ArticlePDF Available

High levels of thioredoxin reductase 1 modulate drug-specific cytotoxic efficacy

Authors:

Abstract and Figures

The selenoprotein thioredoxin reductase 1 (TrxR1) is currently recognized as a plausible anticancer drug target. Here we analyzed the effects of TrxR1 targeting in the human A549 lung carcinoma cell line, having a very high basal TrxR1 expression. We determined the total cellular TrxR activity to be 271.4 +/- 39.5 nmol min(-1) per milligram of total protein, which by far exceeded the total thioredoxin activity (39.2 +/- 3.5 nmol min(-1) per milligram of total protein). Knocking down TrxR1 by approx 90% using siRNA gave only a slight effect on cell growth, irrespective of concurrent glutathione depletion (> or = 98% decrease), and no increase in cell death or distorted cell cycle phase distributions. This apparent lack of phenotype could probably be explained by Trx functions being maintained by the remaining TrxR1 activity. TrxR1 knockdown nonetheless yielded drug-specific modulation of cytotoxic efficacy in response to various chemotherapeutic agents. No changes in response upon exposure to auranofin or juglone were seen after TrxR1 knockdown, whereas sensitivity to 1-chloro-2,4-dinitrobenzene or menadione became markedly increased. In contrast, a virtually complete resistance to cisplatin using concentrations up to 20 microM appeared upon TrxR1 knockdown. The results suggest that high overexpression of TrxR has an impact not necessarily linked to Trx function that nonetheless modulates drug-specific cytotoxic responses.
Content may be subject to copyright.
Original Contribution
High levels of thioredoxin reductase 1 modulate drug-specic cytotoxic efcacy
SoE. Eriksson
a
, Stefanie Prast-Nielsen
a
, Emilie Flaberg
b
, Laszlo Szekely
b
, Elias S.J. Arnér
a,
a
Division of Biochemistry, Department of Medical Biochemistry and Biophysics, Karolinska Institutet, SE-171 77 Stockholm, Sweden
b
Department of Microbiology, Tumor and Cell Biology, Karolinska Institutet, SE-171 77 Stockholm, Sweden
abstractarticle info
Article history:
Received 17 June 2009
Revised 31 August 2009
Accepted 14 September 2009
Available online 17 September 2009
Keywords:
Thioredoxin reductase
Selenoprotein
Cancer
Cytotoxicity
Chemotherapy
Free radicals
The selenoprotein thioredoxin reductase 1 (TrxR1) is currently recognized as a plausible anticancer drug
target. Here we analyzed the effects of TrxR1 targeting in the human A549 lung carcinoma cell line, having a
very high basal TrxR1 expression. We determined the total cellular TrxR activity to be 271.4 ± 39.5 nmol
min
1
per milligram of total protein, which by far exceeded the total thioredoxin activity (39.2 ± 3.5 nmol
min
1
per milligram of total protein). Knocking down TrxR1 by approx 90% using siRNA gave only a slight
effect on cell growth, irrespective of concurrent glutathione depletion (98% decrease), and no increase in
cell death or distorted cell cycle phase distributions. This apparent lack of phenotype could probably be
explained by Trx functions being maintained by the remaining TrxR1 activity. TrxR1 knockdown nonetheless
yielded drug-specic modulation of cytotoxic efcacy in response to various chemotherapeutic agents. No
changes in response upon exposure to auranon or juglone were seen after TrxR1 knockdown, whereas
sensitivity to 1-chloro-2,4-dinitrobenzene or menadione became markedly increased. In contrast, a virtually
complete resistance to cisplatin using concentrations up to 20 μM appeared upon TrxR1 knockdown. The
results suggest that high overexpression of TrxR has an impact not necessarily linked to Trx function that
nonetheless modulates drug-specic cytotoxic responses.
© 2009 Elsevier Inc. All rights reserved.
To protect from oxygen radical-induced damage, cells have
developed multifaceted antioxidant systems [1]. The thioredoxin
(Trx) system is one important enzymatic network for antioxidant
defense and has also a number of additional redox regulatory
functions [1,2]. Trx reduces and thereby supports the activity of
several proteins, e.g., antioxidant peroxiredoxins or methionine
sulfoxide reductase and redox-regulated transcription factors or
other signaling molecules [2,3]. The Trx system also supports
deoxyribonucleotide synthesis, with reduced Trx regenerating a
dithiol in ribonucleotide reductase oxidized upon each cycle of
catalysis [4]. For its enzymatic function, Trx must be reduced by
TrxR (EC 1.8.1.9). In human, three TrxR-encoding genes are found, i.e.,
TXNRD1, encoding TrxR1, which is the major TrxR form in most cells
[5,6];TXNRD2, encoding TrxR2, which is predominantly found in
mitochondria [7]; and TXNRD3, for thioredoxin glutathione reductase
(TGR) mainly expressed in testis [8]. Both TrxR1 and TrxR2 are
essential for embryonic development as shown in knockout mouse
models [9,10], as are their principal substrates Trx1 and Trx2 [11,12].
TrxR proteins are the only enzymes known to reduce Trx and are
thereby believed to be essential for all Trx-dependent reduction
processes. Being selenoproteins, mammalian TrxR's and thereby the
complete Trx systems are fully dependent upon selenium [13,14].
Increasing selenium availability generally results in increased TrxR1
activity, until saturation levels are reached [1518].
With regard to cancer development and treatment, most studies
have focused on the potential importance of TrxR1 as a drug target,
as reviewed elsewhere [1922] and as shall briey be introduced
here. Several observations show that both TrxR1 and Trx1 are often
overexpressed in various cancer forms under basal selenium supply
[21]. Furthermore, TrxR1 has been shown to promote and even be
essential for tumor growth in xenograft cancer models [23,24].
Cancer cells with lowered TrxR1 expression have also been shown to
be more sensitive to UV irradiation or to low doses of cadmium
[25,26]. It is not only for its growth-promoting properties that TrxR1
has been identied as a potential target for anticancer therapy, but
also because of its selenium-dependent enzymological properties.
The Sec residue is an integral part of the TrxR1 active site [27,28].
With a high reactivity of Sec and an easily accessible active site,
TrxR1 has a broad substrate specicity, reducing not only Trx1, but
also other agents, e.g., 5,5-dithiobis[2-nitrobenzoic acid] (DTNB),
Free Radical Biology & Medicine 47 (2009) 16611671
Abbreviations: Trx, thioredoxin; TrxR, thioredoxin reductase; DTNB, 5,5-dithiobis
[2-nitrobenzoic acid]; Sec, selenocysteine; DNCB, 1-chloro-2,4-dinitrobenzene; cDDP,
cisplatin; Oxa, oxaliplatin; SecTRAP, selenium-compromised thioredoxin reductase-
derived apoptotic protein; A549, human lung carcinoma cell line; PEST, penicillin and
streptomycin; siRNA, small inhibitory RNA; FCS, fetal calf serum; GR, glutathione
reductase; GSH, glutathione; GSSG, glutathione disulde; PI, propidium iodide; BSO, L-
buthionine sulfoximine; Grx, glutaredoxin; ROS, reactive oxygen species; PBS,
phosphate-buffered saline; NADPH, nicotinamide adenine dinucleotide phosphate.
Corresponding author. Fax: +46 8 31 15 51.
E-mail address: Elias.Arner@ki.se (E.S.J. Arnér).
0891-5849/$ see front matter © 2009 Elsevier Inc. All rights reserved.
doi:10.1016/j.freeradbiomed.2009.09.016
Contents lists available at ScienceDirect
Free Radical Biology & Medicine
journal homepage: www.elsevier.com/locate/freeradbiomed
lipoic acid, lipoamide, selenite, menadione (vitamin K
3
), ubiquinone,
and vitamin C [1,5,2933]. Moreover, the N-terminal CVNVGC motif
can reduce certain substrates directly [34]. In addition, the Sec
residue in reduced (but not oxidized) TrxR can also be rapidly
targeted by electrophilic agents. Several gold compounds are potent
inhibitors of TrxR1, e.g., auranon used in treatment of rheumatoid
arthritis, which inhibits the enzyme even in the nanomolar range
[7,35,36]. Recent results also show that auranon inhibits overall
selenoprotein synthesis [37]. The rst identied inhibitor of TrxR1
was 1-chloro-2,4-dinitrobenzene (DNCB), which also concomitantly
induces an inherent superoxide-producing NADPH oxidase activity in
the derivatized enzyme [38]. Other inhibitors of TrxR1 include
naturally occurring electrophiles such as quinones [39], isothiocya-
nates [40], mercury [41], or various avonoids [42,43], as well as
several clinically used anticancer drugs including arsenicals [44] and
cisplatin (cDDP) [4547]. TrxR1 derivatized with cDDP generates an
enzyme inactive for its normal function, but which may gain a new
pro-oxidant function within a cellular context, capable of inducing a
rapid cell death in the form of selenium-compromised thioredoxin
reductase-derived apoptotic proteins, SecTRAPs [48,49]. All of these
properties combined have contributed to the commonly held view
that TrxR1 may be a prime molecular target for anticancer
chemotherapy [1922].
The purpose of this study was to investigate the impact of an
endogenously high expression of the selenium-dependent TrxR1 in
cancer cells, using knockdown of the enzyme in a human lung
carcinoma cell line (A549) that has among the highest known
overexpression of TrxR1 of all studied cancer cells (www.proteinatlas.
org). By a thoroughanalysis of the phenotype of these cells upon TrxR1
knockdown, we reasoned that insights could be gained regarding the
functional consequences of overexpression of the enzyme in cancer
cells and of its targeting in such cells by chemotherapy. The results
reveal a surprising complexity in the roles of TrxR1 for modulating
cellular responses to different cytotoxic agents.
Experimental procedures
Chemicals and reagents
Recombinant rat TrxR1 was produced as described [50] and
human wt Trx [51] was kindly provided by Arne Holmgren
(Karolinska Institutet, Stockholm, Sweden). Yeast glutathione reduc-
tase (GR) (Cat. No. G3664), reduced glutathione (GSH), and oxidized
glutathione (GSSG) were obtained from SigmaAldrich Chemicals
(Steinheim, Germany). Chemicals used in the drug sensitivity assays
were as follows: auranon from Alexis Biochemicals (San Diego, CA,
USA), cDDP (Platinol) from BristolMyers Squibb Pharmaceuticals
(New York, NY, USA), and DNCB, juglone (5-hydroxy-1,4,-naphtho-
quinone), menadione (2-methyl-1,4-naphtoquinone, vitamin K
3
),
and oxaliplatin (Oxa) all from SigmaAldrich Chemicals. All other
regular chemicals or reagents were of high purity and purchased from
SigmaAldrich Chemicals, unless otherwise specied.
Cell cultures
Human lung carcinoma cells (A549 cells) were obtained from the
American Tissue Culture Collection (CCL-185) and cultivated in
Dulbecco's modied Eagle medium with 4.5 g/l glucose content
(GIBCO/Invitrogen, Carlsbad, CA, USA). The medium was supplemen-
ted with 10% heat-inactivated fetal calf serum (FCS), 2 mM L-
glutamine, 100 μg/ml streptomycin, and 100 units/ml penicillin
(PEST), all from PAA Laboratories (Pasching, Austria). Selenium was
added to the medium in the form of sodium selenite at concentrations
as described in the text. Cells were grown at 37°C in a humidied
atmosphere with 5% CO
2
and kept under logarithmic growth phase for
all experiments unless stated otherwise.
