota, T., Sakurai, M., Yamashita, T.,

Mizukoshi, E., Honda, M., Miyamoto, K., Kubota, T., Kubota, N., Kadowaki, T.,

Kim, H.J., Lee, I.K., Minokoshi, Y., Saito, Y., Takahashi, K., Yamada, Y.,

Takakura, N. & Kaneko, S. (2010) A liver-derived secretory protein,

selenoprotein P, causes insulin resistance. Cell Metab, 12, 483-495.

Morris, R. (1984) Developments of a water-maze procedure for studying spatial learning

in the rat. J Neurosci Methods, 11, 47-60.

59

Nakayama, A., Hill, K.E., Austin, L.M., Motley, A.K. & Burk, R.F. (2007) All regions of

mouse brain are dependent on selenoprotein P for maintenance of selenium. J

Nutr, 137, 690-693.

Peters, M.M., Hill, K.E., Burk, R.F. & Weeber, E.J. (2006) Altered hippocampus

synaptic function in selenoprotein P deficient mice. Mol Neurodegener, 1, 12.

Rayman, M.P. (2000) The importance of selenium to human health. Lancet, 356, 233-

241.

Renko, K., Werner, M., Renner-Muller, I., Cooper, T.G., Yeung, C.H., Hollenbach, B.,

Scharpf, M., Kohrle, J., Schomburg, L. & Schweizer, U. (2008) Hepatic

selenoprotein P (SePP) expression restores selenium transport and prevents

infertility and motor-incoordination in Sepp-knockout mice. Biochem J, 409, 741-

749.

Riese, C., Michaelis, M., Mentrup, B., Gotz, F., Kohrle, J., Schweizer, U. & Schomburg,

L. (2006) Selenium-dependent pre- and posttranscriptional mechanisms are

responsible for sexual dimorphic expression of selenoproteins in murine tissues.

Endocrinology, 147, 5883-5892.

Scharpf, M., Schweizer, U., Arzberger, T., Roggendorf, W., Schomburg, L. & Kohrle, J.

(2007) Neuronal and ependymal expression of selenoprotein P in the human

brain. J Neural Transm, 114, 877-884.

Schomburg, L., Schweizer, U., Holtmann, B., Flohe, L., Sendtner, M. & Kohrle, J. (2003)

Gene disruption discloses role of selenoprotein P in selenium delivery to target

tissues. Biochem J, 370, 397-402.

Schweizer, U., Michaelis, M., Kohrle, J. & Schomburg, L. (2004) Efficient selenium

transfer from mother to offspring in selenoprotein-P-deficient mice enables dose-

dependent rescue of phenotypes associated with selenium deficiency. Biochem J,

378, 21-26.

Schweizer, U., Streckfuss, F., Pelt, P., Carlson, B.A., Hatfield, D.L., Kohrle, J. &

Schomburg, L. (2005) Hepatically derived selenoprotein P is a key factor for

kidney but not for brain selenium supply. Biochem J, 386, 221-226.

Steinert, P., Bachner, D. & Flohe, L. (1998) Analysis of the mouse selenoprotein P gene.

Biol Chem, 379, 683-691.

Stoedter, M., Renko, K., Hog, A. & Schomburg, L. (2010) Selenium controls the sex-

specific immune response and selenoprotein expression during the acute-phase

response in mice. Biochem J, 429, 43-51.

Takahashi, Y., Hursting, S.D., Perkins, S.N., Wang, T.C. & Wang, T.T. (2006) Genistein

affects androgen-responsive genes through both androgen- and estrogen-induced

signaling pathways. Mol Carcinog, 45, 18-25.

Tobe, R., Mihara, H., Kurihara, T. & Esaki, N. (2009) Identification of proteins

interacting with selenocysteine lyase. Biosci Biotechnol Biochem, 73, 1230-1232.

Valentine, W.M., Abel, T.W., Hill, K.E., Austin, L.M. & Burk, R.F. (2008)

Neurodegeneration in mice resulting from loss of functional selenoprotein P or its

receptor apolipoprotein E receptor 2. J Neuropathol Exp Neurol, 67, 68-77.

Valentine, W.M., Hill, K.E., Austin, L.M., Valentine, H.L., Goldowitz, D. & Burk, R.F.

(2005) Brainstem axonal degeneration in mice with deletion of selenoprotein p.

Toxicol Pathol, 33, 570-576.

60

Zhang, Y., Zhou, Y., Schweizer, U., Savaskan, N.E., Hua, D., Kipnis, J., Hatfield, D.L. &

Gladyshev, V.N. (2008) Comparative analysis of selenocysteine machinery and

selenoproteome gene expression in mouse brain identifies neurons as key

functional sites of selenium in mammals. J Biol Chem, 283, 2427-2438.

61

FIGURE LEGENDS

Figure 1: Spontaneous activity and motor coordination is more reduced in male than

female Sepp1-/- mice compared to control mice fed a standard diet. (A) Rearing and

grooming were measured in the cylinder, and total distance traveled in the open field.

Sepp1-/- mice were less active as measured by rearing (left, genotype **p<0.01) and

grooming (center, genotype ***p<0.001), but did not show decreased exploration of the

open field (right). (B) Motor coordination was measured using the pole test and inverted

hang test. Sepp1-/- mice took longer to turn (left, genotype x sex ***p<0.001) and

descend the pole (center, genotype x sex *p<0.05), and had reduced ability to suspend

themselves for two minutes (right, genotype ***p<0.001). Values are expressed as means

± SEM. n=8-12 per group, *p<0.05, **p<0.01, ***p<0.001, compared with control mice.

Figure 2: High Se diet improves spontaneous activity and motor coordination more in

male than female Sepp1-/- mice compared to control mice. (A) Rearing and grooming

were measured in the cylinder, and total distance traveled in the open field. Se-

supplemented Sepp1-/- mice were as active as controls when measured by rearing (left),

grooming (center), and exploration of the open field (right), and no sex differences were

observed. (B) Motor coordination was measured using the pole test and inverted hang

test. Se-supplemented Sepp1-/- mice took longer to turn (left, genotype **p<0.01) and to

descend the pole (center, genotype *p<0.05) when compared with control mice, but were

capable of suspending themselves for two minutes (right). Male and female Sepp1-/-

mice performed similarly except for turn time, and an interaction effect between

genotype and sex was not statistically evident in any test. Values are expressed as means

± SEM. n=8-10 per group, ***p<0.001, compared with control mice.

Figure 3: Generation of Scly-knockout mice. (A) Image from KOMP. A promoterless

trapping cassette was inserted upstream of exon 4 of the mouse Scly locus on

chromosome 1, causing splicing at the cassette and truncation of the endogenous

transcript. The cassette was flanked by FRT sites for conditional excision of the cassette

by breeding with FLP-recombinase transgenic mice in case of embryonic lethality.

62

Presence of the loxP sites flanking exon 4 allowed excision of a coding exon critical for

enzymatic function by breeding with Cre-recombinase transgenic mice. (B) PCR

genotyping of mice tails was performed to detect presence of wild-type allele (1.2 kb) or

knockout allele (233 bp) using primers described in Materials and Methods. (C)

Quantitative RT-PCR analysis of brains from Scly+/+ and Scly-/- mice indicated no

detectable Scly mRNA in homozygous knockout mice.

Figure 4: Spontaneous activity and motor coordination is normal in male and female

Scly-/- mice fed a low Se diet. (A) Rearing and total distance traveled was measured in

the open field. Scly-/- mice were as active as controls when measured by exploration of

the open field (right), however an interaction between genotype and sex affected rearing

activity (left, genotype x sex *p<0.05). (B) Motor coordination was measured using the

pole test. Scly-/- mice and control mice performed similarly to turn (left) and descend the

pole (right). Male and female mice performed similarly, and no genotype effects were

present. Values are expressed as means ± SEM. n=6-7 per group.

Figure 5: Spatial learning and memory is disrupted in Sepp1-/- mice fed a standard diet.

(A) Average escape latency per training day over time was measured in the Morris Water

Maze. An interaction between Sepp1 genotype and training was present (*p<0.05) and

Sepp1-/- mice had significantly longer latency on days 7 and 8 (*p<0.05). (B) In the

probe trial, Sepp1-/- spent more time exploring the opposite (OP) quadrant (*p<0.05) and

we found a significant interaction effect (genotype x quadrant *p<0.05). They also had

non-significant trends toward fewer platform crossings (p=0.0557) (C) and a reduced

swim speed (p>0.05) (D). Values are expressed as means ± SEM. n=8-10 per group.

Figure 6: Spatial learning and memory is not disrupted in Scly-/- mice fed a standard diet.

(A) Average escape latency per training day over time was measured in the Morris Water

Maze. Scly-/- and control mice showed strongly reduced latency over time, and no

interaction between genotype and training was apparent. (B) In the probe trial at the end

of training, Scly-/- and control mice spent equal time exploring the target (TQ), left (LA)

and right adjacent (RA), and the opposite (OP) quadrants (p>0.05). Both genotypes had

63

similar platform crossings (C) and swim speed (D) (p>0.05). Values are expressed as

means ± SEM. n=11-13 per group.

Figure 7: Spatial learning is mildly impaired in Scly-/- mice fed a low Se diet. (A)

Average escape latency per training day over time was measured in the Morris Water

Maze. An interaction between Scly genotype and training was present (*p<0.05), and

Scly-/- mice had significantly longer latency on days 2, 3, 5 and 6 (*p<0.05). (B) In the

probe trial at the end of training, Scly-/- and control mice spent equal time exploring the

target (TQ), adjacent (LA, RA) and opposite (OP) quadrants (p>0.05). Both genotypes

had similar platform crossings (C) and swim speed (D) (p>0.05) in the probe trial. Values

are expressed as means ± SEM. n=13 per group.

Figure 8: Expression of selenoprotein transcripts is increased in Se-deficient Scly-/- mice

brains. (A) Sepp1 mRNA level was measured in brains of Scly-/- mice (n=8). We

additionally measured GPX1 (n=7) (B), GPX4 (n=8) (C), and Sepw1 (n=8) (D) mRNA

expression in brain of Scly-/- mice fed a low Se diet. Sepp1, GPX1, and GPX4 were

increased (*p<0.05, **p<0.01), while Sepw1 was unchanged. Samples were assayed in

triplicate or quadruplicate and values were normalized to 18s rRNA and expressed as

means ± SEM.

Figure 9: Expression of selenoproteins is decreased in Se-deficient Scly-/- mice brains.

(A) GPX1, (B) GPX4, and (C) Sepw1 protein was measured by western blot and

quantified by integrated intensity (n=8). Below each graph is a representative sample of

the blot including loading equivalence determined by tubulin. GPX1, GPX4, and Sepw1

were drastically reduced in Scly-/- mice (***p<0.0001). All selenoprotein values are

normalized to relative amounts of tubulin, and expressed as means ± SEM.

Figure 10: Glutathione peroxidase activity is decreased in Se-deficient Scly-/- mice

brains. Total GPX activity in brain (n=8) (A), liver (n=8) (B), and serum (n=12-14) (C)

was assayed with a coupled reaction measuring NADPH oxidation. Mice were fed the

low Se diet from weaning until time of sacrifice, when tissues were harvested for

64

analysis. Although both brain and liver GPX activity were reduced by approximately half

(***p<0.001), serum GPX activity was not significantly reduced. Values are standardized

to total protein concentration (brain, liver) or volume of serum and expressed as means ±

SEM.

65

FIGURE 1

Normal diet

66

FIGURE 2

High Se diet

67

FIGURE 3

68

FIGURE 4

Low Se diet

69

FIGURE 5

70

FIGURE 6

71

FIGURE 7

72

FIGURE 8

73

FIGURE 9

74

FIGURE 10

75

CHAPTER 3

EXPRESSION OF SELENOPROTEIN W IN NEURONS EXTENDS INTO

PROCESSES AND IS HIGHLY DEPENDENT ON SELENOPROTEIN P

ABSTRACT

During dietary selenium deprivation, bodily selenium is prioritized to the brain to

maintain selenoprotein expression by a process that depends on selenoprotein P.

Selenoprotein W is a small thioredoxin-like protein that is abundant in brain and muscle

tissues. Although peripheral expression of selenoprotein W is reduced by dietary

selenium deficiency, brain expression is maintained, suggesting it has an important

function in nervous tissue. We assessed the regional, cellular, and subcellular expression

of selenoprotein W in brains of wild-type mice and mice lacking selenoprotein P. We

found that selenoprotein W is widespread in neurons, processes, and neuropil of mouse

brain. Pyramidal neurons of somatosensory, motor, piriform, and cingulate cortex, and

CA1 and CA3 of hippocampus express high levels of selenoprotein W. Purkinje neurons

and their heavily branched dendritic arbors in cerebellum also express abundant

selenoprotein W. Analysis of synaptosome fractions indicated that selenoprotein W is

present at synapses, and expression is dramatically reduced in mice lacking selenoprotein

P. Several components of the selenoprotein synthesis machinery were also found in

isolated nerve terminals. These results indicate that widespread neuronal expression of

selenoprotein W relies on selenoprotein P for selenium. They further suggest that

selenoprotein W synthesis may occur in distal comparments of neurons, far removed

from the nucleus.

76

INTRODUCTION

Selenium (Se) is a trace micronutrient that is incorporated into antioxidant enzymes. Se is

unique among trace elements because it is covalently incorporated into proteins as the

amino acid selenocysteine (Sec). Biosynthesis of Sec and insertion of the residues into

polypeptides requires a Sec Insertion Sequence (SECIS) and several specific proteins to

reinterpret in-frame UGA codons in selenoprotein mRNAs as Sec incorporation sites

[reviewed in (Bellinger et al., 2009)]. Of the 25 primate (24 rodent) Sec-containing

selenoproteins that have been identified using bioinformatics, the glutathione peroxidase,

thioredoxin reductase, and iodothyronine deiodinase enzyme families are functionally

characterized.

Selenoprotein W (Sepw1) is the smallest mammalian selenoprotein and is one of the most

widely distributed selenoproteins across species in all domains of life (Lobanov et al.,

2009, Zhang & Gladyshev, 2008). It was putatively identified in the early 1970s as being

absent in muscle of myopathic lambs suffering from White Muscle disease, but was not

purified, cloned, and named until some 20 years later (Vendeland et al., 1993, Vendeland

et al., 1995, Whanger, 2000). White muscle disease is a Se-responsive muscular

dystrophy syndrome in ruminants, and was named because of the appearance of pale and

dry muscle, usually with longitudinal striations or chalky whiteness caused by abnormal

calcium deposition. Leg muscles are typically disrupted first, but all muscles, including

cardiac, can be affected.

Like most of the selenoproteins, Sepw1 is expected to be involved in oxidation-reduction

(redox) reactions. Indeed, it has been shown to act as a glutathione-dependent antioxidant

that protects cells from peroxide-mediated damage (Jeong et al., 2002). However, the

specific antioxidant function of Sepw1 has been disputed (Xiao-Long et al., 2010), and a

prominent role in cell signaling has also been proposed (Hawkes & Alkan, 2010). Sepw1

directly interacts with 14-3-3 proteins (Aachmann et al., 2007, Dikiy et al., 2007), and

siRNA knockdown of Sepw1 expression inhibits cell proliferation in a p53- and

p21-dependent mechanism (Hawkes et al., 2012).

77

In addition to expression in muscle and proliferating myoblasts, the Sepw1 gene is also

highly expressed in developing and adult mouse brain (Gu et al., 2000, Loflin et al.,

2006). However, unlike muscle, dietary Se depletion does not cause a reduction in Sepw1

levels in sheep or rat brain, despite reducing brain Se concentration and GPX activity

(Sun et al., 2001, Whanger, 2001). Selenoprotein P (Sepp1) supplies Se to the brain, and

mice deficient in Sepp1 have greatly reduced levels of Sepw1 mRNA and protein in brain

(Hoffmann et al., 2007). These data suggest that preferential retention of Sepw1 in brain

during dietary Se-deficiency is maintained by Sepp1. Regional analysis of Sepw1 mRNA

expression in mice brains suggests presence in neurons, with high expression in >90% of

brain regions (Zhang et al., 2008). Intriguingly, Sepw1 mRNA has also been identified in

processes of cultured central and peripheral neurons [(Willis et al., 2005, Willis et al.,

2007), see supplemental tables]. However, the protein expression and function of Sepw1

in brain remains largely unexplored.

In this report, we analyzed expression of Sepw1 protein in mouse brain and mouse

brain-derived primary cells. We report that Sepw1 protein expression is observed in

neurons of several brain regions including cortex, hippocampus, and cerebellum. Sepw1

immunoreactivity extends into the processes of these cells, and isolation of nerve

terminals by synaptosome preparations revealed the presence of Sepw1. We have also

identified several components of the selenoprotein synthesis machinery in isolated nerve

terminals. Additionally, expression of Sepw1 in synaptic and non-synaptic fractions was

reduced in Sepp1-deficient mice, despite no change in selenoprotein synthesis machinery.

Taken together these data suggest that Sepw1 is highly expressed in neurons and may be

locally synthesized in distal processes far removed from the nucleus, including synaptic

sites. The widespread neuronal expression of Sepw1 is dependent on Sepp1, although the

cells producing Sepp1 for Sepw1 expression are undetermined. The enzymatic role of

Sepw1 is unclear, but the combined data argue for a role in oxidant-mediated cell

signaling rather than detoxification of oxidizing agents.

78

MATERIALS AND METHODS

Primary cell culture: Glass bottom tissue culture plates (World Precision Instruments,

Sarasota, FL) were coated with 0.1 mg/ml laminin in 0.1 mg/ml poly-L-lysine solution

for 1 h, and then rinsed with PBS. Primary cells from cortex, hippocampus, and

cerebellum were harvested from postnatal day one C57BL/6 mice, gently dispersed by

trituration, and plated on coated dishes. Cultures were maintained at 5.0% CO2 and 5.0%

relative humidity in Neurobasal-A medium (Invitrogen) with 5% fetal bovine serum (FBS)

with the addition of 100 μM glutamate (Invitrogen) to reduce growth of glial cells and

enrich growth of neurons. B27 supplement (Invitrogen) was added to replace FBS after

24 h, and glutamate omitted from the media after 3 days. FBS lots were tested for Se

content (Bodycote, Santa Fe Springs, CA), and the selenium concentration of media

containing 10% FBS was 105 nM as determined using inductively coupled plasma-mass

spectrometry. B27 was tested for Se content (Bodycote) and the selenium concentration

of media containing 2% B27 was 93.8 nM by the same method.

Animals: C57BL/6 mice and genetically modified male mice on a C57BL/6 background

lacking Sepp1 were bred on commercially available diets containing adequate Se (~0.25

ppm). Animals were given food and water ad libitum on a 12-hour light-cycle and group

housed until experimentation. All experiments were conducted on adult mice aged 3-4

months during the light cycle. All animal procedures and experimental protocols were

approved by the University of Hawaii Institutional Animal Care and Use Committee.

Tissue Preparation: Mice were anesthetized with ketamine-xylazine, and sacrificed by

transcardial perfusion. Mice were initially perfused with phosphate-buffered saline (PBS)

to flush out blood, followed by perfusion with 4% paraformaldehyde (PFA) to fix the

tissue. The mice heads were cut off and the brains dissected out. Brains were post-fixed

in 4% PFA overnight, followed by cryoprotecting the tissue in 10% and 30% sucrose for

at least 4 hours each. The brains were then embedded in optimal cutting temperature

(OCT) compound and frozen until time of sectioning. 40 μm sections were cut on a Leica

CM1900 cryostat and saved in cryoprotectant solution, containing 0.1 M phosphate buffer,

79

30% sucrose (w/v), and 30% ethylene glycol (v/v), until time of further experimentation.

Immunohistochemistry: Primary cortical, hippocampal, and cerebellar cultures maintained

for three weeks in vitro were used for immunolabeling. Brain sections stored in

cryoprotectant in the freezer were warmed to room temperature, and sections containing

cortex and hippocampus were selected for analysis. After thorough washing, sections

were blocked in 5% normal goat serum with 0.3% Triton X-100 in PBS. After blocking,

the sections were incubated in diluted primary antibody solution overnight at 4C. The

following antibodies were used: Rabbit-anti-Sepw1 (Rockland) and Mouse-anti-Tuj1

(Covance). A control section where primary antibody was omitted was also included in

the procedure. After washing out primary antibody, sections were incubated in

species-matched secondary antibody. The secondary antibody was directly conjugated to

fluorophores (Alexa Fluor dyes, Invitrogen) for fluorescence imaging. Additonally, some

sections were dual-labeled with a fluorescent Nissl stain to label neurons (Neurotrace,

Invitrogen). Sections were then mounted onto slides, and coverslipped in VectaShield

containing DAPI for fluorescent labeling of nuclei. Additional sections were

colorimetrically developed using 3,3’-diaminobenzidene (DAB), after signal

amplification using the avidin-biotin complex method (Vector), and coverslipped using

Permount.

Synaptosome Preparation: Synaptosomes were prepared by the method of Dunkley

(Dunkley et al., 1986). Mice were anesthetized with Tribromoethanol and sacrificed by

decapitation. The brain was rapidly excised, rinsed in ice-cold 0.32 M sucrose, and

immersed in ice-cold 0.32 M sucrose with 1 mM EDTA. Brain tissue was homogenized

in 5 ml of ice-cold sucrose/EDTA by 10 strokes at 900 rpm using a pre-chilled

Teflon/glass homogenizer. The homogenate was centrifuged at 3,600 rpm for 10 minutes

at 4C in polycarbonate tubes. The resulting supernatant was collected and diluted with

sucrose/EDTA to a total volume of 9 ml, and the pellet was resuspended in sucrose/EDTA

and saved for whole cell lysis. Approximately 3 ml of diluted supernatant was loaded on

the top of a discontinuous three layer Percoll gradient. Three gradients per brain were

made by adding 2 ml of 23% Percoll to each polycarbonate tube, and slowly layering 2

80

ml each of 10% and 3% Percoll sequentially using a peristaltic pump. The gradients with

sample were centrifuged at 20,000 rpm for 5 minutes at 4C to isolate synaptosomes.

Isolated synaptosomes were collected from the interface band between the 23% and 10%

Percoll layers in each gradient, and transferred and pooled directly to a large

polycarbonate centrifuge tube. To wash synaptosomes, 25 ml of HEPES-buffered saline

(HBS) was added to the tube, and was centrifuged at 15,000 rpm for 10 minutes at 4C.

The pellet was resuspended in HBS, and centrifuged at 7,000 rpm for 7 minutes at 4C.

The final pellet was resuspended in HBS for analysis by SDS-PAGE followed by western

blotting for select selenoproteins and related factors.

SDS-PAGE and Western blot: Total protein was extracted from whole-cell lysates by light

sonication in CelLytic MT buffer (Sigma, St. Louis, MO, USA) containing DTT, EDTA,

and protease inhibitors, followed by centrifugation according to the manufacturers’

protocol. Synaptosomes were resuspeneded in CelLytic MT buffer without sonication or

centrifugation. Protein was added to reduced Laemmli buffer, boiled for 10 minutes, and

loaded into 4-20% gradient polyacrylamide gels (Bio-Rad, Hercules, CA, USA).

Following electrophoresis, gel contents were transferred to PVDF membranes, which

were blocked with undiluted Odyssey Blocking Buffer (Li-Cor Biosciences, Lincoln, NE,

USA) for one hour. Membranes were then probed for 90 minutes with one of the

following primary antibodies: Rabbit-anti-GPX4 (AbFrontier, Seoul, Korea),

Rabbit-anti-SEPW1 (Rockland, Gilbertsville, PA, USA), Rabbit-anti-SEPHS2 (Rockland),

Rabbit-anti-SEPHS1 (Rockland), Rabbit-anti-SecP43 (Santa Cruz Biotech, Santa Cruz,

CA, USA), Goat-anti-SBP2 (Everest Biotech, Oxfordshire, UK), Rabbit-anti-EFSec

(AbCam, Cambridge, MA, USA), Mouse-anti-Syntaxin1 (Santa Cruz Biotechnology),

Mouse-anti-PSD95 (Thermo Scientific), Mouse-anti-TBP (AbCam), Mouse-anti-beta

actin (Sigma, St. Louis, MO, USA) and Mouse-anti-alpha Tubulin (Novus, Littleton, CO,

USA). After washing with PBS containing 0.05% tween-20 (PBST), membranes were

incubated in the dark in secondary antibodies labeled with infrared fluorophores (Li-Cor

Biosciences). After further washes in PBST, blots were imaged and quantified with the

Odyssey infrared imaging system (Li-Cor Biosciences).

81

Imaging: Fluorescence imaging of primary cultures and stained sections were performed

on a Zeiss LSM 5 Pascal laser confocal inverted microscope equipped Ar and HeNe

lasers. AlexaFluor-488, -546, and -633 secondary antibodies with directly conjugated

fluorophores were used to detect primary antibody signals. Images were acquired using

the included LSM software, and were analyzed using ImageJ. Bright-field imaging was

performed using an upright Zeiss AxioScope 2 Plus microscope equipped with an ASI

motorized stage and Zeiss Axiocam MRc camera.

Statistical analysis: Data were analyzed using Microsoft Excel (Redmond, WA, USA),

and plotted using GraphPad Prism software (San Diego, CA, USA). Unpaired t-tests

comparing genotypes were performed for protein expression in synaptosome experiments.

The significance criteria were set at p < 0.05 for statistical measures.

82

RESULTS

We sought to determine whether selenoprotein W (Sepw1) is expressed in neurons in

mice. To address this question, we first cultured primary cells harvested from neonatal

mouse brain, and assessed expression of Sepw1 along with a neuron-specific marker,

class III beta-tubulin (Tuj1). As seen in figure 1, primary cultures consisted primarily of

neurons, and Tuj1 immunoreactivity (magenta, middle) showed some overlap with

Sepw1 expression (green, left) in neurites. Primary neuronal cultures derived from

neonatal cortex (Fig. 1A), hippocampus (Fig. 1B) and cerebellum (Fig. 1C) all contained

robust Sepw1 expression, and some colocalization with Tuj1, as indicated by white color

in the merged panels (Fig. 1, right). Sepw1 immunoreactivity was also observed in some

Tuj1-negative cells, possibly indicating expression in glia.

To confirm that Sepw1 was expressed in adult mouse brain, we performed dual label

fluorescent imaging of fixed mouse brain sections using antibodies directed against

Sepw1 and Tuj1 or a fluorescent Nissl stain. In somatosensory cortex we observed robust

Sepw1 expression in large pyramidal neurons and extensive colocalization with Tuj1 (Fig.

2A). In CA1 of hippocampus (Fig. 2B), the pyramidal layer showed similar

immunolabeling of Sepw1. Intriguingly, Sepw1 expression extended into the apical and

basolateral dendrites of most pyramidal neurons, and was apparent in some axonal

compartments as well (Fig. 2, right). To further analyze regional expression of neuronal

Sepw1, we performed immunohistochemistry with DAB development. Extending the

previous results in somatosensory cortex and CA1, Sepw1 expression was observed in

somas and dendrites of motor cortex (Fig. 3A, left) and CA3 of hippocampus (Fig. 3A,

right). Additionally, piriform cortex (Fig. 3B) and cingulate cortex (Fig. 3C) displayed

high Sepw1 immunoreactivity in pyramidal neurons. Purkinje neurons of cerebellum (Fig.

3D), and their highly branched dendritic arbors, also showed abundant expression of

Sepw1. In fact, most neurons appeared to express Sepw1 to some degree and neuropil

generally appeared immunopositive for Sepw1. Conspicuously, large neurons showed

immunoreactivity in neuritic processes. The widespread expression in neurons of adult

mouse brain, along with the punctate staining in varicose segments of cultured neurons,

83

led us to question if Sepw1 is expressed in the synaptic compartment.

To assess if Sepw1 is expressed in synapses, we prepared synaptosomes from adult mice

and performed western blotting of the purified samples. After confirming the absence of

contaminating nuclear proteins by blotting for TATA-binding protein (TBP) (Fig. 4A), we

blotted for Sepw1 and found it to be present in synaptosomes (Fig. 4B). Sepw1 mRNA is

known to be reduced in the brain of Sepp1-/- mice (Hoffman, 2007). To determine if

synaptically expressed Sepw1 is reduced in the absence of Sepp1, we next prepared

synaptosomes from littermate Sepp1-/- and wild-type mice. We observed a dramatic

decrease in Sepw1 expression in synaptosomes isolated from Sepp1-/- mice compared to

control mice (Fig. 5A). Additionally, western blot analysis of Gpx4 showed presence in

wild-type synaptosomes, and slightly reduced expression in Sepp1-/- synaptosomes (Fig.

5B). We used beta actin to control for loading across samples (Fig. 5C). After

normalizing to Actin, quantification of selenoprotein expression revealed that Sepw1 was

significantly reduced to <25% of wild-type levels, while GPX4 was not significantly

reduced in synaptosomes (Fig. 5D-E).We did not observe a qualitative enrichment of

Sepw1 or Gpx4 in synaptosomes versus whole-cell lysate samples, nor did we observe

selective depletion of Sepw1 in either fraction in Sepp1 knockout mice. These findings

indicate that Sepw1 is roughly homogenously distributed throughout neurons, and all

compartments similarly depend on Sepp1 to maintain expression.

In addition to expression in the neuronal somata, Sepw1 mRNA has been detected in

axons, dendrites and neuropil, suggesting that it may be locally translated in neuronal

processes (Cajigas et al., 2012, Willis et al., 2007). However, selenoprotein synthesis is

unique, and requires several additional protein factors beyond the standard translation

machinery. We sought to assess if translation of selenoproteins can occur in distal

processes of neurons. Therefore, we did western blotting of synaptosomes for several

proteins involved in selenoprotein translation after confirming absence of nuclear

contamination by analyzing TBP (Fig. 6A). Both the Sec-specific elongation factor

(EFSec) (Fig. 6B) and the SECIS-binding protein 2 (Sbp2) (Fig. 6D) are absolutely

required for selenoprotein translation, and appeared to be present in synaptosomes in

84

addition to whole cell lysate samples. Selenophosphate, produced by the selenoenzyme

selenophosphate synthetase 2 (SPS2), acts as the Se-donor during selenoprotein

translation. We were able to detect the presence of SPS2 in synaptosomes (Fig. 6C).