Transient knockdown of TrxR1
Small interfering RNA (siRNA) molecules, specically targeting the
TrxR1 mRNA, were obtained from Qiagen (Valencia, CA, USA). For
phenotypic conrmation, two different siRNA sequences were used
(herein named Siseq1 and Siseq2), targeting different areas of the
TrxR1 mRNA: Siseq1, sense 5-(GCAAGACUCUCGAAAUUAU)dTdT-3,
antisense 5-(AUAAUUUCGAGAGUCUUGC)dAdG-3, and Siseq2, sense
5-(CCUGGCAUUUGGUAGUAUA)dTdT-3, antisense 5-(UAUACUAC-
CAAAUGCCAGG)dCdA-3. Siseq1 targets an mRNA region encoding
the N-terminal redox-active CVNVGC motif in the protein and Siseq2
targets a site in the exon covering the 3-untranslated region (3UTR)
of the mRNA downstream of the selenocysteine insertion sequence
element, needed for Sec incorporation [52]. Both siRNA constructs
should thereby knock down all major TrxR1 splice forms (Supple-
mentary Fig. S1). Two different nonsilencing siRNA controls, showing
no apparent homology to any region of the human genome, were
used, i.e., the Alexa 488-labeled AllStar negative control (Qiagen) as
control for transfection efciency and an unlabeled scramble control
(mock) (Cat. No. 1022076; Qiagen) used for all other experiments.
The sequences for the mock siRNA construct were sense 5-
(UUCUCCGAACGUGUCACGU)dTdT-3and antisense 5-(ACGUGA-
CACGUUCGGAGAA)dTdT-3. Untransfected control cells as well as
mock-treated cells were included in all siRNA experiments, as
described in the text. For transfection experiments A549 cells were
seeded in six-well plates at a density of about 30,00032,000 cells per
well 1518 h before transfection. SiRNA transfection was then
performed according to the manufacturer's protocol by mixing 9 μl
transfection reagent (Hiperfect; Qiagen) and 10 nM siRNA duplexes
in a total volume of 100 μl serum-free medium, per sample. Cells
were incubated for 24 h with the transfection complexes in 2.4 ml
medium without antibiotics, whereupon the medium was replaced
with fresh PEST-containing medium and experiments were con-
ducted as described.
Selenium content in medium
Selenium (m/z 78) content in the fetal calf serum utilized in this
study was analyzed with inductively coupled plasma mass spectrom-
etry as described elsewhere [53] and was determined to be 18.7 ng Se
per gram of FCS, i.e., approx 18.7 μg/L. In the medium supplemented
with 10% FCS, the total selenium concentration would thereby be in
the range of 2025 nM. In the experiments as described in the text, an
additional 25 nM selenium in the form of sodium selenite was added
to the cells during seeding to saturate the synthesis of TrxR1 in the cell
cultures. Extensive control experiments veried that this was a
required but sufcient selenium supplementation to reach TrxR1
saturation.
75
Se radioisotope labeling of cellular proteins
To verify knockdown of TrxR1, cells were seeded in six-well plates
together with
75
Se-labeled selenite (Research Reactor Center, Uni-
versity of Missouri, Columbia, MO, USA) added to the medium, using
2μCi/ml in a total volume of 1.5 ml. No additional unlabeled selenite
was added. Using the specication from the distributors the total
concentration of selenite added to the cells was calculated to be
within the range of 20 to 100 nM. Fifteen to eighteen hours after
seeding with [
75
Se]selenite, siRNA was added to the cells in the
presence of 2 μCi/ml
75
Se, according to the same procedures as
described above. Cell extracts were subsequently prepared 48, 72, 96,
and 144 h post-siRNA treatment and 25 μg total protein per sample
was analyzed on 412% NuPAGE BisTris reducing SDSPAGE
(equipment, gels, and buffers from Invitrogen). Gels were stained
with Coomassie blue to visualize total protein. The gels were then
dried and exposed on a phosphor screen. The autoradiography was
1662 S.E. Eriksson et al. / Free Radical Biology & Medicine 47 (2009) 16611671
visualized with a StormScan PhosphoImager (Molecular Dynamics,
Sunnyvale, CA, USA). Intensities of the
75
Se-labeled bands with an
approx molecular weight of 55 kDa were quantied using Bio-Rad's
Quantity One 1-D 4.6.7 (Hercules, CA, USA).
Preparation of cell extracts
Cells to be used for enzymatic assays, chromatography, or
autoradiography were harvested by trypsinization (GIBCO/Invitrogen),
washed with phosphate-buffered saline, pH 7.4 (PBS; GIBCO/
Invitrogen), and centrifuged (800 g for 5 min). The resulting cell
pellets were resuspended in extraction buffer containing 50 mM Tris
HCl, pH 7.5, 2 mM EDTA (Merck, Darmstadt, Germany), 0.5 mM
phenyl methyl sulfonyl uoride, and 0.5% nonionic detergent (Igepal
Ca-630). Cells were lysed by rapid cycles of freezingand thawing using
liquid nitrogen and a 37°C water bath, respectively. The nonsolubilized
fraction was subsequently removed by centrifugation (16,000 g at 4°C
for 5 min) and the supernatant was used as protein source. The
protein concentrations were determined by the Bradford method (Bio-
Rad Laboratories), using bovine serum albumin as a standard.
Anion-exchange chromatography and mass spectrometry
To determine the identity of the remaining TrxR activity in siRNA-
treated cells, the total protein of the corresponding cell extracts was
subjected to chromatography using a strong anion Resource Q column
(1 ml; GE Healthcare, Little Chalfont, UK) on an ÄKTAexplorer 900
purier HPLC system (GE Healthcare). Buffers for the gradient elution
were as follows: buffer A, 50 mM TrisCl and 2 mM EDTA (TE buffer),
pH 7.5, and buffer B, TE buffer, pH 7.5, with 1 M NaCl. The system was
equilibrated with buffer A, whereupon 300 μg of total protein in a total
volume of 0.5 ml was applied to the column, which was then washed
with buffer A and the owthrough was monitored using absorbance at
280 and 254 nm. Separation was performed using a three-step
gradient as shown, with TrxR activity eluted at 50200 mM NaCl.
Fractions of a xed volume (0.5 ml) were continuously collected using
an automated fraction collector (Frac-950; GE Healthcare) and
analyzed for TrxR activity as described. Fractions with the highest
TrxR activity were pooled and concentrated using a 30-kDa cutoff
lter column (Microcon 0.5 ml; Millipore, Bedford, MA, USA).
Concentrated samples were separated by SDSPAGE and subsequent-
ly subjected to mass spectrometry (MS). Recombinant rat TrxR1
protein was used as migration control in the gel. Briey, Coomassie-
stained bands in the size range of 55 kDa were cut out from the gel and
subsequently subjected to in-gel trypsin digestion, peptide extraction,
mass mapping, and database searches carried out at the Protein
Analysis Center, Karolinska Institutet. The details of a typical result are
shown in supplementary Fig. S2.
TrxR and Trx activity assays
TrxR activity was determined using the previously described
end-point Trx-dependent insulin reduction method [54], modied
and applied to microtiter plates [39]. Total cellular protein (5 μg)
was incubated with 20 μM recombinant human wt Trx [51] in the
presence of 297 μM insulin, 1.3 mM NADPH (AppliChem, Darmstadt,
Germany), 85 mM Hepes buffer, pH 7.6, and 13 mM EDTA for
40 min at 37°C, in a total volume of 50 μl. The reaction was then
stopped by addition of 200 μl of 7.2 M guanidineHCl (Acros
Organics, NJ, USA) in 0.2 M TrisHCl, pH 8.0, containing 1 mM
DTNB. The extent of Trx-dependent formed thiols in the reduced
insulin was then determined by measuring absorbance at 412 nm
(extinction coefcient 13,600 M
1
cm
1
)usingaVersaMaxmicro-
plate reader (Molecular Devices, Sunnyvale, CA, USA) with a
background absorbance reference for each sample containing all
components except Trx, incubated and treated in the same manner.
To determine Trx activity in the samples, 5 μg total protein was
incubated with 250 nM rat TrxR1 (22 U/mg) instead of Trx for
60 min, with otherwise the same conditions as described for the TrxR
activity assay. TrxR activity assays on the fractions from the HPLC
purications were also performed with the Trx-coupled insulin
reduction assay using 10 μl of the fractions and incubation for 90 min.
GR activity assay
To monitor NADPH-dependent GSSG reduction, the spectropho-
tometric method previously described [55] and modied for micro-
titer plates [39] was used. In short, 2 μg of total cellular protein was
diluted in deionized water to a nal volume of 80 μl. Then 120 μlof
freshly prepared master mix containing 2 mM GSSG and 200 μM
NADPH in 0.2 M potassium phosphate, pH 7, with 2 mM EDTA (PE
buffer) was added simultaneously to all samples. These assays were
performed in 96-well microtiter plates at 30°C with oxidation of
NADPH determined from the decrease in absorbance at 340 nm for the
rst 3 min using a VersaMax microplate reader and an extinction
coefcient of 6200 M
1
cm
1
. A background sample containing
everything except GSSG was included in each case and its change in
absorbance was subtracted from the sample values.
Cell growth, cell cycle phase, and viability assessments
To measure cell death and analyze the DNA content prole by ow
cytometry, all oating as well as attached cells were recovered. The
cells were xed using addition of ice-cold 70% ethanol and then
incubated at room temperature for 30 min. Subsequently DNA
staining was done by incubating for 30 min at 4°C with PBS containing
20 μg/ml propidium iodide (PI), 0.2 mg/ml RNase, and 0.1% Triton X-
100. Samples were subsequently analyzed in a uorescence-activated
cell sorter (BectonDickinson FACScan, Rutherford, NJ, USA). Data
collected from 10,000 cells per sample were analyzed with Becton
Dickinson Cell Quest Pro version 4.0.2.
GSH depletion
To investigate the importance of GSH-dependent pathways in the
TrxR1-siRNA-treated cells, GSH levels were lowered by preincubation
with 0.25 mM L-buthionine sulfoximine (BSO). After siRNA transfec-
tion BSO was added to the cells and was subsequently present in the
incubation for an additional 48 h before analysis. Total cell numbers
upon treatments were determined using a Bürker hemacytometer
chamber.
Quantication of total GSH and GSSG
Measurements of total intracellular GSH and GSSG concentrations
were performed according to the glutathione reductaseDTNB
recycling assay [56], further improved and modied to a microtiter
plate format [57]. For this, A549 cells were trypsinized, washed with
cold PBS, and centrifuged (800 g for 5 min). The resulting cell pellets
were immediately resuspended in 150 μl ice-cold 10 mM HCl and
lysed by three cycles of freezing and thawing using liquid nitrogen
and a 37°C water bath, respectively. To determine the protein
concentrations in the samples 10-μl aliquots were rst transferred
to new separate tubes. Proteins were subsequently precipitated from
the main samples by adding 30 μl 5% (w/v) 5-sulfosalicylic acid (SSA)
to 120 μl total sample volume. The samples were then incubated for
10 min, after which the proteins were removed by centrifugation for
15 min (8000 g, at 4°C). The resulting supernatant was collected and
stored at 80°C until analysis of glutathione content. To microtiter
plate wells 20 μl of sample, background control, or GSH standard
solution was added. To neutralize the pH, 20 μl of 143 mM NaH
2
PO
4
buffer containing 6.3 mM EDTA, pH 7.5 (stock buffer), was added. A
1663S.E. Eriksson et al. / Free Radical Biology & Medicine 47 (2009) 16611671
solution of 1.1 mM DTNB and 0.35 mM NADPH (prepared in stock
buffer) was subsequently added to all wells (170 μl/well), and after
5 min incubation at room temperature 40 μl yeast GR solution was
added (nal concentration of GR was 1.2 U/ml). The change in
absorbance was followed for 2 min at 412 nm using the Versamax
plate reader and the concentration of GSH in the samples was
calculated using a standard curve generated from GSH solutions of
known concentrations (nal concentrations of GSH were in the range
of 0.1258μM). All GSH standard solutions were diluted in 10 mM HCl
containing 1% SSA, similar to the samples.
Drug sensitivity assays
Cells were cultured and transfected with siRNA as described above.
The cells were trypsinized and subsequently seeded together with the
indicated concentrations of drugs in 384-well plates 48 h after the
siRNA transfection. Depending on the drug used, viability and growth
of cells were examined either 24 or 48 h after incubation at 37°C in 5%
CO
2
. Each well was loaded with a cell suspension containing either
600 cells for the 24-h incubation or 300 cells for the 48-h time point.
Samples were always run in triplicate. After incubation, cells were
stained with a uorescent dye mixture (Vital dye; Biomarker),
staining living cells green and the nuclei of dead cells red. Each well
of the 384-well plate was imaged using a custom-modied automated
microscope system, including a motorized Nikon Diaphot 200
uorescence microscope (Nikon, Japan), a motorized XY table
(Märzhauser, Germany), an ORCA ER cold CCD camera with a detector
array 1344×1024 px (Hamamatsu, Japan), and an X-Cite, 120-W
mercury lamp for uorescence illumination (EXFO, Canada). In this
study a Plan 2.5×/NA 0.08 Pol objective was used. Images were
captured using the program Platefocus_10, developed by two of the
authors (Flaberg and Szekely) in the visual programming language
environment of the Open Lab automator, as a modied version of the
original EFLCM method [58] with the addition of an autofocus
function. For the live and dead discrimination, two images were
captured for each well using either green or red uorescence,
respectively. The number of red and green cells was then counted
using the Analyze Particle function in ImageJ software. A typical
original image with its resulting analysis is shown in Supplementary
Fig. S3.
Statistics
Values are presented as means ± standard error of the mean
(SEM) and represent at least two independent experiments.
Statistical evaluation of data from the cytotoxicity assays was
performed with the MannWhitney test using the GraphPad Prism
computer program, version 5.0 (GraphPad Software, San Diego, CA,
USA). Asterisks denote statistically signicant differences: pb0.05,
⁎⁎pb0.01, and ⁎⁎⁎pb0.001.
Fig. 1. Selenoprotein expression pattern in A549 cells and conrmation of TrxR1 knockdown. (A)
75
Se incorporation experiment. Cells were seeded in the presence of
75
Se-labeled
selenite 1518 h before siRNA treatment. Cells were harvested 48 h and up to 6 days after the transfection, with the resulting cell extracts separated by SDSPAGE that was exposed
to autoradiography. The arrows indicate the 55-, 57-, and 130-kDa bands representing three forms of TrxR1. The plot on the right shows the correlation between the optical density
of the 55-kDa band and TrxR activity in the cellular lysate. In (BD) anion chromatography analyses are shown. Equal protein amounts (300 μg) from (B) mock, (C) TrxR1 Siseq1, or
(D) TrxR1 Siseq2 cell lysates were analyzed on an anion column, with the salt gradient and protein elution prole as shown with all fractions analyzed for TrxR activity (gray bars).
The indicated fractions (asterisks) having the most signicant TrxR activity were pooled, concentrated, and separated by SDSPAGE. Coomassie-stained bands in the size range of
55 kDa from the sample treated with TrxR1 Siseq1 were subsequently subjected to tryptic digestion and MS analysis (for results from MS analysis see Supplementary Fig. S2).
1664 S.E. Eriksson et al. / Free Radical Biology & Medicine 47 (2009) 16611671
Results
Conrming knockdown of TrxR1
The aim of this study was to assess the cellular phenotype upon
siRNA-mediated knockdown of the selenium-dependent TrxR1 in the
highly TrxR1-overexpressing A549 lung carcinoma cells. First an Alexa
488-labeled AllStar negative siRNA (Qiagen) was used as a control for
transfection efciency, and under the utilized transfection conditions
the efciency was at least 95% (data not shown). Specic siRNA-
mediated knockdown of TrxR1 was rst conrmed using
75
Se labeling
of all cellular selenoproteins, which vividly illustrated TrxR as the
Fig. 2. Analysis of TrxR, Trx, and GR activities. (A) Thioredoxin reductase, (B)
thioredoxin, or (C) glutathione reductase activities were measured in cell extracts
from A549 cells 72 h post-siRNA treatment, using the Trx-dependent insulin reduction
assay (A and B) or a direct GSSG reduction assay (C). Means ± SEM are shown and
represent results from at least three independent experiments.
Fig. 3. Cell cycle distribution after TrxR1 knockdown. Cell cycle phase distribution
analysis was performed 72 h after siRNA transfection. Cells were xed and stained with
PI and the samples (10,000 cells per sample) were then analyzed by ow cytometry.
Results (means ± SEM) are presented as percentage of total cell number in each cell
cycle phase.
Fig. 4. GSH depletion in TrxR1 knockdown cells. (A) Cell growth was analyzed by
counting the total number of A549 cells per sample 72 h post-siRNA transfection with
or without BSO treatment. The initial number of cells at 0 h was inthe rangeof 30,000 to
32,000 in each treatment. Values represent means ± SEM of at least three independent
experiments. (B) Measurement of total glutathione concentrations in cells after BSO
treatment was done using the glutathione reductaseDTNB recycling assay. DTNB is
reduced by GSH to TNB
. The resulting GSSG can then be recycled into GSH by GR. The
measured rate of TNB
formation is thereby proportional to the sum of total cellular
glutathione present in the reaction and the amount was calculated using a standard
curve generated from GSH solutions of known concentrations. Before the assay was run,
all protein in the samples was precipitated by adding 5-sulfosalicylic acid. Samples and
standards were all measured in duplicate and values represent means ± SEM of at least
two independent experiments.
Table 1
TrxR activity after drug treatment in cells treated with TrxR1-targeting or mock siRNA
Drug
b
TrxR activity (%)
a
Conc. (μM) Mock Siseq1 (TrxR1)
Untreated 100 ± 0.5 7.6 ± 1.5
DNCB 5 63.1 ± 7.4 3.7 ± 0.2
DNCB 20 3.0 ± 1.4 4.5 ± 1.0
Menadione 5 95.8 ± 4.2 14.9 ± 0.5
Menadione 20 92.5 ± 0.2 7.7 ± 0.6
Auranon 2.5 41.8 ± 5.0 5.3 ± 0.5
Auranon 10 1.8 ± 0.2 2.3 ± 0.1
Cisplatin 15 55.8 ± 4.7 3.5 ± 0.9
Cisplatin 60 36.5 ± 1.0 2.4 ± 0.1
Oxaliplatin 15 93.2 ± 0.3 7.0 ± 2.8
Oxaliplatin 60 57.5 ± 3.0 3.6 ± 0.3
Juglone 1.25 96.0 ± 1.4 16.3 ± 1.5
Juglone 5 64.2 ± 8.6 5.5 ± 0.8
A549 cells seeded in the presence of 25 nM selenite were treated with various TrxR-
interacting drugs, added 48 h after siRNA transfection, whereupon total cellular TrxR
activity was determined using cell lysates in an insulin-coupled Trx reduction assay.
a
TrxR activity measured in untreated mock cells was set to 100%. Values are
means ± SEM of two independent experiments.
b
Cells treated with DNCB, menadione, and auranon were incubated for 24 h,
whereas cisplatin, oxaliplatin, and juglone treatment lasted for 48 h. The same
conditions were used for the cell viability analysis shown in Figs. 5A and 5B.
1665S.E. Eriksson et al. / Free Radical Biology & Medicine 47 (2009) 16611671
most abundant selenoprotein in A549 cells and also that the method
for knockdown was efcient and long-lasting, up to several days (Fig.
1A). We also noted that, together with the very strong 55-kDa band
efciently knocked down by the siRNA, corresponding to the major
TrxR1 form TXNRD1_v1 [59], another
75
Se-labeled band of approx
57 kDa and one weak band at approx 130 kDa were also knocked
down (Fig. 1A). The 57-kDa band could correspond to TXNRD1_v2
[59],whereasthe130-kDa band could possibly be a yet
unidentied form of TrxR1. The intensity of the 55-kDa band
correlated with the total cellular TrxR activity (Fig. 1A). Generally,
the most efcient knockdown with the lowest cellular TrxR activity
and weakest 55-kDa
75
Se-labeled band intensity was seen between
72 and 96 h after siRNA transfection. Cell extracts from control cells
seeded and grown in selenite-supplemented medium (25 nM) had a
total cellular TrxR activity of 271.4 ± 39.5 nmol min
1
mg protein
1
,
whereas TrxR activity in TrxR1 siRNA-treated cells was 31.4 ±
Fig. 5. Drug toxicity in relation to TrxR1 levels. (A and B) Upper four rows show the distribution between live and dead cells (dark gray, live cells, and light gray, dead cells) after the
corresponding treatments as indicated. Means ± SEM from at least two independent experiments done in triplicate are shown. The bottom rows show the percentage of dead cells
of the total cell number per well for mock-or Siseq1-treated cells. Signicant differences in sensitivity are indicated: pb0.05, ⁎⁎pb0.01, or ⁎⁎⁎pb0.001.
1666 S.E. Eriksson et al. / Free Radical Biology & Medicine 47 (2009) 16611671
8.5 nmol min
1
mg protein
1
72 h posttransfection. Cell extracts from
cells grown without additional selenite in the medium had a cellular
TrxR activity in the range of 90150 nmol min
1
mg protein
1
,
whereas TrxR1 knockdown cells in that case had a total TrxR activity
lowered to around 7.515 nmol min
1
mg protein
1
(data not
shown). Using both of the siRNA constructs in combination (a mixture
of either 5+ 5 nM or 10+ 10 nM Siseq1 and Siseq2) did not result in
additional down-regulation of TrxR1 compared to using each siRNA
separately (data not shown). We wondered whether the low albeit
remaining cellular TrxR activity in the knockdown cells could derive
from unaffected mitochondrial TrxR2 expression, which would also
give a
75
Se-labeled protein band of about 55 kDa. Thus, to determine
the identity of the remaining TrxR activity upon knockdown, 300 μgof
total cell extracts from either control or TrxR1-knockdown cells was
Fig. 5 (continued).