Interestingly, EFSec and SPS2 appeared to be enriched in synaptosome fractions

compared to whole cell lysate samples, further suggesting the existence of selenoprotein

translation in this compartment. The Sec-tRNA associated-protein, SecP43, and

selenocysteine lyase (Scly) have been implicated in selenoprotein translation efficiency,

but are not absolutely required (Kurokawa et al., 2011, Squires & Berry, 2008). We were

unable to detect SecP43 in synaptosomes, despite robust expression in whole-cell lysates

(Fig. 6E). Conversely, Scly did appear to be present in synaptosomes in similar

abundance as whole cell lysates (Fig. 6F). Unlike Sepw1, and Gpx4 to a lesser degree,

none of the proteins involved in selenoprotein synthesis were altered in Sepp1-/- mice,

compared to wild-type controls. This is especially curious for SPS2, which is itself a

selenoprotein. The combination of presented data argues that translation of selenoproteins

Sepw1 and Gpx4 in synapses may be possible.

85

DISCUSSION

The results reported herein describe the first characterization of regional selenoprotein W

(Sepw1) localization in mouse brain, as well as selenoprotein and synthesis factor

expression in isolated nerve terminals. Sepw1 is abundantly expressed in neuronal somata

and neuropil, and is expressed along with several selenoprotein synthesis proteins in

synaptosome fractions. In all regions of brain in mice lacking selenoprotein P (Sepp1),

Sepw1 expression is drastically reduced without effect on synthesis factors, indicating

that Sepp1 facilitates Sepw1 synthesis.

Selenium (Se) is a trace micronutrient that is incorporated into the unique amino acid,

selenocysteine (Sec). Sec-containing selenoproteins are typically oxidoreductase enzymes

that play crucial roles in reducing reactive oxygen species and oxidized macromolecules.

A selenoprotein that is widely distributed across all domains of life, Sepw1, is particularly

abundant in brain and muscle of mammals (Gu et al., 2000). Sepw1 mRNA expression is

observed in cephalic neural folds and somites in developing rodents, with continued high

expression as they become the adult brain and skeletal muscles (Loflin et al., 2006).

Sepw1 was initially identified due to its absence in muscle of myopathic Se-deficient

lambs, however brain expression of Sepw1, unlike in muscle, is not depleted by dietary

Se deficiency (Whanger, 2001). However, Sepp1-deficient mice show reduced Sepw1

mRNA and protein in brain (Hoffmann et al., 2007).

Sepw1 is the smallest described mammalian selenoprotein at ~10 kDa and contains an

N-terminal thioredoxin-like Cys-X-X-Sec redox motif, where X is any amino acid

(Lobanov et al., 2009). As with all selenoproteins, the Sec residue is encoded by a UGA

codon in the mRNA. A Sec Insertion Sequence (SECIS) in the 3'UTR of the mRNA, the

SECIS binding protein SBP2, and the Sec-specific elongation factor EFSec help to

bypass translation termination and incorporate Sec during translation [reviewed in

(Squires & Berry, 2008)]. Sepw1 also has another conserved Cys residue in the

N-terminal region that is known to bind glutathione (GSH) (Beilstein et al., 1996, Gu et

al., 1999). Antioxidant function attributed to Sepw1 is GSH dependent. In vitro

86

experimental studies, which increased or decreased Sepw1 expression, have demonstrated

elevated and reduced resistance to oxidizing agents but only in the presence of reduced

GSH (Jeong et al., 2002). However, other studies demonstrate that siRNA knockdown of

Sepw1 causes increased enzyme activities of glutathione peroxidase, superoxide

dismutase, and catalase and total antioxidative capability and glutathione level in cultured

muscle cells, which prevents oxidant-induced apoptosis (Xiao-Long et al., 2010). These

authors suggested a role for Sepw1 in the antioxidative system that is not direct peroxide

detoxification. Therefore, the in situ enzymatic role of Sepw1 has remained elusive.

Sepw1 mRNA rapidly declines in response to peroxide, suggesting that it has a role in

oxidative metabolism. Similar to the metabolic enzyme glyceraldehyde phosphate

dehydrogenase (GAPDH), oxidative inactivation of Sepw1 may be involved in rerouting

carbohydrate flux from glycolysis to the pentose phosphate pathway, stimulating NADPH

generation and reducing the intracellular pool of GSH (Loflin et al., 2006).

A pull down experiment indicated that Sepw1 interacts with the cytoskeletal microtubule

protein tubulin (Dikiy et al., 2007). Our data show robust colocalization of Sepw1 with

the neuron-specific beta tubulin, Tuj1. Sepw1 was additionally shown to

immunoprecipitate specifically with the beta and gamma isoforms of the 14-3-3 family of

scaffolding proteins (Aachmann et al., 2007). A computational study explored a putative

reaction mechanism, whereby Sepw1 regulates the oxidation state of a conserved and

solvent exposed Cys residue of 14-3-3 beta and gamma. Sepw1 is suggested to reduce

oxidized Cys-Sulfenic acid of 14-3-3 back to its parental thiol using the Cys-X-X-Sec

motif in combination with the bound GSH moiety (Musiani et al., 2010). 14-3-3 proteins

are multifunctional proteins that coordinate the interaction of kinases and phosphatases

with other regulatory proteins, thereby affecting phosphorylation-dependent cellular

processes (Fu et al., 2000). Like SEPW1 gene expression, the YWHAG gene for 14-3-3

gamma, is highly expressed in brain, skeletal muscle, and heart in humans (Horie et al.,

1999).

Sepw1 has also been implicated in regulating growth factor-stimulated control of cell

cycle-entry in epithelial cells. Knockdown of Sepw1 by siRNA in breast and prostate

87

epithelial cells inhibits EGF-stimulated G1/S transition via nuclear accumulation of p53,

leading to induction of p21 and G1 arrest (Hawkes & Alkan, 2011, Hawkes et al., 2012).

Intriguingly, the stability of the EGFR as well as tyrosine phosphorylation was decreased

by knockdown of Sepw1 (Hawkes, personal communication). EGFR tyrosine

phosphorylation and reduction of the EGFR turnover rate is a process that depends on

EGF-induced hydrogen peroxide production, which may oxidatively inactivate protein

tyrosine phosphatases (Deyulia & Carcamo, 2005, Deyulia et al., 2005). Since 14-3-3

proteins are established in regulating phosphorylation-mediated cell signaling, Sepw1

may function in oxidative signal transduction reactions from receptors to target proteins

via reactive oxygen intermediates.

High muscle expression of Sepw1 mRNA is associated with myoblasts, and expression is

decreased in differentiated myotubes (Loflin et al., 2006). Thus, the abundance of Sepw1

mRNA and protein in post-mitotic neurons is mysterious. Sepw1 mRNA and protein are

widely expressed in neurons, including apparent expression in axonal and dendritic

compartments (Cajigas et al., 2012, Willis et al., 2005, Willis et al., 2007). Whether

translation of Sepw1 occurs in these distal cellular compartments is uncertain.

Selenoprotein translation in mammals specifically requires the proteins SBP2 and EFSec,

in addition to the standard translation machinery. Both of these proteins were identified in

synaptosomes, along with SPS2 and SCLY which are important in Sec metabolism. Thus,

the only major protein involved in selenoprotein translation that was not investigated in

this study is the Sec-synthetase enzyme, SepSecS. SepSecS is also known as soluble liver

antigen/liver pancreas antigen (SLA/LP) and is required to generate the Sec-loaded

tRNASec (Palioura et al., 2010). We were unable to test for the presence of SepSecS in

synaptosomes.

Selenoprotein mRNAs are thought to be packaged into mRNP complexes, which aid in

preventing nonsense codon-mediated decay (NMD) of transcripts with a Sec-specifying

UGA that could be interpreted as a premature termination codon. Two mRNA binding

proteins important for nervous system function, DJ-1/Park7 and Staufen 2, have been

experimentally demonstrated to bind Sepw1 mRNA (Blackinton et al., 2009, Furic et al.,

88

2008, Maher-Laporte & Desgroseillers, 2010, Van Der Brug et al., 2008). DJ-1 is a

multifunctional redox-sensitive protein that is associated with Parkinson’s disease, other

neurodegenerative disorders, and cancer (Kahle et al., 2009). Two isoforms of staufen,

Stau1 and Stau2, are known to direct subcellular localization of mRNAs, and have

recently been implicated in synaptic plasticity (Lebeau et al., 2011a, Lebeau et al., 2011b).

Both DJ-1 and Stau2 proteins have shown varying degrees of localization to synapses,

axons and dendrites, further suggesting the local regulation of Sepw1 expression in distal

compartments of neurons (Jeong et al., 2007, Olzmann et al., 2007, Price et al., 2006,

Usami et al., 2011). Lastly, Stau2 is reported to interact with the key NMD protein Upf1,

which is implicated in selenoprotein translation because their mRNAs contain in-frame

termination codons. Intriguingly, Stau2 may upregulate protein expression of target

mRNAs in a Upf1-dependent mechanism, while upregulation of the target mRNAs is

Upf1-independent (Miki et al., 2011). These mechanisms may ultimately provide high

specificity and sensitivity for the local expression of Sepw1 in neurons.

In sum, we have shown that Sepw1 expression in mouse is abundant in neurons of several

brain regions, including cingulate and piriform cortex, hippocampus, and Purkinje cells

of cerebellum. We also showed Sepw1 expression in neuronal processes and some

colocalization with tubulin. Analysis of isolated nerve terminals further revealed the

presence of Sepw1 and much of the selenoprotein synthesis machinery in synaptic

compartments. Sepw1 expression in synaptosomes and whole-cell lysates of brain was

drastically reduced in Sepp1 knockout mice. Combined with previous reports

documenting Sepw1 mRNA expression in neuronal processes and association of Sepw1

transcripts with mRNA-binding proteins, regulation of Sepw1 expression in pre- and/or

post-synaptic compartments is suggested and warrants further investigation.

89

REFERENCES

Aachmann, F.L., Fomenko, D.E., Soragni, A., Gladyshev, V.N. & Dikiy, A. (2007)

Solution structure of selenoprotein W and NMR analysis of its interaction with

14-3-3 proteins. J Biol Chem, 282, 37036-37044.

Beilstein, M.A., Vendeland, S.C., Barofsky, E., Jensen, O.N. & Whanger, P.D. (1996)

Selenoprotein W of rat muscle binds glutathione and an unknown small molecular

weight moiety. J Inorg Biochem, 61, 117-124.

Bellinger, F.P., Raman, A.V., Reeves, M.A. & Berry, M.J. (2009) Regulation and function

of selenoproteins in human disease. Biochem J, 422, 11-22.

Blackinton, J., Kumaran, R., van der Brug, M.P., Ahmad, R., Olson, L., Galter, D., Lees,

A., Bandopadhyay, R. & Cookson, M.R. (2009) Post-transcriptional regulation of

mRNA associated with DJ-1 in sporadic Parkinson disease. Neuroscience letters,

452, 8-11.

Cajigas, I.J., Tushev, G., Will, T.J., Tom Dieck, S., Fuerst, N. & Schuman, E.M. (2012)

The local transcriptome in the synaptic neuropil revealed by deep sequencing and

high-resolution imaging. Neuron, 74, 453-466.

DeYulia, G.J., Jr. & Carcamo, J.M. (2005) EGF receptor-ligand interaction generates

extracellular hydrogen peroxide that inhibits EGFR-associated protein tyrosine

phosphatases. Biochemical and biophysical research communications, 334, 38-42.

DeYulia, G.J., Jr., Carcamo, J.M., Borquez-Ojeda, O., Shelton, C.C. & Golde, D.W. (2005)

Hydrogen peroxide generated extracellularly by receptor-ligand interaction

facilitates cell signaling. Proceedings of the National Academy of Sciences of the

United States of America, 102, 5044-5049.

Dikiy, A., Novoselov, S.V., Fomenko, D.E., Sengupta, A., Carlson, B.A., Cerny, R.L.,

Ginalski, K., Grishin, N.V., Hatfield, D.L. & Gladyshev, V.N. (2007) SelT, SelW,

SelH, and Rdx12: genomics and molecular insights into the functions of

selenoproteins of a novel thioredoxin-like family. Biochemistry, 46, 6871-6882.

Dunkley, P.R., Jarvie, P.E., Heath, J.W., Kidd, G.J. & Rostas, J.A. (1986) A rapid method

for isolation of synaptosomes on Percoll gradients. Brain research, 372, 115-129.

Fu, H., Subramanian, R.R. & Masters, S.C. (2000) 14-3-3 proteins: structure, function,

and regulation. Annu Rev Pharmacol Toxicol, 40, 617-647.

Furic, L., Maher-Laporte, M. & DesGroseillers, L. (2008) A genome-wide approach

identifies distinct but overlapping subsets of cellular mRNAs associated with

Staufen1- and Staufen2-containing ribonucleoprotein complexes. RNA, 14,

324-335.

Gu, Q.P., Beilstein, M.A., Barofsky, E., Ream, W. & Whanger, P.D. (1999) Purification,

characterization, and glutathione binding to selenoprotein W from monkey muscle.

Arch Biochem Biophys, 361, 25-33.

Gu, Q.P., Sun, Y., Ream, L.W. & Whanger, P.D. (2000) Selenoprotein W accumulates

primarily in primate skeletal muscle, heart, brain and tongue. Molecular and

cellular biochemistry, 204, 49-56.

Hawkes, W.C. & Alkan, Z. (2010) Regulation of redox signaling by selenoproteins. Biol

Trace Elem Res, 134, 235-251.

Hawkes, W.C. & Alkan, Z. (2011) Delayed cell cycle progression from SEPW1 depletion

90

is p53- and p21-dependent in MCF-7 breast cancer cells. Biochemical and

biophysical research communications, 413, 36-40.

Hawkes, W.C., Printsev, I. & Alkan, Z. (2012) Selenoprotein W depletion induces a p53-

and p21-dependent delay in cell cycle progression in RWPE-1 prostate epithelial

cells. Journal of cellular biochemistry, 113, 61-69.

Hoffmann, P.R., Hoge, S.C., Li, P.A., Hoffmann, F.W., Hashimoto, A.C. & Berry, M.J.

(2007) The selenoproteome exhibits widely varying, tissue-specific dependence

on selenoprotein P for selenium supply. Nucleic Acids Res, 35, 3963-3973.

Horie, M., Suzuki, M., Takahashi, E. & Tanigami, A. (1999) Cloning, expression, and

chromosomal mapping of the human 14-3-3gamma gene (YWHAG) to 7q11.23.

Genomics, 60, 241-243.

Jeong, D., Kim, T.S., Chung, Y.W., Lee, B.J. & Kim, I.Y. (2002) Selenoprotein W is a

glutathione-dependent antioxidant in vivo. FEBS letters, 517, 225-228.

Jeong, J.H., Nam, Y.J., Kim, S.Y., Kim, E.G., Jeong, J. & Kim, H.K. (2007) The transport

of Staufen2-containing ribonucleoprotein complexes involves kinesin motor

protein and is modulated by mitogen-activated protein kinase pathway. Journal of

neurochemistry, 102, 2073-2084.

Kahle, P.J., Waak, J. & Gasser, T. (2009) DJ-1 and prevention of oxidative stress in

Parkinson's disease and other age-related disorders. Free Radic Biol Med, 47,

1354-1361.