1667S.E. Eriksson et al. / Free Radical Biology & Medicine 47 (2009) 16611671
subjected to anion-exchange chromatography followed by MS analysis
(Figs. 1B, 1C, and 1D). This analysis identied a remaining presence of
low levels of TrxR1 in the knockdown cells (results shown in
Supplementary Fig. S2), whereas no TrxR2 could be detected in the
analyzed fractions under the conditions used. Thus, the low remaining
TrxR activity and
75
Se-labeled 55-kDa band in the siRNA-treated cells
(Fig. 1A) mainly represent a low remaining TrxR1 activity, although
the additional presence of low amounts of TrxR2 cannot be excluded.
Next we found that knocking down the cellular TrxR1 activity to about
10% of controls (Fig. 2A) gave no apparent effects on total cellular Trx
activity (Fig. 2B) or GR activity (Fig. 2C) as determined in crude cell
extracts. The cellular Trx activity, measured under saturating condi-
tions and thus representing the maximal Trx capacity, was in the
range of 39.1 ± 3.3 nmol min
1
mg total protein
1
.Notably,the
control A549 cells would thereby have a capacity for about sevenfold
higher TrxR activity than their maximal Trx activity (cf. control cells in
Figs. 2A and 2B). This analysis also showed that in the TrxR1-
knockdown cells, the cellular TrxR1 activity was lowered to the same
range as the total cellular Trx activity found in these cells.
Cell growth, cell cycle phase, and viability assessments
With the TrxR1 activity lowered to about 10% of controls, the
A549 cells showed no signicant effects on cell growth compared to
mock, either with or without the addition of 25 nM selenite to the
growth medium (data not shown). However, the transfection
procedures as such caused a general decrease in cell proliferation
of approx 30%. Analysis of cell cycle phases also showed no
signicant changes upon TrxR1 knockdown (Fig. 3). Furthermore,
the fraction of cells with a subdiploid DNA content (sub-G1),
identifying dead cells, was less than 4% in both mock and TrxR1-
knockdown cells (Fig. 3). Surprisingly, we found that even TrxR1-
knockdown cells subjected to GSH depletion for 48 h with BSO
showed no alterations in viability or cell growth compared to mock
controls (Fig. 4A), although it gave about 98% reduction in
intracellular glutathione levels (Fig. 4B). The results suggest a
tremendous reserve capacity for ribonucleotide reduction and
basal support of Trx-and GSH-dependent reduction pathways in
A549 cells. The mean concentration of total glutathione measured in
untreated control cells (165 ± 21 nmol per milligram of total
protein) was not signicantly different from that in Siseq1-treated
cells (176 ± 9 nmol per milligram of total protein in). After BSO
treatment, the mean values of total glutathione were typically in the
range of 2.5 to 5 nmol per milligram of total protein (Fig. 4B).
Considering the lack of an apparent phenotype upon knocking
down TrxR1 in A549 cells, we next wished to study the potential
effects on the sensitivity of these cells to drugs, using a selection of
compounds known to target this enzyme.
Effects of TrxR1 knockdown on cellular drug sensitivity
For analyses of drug sensitivity, we treated the cells with DNCB,
which may irreversibly derivatize the active site of TrxR and
concomitantly induce an NADPH-oxidase activity in the inhibited
enzyme [38]. Furthermore, we treated cells with menadione, a
nondisulde substrate of TrxR [32] that may redox cycle with TrxR,
leading to the production of pro-oxidant species, as similarly shown
with juglone [39], which, however, furthermore inhibits the enzyme
[39] and was also included in our analysis. We furthermore wished to
study the cytotoxic effects of auranon in these cells as well as those
of the two platinum-based anticancer drugs cDDP and Oxa, which
have also been shown to target the enzyme [4547]. We found that
DNCB and auranon were obvious potent inhibitors of total cellular
TrxR activity, whereas juglone, cDDP, and Oxa had intermediary
effects and menadione did not inhibit the cellular activity of the
enzyme under the conditions used here and as analyzed in crude cell
extracts (Table 1). Analyzing the potential cell death induced by these
compounds revealed that TrxR1 knockdown signicantly sensitized
A549 cells to treatment with DNCB (in the concentration range of 2.5
10 μM) or menadione (510 μM), whereas there was no difference in
sensitivity to auranon upon TrxR1 knockdown (Fig. 5A). In contrast,
the cells surprisingly became less susceptible to the cytotoxicity of
cDDP (usingb20 μM) upon TrxR1 knockdown, whereas there was no
change in the sensitivity to Oxa or juglone (Fig. 5B).
Discussion
In this study we found that knocking down the endogenous
overexpression of TrxR in a cancer cell line using siRNA leads to
diverse, complex, and unexpected cellular phenotypes. The basal
nonstressed growth of A549 cells was unaffected by an efcient
(approx 90%) TrxR1 knockdown, even in combination with GSH
depletion, illustrating that these cells have a tremendous reserve
capacity in vital redox systems. We also found that the very high
TrxR1 activity in A549 cells exceeded by severalfold the total Trx-
reducing capacity, suggesting that TrxR1 might carry out additional
functions beyond Trx reduction in these cells. Previous studies have
shown that TrxR1 can be highly overexpressed in several forms of
cancer and cancer cell lines [60] (www.proteinatlas.org) and many
studies have indicated that TrxR1 may be a potential target for
anticancer therapy and cancer prevention [2022,61]. The Trx system
is clearly involved in many important cellular functions, e.g., cell
proliferation, through the disulde reduction of ribonucleotide
reductase needed for deoxyribonucleotide synthesis [4], and antiox-
idant defense, supporting the activity of, e.g., methionine sulfoxide
reductase and peroxiredoxins [1]. The Trx system also has direct
antiapoptotic functions, e.g., with reduced Trx1 inhibiting apoptosis
signal-regulating kinase 1 [62] and with TrxR1 playing a potential role
in normal p53 maturation [63]. In the latter study it was shown that
modifying TrxR1 with electrophilic lipids resulted in changes in p53
conformation, whereas lowering the TrxR1 levels in the cells before
treatment with the electrophilic agents protected the p53 conforma-
tion [63]. It was also found that certain forms of TrxR1, such as the
enzyme derivatized by cDDP, could gain cell death-inducing proper-
ties in the form of SecTRAPs [48,49]. Based on these prior ndings and
the many important endogenous functions of TrxR1 we were at rst
surprised that we found only mild effects on the basal cell growth of
A549 cells upon TrxR1 knockdown. Because ribonucleotide reductase
function can be supported by either the Trx or the glutaredoxin (Grx)
system, the latter being dependent upon GSH [64], we initially
reasoned that the Grx system supported the DNA precursor synthesis
upon TrxR1 knockdown. However, depleting the GSH levels by BSO in
the TrxR1-knockdown cells did not cause any signicant additional
impact on the cell growth. The answer is likely to be found in the total
cellular capacity of the Trx system in these cells. Knocking down
TrxR1 gave about 90% lower activity compared to the controls but,
importantly, the remaining 31.4 ± 8.5 nmol min
1
mg protein
1
of
cellular TrxR1 activity was in the same range as the total Trx capacity
and could thus be enough to sustain the normal ribonucleotide
synthesis and other Trx-dependent enzyme systems. Supposing a
duration of 8 h for S phase in these cells, 12.5 × 10
6
ribonucleotides
would have to be reduced per minute for the synthesis of a complete
genome of 6×10
9
deoxyribonucleotides. Assuming that a typical
A549 cell contains about 1 ng of total protein [60,65], the hereby
determined Trx activity of 39.1 nmol min
1
mg protein
1
could
support about 23.5×10
9
reactions per minute and per cell, i.e.,
signicantly more than theoretically needed for Trx support of DNA
synthesis. This can explain the lack of effects from TrxR knockdown in
these cells and may explain an earlier study showing that down-
regulation of TrxR in cancer cells from mouse resulted in no changes
in deoxyribonucleotide pools [24]. With this in mind, it should be
emphasized that unless compartmentalization effects or other factors
1668 S.E. Eriksson et al. / Free Radical Biology & Medicine 47 (2009) 16611671
exist that increase the intracellular Trx activity beyond that of the
saturating conditions of our enzyme assay, the total TrxR1 capacity in
the A549 cells was about sevenfold higher than their total Trx
capacity. With the remaining 10% TrxR1 activity in the knockdown
cells therefore likely to be enough to sustain most if not all of the Trx1
functions, what other types of functions, if any, would the very high
levels of TrxR1 in A549 cells have? The reactive C-terminal catalytic
selenocysteine residue in TrxR1 clearly makes the enzyme capable of
reducing several other substrates in vitro in addition to Trx and
thereby TrxR could also have additional cellular substrates apart from
Trx [1]. It should also be noted that some splice forms, such as
TXNRD1_v3, lack Trx-reducing activity, which could complicate the
picture even further [59,66,67]. Hateld and co-workers showed in a
mouse model that xenografts with TrxR1-knockdown cells could not
form tumors in contrast to control cells [23]. Such effects could hardly
be due to impaired replicative capacity, because the cells grew in
culture [23], but could potentially be due to knockdown of TrxR1-
dependent cellcell interactions not detectable under normal cell
culture conditions. It should be noted that our siRNA constructs
covering either part of the active site of the enzyme (its CVNVGC
motif) or the 3UTR exon (Supplementary Fig. S1) knocked down at
least three
75
Se-labeled forms of TrxR1 present in A549 cells, even if
the classical 55-kDa form was the most prominent one (Fig. 1). Future
studies with specic siRNA constructs for the various splice forms
would be needed to distinguish between these forms with regard to
any observed phenotypes.
Several previous studies have found a close correlation between
induction of cell death and TrxR1 inhibition by various types of
cytotoxic compounds [20,22,37]. However, our results clearly reveal
that impairment of TrxR1 activity solely by knocking down the
enzyme is not enough to induce cell death in highly overexpressing
A549 cells and, also, that even if TrxR1 obviously affects the extent of
certain drug-induced toxicities, such effects may be due to more
intricate mechanisms than just a lowered TrxR1 activity. The A549
cells with endogenous overexpression of TrxR1 had a markedly
increased sensitivity to the platinum anticancer drug cDDP compared
to TrxR1 knockdown, but they had higher resistance to DNCB and
menadione. The effects of other agents targeting TrxR, including
auranon, considered to be a potent TrxR1-specic inhibitor, were
unrelated to the expression levels of TrxR1. The cytotoxic effects seen
upon the use of different TrxR inhibitors could, naturally, involve
different off-target effects, but potentially they could also involve
different direct effects on TrxR1. Our group has previously shown that
forms of TrxR1 derivatized with cDDP, or being truncated at the
position of the Sec residue, are completely devoid of Trx reduction
capacity but can show a gain of function in the form of pro-oxidant
cell-killing SecTRAPs [48,49]. Such mechanisms would agree with our
ndings herein, showing that A549 cells with their endogenously up-
regulated levels of TrxR1 have a signicantly increased sensitivity to
cDDP compared to TrxR1-knockdown cells. It is thus possible that in
A549 cells, where the TrxR1 capacity was found to exceed by far the
activity needed for Trx reduction, the selenoprotein may easily reside
in its reduced selenolate-exposing form [68] (because of less need for
reduction of substrates such as Trx) and thereby could more easily
form lethal SecTRAPs upon exposure to cDDP. If this is indeed the case,
it would suggest that cells having a TrxR1 activity higher than that
needed for Trx reduction would be particularly sensitive to cDDP
because they could easily form SecTRAPs. This concept is supported by
our ndings but should be further addressed in future studies. In
contrast to the ndings with cDDP, knockdown of TrxR1 in A549 cells
made these cells more sensitive to DNCB and menadione, two
compounds known to act as stressors capable of inducing ROS
production [1,69]. In these cases it seemed like the antioxidant
functions of highly overexpressed TrxR1 could be important for cell
protection against these compounds. DNCB is a very good inhibitor of
TrxR1 [38], whereas menadione is not an inhibitor but rather a pure
substrate [32]. Exposure to menadione has also been shown to result
in up-regulation of TrxR1 in a hepatic cancer cell line [69], probably
owing to an induced oxidative stress with a nuclear factor erythroid-
2-related factor 2-dependent response [70]. It should be noted that
although DNCB is a clear inhibitor of TrxR1 [38], as also shown in our
study (Table 1), it is furthermore a model substrate of glutathione S-
transferases and lowers GSH levels in cells [71]. However, our results
from the combination of TrxR1 knockdown with GSH depletion by
BSO treatment suggested that this combination as such is not
necessarily lethal to A549 cells. Thus, our results clearly demonstrate
that although TrxR1 targeting is important and promising as an
anticancer chemotherapeutic principle, the actual cytotoxic prole
and TrxR1-related mechanisms underlying drug efcacy need to be
highly drug-specic. In the protective effects of TrxR1 against DNCB-
or menadione-induced cell death, it should again be emphasized that
this protection is not necessarily linked to Trx-dependent pathways,
because TrxR1 activity also in the knockdown cells should be
sufcient to sustain the overall Trx capacity in these cells. It is in
this case possible that reduction of low-molecular-weight antiox-
idants or peroxides could explain the protective effects of the very
high TrxR1 levels [1]. To conclude, this study shows that high
overexpression of TrxR1 in cancer cells increases the cytotoxic efcacy
of drugs such as cisplatin, whereas it protects the cells from other
drugs such as DNCB or menadione. It seems clear that TrxR1 is an
interesting target for anticancer chemotherapy, but the molecular
mechanisms underlying such therapy may be more complex than
hitherto fully understood.
Acknowledgments
Determination of selenium content in the fetal calf serum was
kindly performed by Marie Vahter and collaborators, Karolinska
Institutet. Oxaliplatin used in the drug sensitivity assays was a kind
gift from Maria Shoshan, and human thioredoxin was kindly provided
by Arne Holmgren, both at Karolinska Institutet. The kind help of Qing
Cheng and Olle Rengby in the production of TrxR1 is also
acknowledged. This study was supported by funding from the
Swedish Research Council (Medicine), the Swedish Cancer Society,
Hedlunds Stiftelse, Knut and Alice Wallenbergs Stiftelse, and
Karolinska Institutet.
Appendix A. Supplementary data
Supplementary data associated with this article can be found, in
the online version, at doi:10.1016/j.freeradbiomed.2009.09.016.
References
[1] Nordberg, J.; Arnér, E. S. J. Reactive oxygen species, antioxidants, and the
mammalian thioredoxin system. Free Radic. Biol. Med. 31:12871312; 2001.
[2] Arner, E. S.; Holmgren, A. Physiological functions of thioredoxin and thioredoxin
reductase. Eur. J. Biochem. 267:61026109; 2000.
[3] Berndt, C.; Lillig, C. H.; Holmgren, A. Thiol-based mechanisms of the thioredoxin
and glutaredoxin systems: implications for diseases in the cardiovascular system.
Am. J. Physiol. Heart Circ. Physiol. 292:H12271236; 2007.
[4] Nordlund, P.; Reichard, P. Ribonucleotide reductases. Annu. Rev. Biochem. 75:
681706; 2006.
[5] Luthman, M.; Holmgren, A. Rat liver thioredoxin and thioredoxin reductase:
purication and characterization. Biochemistry 21:66286633; 1982.
[6] Rundlöf, A. K.; Arnér, E. S. J. Regulation of the mammalian selenoprotein
thioredoxin reductase 1 in relation to cellular phenotype, growth, and signaling
events. Antioxid. Redox Signal. 6:4152; 2004.
[7] Rigobello, M. P.; Callegaro, M. T.; Barzon, E.; Benetti, M.; Bindoli, A. Purication of
mitochondrial thioredoxin reductase and its involvement in the redox regulation
of membrane permeability. Free Radic. Biol. Med. 24:370376; 1998.
[8] Sun, Q. A.; Su, D.; Novoselov, S. V.; Carlson, B. A.; Hateld, D. L.; Gladyshev, V. N.
Reaction mechanism and regulation of mammalian thioredoxin/glutathione
reductase. Biochemistry 44:1452814537; 2005.
[9] 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. Cytoplasmic thioredoxin reductase is essential for embryogenesis
but dispensable for cardiac development. Mol. Cell. Biol. 25:19801988; 2005.
1669S.E. Eriksson et al. / Free Radical Biology & Medicine 47 (2009) 16611671
[10] 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. Essential role for mitochondrial thioredoxin
reductase in hematopoiesis, heart development, and heart function. Mol. Cell. Biol.
24:94149423; 2004.
[11] Nonn, L.; Williams, R. R.; Erickson, R. P.; Powis, G. The absence of mitochondrial
thioredoxin 2 causes massive apoptosis, exencephaly, and early embryonic
lethality in homozygous mice. Mol. Cell. Biol. 23:916922; 2003.
[12] Matsui, M.; Oshima, M.; Oshima, H.; Takaku, K.; Maruyama, T.; Yodoi, J.; Taketo,
M. M. Early embryonic lethality caused by targeted disruption of the mouse
thioredoxin gene. Dev. Biol. 178:179185; 1996.
[13] Zhong, L.; Holmgren, A. Essential role of selenium in the catalytic activities of
mammalian thioredoxin reductase revealed by characterization of recombinant
enzymes with selenocysteine mutations. J. Biol. Chem. 275:1812118128; 2000.
[14] Gromer, S.; Eubel, J. K.; Lee, B. L.; Jacob, J. Human selenoproteins at a glance. Cell.
Mol. Life Sci. 62:24142437; 2005.
[15] Berggren, M. M.; Mangin, J. F.; Gasdaka, J. R.; Powis, G. Effect of selenium on rat
thioredoxin reductase activity: increase by supranutritional selenium and
decrease by selenium deciency. Biochem. Pharmacol. 57:187193; 1999.
[16] Nalvarte, I.; Damdimopoulos, A. E.; Nystom, C.; Nordman, T.; Miranda-Vizuete, A.;
Olsson, J. M.; Eriksson, L.; Bjornstedt, M.; Arner, E. S. J.; Spyrou, G. Overexpression
of enzymatically active human cytosolic and mitochondrial thioredoxin reductase
in HEK-293 cellseffect on cell growth and differentiation. J. Biol. Chem. 279:
5451054517; 2004.
[17] Hill, K. E.; McCollum, G. W.; Boeglin, M. E.; Burk, R. F. Thioredoxin reductase
activity is decreased by selenium deciency. Biochem. Biophys. Res. Commun. 234:
293295; 1997.
[18] Crosley, L. K.; Meplan, C.; Nicol, F.; Rundlof, A. K.; Arner, E. S. J.; Hesketh, J. E.;
Arthur, J. R. Differential regulation of expression of cytosolic and mitochondrial
thioredoxin reductase in rat liver and kidney. Arch. Bio chem. Bioph ys. 459:178188;
2007.
[19] Papp, L. V.; Lu, J.; Holmgren, A.; Khanna, K. K. From selenium to selenoproteins:
synthesis, identity, and their role in human health. Antioxid. Redox Signal. 9:
775806; 2007.
[20] Urig, S.; Becker, K. On the potential of thioredoxin reductase inhibitors for cancer
therapy. Semin. Cancer Biol. 16:452465; 2006.
[21] Gromer, S.; Urig, S.; Becker, K. The thioredoxin systemfrom science to clinic.
Med. Res. Rev. 24:4089; 2004.
[22] Arner, E. S. J.; Holmgren, A. The thioredoxin system in cancer. Semin. Cancer Biol.
16:420426; 2006.
[23] Yoo, M. H.; Xu, X. M.; Carlson, B. A.; Gladyshev, V. N.; Hateld, D. L. Thioredoxin
reductase 1 deciency reverses tumor phenotype and tumorigenicity of lung
carcinoma cells. J. Biol. Chem 281:1300513008; 2006.
[24] Yoo, M. H.; Xu, X. M.; Carlson, B. A.; Patterson, A. D.; Gladyshev, V. N.; Hateld, D. L.
Targeting thioredoxin reductase 1 reduction in cancer cells inhibits self-sufcient
growth and DNA replication. PLoS ONE e1112:2; 2007.
[25] Yoo, M. H.; Xu, X. M.; Turanov, A. A.; Carlson, B. A.; Gladyshev, V. N.; Hateld, D. L.
A new strategy for assessing selenoprotein function: siRNA knockdown/knock-in
targeting the 3-UTR. RNA 13:921929; 2007.
[26] Nishimoto, M.; Sakaue, M.; Hara, S. Short-interfering RNA-mediated silencing of
thioredoxin reductase 1 alters the sensitivity of HeLa cells toward cadmium.
Biol. Pharm. Bull. 29:543546; 2006.
[27] Zhong, L.; Arnér, E. S. J.; Holmgren, A. Structure and mechanism of mammalian
thioredoxin reductase: the active site is a redox-active selenolthiol/selenenyl-
sulde formed from the conserved cysteineselenocysteine sequence. Proc. Natl.
Acad. Sci. USA 97:58545859; 2000.
[28] Cheng, Q.; Sandalova, T.; Lindqvist, Y.; Arner, E. S. Crystal structure and catalysis of
the selenoprotein thioredoxin reductase 1. J. Biol. Chem. 284:39984008; 2009.
[29] Cenas, N.; Nivinskas, H.; Anusevicius, Z.; Sarlauskas, J.; Lederer, F.; Arnér, E. S. J.
Interactions of quinones with thioredoxin reductase: a challenge to the
antioxidant role of the mammalian selenoprotein. J. Biol. Chem. 279:25832592;
2004.
[30] May, J. M.; Mendiratta, S.; Hill, K. E.; Burk, R. F. Reduction of dehydroascorbate
to ascorbate by the selenoenzyme thioredoxin reductase. J. Biol. Chem. 272:
2260722610; 1997.
[31] Arnér, E. S. J.; Nordberg, J.; Holmgren, A. Efcient reduction of lipoamide and lipoic
acid by mammalian thioredoxin reductase. Biochem. Biophys. Res. Commun. 225:
268274; 1996.
[32] Holmgren, A. Thioredoxin. Annu. Rev. Biochem. 54:237271; 1985.
[33] Björnstedt, M.; Kumar, S.; Björkhem, L.; Spyrou,G.; Holmgren, A. Selenium and the
thioredoxin and glutaredoxin systems. Biomed. Environ. Sci. 10:271279; 1997.
[34] Lothrop, A. P.; Ruggles, E. L.; Hondal, R. J. No selenium required: reactions
catalyzed by mammalian thioredoxin reductase that are independent of a
selenocysteine residue. Biochemistry 48:62136223; 2009.
[35] Omata, Y.; Folan, M.; Shaw, M.; Messer, R. L.; Lockwood, P. E.; Hobbs, D.;
Bouillaguet, S.; Sano, H.; Lewis, J. B.; Wataha, J. C. Sublethal concentrations of
diverse gold compounds inhibit mammalian cytosolic thioredoxin reductase
(TrxR1). Toxicol. In Vitro 20:882890; 2006.