Kurokawa, S., Takehashi, M., Tanaka, H., Mihara, H., Kurihara, T., Tanaka, S., Hill, K.,

Burk, R. & Esaki, N. (2011) Mammalian selenocysteine lyase is involved in

selenoprotein biosynthesis. Journal of nutritional science and vitaminology, 57,

298-305.

Lebeau, G., DesGroseillers, L., Sossin, W. & Lacaille, J.C. (2011a) mRNA binding

protein staufen 1-dependent regulation of pyramidal cell spine morphology via

NMDA receptor-mediated synaptic plasticity. Molecular brain, 4, 22.

Lebeau, G., Miller, L.C., Tartas, M., McAdam, R., Laplante, I., Badeaux, F.,

DesGroseillers, L., Sossin, W.S. & Lacaille, J.C. (2011b) Staufen 2 regulates

mGluR long-term depression and Map1b mRNA distribution in hippocampal

neurons. Learn Mem, 18, 314-326.

Lobanov, A.V., Hatfield, D.L. & Gladyshev, V.N. (2009) Eukaryotic selenoproteins and

selenoproteomes. Biochim Biophys Acta, 1790, 1424-1428.

Loflin, J., Lopez, N., Whanger, P.D. & Kioussi, C. (2006) Selenoprotein W during

development and oxidative stress. J Inorg Biochem, 100, 1679-1684.

Maher-Laporte, M. & DesGroseillers, L. (2010) Genome wide identification of

Staufen2-bound mRNAs in embryonic rat brains. BMB reports, 43, 344-348.

Miki, T., Kamikawa, Y., Kurono, S., Kaneko, Y., Katahira, J. & Yoneda, Y. (2011) Cell

type-dependent gene regulation by Staufen2 in conjunction with Upf1. BMC

molecular biology, 12, 48.

Musiani, F., Ciurli, S. & Dikiy, A. (2010) The interaction of selenoprotein W with 14-3-3

proteins: a computational approach. J Proteome Res.

Olzmann, J.A., Bordelon, J.R., Muly, E.C., Rees, H.D., Levey, A.I., Li, L. & Chin, L.S.

(2007) Selective enrichment of DJ-1 protein in primate striatal neuronal processes:

implications for Parkinson's disease. The Journal of comparative neurology, 500,

585-599.

91

Palioura, S., Herkel, J., Simonovic, M., Lohse, A.W. & Soll, D. (2010) Human SepSecS

or SLA/LP: selenocysteine formation and autoimmune hepatitis. Biological

chemistry, 391, 771-776.

Price, T.J., Flores, C.M., Cervero, F. & Hargreaves, K.M. (2006) The RNA binding and

transport proteins staufen and fragile X mental retardation protein are expressed

by rat primary afferent neurons and localize to peripheral and central axons.

Neuroscience, 141, 2107-2116.

Squires, J.E. & Berry, M.J. (2008) Eukaryotic selenoprotein synthesis: mechanistic

insight incorporating new factors and new functions for old factors. IUBMB Life,

60, 232-235.

Sun, Y., Butler, J.A. & Whanger, P.D. (2001) Glutathione peroxidase activity and

selenoprotein W levels in different brain regions of selenium-depleted rats(1). J

Nutr Biochem, 12, 88-94.

Usami, Y., Hatano, T., Imai, S., Kubo, S., Sato, S., Saiki, S., Fujioka, Y., Ohba, Y., Sato, F.,

Funayama, M., Eguchi, H., Shiba, K., Ariga, H., Shen, J. & Hattori, N. (2011)

DJ-1 associates with synaptic membranes. Neurobiology of disease, 43, 651-662.

van der Brug, M.P., Blackinton, J., Chandran, J., Hao, L.Y., Lal, A., Mazan-Mamczarz, K.,

Martindale, J., Xie, C., Ahmad, R., Thomas, K.J., Beilina, A., Gibbs, J.R., Ding, J.,

Myers, A.J., Zhan, M., Cai, H., Bonini, N.M., Gorospe, M. & Cookson, M.R.

(2008) RNA binding activity of the recessive parkinsonism protein DJ-1 supports

involvement in multiple cellular pathways. Proceedings of the National Academy

of Sciences of the United States of America, 105, 10244-10249.

Vendeland, S.C., Beilstein, M.A., Chen, C.L., Jensen, O.N., Barofsky, E. & Whanger, P.D.

(1993) Purification and properties of selenoprotein W from rat muscle. J Biol

Chem, 268, 17103-17107.

Vendeland, S.C., Beilstein, M.A., Yeh, J.Y., Ream, W. & Whanger, P.D. (1995) Rat

skeletal muscle selenoprotein W: cDNA clone and mRNA modulation by dietary

selenium. Proceedings of the National Academy of Sciences of the United States

of America, 92, 8749-8753.

Whanger, P.D. (2000) Selenoprotein W: a review. Cellular and molecular life sciences :

CMLS, 57, 1846-1852.

Whanger, P.D. (2001) Selenium and the brain: a review. Nutritional neuroscience, 4,

81-97.

Willis, D., Li, K.W., Zheng, J.Q., Chang, J.H., Smit, A.B., Kelly, T., Merianda, T.T.,

Sylvester, J., van Minnen, J. & Twiss, J.L. (2005) Differential transport and local

translation of cytoskeletal, injury-response, and neurodegeneration protein

mRNAs in axons. The Journal of neuroscience : the official journal of the Society

for Neuroscience, 25, 778-791.

Willis, D.E., van Niekerk, E.A., Sasaki, Y., Mesngon, M., Merianda, T.T., Williams, G.G.,

Kendall, M., Smith, D.S., Bassell, G.J. & Twiss, J.L. (2007) Extracellular stimuli

specifically regulate localized levels of individual neuronal mRNAs. The Journal

of cell biology, 178, 965-980.

Xiao-Long, W., Chuan-Ping, Y., Kai, X. & Ou-Jv, Q. (2010) Selenoprotein W depletion in

vitro might indicate that its main function is not as an antioxidative enzyme.

Biochemistry. Biokhimiia, 75, 201-207.

Zhang, Y. & Gladyshev, V.N. (2008) Trends in selenium utilization in marine microbial

92

world revealed through the analysis of the global ocean sampling (GOS) project.

PLoS genetics, 4, e1000095.

Zhang, Y., Zhou, Y., Schweizer, U., Savaskan, N.E., Hua, D., Kipnis, J., Hatfield, D.L. &

Gladyshev, V.N. (2008) Comparative analysis of selenocysteine machinery and

selenoproteome gene expression in mouse brain identifies neurons as key

functional sites of selenium in mammals. J Biol Chem, 283, 2427-2438.

93

FIGURE LEGENDS

Figure 1: Sepw1 is expressed in cell bodies and processes of cultured neurons. Primary

cultures derived from neonatal mouse cortex (A), hippocampus (B), and cerebellum (C)

were grown on coverslips for 3 weeks and subsequently double immunolabeled for the

presence of Sepw1 (in green, left) and a neuronal marker, Tuj1 (in magenta, center).

Merged images show sporadic colocalization between Sepw1 and Tuj1 (in white, right) in

neuronal somata and neurites in cultures from all three braion regions.

Figure 2: Sepw1 is expressed in cell bodies and processes of pyramidal neurons in cortex

and hippocampus. Fixed, frozen mouse brain sections were immunolabeled for Sepw1 in

somatosensory cortex (A) and CA1 region of hippocampus (B). Cortex sections were

additionally dual immunolabled with Tuj1, while hippocampus sections were labeled with

a fluorescent Nissl stain. Large pyramidal neurons in both regions strongly displayed

Sepw1 immunoreactivity, which prominently extended into the apical and basolateral

dendrites. A few proximal axonal segments also showed Sepw1 immunoreactivity.

Figure 3: Regional expression of Sepw1 in neurons of mouse brain. (A) Motor cortex

(left) and CA3 region of hippocampus (right) displayed prominent Sepw1 staining in cell

bodies and processes. (B-D) Low magnification (left) and high magnification (right)

photomicrographs of piriform cortex (B), cingulate cortex (C), and cerebellum (D) show

Sepw1 immunoreactivity in neurons. Large pyramidal (B-C) and Purkinje (D) neurons

show high expression in dendritic arbors.

Figure 4: Sepw1 is present in isolated nerve terminals. Synaptosomes were prepared and

analyzed by SDS-PAGE followed by Western blot. Lanes 1,2 and 3,4 represent two

independent preparations from wild-type C57 mice run in duplicate. (A) Analysis of

TATA-binding protein (TBP) in synaptosome fractions indicated that the preparations

were free of nuclear contamination. (B) Western blot showed the presence of Sepw1 in

synaptosome fractions.

94

Figure 5: Sepw1 expression in isolated nerve terminals is greatly reduced in mice lacking

Sepp1. Additional synaptosome fractions were prepared from Sepp1-/- mice (KO) and

wild-type littermate conrols (WT). These fractions were analyzed in comparison to whole

cell lysate (WCL) and mitochondrial (Mito) fractions. Western blotting for Sepw1 (A)

and Gpx4 (B) revealed the presence of both selenoproteins, with Actin (C) used as a

loading-control. Quantitation of synaptosomal expression of Sepw1 (D) and Gpx4 (E)

revealed that Sepw1 was significantly decreased (p<0.01), while Gpx4 was not.

Figure 6: Several selenoprotein synthesis factors are present in isolated nerve terminals.

(A) TBP was analyzed to confirm that nuclear proteins, where selenoprotein synthesis

factors are known to be present, are not contaminating the synaptosome fractions. (B)

EFSec was found to be present in synaptosomes and qualitatively enriched compared to

whole cell lysates, and absent from mitochondria. (C) SPS2, the enzyme that generates

selenophosphate, showed a similar expression pattern, being apparently increased in

isolated nerve terminals, and absent from mitochondria. (D) SBP2, which is required for

selenoprotein translation, was found in similarly low abundance in synaptosomes and

whole cell lysates, but not mitochondria. (E) The tRNASec-associated protein SecP43

was found only in whole cell lysates, where as Sec lyase (Scly) (F) was found in both

synaptosome and whole cell lysate fracions.

95

FIGURE 1

96

FIGURE 2

97

FIGURE 3

98

FIGURE 4

99

FIGURE 5

100

FIGURE 6

101

CHAPTER 4

METHAMPHETAMINE-INDUCED ALTERATIONS IN SELENOPROTEIN

EXPRESSION IN MICE

ABSTRACT

Selenium is an essential micronutrient in mammals. It is primarily incorporated into the

amino acid selenocysteine, in a family of proteins termed selenoproteins. Some

selenoproteins are known antioxidant enzymes, and many of them are expressed in the

brain. Methamphetamine is a highly addictive drug with potentially neurotoxic side

effects. Previous studies have demonstrated modulatory effects of dietary selenium levels

on methamphetamine-induced neurotoxicity. Specifically, dietary selenium deficiency

promotes excessive neurotoxicity after methamphetamine administration, while dietary

selenium supplementation prevents it. Although the activity of cellular glutathione

peroxidase, a well characterized selenoprotein, is largely attributed to the effects of

dietary selenium, other selenoproteins may also influence the extent of

methamphetamine-induced neurotoxicity. To determine what selenoproteins may be

adversely impacted by methamphetamine, we screened for changes in brain selenoprotein

transcripts after exposure to methamphetamine in mice. We found an upregulation of

select selenoprotein mRNAs without effect on protein levels, possibly indicating that

methamphetamine causes increased transcription of selenoprotein genes, but an inability

to upregulate selenoprotein translation. We also looked at methamphetamine-induced

toxicity in mice lacking selenoprotein P, a selenium transport protein important for brain

selenium supply. Selenoprotein P knockout mice did not display exacerbated toxicity

after methamphetamine administration, compared to wild-type mice. These results may

indicate that methamphetamine alters the expression profile of selenoproteins in brain,

but less than selenoprotein P.

102

INTRODUCTION

Methamphetamine (MA) is a psychostimulant drug with toxic side effects. Long term use

of MA leads to striatal depletion of dopamine (DA), metabolites, and proteins involved in

DA synthesis and transport (Cadet & Krasnova, 2009). It has also been associated with

persistent neurotoxicity towards dopaminergic terminals and striatal neurons (Tulloch et

al., 2011). Although the exact cellular and molecular mechanisms of toxicity are not fully

elucidated, several reports indicate a prominent role for oxidative signaling via reactive

oxygen species (ROS) including hydrogen peroxide and peroxynitrite (Cadet &

Brannock, 1998).

Selenium (Se) is a micronutrient that is incorporated into oxidoreductase enzymes with

potent antioxidant activity and prominent roles in cellular oxidation-reduction (redox)

balance (Kryukov et al., 2003). These enzymes use Se in the form of selenocysteine

(Sec). Sec is a unique amino acid because it is encoded by the UGA codon in mRNAs,

which typically signals for termination of protein synthesis. To recode the UGA for Sec

incorporation, the 3' untranslated regions of selenoprotein mRNAs are endowed with a

Sec Insertion Sequence (SECIS). The SECIS element is a stem-loop structure that binds

to specific proteins, SECIS binding protein 2 and the Sec-specific elongation factor,

which help bypass termination and promote insertion of Sec [reviewed in (Bellinger et

al., 2009)].

Sec is a highly reactive residue, thereby rendering enzymatic selenoproteins with high

catalytic activity. Cellular glutathione peroxidase 1 was the first discovered

selenoenzyme, and rapidly reacts with hydrogen peroxide and reduced glutathione

(GSH), releasing water and glutathione disulfide (GSSG) (Rotruck et al., 1973).

Subsequently, four other GPX isozymes were determined to be selenoproteins in humans.

GPX4 has high affinity for lipid and organic hydroperoxides, is essential for neuronal

survival, and is required for embryonic development (Seiler et al., 2008, Yoo et al.,

2012). GPX4 is associated with neuromelanin in dopaminergic neurons of human brain,

and in dystrophic axons of Parkinson's disease patients (Bellinger et al., 2011).

103

Previous studies have shown that dietary Se status can profoundly alter the metabolism of

dopamine in brain and MA-induced toxicity to dopaminergic neurons. Specifically, a Se

deficient diet increases DA turnover in the substantia nigra, hippocampus and prefrontal

cortex (Castano et al., 1995, Castano et al., 1997, Castano et al., 1993). Further, Se

supplementation can prevent the MA-induced depletion of TH, DA and DA metabolites

in the nigrostriatal pathway (Imam & Ali, 2000, Imam et al., 1999, Kim et al., 1999).

GPX was suggested to be specifically inactivated by MA-induced production of

peroxynitrite, and was generally attributed as the cause for protective effects of Se

supplementation. However, individual selenoproteins were not examined. In this

research, we investigated the regulation of selenoproteins in response to MA by mRNA

and protein expression to assess the contribution of select selenoproteins.

104

MATERIALS AND METHODS

Animals: Male and female C57BL/6 mice, and genetically modified mice on the same

background lacking Sepp1 were bred on commercially available diets containing

adequate Se (~0.25 ppm). Animals were given food and water ad libitum on a 12-hour

light-cycle and group housed until experimentation. All experiments were conducted on

adult mice aged 3-4 months during the light cycle. Male and female mice of both

genotypes were used in approximately equal numbers to examine sex differences present

in the animals. All animal procedures and experimental protocols were approved by the

University of Hawaii Institutional Animal Care and Use Committee.

Animal Procedures: Mice were administered methamphetamine by intraperitoneal (i.p.)

injection with various doses depending on the experiment. In initial studies,

methamphetamine was chronically administered to C57/Bl6 mice 5 days per week for 2

or 8 weeks, with the dosage escalating from 2mg/kg to 20 mg /kg and then 3 mg/kg to 30

mg/kg. In two subsequent studies using Sepp1-/- mice and wild-type littermate controls,

two different paradigms were used. The first study used two i.p. injections of 10 mg/kg

administered approximately 4 hours apart, while the second used a single bolus of 40

mg/kg. For qPCR and Western blot analysis, animals were sacrificed 24 hours after the

final injection. For histological analysis, animals were sacrificed 72 hours after the final

injection.

Tissue Preparation: For biochemistry, mice were sacrificed by CO2 asphyxiation, and the

brains rapidly excised, washed in PBS and snap-frozen in liquid nitrogen. The striatum

and ventral mesencephalon were dissected from frozen tissue in a cold chamber

maintained at -20C. Tissue was then cut into small pieces, using a razor blade, and

collected into pre-chilled tubes. Tissue was homogenized by sonication in the appropriate

buffer for the ensuing experiment. For histology, mice were anesthetized with ketamine-

xylazine, and sacrificed by transcardial perfusion. Mice were initially perfused with

phosphate-buffered saline (PBS) to flush out blood, followed by perfusion with 4%

paraformaldehyde (PFA) to fix the tissue. The mice heads were cut off and the brains

dissected out. Brains were post-fixed in 4% PFA overnight, followed by cryoprotecting

105

the tissue in 10% and 30% sucrose for at least 4 hours each. The brains were then

embedded in optimal cutting temperature (OCT) compound and frozen until time of

sectioning. 40 μm sections were cut on a Leica CM1900 cryostat and saved in

cryoprotectant solution, containing 0.1 M phosphate buffer, 30% sucrose (w/v), and 30%

ethylene glycol (v/v), until time of further experimentation.

Quantitative RT-PCR: Total RNA from tissue was prepared by Trizol extraction

(Invitrogen, Carlsbad, CA, USA) followed by purification using the RNeasy kit (QIAgen,

Valencia, CA, USA). Concentration and purity of extracted RNA and synthesized cDNA

was determined using A260/A280 ratio measured on an ND1000 Spectrophotometer

(NanoDrop Technologies, Wilmington, DE, USA). Synthesis of cDNA was carried out

using High Capacity cDNA Reverse Transcription Kit (Applied Biosystems, Foster City,

CA, USA), with 1 g RNA per 20 l reaction. For real-time PCR, 100 ng of the cDNA

was used in 5 l reactions with Platinum SYBR Green qPCR SuperMix-UDG

(Invitrogen). Reactions were carried out in triplicate or quadruplicate in a LightCycler

480 II thermal cycler (Roche, Indianapolis, IN, USA). Cycling conditions followed the

manufacturers suggestions in the SYBR Green kit instructions. All qPCR results were

normalized to 18S rRNA expression as a housekeeping gene and analyzed using Absolute

Quantification Software (Roche).

SDS-PAGE and Western blot: Total protein was extracted from powdered mouse tissues

by light sonication in CelLytic MT buffer (Sigma, St. Louis, MO, USA), followed by

centrifugation according to the manufacturers’ protocol. Protein was added to reduced

Laemmli buffer, boiled for 10 minutes, and loaded into 4-20% gradient polyacrylamide

gels (Bio-Rad, Hercules, CA, USA). Following electrophoresis, gel contents were

transferred to PVDF membranes, which were blocked with 5% milk in PBS containing

0.05% tween-20 (PBST) for one hour. Membranes were then probed for 90 minutes at

room temperature or overnight at +4C with the following primary antibodies: Rabbit-

anti-Tyrosine Hydroxylase (TH) (Pel-Freez, Rogers, AR, USA), Rabbit-anti-GPX4

(AbNova, Taipei, Taiwan), Rabbit-anti-SEPW1 (Rockland, Gilbertsville, PA, USA),

Rabbit-anti-SEPHS2 (Rockland) and Mouse-anti-beta actin (Sigma, St. Louis, MO,

USA). After washing with PBST, membranes were incubated in secondary antibodies

106

conjugated to horseradish peroxidase. After further washes in PBST, blots were

visualized by enhanced chemiluminesence reaction followed by exposure to film.

Immunohistochemistry: Brain sections stored in cryoprotectant in the freezer were

warmed to room temperature, and sections containing striatum and substantia nigra were

selected for analysis. After thorough washing, sections were blocked in 5% normal goat

serum with 0.3% Triton X-100 in PBS. After blocking, the sections were incubated in

diluted primary antibody solution overnight at +4C. The following antibodies were used:

Rabbit-anti-TH (Pel-Freez), Rabbit-anti-GPX4 (AbNova), and Rabbit-anti-SEPW1

(Rockland). A control section where primary antibody was omitted was also included in

the procedure. After washing out primary antibody, sections were incubated in species-

matched secondary antibody. The secondary antibody was either directly conjugated to

fluorophores (Alexa Fluor dyes, Invitrogen) for fluorescence imaging, or was biotinylated

for colorimetric visualization. After incubation in biotinylated secondary antibody, the

sections were then incubated in Avidin-Biotin Complex solution, followed by

visualization with 3,3-Diaminobenzidine (DAB) (Vector Labs, Burlingame, CA, USA).

Additionally, some sections were dual-labeled with a fluorescent Nissl stain to label

neurons (Neurotrace, Invitrogen), or phalloidin to label filamentous actin (Invitrogen).

Sections were then mounted onto slides, and either coverslipped in VectaShield

containing DAPI for fluorescence, or air-dried overnight and coverslipped using

Permount after ethanol dehydration and Xylene clearing.

TUNEL Histology: Apoptotic cell death in sections was assessed using an In Situ Cell

Death Detection Kit (Roche), using the Terminal deoxynucletidyl transferase (TdT)

dUTP Nick End Labeling (TUNEL) method. Sections were thoroughly washed, and then

blocked with 3% hydrogen peroxide in methanol for 10 minutes at room temperature.

Sections were then permeabilized by incubation in 0.5% Triton X-100 at +65C for 30

minutes. After washing in PBS, TUNEL labeling was performed for 60 minutes at +37C

in the dark. A negative control section lacking TdT, and a positive control section that

was incubated in DNAse I was also included. The fluorescent label was then converted

by incubating in POD conversion solution included in the kit for 30 minutes at +37C, and

was visualized by DAB reaction. The sections were mounted and air-dried overnight. The

107

sections were then counterstained with 0.5% Cresyl Violet after ethanol dehydration,

cleared in Xylene, and coverslipped using Permount.

Imaging: Both fluorescence and bright-field imaging of stained sections were acquired on

an upright Zeiss Axioscope 2 Plus microscope equipped with an ASI motorized stage and

Zeiss Axiocam MRc camera. Fluorescence imaging was done with an ultraviolet lamp

and appropriate filter sets for the fluorophores used, while bright-field imaging used

standard white-light illumination. Images were acquired using the included AxioVision

software, and were analyzed using ImageJ.

Statistical analysis: Data were analyzed using Microsoft Excel (Redmond, WA, USA),

and plotted using GraphPad Prism software (San Diego, CA, USA). Two-way ANOVA

was used to determine an interaction between MA administration and selenoprotein

expression. Post hoc test using Bonferroni correction for multiple comparisons was used

to determine significance between individual groups. Unpaired t-tests comparing control

and experimental groups were performed for some experiments. The significance criteria

were set at p < 0.05 for all statistical measures.

108

RESULTS

We sought to determine whether methamphetamine (MA) causes an alteration in

selenoprotein transcript expression in the substantia nigra and striatum of mice brains. To

achieve this, we administered MA to C57BL/6 mice for a period of 2 weeks, and

harvested their brains for analysis by quantitative real-time polymerase chain reaction

(qPCR). The majority of selenoprotein mRNAs were analyzed, and two-way ANOVA

revealed that MA had an effect on selenoprotein transcript expression in midbrain

(F(1,92)=6.329, p=0.0136) and striatum (F(1,92)=6.854, p=0.0103). Qualitatively,

selenoprotein expression appeared to increase in response to MA administration in both

brain regions (Fig. 1). Individual selenoprotein transcripts of interest were further

analyzed by t-test to compare groups. We found that selenoprotein W (Sepw1)

(t(4)=3.083, p=0.0368) and selenophosphate synthetase 2 (Sps2) (t(4)=2.809, p=0.0484)

mRNAs were significantly increased in midbrain, but not in striatum of MA-treated mice.

However, we did not find a significant increase in mRNA expression of glutathione

peroxidase 4 (Gpx4) or selenoprotein P (Sepp1) in either brain region (Fig. 2). These

results led us to question if selenoprotein expression was also increased in the mice

brains. In contrast to the mRNA data, western blot analysis of Sepw1, Sps2, and Gpx4

did not reveal an increase in selenoprotein expression in response to MA in either brain

region (Fig. 3).

We next sought to determine if prolonged, higher-dose MA exposure has a more dramatic

effect on selenoprotein expression. When mice were exposed to elevated doses of MA for

8 weeks, we found that selenoprotein mRNAs in midbrain were not increased, unlike

with short term MA administration. However, Sepw1 and Sps2 were decreased in

midbrain, contrasting their previously seen upregulation (Fig. 4, top). This may indicate

that expression of Sepw1 and Sps2 transcripts are regulated in midbrain by MA. In

striatum, qPCR showed that selenoprotein mRNA levels were dramatically reduced (Fig.

4, bottom). High-dose exposure to MA is known to cause degeneration of dopaminergic

terminals and medium spiny neurons in striatum (Zhu et al., 2006), which may explain

the reduction of selenoprotein transcripts in this region.

109

To determine if selenoprotein P augments selenoprotein expression in the context of MA-

induced changes, we next administered MA to Sepp1-/- and wild-type littermate control

mice. When administering two injections of 10 mg/kg MA and sacrificing 24 hours after

the second injection, we did not detect any major changes in selenoprotein expression due

to MA. However, the Sepp1-/- animals had reduced expression of several selenoprotein

transcripts, most notably with Sepw1 (data not shown).

As we observed little change in selenoprotein mRNA expression with 2 injections of 10

mg/kg MA, we next administered high-dose MA at 40 mg/kg. We investigated protein

expression by western blot and immunohistochemistry, and also performed TUNEL

histology to investigate cell death. Western blot analysis of striatum revealed that Sepp1-

/- mice have reduced levels of Sepw1 and Gpx4, but not Sephs2, and that MA did not

significantly influence the expression of selenoproteins in wild-type or knockout mice

(Fig. 5). Similar results were obtained in midbrain (data not shown). Tyrosine

hydroxylase (TH) expression was investigated by western blotting and by histology. MA

caused a slight upregulation of TH in midbrain lysates when measured by Western blot

(data not shown). However, an apparent decrease in TH staining in substantia nigra was

observed in brain sections (Fig. 6). TH expression did not show much change in striatum

in response to MA either by western blot or immunohistochemistry (data not shown).

Sepp1-/- mice did not respond differently than wild-type mice in terms of TH expression

in midbrain or striatum. In congruence with these findings, TUNEL histology revealed an

absence of apoptotic cell death in Sepp1-/- and wild-type mice exposed to MA (Fig. 7).

110

DISCUSSION

The results presented here show that MA affects both the short- and long-term mRNA

expression of selenoproteins in mammalian brain. These data are in agreement with

reports indicating that Se deficiency potentiates the neurotoxic profile of MA exposure,

while Se repletion attenuates it (Imam et al., 1999, Kim et al., 1999). Neurons appear to

be the key site of Se utilization, although these cells possess fairly low GPX activity

(Zhang et al., 2008). This suggests that selenoproteins in addition to GPX1 are involved

in protection from MA-induced neurotoxicity.

GPX4 is a phospholipid hydroperoxidase that is particularly important for neuronal

development and survival (Seiler et al., 2008). Conditional deletion of GPX4 in neurons

of adult mice causes rapid lipoxygenase-mediated apoptosis (Yoo et al., 2012). GPX4 is

found in dopaminergic neurons of human brain, where it has a conspicuous association

with neuromelanin. It has also been found in dystrophic neurites of individuals with

Parkinson’s disease (Bellinger et al., 2011). These findings strongly implicate GPX4 as a

necessary component of dopamine neurons.

Selenoprotein P (Sepp1) is an extracellular selenoprotein that functions in Se distribution,

and has roles in preventing neurodegeneration and deficits in synaptic plasticity (Caito et

al., 2011, Peters et al., 2006, Valentine et al., 2008, Valentine et al., 2005). The finding

that Sepp1-deficient mice did not display enhanced sensitivity to neurotoxicity in

dopaminergic brain regions has not been reported, but is probably due to the severe

pathology associated with deletion of Sepp1 alone. Although TUNEL labeling of

apoptotic cells revealed no increase in cell death in Sepp1 knockout mice, the mice are

unquestionably under more cellular stress than wild-type MA-treated control

counterparts.

Selenoprotein W (Sepw1) is a small thioredoxin-like protein that has a role in mediating

cell signaling and is particularly abundant in muscle and brain of mammals (Gu et al.,

2000, Whanger, 2009). Interestingly, dietary Se deficiency causes depletion of Sepw1 in

muscle, but not brain (Whanger, 2001). Maintenance of Sepw1 expression during dietary

Se deprivation is likely due to the Se-transport function of Sepp1, since mice deficient in

111

Sepp1 have drastically reduced expression of Sepw1 mRNA and protein in brain

(Hoffmann et al., 2007). Sepw1 mRNA expression was altered by both short and long

term MA administration, albeit in different directions, perhaps making it a specific target

in MA-induced neurotoxicity.

Lastly, selenophosphate synthetase 2 (Sps2) is a selenoprotein that generates

selenophosphate for selenocysteine biosynthesis (Xu et al., 2007). Although we initially

found that Sephs2 mRNA was acutely upregulated by MA, the effect on protein was

minimal. However, the mRNA was downregulated upon long-term MA treatment, similar

to Sepw1. Transcriptional regulation of Sepw1 and Sps2 suggests these two

selenoproteins may be specifically targeted in midbrain during MA exposure. Further, the

activity of the Sps2 enzyme was not assessed, and selenophosphate production may

indeed be compromised by MA treatment.

Collectively these data suggest that selenoproteins are neuroprotective in the context of

MA exposure. The reduction of selenoprotein mRNA expression in striatum after chronic

MA exposure indicates that Se availability may be limiting under MA-induced stress, and

therapeutic supplementation with dietary Se may be helpful. We did not detect dramatic

changes in selenoprotein expression by western blot, but it remains possible that the Sec

residue is oxidatively modified and that enzymatic catalysis is inactivated or reduced by

high levels of MA-induced ROS production.

112

REFERENCES

Bellinger, F.P., Bellinger, M.T., Seale, L.A., Takemoto, A.S., Raman, A.V., Miki, T.,

Manning-Bog, A.B., Berry, M.J., White, L.R. & Ross, G.W. (2011) Glutathione

Peroxidase 4 is associated with Neuromelanin in Substantia Nigra and Dystrophic

Axons in Putamen of Parkinson's brain. Molecular neurodegeneration, 6, 8.