[36] Gromer, S.; Wissing, J.; Behne, D.; Ashman, K.; Schirmer, R. H.; Flohe, L.; Becker,
K. A. hypothesis on the catalytic mechanism of the selenoenzyme thioredoxin
reductase. Biochem. J 332 (Pt 2):591592; 1998.
[37] Talbot, S.; Nelson, R.; Self, W. T. Arsenic trioxide and auranon inhibit
selenoprotein synthesis: implications for chemotherapy for acute promyelocytic
leukaemia. Br. J. Pharmacol. 154:940948; 2008.
[38] Arnér, E. S. J.; Bjornstedt, M.; Holmgren, A. 1-Chloro-2,4-dinitrobenzene is an
irreversible inhibitor of human thioredoxin reductase: loss of thioredoxin
disulde reductase activity is accompanied by a large increase in NADPH oxidase
activity. J. Biol. Chem. 270:34793482; 1995.
[39] Cenas, N.; Prast, S.; Nivinskas, H.; Sarlauskas, J.; Arner, E. S. Interactions of
nitroaromatic compounds with the mammalian selenoprotein thioredoxin reduc-
tase and the relation to induction of apoptosis in human cancer cells. J. Biol. Chem.
281:55935603; 2006.
[40] Brown, K. K.; Eriksson, S. E.; Arner, E. S. J.; Hampton, M. B. Mitochondrial
peroxiredoxin 3 is rapidly oxidized in cells treated with isothiocyanates. Free
Radic. Biol. Med. 45:494502; 2008.
[41] Carvalho, C. M.; Chew, E. H.; Hashemy, S. I.; Lu, J.; Holmgren, A. Inhibition of the
human thioredoxin system: a molecular mechanism of mercury toxicity. J. Biol.
Chem. 283:1191311923; 2008.
[42] Wallenborg, K.; Vlachos, P.; Eriksson, S.; Huijbregts, L.; Arner, E. S.; Joseph, B.;
Hermanson, O. Red wine triggers cell death and thioredoxin reductase inhibition:
effects beyond resveratrol and SIRT1. Exp. Cell Res. 315:13601371; 2009.
[43] Lu, J.; Papp, L. V.; Fang, J.; Rodriguez-Nieto, S.; Zhivotovsky, B.; Holmgren, A.
Inhibition of mammalian thioredoxin reductase by some avonoids: implications
for myricetin and quercetin anticancer activity. Cancer Res. 66:44104418; 2006.
[44] Lu, J.; Chew, E. H.; Holmgren, A. Targeting thioredoxin reductase is a basis for
cancer therapy by arsenic trioxide. Proc. Natl. Acad. Sci. USA 104:1228812293;
2007.
[45] Hellberg, V.; Wallin, I.; Eriksson, S.; Hernlund, E.; Jerremalm, E.; Berndtsson, M.;
Eksborg, S.; Arner, E. S.; Shoshan, M.; Ehrsson, H.; Laurell, G. Cisplatin and
oxaliplatin toxicity: importance of cochlear kinetics as a determinant for
ototoxicity. J. Natl. Cancer Inst. 101:3747; 2009.
[46] Witte, A. B.; Anestal, K.; Jerremalm, E.; Ehrsson, H.; Arner, E. S. J. Inhibition of
thioredoxin reductase but not of glutathione reductase by the major classes of
alkylating and platinum-containing anticancer compounds. Free Radic. Biol. Med.
39:696703; 2005.
[47] Arnér, E. S. J.; Nakamura, H.; Sasada, T.; Yodoi, J.; Holmgren, A.; Spyrou, G. Analysis
of the inhibition of mammalian thioredoxin, thioredoxin reductase, and
glutaredoxin by cis-diamminedichloroplatinum (II) and its major metabolite,
the glutathioneplatinum complex. Free Radic. Biol. Med. 31:11701178; 2001.
[48] Anestal, K.; Arner, E. S. Rapid induction of cell death by selenium-compromised
thioredoxin reductase 1 but not by the fully active enzyme containing
selenocysteine. J. Biol. Chem. 278:1596615972; 2003.
[49] Anestal, K.; Prast-Nielsen, S.; Cenas, N.; Arner, E. S. Cell death by SecTRAPs:
thioredoxin reductase as a prooxidant killer of cells. PLoS ONE e1846:3; 2008.
[50] Rengby, O.; Johansson, L.; Carlson, L. A.; Serini, E.; Vlamis-Gardikas, A.; Karsnas, P.;
Arnér, E. S. J. Assessment of production conditions for efcient use of Escherichia
coli in high-yield heterologous recombinant selenoprotein synthesis. Appl.
Environ. Microbiol. 70:51595167; 2004.
[51] Ren, X.; Björnstedt, M.; Shen, B.; Ericson, M. L.; Holmgren, A. Mutagenesis of
structural half-cystine residues in human thioredoxin and effects on the
regulation of activity by selenodiglutathione. Biochemistry 32:97019708; 1993.
[52] Zhong, L.; Arnér, E. S. J.; Ljung, J.; Åslund, F.; Holmgren, A. Rat and calf thioredoxin
reductase are homologous to glutathione reductase with a carboxyl-terminal
elongation containing a conserved catalytically active penultimate selenocysteine
residue. J. Biol. Chem. 273:85818591; 1998.
[53] Rengby, O.; Cheng, Q.; Vahter, M.; Jornvall, H.; Arner, E. S. Highly active dimeric
and low-activity tetrameric forms of selenium-containing rat thioredoxin
reductase 1. Free Radic. Biol. Med. 46:893904; 2009.
[54] Arnér, E. S. J.; Zhong, L.; Holmgren, A. Preparation and assay of mammalian
thioredoxin and thioredoxin reductase. Methods Enzymol. 300:226239; 1999.
[55] Carlberg, I.; Mannervik, B. Glutathione reductase. Methods Enzymol. 113:484490;
1985.
[56] Tietze, F. Enzymic method for quantitative determination of nanogram amounts
of total and oxidized glutathione: applications to mammalian blood and other
tissues. Anal. Biochem. 27:502522; 1969.
[57] Vandeputte, C.; Guizon, I.; Genestie-Denis, I.; Vannier, B.; Lorenzon, G. A
microtiter plate assay for total glutathione and glutathione disulde contents in
cultured/isolated cells: performance study of a new miniaturized protocol. Cell
Biol. Toxicol. 10:415421; 1994.
[58] Flaberg, E.; Sabelstrom, P.; Strandh, C.; Szekely, L. Extended eld laser confocal
microscopy (EFLCM): combining automated gigapixel image capture with in silico
virtual microscopy. BMC Med. Imaging 8:13; 2008.
[59] Rundlöf, A. K.; Janard, M.; Miranda-Vizuete, A.; Arnér, E. S. J. Evidence for
intriguingly complex transcription of human thioredoxin reductase 1. Free Radic.
Biol. Med. 36:641656; 2004.
[60] Berggren, M.; Gallegos, A.; Gasdaska, J. R.; Gasdaska, P. Y.; Warneke, J.; Powis, G.
Thioredoxin and thioredoxin reductase gene expression in human tumors and cell
lines, and the effects of serum stimulation and hypoxia. Anticancer Res. 16:
34593466; 1996.
[61] Yoo, M. H.; Xu, X. M.; Carlson, B. A.; Gladyshev, V. N.; Hateld, D. L. Thioredoxin
reductase 1 deciency reverses tumor phenotype and tumorigenicity of lung
carcinoma cells. J. Biol. Chem. 281:1300513008; 2006.
[62] Saitoh, M.; Nishitoh, H.; Fujii, M.; Takeda, K.; Tobiume, K.; Sawada, Y.; Kawabata,
M.; Miyazono, K.; Ichijo, H. Mammalian thioredoxin is a direct inhibitor of
apoptosis signal-regulating kinase (ASK) 1. EMBO J. 17:25962606; 1998.
[63] Cassidy, P. B.; Edes, K.; Nelson, C. C.; Parsawar, K.; Fitzpatrick, F. A.; Moos, P. J.
Thioredoxin reductase is required for the inactivation of tumor suppressor p53
and for apoptosis induced by endogenous electrophiles. Carcinogenesis 27:
25382549; 2006.
[64] Avval, F. Z.; Holmgren, A. Molecular mechanisms of thioredoxin and glutaredoxin
as hydrogen donors for mammalian S phase ribonucleotide reductase. J. Biol.
Chem. 284:82338240; 2009.
1670 S.E. Eriksson et al. / Free Radical Biology & Medicine 47 (2009) 16611671
[65] Fujiwara, N.; Fujii, T.; Fujii, J.; Taniguchi, N. Functional expression of rat
thioredoxin reductase: selenocysteine insertion sequence element is essential
for the active enzyme. Biochem. J. 340 (Pt. 2):439444; 1999.
[66] Dammeyer, P.; Damdimopoulos, A. E.; Nordman, T.; Jimenez, A.; Miranda-Vizuete,
A.; Arner, E. S. Induction of cell membrane protrusions by the N-terminal
glutaredoxin domain of a rare splice variant of human thioredoxin reductase 1.
J. Biol. Chem. 283:28142821; 2008.
[67] Su, D.; Gladyshev, V. N. Alternative splicing involving the thioredoxin reduc-
tase module in mammals: a glutaredoxin-containing thioredoxin reductase 1.
Biochemistry 43:1217712188; 2004.
[68] Sandalova, T.; Zhong, L.; Lindqvist, Y.; Holmgren, A.; Schneider, G. Three-
dimensional structure of a mammalian thioredoxin reductase: implications for
mechanism and evolution of a selenocysteine-dependent enzyme. Proc. Natl.
Acad. Sci. USA 98:95339538; 2001.
[69] Jung, H. I.; Lim, H. W.; Kim, B. C.; Park, E. H.; Lim, C. J. Differential thioredoxin
reductase activity from human normal hepatic and hepatoma cell lines. Yonsei
Med. J. 45:263272; 2004.
[70] Singh, A.; Boldin-Adamsky, S.; Thimmulappa, R. K.; Rath, S. K.; Ashush, H.; Coulter,
J.; Blackford, A.; Goodman, S. N.; Bunz, F.; Watson, W. H.; Gabrielson, E.; Feinstein,
E.; Biswal, S. RNAi-mediated silencing of nuclear factor erythroid-2-related factor
2 gene expression in non-small cell lung cancer inhibits tumor growth and
increases efcacy of chemotherapy. Cancer Res. 68:79757984; 2008.
[71] Armstrong, R. N. Structure, catalytic mechanism, and evolution of the glutathione
transferases. Chem. Res. Toxicol. 10:218; 1997.
1671S.E. Eriksson et al. / Free Radical Biology & Medicine 47 (2009) 16611671
... [15,46,47] Therefore, we exemplary evaluated the inhibition of 1 and 2 and compared them to one of the most potent gold-based TrxR inhibitors, auranofin, in an end-point insulin reduction-based and DTNB-coupled microplate reader assay in cellulo (Table 2). [48] The A549 lung carcinoma cells were selected for these experiments due to their high basal expression of TrxR compared to the other cancer cell lines, [49] and, most notably, for the marked antiproliferative effects of 2 towards these cells. According to the enzyme inhibition data, whereas ligand 1 alone is not able to exhibit TrxR inhibition in the tested range of concentrations, gold complex 2 shows an IC 50 of 0.57 μM, which is only slightly less active (1.9-fold) than the reference auranofin (IC 50 0.32 μM) (SI, Figure S6) but more active than other gold complexes that we had recently studied with the same assay. ...
Article
Full-text available
A dinuclear gold(I) complex featuring a strongly donating bis‐N‐heterocyclic imine ligand was synthesised and characterised by different methods, including single crystal X‐ray diffraction (SC‐XRD) analysis. The compound has been tested for its antiproliferative effects in a panel of human cancer cell lines in vitro, showing highly selective anticancer effects, particularly against human A549 non‐small cell lung cancer cells (NSCLC), with respect to non‐tumorigenic cells (VERO). The accumulation of the compound in A549 and VERO cells was studied by high‐resolution continuum source atomic absorption spectrometry (HRCS‐AAS), revealing that the anticancer effects are not particularly related to the different amounts of gold taken up by the cells over 72 h. Enzyme inhibition studies to evaluate the activity of the seleno‐enzyme thioredoxin reductase (TrxR) in cancer cell extracts show that the gold(I) compound is a potent inhibitor (IC50=0.567±0.208 μM), while the free ligand is ineffective. This result correlates with the observed compound's selectivity towards A549 cells overexpressing the enzyme.
... Then, it is important to pay attention to the selenium concentration in the cell culture medium when looking at the activity of selenoproteins (such as TrxR or glutathione peroxidase, GPx) [57]. Few studies have been carried out on the correlation between the quantity of Se and their activity, and in turn, on drugs' cytotoxicity [58][59][60]. As an example, in 2012, Arnér and collaborators brought to light the subtle effect of selenium concentration on the antiproliferative activity of cisplatin. ...
Article
Full-text available
Simple Summary The identification of biological targets is an essential step in deciphering the mechanism of action of anticancer drugs. In this review, we chose to study the relationship between the inhibition of thioredoxin reductase (TrxR), a key enzyme in maintaining the redox balance of cells, and the cytotoxic effects of two groups of organometallic complexes. The first group is essentially composed of Au(I) and Au(III) complexes and the second one comprises metallocifens (organometallic complexes derived from tamoxifen). The results show that these two groups interact differently with TrxR at the molecular level. Even if the contribution of TrxR inhibition to the cytotoxicity of complexes is clearly established for many of them, the number of complexes for which TrxR inhibition plays a predominant role appears quite limited. Eventually, the antiproliferative activity of most of the complexes appears to stem from the interaction with several targets, a favorable strategy to tackle MDR tumors. Abstract Cancers classified as multidrug-resistant (MDR) are a family of diseases with poor prognosis despite access to increasingly sophisticated treatments. Several mechanisms explain these resistances involving both tumor cells and their microenvironment. It is now recognized that a multi-targeting approach offers a promising strategy to treat these MDR tumors. Inhibition of thioredoxin reductase (TrxR), a key enzyme in maintaining redox balance in cells, is a well-identified target for this approach. Auranofin was the first inorganic gold complex to be described as a powerful inhibitor of TrxR. In this review, we will first recall the main results obtained with this metallodrug. Then, we will focus on organometallic complexes reported as TrxR inhibitors. These include gold(I), gold(III) complexes and metallocifens, i.e., organometallic complexes of Fe and Os derived from tamoxifen. In these families of complexes, similarities and differences in the molecular mechanisms of TrxR inhibition will be highlighted. Finally, the possible relationship between TrxR inhibition and cytotoxicity will be discussed and put into perspective with their mode of action.
... This excess of TrxR could be necessary to perform other functions, or in case of exposure to inhibition factors. When TrxR was inhibited by 90% in HeLa (human cell line) cells, Trx1 remained in a reduced state (Eriksson et al. 2009). Small interfering RNA knocked down TrxR1 by 90% in A549 (human epithelial cell) cells but did not seem to affect Trx activity or inhibit cell growth (Watson et al. 2008). ...
Article
Full-text available
Mercury (Hg) exposure has not been examined in many recreational nearshore fish species that are commonly consumed around the Hawaiian Islands. Specific gene transcripts, such as metallothionein (MET) and thioredoxin reductase (TrxR), can be used to examine Hg exposure responses in aquatic organisms. This study measured total mercury (THg) in four species from two groups of Hawaiian nearshore fishes: giant trevally (Caranx ignobilis, n = 13), bluefin trevally (C. melampygus, n = 4), sharp jaw bonefish (Albula virgata, n = 2), and round jaw bonefish (A. glossodonta, n = 19). Total Hg accumulation and abundance profiles of MET and TrxR were evaluated for muscle, liver, and kidney tissues. Total Hg in round jaw bonefish and giant trevally tissues accumulated with length and calculated age. In round jaw bonefish tissues, mean THg was greater in kidney (1156 ng/g wet mass (wm)) than liver (339 ng/g wm) and muscle (330 ng/g wm). Giant trevally muscle (187 ng/g wm) and liver (277 ng/g wm) mean THg did not differ significantly. Fish species in this study were compared to commercial and local fish species with state and federal muscle tissue consumption advisories based on THg benchmarks developed by the U.S. Food and Drug Administration (FDA) and Environmental Protection Agency (EPA). Both bonefishes had mean muscle THg that exceeded benchmarks suggesting consumption advisories should be considered. MET transcript in round jaw bonefish kidney tissue and kidney THg exhibited a marginally significant positive correlation, while TrxR transcript in liver tissue negatively correlated with increasing liver THg. These results contribute to our understanding of Hg exposure associated health effects in fish.
... Having found that RSL3 inhibits TXNRD1 with similar efficacy as the previously developed TRi-1 and TRi-2 compounds [17], we next compared these three compounds with regards to their efficacies in eliciting cell death, asking whether all three compounds would display similar ferroptotic features in the cell death that they trigger. First, we compared the cytotoxic effects of the three compounds in lung adenocarcinoma A549 cells, which have an unusually high activity and expression of TXNRD1 that is also known to be able to modulate the cytotoxic efficacy of different anticancer drugs [34,35]. Additional expression information for the cell lines used can be found in Extended Data Table 1. ...
Article
Full-text available
Ferroptosis is defined as cell death triggered by iron-dependent lipid peroxidation that is preventable by antioxidant compounds such as ferrostatin-1. Endogenous suppressors of ferroptosis include FSP-1 and the selenoprotein GPX4, the latter of which directly enzymatically reduces lipid hydroperoxides. Small molecules that trigger ferroptosis include RSL3, ML162, and ML210; these compounds are often used in studies of ferroptosis and are generally considered as GPX4 inhibitors. Here, we found that RSL3 and ML162 completely lack capacity of inhibiting the enzymatic activity of recombinant selenoprotein GPX4. Surprisingly, these compounds were instead found to be efficient inhibitors of another selenoprotein, TXNRD1. Other known inhibitors of TXNRD1, including auranofin, TRi-1 and TRi-2, are also efficient inducers of cell death but that cell death could not be suppressed with ferrostatin-1. Our results collectively suggest that prior studies using RSL3 and ML162 may need to be reevaluated in the context of ferroptosis with regards to additional enzyme targets and mechanisms of action that may be involved.
Article
Full-text available
Reactive oxygen species (ROS) constitute a spectrum of oxygenic metabolites crucial in modulating pathological organism functions. Disruptions in ROS equilibrium span various diseases, and current insights suggest a dual role for ROS in tumorigenesis and the immune response within cancer. This review rigorously examines ROS production and its role in normal cells, elucidating the subsequent regulatory network in inflammation and cancer. Comprehensive synthesis details the documented impacts of ROS on diverse immune cells. Exploring the intricate relationship between ROS and cancer immunity, we highlight its influence on existing immunotherapies, including immune checkpoint blockade, chimeric antigen receptors, and cancer vaccines. Additionally, we underscore the promising prospects of utilizing ROS and targeting ROS modulators as novel immunotherapeutic interventions for cancer. This review discusses the complex interplay between ROS, inflammation, and tumorigenesis, emphasizing the multifaceted functions of ROS in both physiological and pathological conditions. It also underscores the potential implications of ROS in cancer immunotherapy and suggests future research directions, including the development of targeted therapies and precision oncology approaches. In summary, this review emphasizes the significance of understanding ROS‐mediated mechanisms for advancing cancer therapy and developing personalized treatments.
Article
Full-text available
Selenocysteine is a catalytic residue at the active site of all selenoenzymes in bacteria and mammals, and it is incorporated into the polypeptide backbone by a co-translational process that relies on the recoding of a UGA termination codon into a serine/selenocysteine codon. The best-characterized selenoproteins from mammalian species and bacteria are discussed with emphasis on their biological function and catalytic mechanisms. A total of 25 genes coding for selenoproteins have been identified in the genome of mammals. Unlike the selenoenzymes of anaerobic bacteria, most mammalian selenoenzymes work as antioxidants and as redox regulators of cell metabolism and functions. Selenoprotein P contains several selenocysteine residues and serves as a selenocysteine reservoir for other selenoproteins in mammals. Although extensively studied, glutathione peroxidases are incompletely understood in terms of local and time-dependent distribution, and regulatory functions. Selenoenzymes take advantage of the nucleophilic reactivity of the selenolate form of selenocysteine. It is used with peroxides and their by-products such as disulfides and sulfoxides, but also with iodine in iodinated phenolic substrates. This results in the formation of SeX bonds (X = O, S, N, or I) from which a selenenylsulfide intermediate is invariably produced. The initial selenolate group is then recycled by thiol addition. In bacterial glycine reductase and D-proline reductase, an unusual catalytic rupture of selenium-carbon bonds is observed. The exchange of selenium for sulfur in selenoproteins, and information obtained from model reactions, suggest that a generic advantage of selenium compared with sulfur relies on faster kinetics and better reversibility of its oxidation reactions.
Chapter
Fluorescent chemosensors have been widely applied in many diverse fields such as biology, physiology, pharmacology, and environmental sciences. The interdisciplinary nature of chemosensor research has continued to grow over the last 25 years to meet the increasing needs of monitoring our environment and health. More recently, a large range of fluorescent chemosensors have been established for the detection of biologically and/or environmentally important species, and are increasingly being used to solve biological problems. The use of these molecules as imaging probes to diagnose and treat disease is gaining momentum with clear future applications. This book will bring together world-leading experts to describe the current state of play in the field and introduce the cutting-edge research and possible future directions into fluorescent chemosensors design. Chapters focus on the basic principles involved in the design of chemosensors for specific analytes, problems, and challenges in the field. Concentrating on advanced techniques and methods, the book will be of use for academics and researchers across a number of disciplines, with international appeal.
Chapter
Full-text available
Publisher Summary Glutathione reductase is a flavoprotein catalyzing the NADPH-dependent reduction of glutathione disulfide (GSSG) to glutathione (GSH). The reaction is essential for the maintenance of glutathione levels. Glutathione has a major role as a reductant in oxidation–reduction processes, and serves in detoxication and several other cellular functions of great importance. A purification method of this enzyme from calf liver and rat liver is described in this chapter. Similar methods are used for the purification of the enzyme from yeast, porcine, and human erythrocytes. All the steps are carried out at about 5 ° . The purification method from calf liver consists of various steps including preparation of cytosol fraction, chromatography on DEAE-sephadex, precipitation with ammonium sulfate, and chromatography on hydroxyapatite. The purification of glutathione reductase from rat liver is usually combined with the preparation of glutathione transferases, thioltransferase, and glyoxalase I.