Bellinger, F.P., Raman, A.V., Reeves, M.A. & Berry, M.J. (2009) Regulation and

function of selenoproteins in human disease. Biochem J, 422, 11-22.

Cadet, J.L. & Brannock, C. (1998) Free radicals and the pathobiology of brain dopamine

systems. Neurochemistry international, 32, 117-131.

Cadet, J.L. & Krasnova, I.N. (2009) Molecular bases of methamphetamine-induced

neurodegeneration. International review of neurobiology, 88, 101-119.

Caito, S.W., Milatovic, D., Hill, K.E., Aschner, M., Burk, R.F. & Valentine, W.M.

(2011) Progression of neurodegeneration and morphologic changes in the brains

of juvenile mice with selenoprotein P deleted. Brain research, 1398, 1-12.

Castano, A., Ayala, A., Rodriguez-Gomez, J.A., de la Cruz, C.P., Revilla, E., Cano, J. &

Machado, A. (1995) Increase in dopamine turnover and tyrosine hydroxylase

enzyme in hippocampus of rats fed on low selenium diet. Journal of neuroscience

research, 42, 684-691.

Castano, A., Ayala, A., Rodriguez-Gomez, J.A., Herrera, A.J., Cano, J. & Machado, A.

(1997) Low selenium diet increases the dopamine turnover in prefrontal cortex of

the rat. Neurochemistry international, 30, 549-555.

Castano, A., Cano, J. & Machado, A. (1993) Low selenium diet affects monoamine

turnover differentially in substantia nigra and striatum. Journal of

neurochemistry, 61, 1302-1307.

Gu, Q.P., Sun, Y., Ream, L.W. & Whanger, P.D. (2000) Selenoprotein W accumulates

primarily in primate skeletal muscle, heart, brain and tongue. Molecular and

cellular biochemistry, 204, 49-56.

Hoffmann, P.R., Hoge, S.C., Li, P.A., Hoffmann, F.W., Hashimoto, A.C. & Berry, M.J.

(2007) The selenoproteome exhibits widely varying, tissue-specific dependence

on selenoprotein P for selenium supply. Nucleic Acids Res, 35, 3963-3973.

Imam, S.Z. & Ali, S.F. (2000) Selenium, an antioxidant, attenuates methamphetamine-

induced dopaminergic toxicity and peroxynitrite generation. Brain research, 855,

186-191.

Imam, S.Z., Newport, G.D., Islam, F., Slikker, W., Jr. & Ali, S.F. (1999) Selenium, an

antioxidant, protects against methamphetamine-induced dopaminergic

neurotoxicity. Brain research, 818, 575-578.

Kim, H.C., Jhoo, W.K., Choi, D.Y., Im, D.H., Shin, E.J., Suh, J.H., Floyd, R.A. & Bing,

G. (1999) Protection of methamphetamine nigrostriatal toxicity by dietary

selenium. Brain research, 851, 76-86.

Kryukov, G.V., Castellano, S., Novoselov, S.V., Lobanov, A.V., Zehtab, O., Guigo, R. &

Gladyshev, V.N. (2003) Characterization of mammalian selenoproteomes.

Science, 300, 1439-1443.

Peters, M.M., Hill, K.E., Burk, R.F. & Weeber, E.J. (2006) Altered hippocampus

synaptic function in selenoprotein P deficient mice. Molecular

neurodegeneration, 1, 12.

113

Rotruck, J.T., Pope, A.L., Ganther, H.E., Swanson, A.B., Hafeman, D.G. & Hoekstra,

W.G. (1973) Selenium: biochemical role as a component of glutathione

peroxidase. Science, 179, 588-590.

Seiler, A., Schneider, M., Forster, H., Roth, S., Wirth, E.K., Culmsee, C., Plesnila, N.,

Kremmer, E., Radmark, O., Wurst, W., Bornkamm, G.W., Schweizer, U. &

Conrad, M. (2008) Glutathione peroxidase 4 senses and translates oxidative stress

into 12/15-lipoxygenase dependent- and AIF-mediated cell death. Cell Metab, 8,

237-248.

Tulloch, I., Afanador, L., Mexhitaj, I., Ghazaryan, N., Garzagongora, A.G. & Angulo,

J.A. (2011) A single high dose of methamphetamine induces apoptotic and

necrotic striatal cell loss lasting up to 3 months in mice. Neuroscience, 193, 162-

169.

Valentine, W.M., Abel, T.W., Hill, K.E., Austin, L.M. & Burk, R.F. (2008)

Neurodegeneration in mice resulting from loss of functional selenoprotein P or its

receptor apolipoprotein E receptor 2. Journal of neuropathology and experimental

neurology, 67, 68-77.

Valentine, W.M., Hill, K.E., Austin, L.M., Valentine, H.L., Goldowitz, D. & Burk, R.F.

(2005) Brainstem axonal degeneration in mice with deletion of selenoprotein p.

Toxicologic pathology, 33, 570-576.

Whanger, P.D. (2001) Selenium and the brain: a review. Nutritional neuroscience, 4, 81-

97.

Whanger, P.D. (2009) Selenoprotein expression and function-selenoprotein W. Biochim

Biophys Acta, 1790, 1448-1452.

Xu, X.M., Carlson, B.A., Irons, R., Mix, H., Zhong, N., Gladyshev, V.N. & Hatfield,

D.L. (2007) Selenophosphate synthetase 2 is essential for selenoprotein

biosynthesis. Biochem J, 404, 115-120.

Yoo, S.E., Chen, L., Na, R., Liu, Y., Rios, C., Van Remmen, H., Richardson, A. & Ran,

Q. (2012) Gpx4 ablation in adult mice results in a lethal phenotype accompanied

by neuronal loss in brain. Free Radic Biol Med, 52, 1820-1827.

Zhang, Y., Zhou, Y., Schweizer, U., Savaskan, N.E., Hua, D., Kipnis, J., Hatfield, D.L. &

Gladyshev, V.N. (2008) Comparative analysis of selenocysteine machinery and

selenoproteome gene expression in mouse brain identifies neurons as key

functional sites of selenium in mammals. J Biol Chem, 283, 2427-2438.

Zhu, J.P., Xu, W. & Angulo, J.A. (2006) Methamphetamine-induced cell death: selective

vulnerability in neuronal subpopulations of the striatum in mice. Neuroscience,

140, 607-622.

114

FIGURE LEGENDS

Figure 1: Expression of selenoprotein mRNAs is altered by methamphetamine. Following

two weeks of treatment with methamphetamine or vehicle, mice brains were analyzed for

expression of selenoprotein transcripts by qPCR. The regions principally affected by

methamphetamine were dissected from frozen brains. Methamphetamine administration

caused significant variance in expression of selenoprotein mRNAs both midbrain (top)

and striatum (bottom). Data are presented as fold-change relative to PBS treated-mice for

the sake of clarity, but raw data was analyzed by two-way ANOVA.

Figure 2: Selenoprotein W and selenophosphate synthetase 2 mRNA are upregulated in

midbrain after two weeks of methamphetamine administration. Select selenoprotein

mRNAs of interest were analyzed by t-test to determine significance between treatment

groups. Sepw1 and Sps2 had significantly elevated expression in midbrain (n=3, p<0.05),

but not in striatum of methamphetamine-treated mice. Trends towards increased

expression of Sepp1 and Gpx4 did not reach statistical significance in either brain region.

Figure 3: Selenoprotein expression following two weeks of methamphetamine

administration is not changed in mice brains. SDS-PAGE and Western blotting for

Sepw1 and Gpx4 was performed on the same samples as assessed by qPCR (n=3).

Qualitatively, expression of Sepw1 and Gpx4 appeared to be reduced, however

considerable variation and low sample size resulted in no statistically significant

alteration.

Figure 4: Long term methamphetamine administration dramatically reduces selenoprotein

mRNAs in striatum. Following eight weeks of treatment with methamphetamine or

vehicle, mice brains were analyzed by qPCR for expression of selenoprotein transcripts

(n=3-4). Sepw1 and Sps2 showed reduced expression in midbrain of methamphetamine

115

treated mice, but none of the other selenoproteins differed in midbrain, compared to

vehicle treated animals. In contrast, reduced expression of nearly all selenoprotein

mRNAs was observed in striatum.

Figure 5: Sepp1 impacts brain expression of selenoproteins more than methamphetamine.

Sepp1-deficient mice were administered a high-dose bolus of methamphetamine, and

expression of selenoproteins in striatum was investigated by Western blot (n=3-4).

Expression of Sepw1 and Gpx4 was reduced in brains of Sepp1 knockout mice, and did

not differ between methamphetamine and vehicle treatment groups. Expression of Sps2

was stable across all experimental groups.

Figure 6: Methamphetamine reduces tyrosine hydroxylase expression in substantia nigra

independently of Sepp1. Sepp1 knockout (KO) and wild-type (WT) control mice were

administered a bolus of MA or PBS as a vehicle control, and the brains were

subsequently analyzed by immunohistochemistry for the presence of tyrosine

hydroxylase (TH). An apparent decrease in TH-immunoreactivity in neurons of the

substantia nigra was observed in MA-treated animals. Sepp1 KO mice had similar

expression of TH as WT control mice, under both MA and PBS treatment conditions.

Figure 7: Apoptotic cell death is not increased in methamphetamine-treated selenoprotein

P knockout mice. Following a high-dose bolus administration of MA or vehicle control,

TUNEL labeling was performed on brain sections from Sepp1 KO and WT littermate

mice to investigate apoptosis. When compared to the negative (C) and positive (D)

control sections, both WT (A) and KO (B) methamphetamine-treated mice showed

virtually no TUNEL positive cells, indicating an absence of apoptotic cell death.

116

FIGURE 1

117

FIGURE 2

118

FIGURE 3

119

FIGURE 4

120

FIGURE 5

121

FIGURE 6

122

FIGURE 7

123

CHAPTER 5

CONCLUSION

SUMMARY AND DISCUSSION

The overall goal of the research presented in this dissertation was to gain a deeper

understanding of selenoprotein expression and function in the mammalian brain. The first

section investigates the neurological consequences of genetic disruption of selenocysteine

lyase (SCLY), particularly in comparison to the behavioral phenotype of selenoprotein P

(SEPP1) knockout mice. The expression and activity of selenoproteins in SCLY

knockout mice was also analyzed. The second section describes the cellular and

subcellular expression of a relatively uncharacterized selenoprotein that is abundant in

brain, selenoprotein W (SEPW1). SEPW1 expression in brain was also examined in

SEPP1 knockout mice. The final section assesses and describes selenoprotein expression

in mouse brain after exposure to the drug methamphetamine. Methamphetamine-induced

pathology and alterations in proteins relevant to dopamine metabolism were also

investigated in the brains of mice lacking SEPP1.

The local and global distribution of bodily Se in mammals is a complex phenomenon that

is still mechanistically unclear. SEPP1 is certainly involved in the hierarchy of Se

distribution that prioritizes Se to the brain when dietary availability is limiting, however

the kinetics of Se utilization in the brain are unknown (Burk & Hill, 2009). The form of

Se in SEPP1 is Sec, but the presumed substrate for selenoprotein synthesis is selenide.

Therefore a complementary system to break down Sec is necessary to complete the Se

cycle and utilize Sec from SEPP1. To this end, SCLY has been proposed to close the loop

and recycle Sec residues within organisms as a source of Se (Mihara et al., 2000). To

investigate if SCLY is necessary to recycle Sec residues, we determined if genetic

deletion of SEPP1 and SCLY in mice produces similar phenotypes.

124

On the contrary, we found that SCLY knockout mice display a minimal neurological

phenotype when contrasted to SEPP1 knockout mice. When given a Se-adequate diet, no

behavioral impairment was observed. Even when challenged with a Se-poor diet, the

SCLY knockout animals have only a slight learning impairment, and no motor

phenotype. This is highly discordant with the severe motor phenotype observed in SEPP1

knockout mice. This finding suggests that mice have one or more additional enzymes that

can break down Sec and provide usable Se, preventing the SEPP1 knockout phenotype.

The cysteine desulfurase enzymes NFS1 is known to have selenocysteine lyase acitivity,

and may be functionally compensating for the lack of SCLY (Lacourciere et al., 2000).

Despite a very mild behavioral phenotype in SCLY knockout mice fed a low Se diet, the

animals had drastically reduced expression of selenoproteins in brain, and GPX activity

in brain and liver was similarly reduced. This finding confirms that SCLY contributes to

the Se pool for selenoprotein translation, at least when dietary Se is limiting. Thus, both

SCLY and complementary enzymes such as NFS1 are likely to be involved in utilizing

Sec for synthesis of selenoproteins.

It is clear that Se is prioritized to the brain during dietary deficiency, indicating that

selenoproteins are important for nervous system function [reviewed in (Chen & Berry,

2003)]. Most of the discovered selenoproteins are found in the brain, but their cellular

functions and enzymatic activity remain elusive. SEPW1 is a small selenoprotein that is

highly expressed in muscle and nervous tissues, and is peculiarly absent in muscle of Se-

deficient lambs suffering a myopathy called White muscle disease. Unlike in muscle and

most tissues, SEPW1 expression is not diminished in brains of Se-deficient mammals

(Whanger, 2001). SEPW1 contains an N-terminal C-X-X-U motif in a thioredoxin-like

fold, where C is cysteine, U is Sec, and X is any amino acid. This structure is highly

conserved in thioredoxin superfamily proteins and usually is indicative of a redox

function (Dikiy et al., 2007). Biochemical and computational experiments indicate that

SEPW1 interacts with 14-3-3 beta and gamma proteins, possibly regulating the oxidation

state of a conserved cysteine residue (Aachmann et al., 2007, Musiani et al., 2010). 14-3-

3 proteins are multifunctional scaffolding proteins that coordinate the interaction of

receptors, kinases, and phosphatases, thereby modulating cell signaling. Using siRNA to

125

reduce SEPW1, it has been demonstrated that cell cycle progression and proliferation

involves SEPW1. Knockdown of SEPW1 in prostate and breast epithelial cells results in

G1 arrest, and this effect is dependent on the tumor suppressor protein p53 (Hawkes &

Alkan, 2011, Hawkes et al., 2012). These results suggest that SEPW1 has an important

role in cell signaling related to cellular proliferation, differentiation, and death. However

the functional role of SEPW1 in the brain is unknown.

To investigate the expression of SEPW1 in the brain, we analyzed the localization of

SEPW1 in cultured cells, in mice brain sections, and in isolated nerve terminals from

mice brains. Intriguingly, SEPW1 is highly expressed in neurons in culture and adult

mouse brain, and expression is apparent in the processes of large neurons in multiple

brain regions. Analysis of isolated nerve terminals, or synaptoneurosomes, revealed the

presence of SEPW1. Previous studies from other labs have indicated that SEPW1 mRNA

is found in processes of neurons as well. This suggested that SEPW1 may be locally

translated in these subcellular neuronal compartments distal to the nucleus. However

selenoprotein synthesis requires specialized factors, some of which exhibit shuttling

between the nucleus and cytoplasm. Synaptoneurosomes contained several of these

factors, importantly including EFSec and SBP2, which are both required for

selenoprotein synthesis. Additionally, we found that SEPW1 expression is drastically

reduced in the brains of SEPP1 knockout mice, unlike in animals with dietary Se

deprivation. Similar results were obtained for other selenoproteins, but not for synthesis

factors. In sum, these results indicate that SEPW1 expression is widespread in neurons, it

may be synthesized locally, and its expression is dependent on SEPP1. The Sec residues

of SEPP1 are likely to supply Se for synthesis of SEPW1 and other selenoproteins, but a

signaling function of SEPP1 that stimulates transcription and translation of SEPW1

specifically cannot be ruled out.

The dopaminergic neurotransmitter system consists of neurons emanating from midbrain,

which synthesize dopamine and release it in several regions throughout the brain. The

two primary populations of dopaminergic neurons have cell somas in the substantia nigra

pars compacta and ventral tegmental area, located in the ventral mesencephalon. Neurons

126

in the substantia nigra project mainly to the caudate nucleus and putamen, collectively

referred to as the striatum. The nigrostriatal pathway facilitates the initiation and smooth

control of movement. Extensive degeneration of nigrostriatal neurons is a canonical

feature of Parkinson's disease, and is responsible for motor symptoms including rigidity

and bradykinesia (Cookson, 2005). Parkinson's disease is a complex disorder with genetic

and environmental causes, and positive effects of Se-supplementation are reported in

animal models of the disease (Khan, 2010, Zafar et al., 2003). Research in our lab has

shown an association of GPX4 with neuromelanin in dopaminergic neurons of the

substantia nigra pars compacta (Bellinger et al., 2011). Lewy bodies, a pathological

hallmark of Parkinson's disease, are associated with SEPP1 in post-mortem tissue from

human patients (Bellinger, et al., unpublished).

The population of dopaminergic neurons in the ventral tegmental area mainly project to

the nucleus accumbens via the amygdala and hippocampus, in addition to frontal cortex.

These two pathways are referred to as the mesolimbic and mesocortical tracts

respectively, and are involved in diverse cognitive functions including pleasure and

reward. Both positive and negative deregulation of these fibers is observed in

schizophrenia and mood disorders, which are associated with Se-binding proteins and Se-

status (Benton, 2002, Glatt et al., 2005). Although these three primary dopaminergic

pathways are approximately separated anatomically, the distinction at the cellular level is

less discrete and considerable overlap between projection targets is observed.

Methamphetamine is a highly addictive psychostimulant drug of abuse that stimulates

dopamine release and reversal of the plasma and vesicular membrane dopamine

transporters, causing high levels of synaptic, cytosolic and extracellular dopamine (Cadet

& Krasnova, 2009). This is accompanied by regional changes in brain glucose

metabolism, the blood-brain barrier, and myelination (Bowyer et al., 2008, Volkow et al.,

2001a). Chronic abuse of methamphetamine is neurotoxic in humans, leading to loss of

dopaminergic terminals in the striatum and cortex (Sekine et al., 2003, Volkow et al.,

2001b). The exact cause of neurotoxicity is debated, but oxidative processes are strongly

implicated. In rodent models, Se supplementation can alleviate or prevent

127

methamphetamine-induced neurotoxicity, but the contributions of specific selenoproteins

are uncertain. Dietary Se deficiency elevates and prolongs high potassium-induced

dopamine release in the striatum, and increases the turnover rate of dopamine in the

substantia nigra, prefrontal cortex and hippocampus (Castano et al., 1995, Castano et al.,

1997, Castano et al., 1993, Watanabe et al., 1997). Additionally, Se deficiency causes

induction of tyrosine hydroxylase and dopamine transporter mRNA in nigrostriatal

neurons, with concomitant increases in dopamine synthesis and uptake (Romero-Ramos

et al., 2000). Conversely, dietary Se supplementation reduces the activity of the

dopamine catabolic enzyme, monoamine oxidase (MAO) (Tang et al., 2008). Because

dopamine, and related precursors and breakdown products, can spontaneously oxidize to

electron-deficient quinones, oxidative processes partially determine survival of

dopaminergic neurons (Chen et al., 2008). The antioxidant activity of selenoproteins such

as GPX4 can inhibit oxidant-induced apoptosis (Seiler et al., 2008). Additionally,

dopamine deamination by MAO generates hydrogen peroxide, and MAO-catalyzed

peroxide generation is coupled to the enzymatic activity of the selenoprotein GPX1

(Maker et al., 1981). These studies suggest that the function and survival of dopamine

neurons rely on brain Se and one or more selenoproteins. Therefore toxicity caused by

methamphetamine may be partially due to deleterious effects on selenoproteins, and Se

supplementation may be of therapeutic benefit.

To determine if methamphetamine causes an alteration in selenoprotein expression, an

investigation of selenoprotein transcripts and proteins in striatum and midbrain after

methamphetamine administration in mice was performed. We found that selenoprotein

transcripts generally increased after acute methamphetamine exposure. Based on altered

SEPW1 and SPS2 mRNA levels in both MA-treatment paradigms, selenoprotein

expression may be altered in midbrain by MA. It seems probable that the oxidative

burden caused by methamphetamine is the primary cause of alterations in SEPW1 and

SPS2 transcripts. However in the same samples, SEPW1 protein was unchanged. This

may indicate that methamphetamine causes oxidation and degradation of the proteins, or

that the mRNA translation is inefficient, and the transcriptional increase is compensatory

to normalize protein expression.

128

OXIDATIVE STRESS AND SELENOPROTEINS

An overarching theme surrounding the functions of selenium and selenoproteins in cell

biology is the relationship with oxidative stress and redox control systems. Oxidative

stress is a broadly used term, but generally signifies an imbalance in prooxidants and

antioxidants in cells and their environment, causing disruption of redox control and

macromolecular damage. Free radicals and nonradical oxidants operate differently as

prooxidants, because free radicals are promiscuously reactive while nonradical oxidants

typically react with specific targets. Additionally, macromolecular damage is a divergent

outcome from disrupted redox control and signaling. Macromolecular damage to DNA,

proteins, and lipids is caused by redox reactions with free radicals, which contain an

unpaired electron and are scarce in biological systems due to enzymes and chemicals that

rapidly scavenge them and typically generate nonradical oxidants (Jones, 2006). For

example the superoxide radical (O2-) is rapidly converted by superoxide dismutase to the

nonradical oxidant hydrogen peroxide (H2O2). Superoxide can also react with nitric oxide

(NO), another free radical, to produce the nonradical oxidant peroxynitrite (ONO2-). Free

radical-mediated macromolecular damage is a common end-point observed in aging and

disease. However, mechanistically it is likely that selenoproteins function in redox

systems control and signaling.

Due to rapid conversion in organisms, free radicals are quantitatively miniscule compared

to nonradical two-electron oxidants such as hydrogen peroxide, fatty acid

hydroperoxides, disulfides, peroxynitrite, aldehydes, and quinones. In prototypical redox

control systems, these nonradical oxidants reversibly react with protein thiols to elicit a

redox signal that functions in physiological regulation. The mechanisms of reversible

thiol oxidation as a biological control system generally include reaction with an active

site cysteine, modification of a distal allosteric cysteine, and alteration of macromolecular

interactions (Jones, 2008). If the relevant cysteine is in an active site such modifications

may serve as a simple binary on-off “switch” to activate or inactivate the protein.

Another scenario in which the cysteine residue is not in the active site allows for

129

allosteric regulation, which can throttle protein function along a continuous gradient, like

a “dimmer”. Additionally, oxidation of cysteine thiols is known to control intra- and

inter-molecular protein interactions by formation of disulfides. Among other things,

disulfide crosslinking is well established in maintaining proper protein structure,

regulating the viscosity of mucus, and connecting actin filaments and tethering proteins

to the cytoskeleton (Farah & Amberg, 2007, Moriarty-Craige & Jones, 2004).

The GSH and TXN redox couples are central to a global system of connected elements

involving reversible oxidation of proteins containing cysteine, methionine, and Sec

residues. GPX and TXNRD selenoproteins, being critical effectors of peroxide and

thioredoxin reduction, have a direct role in the redox systems biochemistry of organisms.

Oxidative stress in the form of disrupted redox control arises when cysteine (or Sec or

methionine) residues become abnormally oxidized or irreversibly modified, stripping the

reversible redox reactivity that is required for physiological processes. Several covalent

and noncovalent modifications to distal or active site cysteine residues modify protein

structure and function. A redox-active cysteine residue can become reversibly oxidized to

sulfenic acid, which can be reduced step-wise by two molecules of reduced GSH.

Additionally, cysteinyl-S-glutathione conjugates are semi-stable in certain proteins, and

S-glutathionylation can lead to altered protein function as observed with glyceraldehyde

3-phosphate dehydrogenase (GAPDH) and caspase 3 (Mohr et al., 1999, Sykes et al.,

2007). If two neighboring reduced cysteine residues (dithiol) are reactive, the pair can be

reversibly oxidized to a form a cysteine disulfide bridge, as with TXN. The gaseous

molecules, NO and hydrogen sulfide (H2S), also react with cysteine residues to produce

modified derivatives that are functionally relevant. For example, S-nitrosylation of

Cys118 in Ras activates the protein by stimulating guanine nucleotide exchange.

Similarly sulfhydration, the addition of a sulfhydryl group from hydrogen sulfide gas to a

cysteine thiol, has been demonstrated to augment the activity of GAPDH (Lander et al.,

1996, Mustafa et al., 2009). Proteins often contain multiple cysteine residues that can be

reversibly modified (e.g. dithiol-disulfide, S-glutathionylation, S-nitrosylation, S-

sulfhydration, etc.) to affect protein function and/or interactions. Therefore

130

selenoproteins, with Sec, cysteine and methionine residues, are integral in redox control

systems and may be the most redox-active proteins in cells and organisms.

REDOX SYSTEMS AND SELENOPROTEINS IN THE BRAIN

The brain is an extremely oxygen-hungry organ. Therefore perhaps unsurprisingly,

oxidative stress is consistently implicated in brain dysfunction and disease. Despite being

roughly 2% of total body weight, it consumes approximately 20% of total body oxygen.

In other words, the rate of oxygen consumption by the brain is about ten times higher

than the rest of the body per gram of tissue. This staggering demand is highlighted by the

fact that oxygen deprivation to the brain causes loss of consciousness, brain damage, and

death within minutes (Lutz et al., 2003). The chemical reactivity between selenium,

oxygen, and sulfur produces a unique relationship between these chalcogen elements and

the mammalian nervous system.

Due to the high oxygen consumption of nervous tissue, reactive byproducts of oxygen

metabolism are continuously generated. Surprisingly however, antioxidant systems are

relatively sparse compared to other tissues. In particular, the brain has less catalase

activity and is more reliant on GPX1 for reduction of hydrogen peroxide (Dringen, 2000,

Dringen et al., 2000). However GPX activity is also low in brain compared to other

organs. The abundance of peroxide despite a relative lack of reducing capacity suggests

that peroxide may serve as an important regulatory molecule in the brain, while also

being a potentially toxic molecule when unregulated. Peroxide-mediated signaling via

effects on protein kinases, phosphatases, and transcription factors in neurons, glia, and

vascular cells is accepted, if not well understood (Bao et al., 2009). Nitric oxide is

another reactive product of oxygen metabolism with potent signaling properties. Three

isoforms of nitric oxide synthase, endothelial, inducible and neuronal, generate nitric

oxide in the presence of oxygen and arginine. The apposition of these isoforms in cell

types on both sides of the blood-brain barrier suggest that bidirectional regulation of

oxygen metabolism across the cerebral vasculature to the brain parenchyma is at least

partially mediated by nitric oxide (Zonta et al., 2003). Intriguingly SEPP1 is present on

131

both sides of the blood-brain barrier and scavenges nitric oxide and the reaction product

with superoxide, peroxynitrite (Arteel et al., 1998). Furthermore, SEPP1 knockout mice

exhibit degeneration of neurons in areas of the brain with high rates of blood perfusion

and glucose oxidation, such as the inferior colliculus (Valentine et al., 2008).

The nervous system also contains large quantities of oxidizable lipids and metals,

rendering it vulnerable to disruption of redox signaling and oxidant-mediated damage.

Genetic deletion of GPX4 in neurons causes rapid neurodegeneration, suggesting that

reduction of lipid hydroperoxides is essential for neuronal survival (Seiler et al., 2008,

Yoo et al., 2012). SEPP1 facilitates expression of GPX4 and other selenoproteins in

brain, and the neurodegeneration seen in SEPP1 knockout mice may be due to

insufficient GPX4 production. Additionally the metal-chelating properties of SEPP1 may

limit the availability of oxidizable metals, such as iron and copper, to catalyze the

nonspecific production of radical hydroxyl ions, which are highly reactive and

indiscriminately target nearby molecules (Sidenius et al., 1999). Collectively, these data

imply that SEPP1 promotes stable oxygen metabolism directly and indirectly to support a

healthy nervous system. While it directly reduces extracellular oxidants relevant to

neurovascular physiology, it indirectly stimulates the expression of other selenoproteins

involved in oxidative regulation of neuronal survival.

From a systems biology perspective, cellular redox status is understood in terms of redox

potential. The redox potential (Eh), or electron motive force, for an oxidation/reduction

couple depends on the inherent tendency of the molecule to accept/donate electrons

relative to the standard hydrogen electrode, and the concentrations of the acceptor and

donor (reduced and oxidized species of the couple). In biological systems the redox

potential of the major thiol/disulfide couples (GSH and TXN) is maintained at stable

nonequilibrium conditions, or in other words the mean Eh of the GSH and TXN couples

in subcellular compartments is nonzero. These redox couples, along with the Cys/CySS

(cysteine/cystine) couple and protein Cys residues involved in redox signaling, are

neither in equilibrium with each other nor the NADPH/NADP+ couple. In addition to

nonequilibrated redox potentials between couples within a subcellular compartment,

132

redox potential is also not equilibrated across subcellular and extracellular compartments.

The mitochondria and nucleus tend to be highly reduced (i.e. redox potential farther from

zero), while the extracellular compartment is relatively oxidized (i.e. redox potential

closer to zero), and the cytoplasm, endoplasmic reticulum, and lysosomes display

intermediate values. The steady-state Eh for thiol-disulfide couples range from -400 mV

for NADPH/NADP+ to -60 mV for plasma Cys/CySS (Kemp et al., 2008).

With an intracellular reduced GSH concentration of 1 mM, oxidation of just 18 µM of

GSH to GSSG (e.g. by GPX) will lower the GSH/GSSG redox potential (ΔEh) by ~60

mV. This relatively shallow redox potential gradient is sufficient to distinguish between

proliferating and apoptotic cells, and can theoretically drive a 100-fold change in the

dithiol:disulfide ratio in proteins with a reactive dithiol motif (Schafer & Buettner, 2001).

Therefore the current through a redox circuit need only be a fraction of the total electron

transfer in cells, provided that spatial or catalytic mechanisms are able to control reaction

rates. Under aerobic conditions oxygen and peroxide are always present in cells, therefore

coupling of electron transfer to oxidase or peroxidase (e.g. GPX) catalyzed reactions can

provide an additional energetic driving force to maintain function of low-current redox

control circuits. Kinetic and spatial insulation of reactive thiol and selenol couples allows

biological systems to be highly responsive and dynamic. Because noncatalyzed

oxidation-reduction and thiol-disulfide exchange reactions are relatively slow in the

reducing cellular environment, this enzymatic insulation concurrently provides specificity

and directionality to electron transfer reactions (Kemp et al., 2008).

S-glutathionylated Sec and diselenide bridges have not been described in mammals in

vivo, but transient and stable selenenyl-sulfide linkages are observed in selenoproteins as

either active intermediates or structural features, respectively. For example the reaction

mechanisms for GPX and TXNRD enzymes involve a transient mixed selenenyl-sulfide

bond, while rat selenoprotein P (SEPP1) has two such linkages that appear to be a

structural feature (Hatfield, 2006, Ma et al., 2005). Transient nitrosylation and

sulfhydration of Sec is thermodynamically feasible, but whether it occurs in cells and

organisms with physiological relevance is unknown. Sec is a low concentration amino

133

acid and free Sec is virtually absent from cells, implying that Sec/selenocystine and

Sec/cysteine redox couples are unlikely to be of functional importance. However

selenoproteins are present and may be maintained in a nonequilibrium steady-state

favoring the reduced form. Kinetic (i.e. reaction rate and substrate specificity) and spatial

(i.e. subcellular and tissue distribution) insulation of the various selenoproteins has been

described. Whether selenoproteins represent kinetically limiting sites of redox control

remains an important unanswered question. The standard reduction potential (E0) of

selenoproteins is unexplored to date. Determining whether the mean E0 of selenoproteins,

particularly SEPW1 and the other TXN-like selenoproteins, is maintained in a

nonequilibrium steady-state will help determine if they are central redox control points.

The Sec amino acid has known kinetic advantages compared to cysteine in catalysis, and

GPX and TXNRD selenoproteins occupy pivotal positions in respect to redox circuitry.

Therefore it is probable that more selenoproteins will be identified as kinetically limiting

redox control points that are rapidly disrupted during oxidative metabolism in the nervous

system. Several selenoproteins have unknown function but, given the reactivity of Sec

with oxygen and cysteine, they are aptly structured to transduce oxidant signals by

forming cysteine sulfenic acids and disulfide bonds in downstream target proteins, even

in the predominantly reducing cellular environment. Clarifying the role of selenoproteins

in oxygen metabolism will promote better therapies for developmental and neurological

diseases. Timely and localized administration of selenium in one form or another may in

fact be a cheap and effective treatment for a broad range of diseases.

134

REFERENCES

Aachmann, F.L., Fomenko, D.E., Soragni, A., Gladyshev, V.N. & Dikiy, A. (2007)

Solution structure of selenoprotein W and NMR analysis of its interaction with

14-3-3 proteins. J Biol Chem, 282, 37036-37044.

Arteel, G.E., Mostert, V., Oubrahim, H., Briviba, K., Abel, J. & Sies, H. (1998)

Protection by selenoprotein P in human plasma against peroxynitrite-mediated

oxidation and nitration. Biological chemistry, 379, 1201-1205.

Bao, L., Avshalumov, M.V., Patel, J.C., Lee, C.R., Miller, E.W., Chang, C.J. & Rice,

M.E. (2009) Mitochondria are the source of hydrogen peroxide for dynamic

brain-cell signaling. The Journal of neuroscience : the official journal of the

Society for Neuroscience, 29, 9002-9010.

Bellinger, F.P., Bellinger, M.T., Seale, L.A., Takemoto, A.S., Raman, A.V., Miki, T.,

Manning-Bog, A.B., Berry, M.J., White, L.R. & Ross, G.W. (2011) Glutathione

Peroxidase 4 is associated with Neuromelanin in Substantia Nigra and Dystrophic

Axons in Putamen of Parkinson's brain. Molecular neurodegeneration, 6, 8.

Benton, D. (2002) Selenium intake, mood and other aspects of psychological functioning.

Nutritional neuroscience, 5, 363-374.

Bowyer, J.F., Robinson, B., Ali, S. & Schmued, L.C. (2008) Neurotoxic-related changes

in tyrosine hydroxylase, microglia, myelin, and the blood-brain barrier in the

caudate-putamen from acute methamphetamine exposure. Synapse, 62, 193-204.

Burk, R.F. & Hill, K.E. (2009) Selenoprotein P-expression, functions, and roles in

mammals. Biochim Biophys Acta, 1790, 1441-1447.

Cadet, J.L. & Krasnova, I.N. (2009) Molecular bases of methamphetamine-induced

neurodegeneration. International review of neurobiology, 88, 101-119.

Castano, A., Ayala, A., Rodriguez-Gomez, J.A., de la Cruz, C.P., Revilla, E., Cano, J. &

Machado, A. (1995) Increase in dopamine turnover and tyrosine hydroxylase

enzyme in hippocampus of rats fed on low selenium diet. Journal of neuroscience

research, 42, 684-691.

Castano, A., Ayala, A., Rodriguez-Gomez, J.A., Herrera, A.J., Cano, J. & Machado, A.

(1997) Low selenium diet increases the dopamine turnover in prefrontal cortex of

the rat. Neurochemistry international, 30, 549-555.

Castano, A., Cano, J. & Machado, A. (1993) Low selenium diet affects monoamine

turnover differentially in substantia nigra and striatum. Journal of

neurochemistry, 61, 1302-1307.

Chen, J. & Berry, M.J. (2003) Selenium and selenoproteins in the brain and brain

diseases. Journal of neurochemistry, 86, 1-12.

Chen, L., Ding, Y., Cagniard, B., Van Laar, A.D., Mortimer, A., Chi, W., Hastings, T.G.,

Kang, U.J. & Zhuang, X. (2008) Unregulated cytosolic dopamine causes

neurodegeneration associated with oxidative stress in mice. The Journal of

neuroscience : the official journal of the Society for Neuroscience, 28, 425-433.

Cookson, M.R. (2005) The biochemistry of Parkinson's disease. Annual review of

biochemistry, 74, 29-52.

Dikiy, A., Novoselov, S.V., Fomenko, D.E., Sengupta, A., Carlson, B.A., Cerny, R.L.,

Ginalski, K., Grishin, N.V., Hatfield, D.L. & Gladyshev, V.N. (2007) SelT, SelW,

135

SelH, and Rdx12: genomics and molecular insights into the functions of

selenoproteins of a novel thioredoxin-like family. Biochemistry, 46, 6871-6882.

Dringen, R. (2000) Metabolism and functions of glutathione in brain. Progress in

neurobiology, 62, 649-671.

Dringen, R., Gutterer, J.M. & Hirrlinger, J. (2000) Glutathione metabolism in brain

metabolic interaction between astrocytes and neurons in the defense against

reactive oxygen species. Eur J Biochem, 267, 4912-4916.

Farah, M.E. & Amberg, D.C. (2007) Conserved actin cysteine residues are oxidative

stress sensors that can regulate cell death in yeast. Mol Biol Cell, 18, 1359-1365.

Glatt, S.J., Everall, I.P., Kremen, W.S., Corbeil, J., Sasik, R., Khanlou, N., Han, M.,

Liew, C.C. & Tsuang, M.T. (2005) Comparative gene expression analysis of

blood and brain provides concurrent validation of SELENBP1 up-regulation in

schizophrenia. Proceedings of the National Academy of Sciences of the United

States of America, 102, 15533-15538.

Hatfield, D.L. (2006) Selenium : its molecular biology and role in human health,

Springer, New York, NY.

Hawkes, W.C. & Alkan, Z. (2011) Delayed cell cycle progression from SEPW1 depletion

is p53- and p21-dependent in MCF-7 breast cancer cells. Biochemical and

biophysical research communications, 413, 36-40.

Hawkes, W.C., Printsev, I. & Alkan, Z. (2012) Selenoprotein W depletion induces a p53-

and p21-dependent delay in cell cycle progression in RWPE-1 prostate epithelial

cells. Journal of cellular biochemistry, 113, 61-69.

Jones, D.P. (2006) Redefining oxidative stress. Antioxid Redox Signal, 8, 1865-1879.

Jones, D.P. (2008) Radical-free biology of oxidative stress. Am J Physiol Cell Physiol,

295, C849-868.

Kemp, M., Go, Y.M. & Jones, D.P. (2008) Nonequilibrium thermodynamics of

thiol/disulfide redox systems: a perspective on redox systems biology. Free Radic

Biol Med, 44, 921-937.

Khan, H.A. (2010) Selenium partially reverses the depletion of striatal dopamine and its

metabolites in MPTP-treated C57BL mice. Neurochemistry international, 57,

489-491.

Lacourciere, G.M., Mihara, H., Kurihara, T., Esaki, N. & Stadtman, T.C. (2000)

Escherichia coli NifS-like proteins provide selenium in the pathway for the

biosynthesis of selenophosphate. J Biol Chem, 275, 23769-23773.

Lander, H.M., Milbank, A.J., Tauras, J.M., Hajjar, D.P., Hempstead, B.L., Schwartz,

G.D., Kraemer, R.T., Mirza, U.A., Chait, B.T., Burk, S.C. & Quilliam, L.A.

(1996) Redox regulation of cell signalling. Nature, 381, 380-381.

Lutz, P.L., Nilsson, G.r.E. & Prentice, H.M. (2003) The brain without oxygen : causes of

failure--physiological and molecular mechanisms for survival, Kluwer Academic

Pub., Dordrecht ; Boston.

Ma, S., Hill, K.E., Burk, R.F. & Caprioli, R.M. (2005) Mass spectrometric determination

of selenenylsulfide linkages in rat selenoprotein P. J Mass Spectrom, 40, 400-404.

Maker, H.S., Weiss, C., Silides, D.J. & Cohen, G. (1981) Coupling of dopamine

oxidation (monoamine oxidase activity) to glutathione oxidation via the

generation of hydrogen peroxide in rat brain homogenates. Journal of

neurochemistry, 36, 589-593.

136

Mihara, H., Kurihara, T., Watanabe, T., Yoshimura, T. & Esaki, N. (2000) cDNA

cloning, purification, and characterization of mouse liver selenocysteine lyase.

Candidate for selenium delivery protein in selenoprotein synthesis. J Biol Chem,

275, 6195-6200.

Mohr, S., Hallak, H., de Boitte, A., Lapetina, E.G. & Brune, B. (1999) Nitric oxide-

induced S-glutathionylation and inactivation of glyceraldehyde-3-phosphate

dehydrogenase. J Biol Chem, 274, 9427-9430.

Moriarty-Craige, S.E. & Jones, D.P. (2004) Extracellular thiols and thiol/disulfide redox

in metabolism. Annu Rev Nutr, 24, 481-509.

Musiani, F., Ciurli, S. & Dikiy, A. (2010) The interaction of selenoprotein W with 14-3-3

proteins: a computational approach. J Proteome Res.

Mustafa, A.K., Gadalla, M.M., Sen, N., Kim, S., Mu, W., Gazi, S.K., Barrow, R.K.,

Yang, G., Wang, R. & Snyder, S.H. (2009) H2S signals through protein S-

sulfhydration. Sci Signal, 2, ra72.

Romero-Ramos, M., Venero, J.L., Cano, J. & Machado, A. (2000) Low selenium diet

induces tyrosine hydroxylase enzyme in nigrostriatal system of the rat. Brain

research. Molecular brain research, 84, 7-16.

Schafer, F.Q. & Buettner, G.R. (2001) Redox environment of the cell as viewed through

the redox state of the glutathione disulfide/glutathione couple. Free Radic Biol

Med, 30, 1191-1212.

Seiler, A., Schneider, M., Forster, H., Roth, S., Wirth, E.K., Culmsee, C., Plesnila, N.,

Kremmer, E., Radmark, O., Wurst, W., Bornkamm, G.W., Schweizer, U. &

Conrad, M. (2008) Glutathione peroxidase 4 senses and translates oxidative stress

into 12/15-lipoxygenase dependent- and AIF-mediated cell death. Cell Metab, 8,

237-248.

Sekine, Y., Minabe, Y., Ouchi, Y., Takei, N., Iyo, M., Nakamura, K., Suzuki, K.,

Tsukada, H., Okada, H., Yoshikawa, E., Futatsubashi, M. & Mori, N. (2003)

Association of dopamine transporter loss in the orbitofrontal and dorsolateral

prefrontal cortices with methamphetamine-related psychiatric symptoms. The

American journal of psychiatry, 160, 1699-1701.

Sidenius, U., Farver, O., Jons, O. & Gammelgaard, B. (1999) Comparison of different

transition metal ions for immobilized metal affinity chromatography of

selenoprotein P from human plasma. Journal of chromatography. B, Biomedical

sciences and applications, 735, 85-91.

Sykes, M.C., Mowbray, A.L. & Jo, H. (2007) Reversible glutathiolation of caspase-3 by

glutaredoxin as a novel redox signaling mechanism in tumor necrosis factor-

alpha-induced cell death. Circ Res, 100, 152-154.

Tang, Y.L., Wang, S.W. & Lin, S.M. (2008) Both inorganic and organic selenium

supplements can decrease brain monoamine oxidase B enzyme activity in adult

rats. The British journal of nutrition, 100, 660-665.

Valentine, W.M., Abel, T.W., Hill, K.E., Austin, L.M. & Burk, R.F. (2008)

Neurodegeneration in mice resulting from loss of functional selenoprotein P or its

receptor apolipoprotein E receptor 2. Journal of neuropathology and experimental

neurology, 67, 68-77.

Volkow, N.D., Chang, L., Wang, G.J., Fowler, J.S., Franceschi, D., Sedler, M.J., Gatley,

S.J., Hitzemann, R., Ding, Y.S., Wong, C. & Logan, J. (2001a) Higher cortical

137

and lower subcortical metabolism in detoxified methamphetamine abusers. The

American journal of psychiatry, 158, 383-389.

Volkow, N.D., Chang, L., Wang, G.J., Fowler, J.S., Leonido-Yee, M., Franceschi, D.,

Sedler, M.J., Gatley, S.J., Hitzemann, R., Ding, Y.S., Logan, J., Wong, C. &

Miller, E.N. (2001b) Association of dopamine transporter reduction with

psychomotor impairment in methamphetamine abusers. The American journal of

psychiatry, 158, 377-382.

Watanabe, C., Kasanuma, Y. & Satoh, H. (1997) Deficiency of selenium enhances the

K+-induced release of dopamine in the striatum of mice. Neuroscience letters,

236, 49-52.

Whanger, P.D. (2001) Selenium and the brain: a review. Nutritional neuroscience, 4, 81-

97.

Yoo, S.E., Chen, L., Na, R., Liu, Y., Rios, C., Van Remmen, H., Richardson, A. & Ran,

Q. (2012) Gpx4 ablation in adult mice results in a lethal phenotype accompanied

by neuronal loss in brain. Free Radic Biol Med, 52, 1820-1827.

Zafar, K.S., Siddiqui, A., Sayeed, I., Ahmad, M., Salim, S. & Islam, F. (2003) Dose-

dependent protective effect of selenium in rat model of Parkinson's disease:

neurobehavioral and neurochemical evidences. Journal of neurochemistry, 84,

438-446.

Zonta, M., Angulo, M.C., Gobbo, S., Rosengarten, B., Hossmann, K.A., Pozzan, T. &

Carmignoto, G. (2003) Neuron-to-astrocyte signaling is central to the dynamic

control of brain microcirculation. Nature neuroscience, 6, 43-50.