Article
Full-text available
Reactive oxygen species (ROS) are known mediators of intracellular signaling cascades. Excessive production of ROS may, however, lead to oxidative stress, loss of cell function, and ultimately apoptosis or necrosis. A balance between oxidant and antioxidant intracellular systems is hence vital for cell function, regulation, and adaptation to diverse growth conditions. Thioredoxin reductase (TrxR) in conjunction with thioredoxin (Trx) is a ubiquitous oxidoreductase system with antioxidant and redox regulatory roles. In mammals, extracellular forms of Trx also have cytokine-like effects. Mammalian TrxR has a highly reactive active site selenocysteine residue resulting in a profound reductive capacity, reducing several substrates in addition to Trx. Due to the reactivity of TrxR, the enzyme is inhibited by many clinically used electrophilic compounds including nitrosoureas, aurothioglucose, platinum compounds, and retinoic acid derivatives. The properties of TrxR in combination with the functions of Trx position this system at the core of cellular thiol redox control and antioxidant defense. In this review, we focus on the reactions of the Trx system with ROS molecules and different cellular antioxidant enzymes. We summarize the TrxR-catalyzed regeneration of several antioxidant compounds, including ascorbic acid (vitamin C), selenium-containing substances, lipoic acid, and ubiquinone (Q10). We also discuss the general cellular effects of TrxR inhibition. Dinitrohalobenzenes constitute a unique class of immunostimulatory TrxR inhibitors and we consider the immunomodulatory effects of dinitrohalobenzene compounds in view of their reactions with the Trx system.
Article
Full-text available
Confocal laser scanning microscopy has revolutionized cell biology. However, the technique has major limitations in speed and sensitivity due to the fact that a single laser beam scans the sample, allowing only a few microseconds signal collection for each pixel. This limitation has been overcome by the introduction of parallel beam illumination techniques in combination with cold CCD camera based image capture. Using the combination of microlens enhanced Nipkow spinning disc confocal illumination together with fully automated image capture and large scale in silico image processing we have developed a system allowing the acquisition, presentation and analysis of maximum resolution confocal panorama images of several Gigapixel size. We call the method Extended Field Laser Confocal Microscopy (EFLCM). We show using the EFLCM technique that it is possible to create a continuous confocal multi-colour mosaic from thousands of individually captured images. EFLCM can digitize and analyze histological slides, sections of entire rodent organ and full size embryos. It can also record hundreds of thousands cultured cells at multiple wavelength in single event or time-lapse fashion on fixed slides, in live cell imaging chambers or microtiter plates. The observer independent image capture of EFLCM allows quantitative measurements of fluorescence intensities and morphological parameters on a large number of cells. EFLCM therefore bridges the gap between the mainly illustrative fluorescence microscopy and purely quantitative flow cytometry. EFLCM can also be used as high content analysis (HCA) instrument for automated screening processes.
Article
Full-text available
We have determined the sequence of 23 peptides from bovine thioredoxin reductase covering 364 amino acid residues. The result was used to identify a rat cDNA clone (2.19 kilobase pairs), which contained an open reading frame of 1496 base pairs encoding a protein with 498 residues. The bovine and rat thioredoxin reductase sequences revealed a close homology to glutathione reductase including the conserved active site sequence (Cys-Val-Asn-Val-Gly-Cys). This also confirmed the identity of a previously published putative human thioredoxin reductase cDNA clone. Moreover, one peptide of the bovine enzyme contained a selenocysteine residue in the motif Gly-Cys-SeCys-Gly (where SeCys represents selenocysteine). This motif was conserved at the carboxyl terminus of the rat and human enzymes, provided that TGA in the sequence GGC TGC TGA GGT TAA, being identical in both cDNA clones, is translated as selenocysteine and that TAA confers termination of translation. The 3'-untranslated region of both cDNA clones contained a selenocysteine insertion sequence that may form potential stem loop structures typical of eukaryotic selenocysteine insertion sequence elements required for the decoding of UGA as selenocysteine. Carboxypeptidase Y treatment of bovine thioredoxin reductase after reduction by NADPH released selenocysteine from the enzyme with a concomitant loss of enzyme activity measured as reduction of thioredoxin or 5,5'-dithiobis(2-nitrobenzoic acid). This showed that the carboxyl-terminal motif was essential for the catalytic activity of the enzyme.
Article
Full-text available
Mammalian thioredoxin reductases (TrxR) are homodimers, homologous to glutathione reductase (GR), with an essential selenocysteine (SeCys) residue in an extension containing the conserved C-terminal sequence -Gly-Cys-SeCys-Gly. In the oxidized enzyme, we demonstrated two nonflavin redox centers by chemical modification and peptide sequencing: one was a disulfide within the sequence -Cys59-Val-Asn-Val-Gly-Cys64, identical to the active site of GR; the other was a selenenylsulfide formed from Cys497-SeCys498 and confirmed by mass spectrometry. In the NADPH reduced enzyme, these centers were present as a dithiol and a selenolthiol, respectively. Based on the structure of GR, we propose that in TrxR, the C-terminal Cys497-SeCys498 residues of one monomer are adjacent to the Cys59 and Cys64 residues of the second monomer. The reductive half-reaction of TrxR is similar to that of GR followed by exchange from the nascent Cys59 and Cys64 dithiol to the selenenylsulfide of the other subunit to generate the active-site selenolthiol. Characterization of recombinant mutant rat TrxR with SeCys498 replaced by Cys having a 100-fold lower kcat for Trx reduction revealed the C-terminal redox center was present as a dithiol when the Cys59-Cys64 was a disulfide, demonstrating that the selenium atom with its larger radius is critical for formation of the unique selenenylsulfide. Spectroscopic redox titrations with dithionite or NADPH were consistent with the structure model. Mechanisms of TrxR in reduction of Trx and hydroperoxides have been postulated and are compatible with known enzyme activities and the effects of inhibitors, like goldthioglucose and 1-chloro-2,4-dinitrobenzene.
Article
Thioredoxin reductase is a newly identified selenocysteine-containing enzyme that catalyzes the NADPH-dependent reduction of the redox protein thioredoxin. Thioredoxin stimulates cell growth, is found in dividing normal cells, and is over-expressed in a number of human cancers. Redox activity is essential for the growth effects of thioredoxin; thus, thioredoxin reductase could be involved in regulating cell growth through its reduction of thioredoxin. In rats fed a selenium-deficient diet (<0.01 ppm) for up to 98 days, thioredoxin reductase activity was decreased, compared with that of rats fed a normal selenium diet (0.1 ppm), in lung, liver, and kidney, while thioredoxin reductase activity in the spleen and prostate was unaltered. Rats fed a high selenium diet (1.0 ppm) exhibited a 1.5-fold increase in kidney and a 2.0-fold increase in lung thioredoxin reductase activity that began to return to control values after 20 and 69 days, respectively. Liver showed a 2.1-fold increase in thioredoxin reductase activity at 20 days only. Thioredoxin reductase protein levels measured by western blotting using an antibody to human thioredoxin reductase were decreased in rats fed the selenium-deficient diet and did not increase in rats fed the high selenium diet. Rat thioredoxin reductase was shown to incorporate ⁷⁵Selenium. Thus, in some tissues at least, the increase in thioredoxin reductase activity of rats fed a high selenium diet appears to be due to an increase in the specific activity of the enzyme, possibly caused by increased selenocysteine incorporation without an increase in thioredoxin reductase protein synthesis.
Article
Human thioredoxin reductase is a dimeric enzyme that catalyzes reduction of the disulfide in oxidized thioredoxin by a mechanism involving transfer of electrons from NADPH via FAD to a redox-active disulfide. 1-Chloro-2,4-dinitrobenzene (DNCB) is an alkylating agent used for depleting intracellular GSH and also showing distinct immunomodulatory properties. We have discovered that low concentrations of DNCB completely inactivated human or bovine thioredoxin reductase, with a second order rate constant in excess of 200 M s, which is almost 10,000-fold faster than alkylation of GSH. Total inactivation of 50 nM reduced thioredoxin reductase was obtained by 100 μM DNCB after 5 min of incubation at 20°C also in the presence of 1 mM GSH. The inhibition occurred with enzyme only in the presence of NADPH and persisted after removal of DNCB, suggesting alkylation of the active site nascent thiols as the mechanism of inactivation. Thioredoxin reductase modified by DNCB lacked reducing activity with oxidized thioredoxin, 5,5′-dithiobis-(2-nitrobenzoic acid), or sodium selenite. However, the DNCB-modified enzyme oxidized NADPH at a rate of 4.7 nmol/min/nmol of enzyme in the presence of atmospheric oxygen. This activity was not dependent on the presence of DNCB in solution and constituted a 34-fold increase of the inherent low NADPH oxidase activity of the native enzyme. DNCB is a specific inhibitor of mammalian thioredoxin reductase, which reacted 100-fold faster than glutathione reductase. The inactivation of the disulfide reducing activity of thioredoxin reductase and thioredoxin with a concomitant large increase of the NADPH oxidase activity producing reactive oxygen intermediates may mediate effects of DNCB on cells in vivo.
Article
Thioredoxins belong to a widely distributed group of small proteins with strong reducing activities mediated by a consensus redox-active dithiol (Cys-Gly-Pro-Cys). Thioredoxin was first isolated as a hydrogen donor for enzymatic synthesis of deoxyribonucleotides by ribonucleotide reductase inEscherichia coli.Recent studies have revealed a variety of roles that thioredoxin plays in transcription, growth control, and immune function. In this report, we describe the phenotype of mice carrying a targeted disruption of the thioredoxin gene (Txn). Heterozygotes are viable, fertile, and appear normal. In contrast, homozygous mutants die shortly after implantation, and the concepti were resorbed prior to gastrulation. When preimplantation embryos were placed in culture, the inner cell mass cells of the homozygous embryos failed to proliferate. These results indicate thatTxnexpression is essential for early differentiation and morphogenesis of the mouse embryo.
Article
Several studies have demonstrated a correlation between cellular toxicity of cis-diamminedichloroplatinum (II) (cisplatin, CDDP) and inhibited intracellular activity of the thioredoxin system, i.e., thioredoxin (Trx), thioredoxin reductase (TrxR), and NADPH. Conversely, increased cellular activity of the Trx system confers resistance to CDDP. In this study, we have analyzed the interaction of CDDP with Trx and TrxR in order to clarify the mechanism. The inhibition with time-dependent kinetics by CDDP of NADPH-reduced (but not oxidized) TrxR was irreversible, strongly suggesting covalent modification of the reduced selenocysteine-containing active site. Assuming second order kinetics, the rate constant of TrxR inhibition by CDDP was 21 ± 3 M−1s−1. Transplatin was found to be an even more efficient inhibitor, with a second order rate constant of 84 ± 22 M−1s−1, whereas carboplatin (up to 1 mM) gave no inhibition of the enzyme under the same conditions. Escherichia coli Trx or human or bacterial glutaredoxin (Grx) activities were in comparison only slightly or not at all inhibited by either CDDP, transplatin, or carboplatin. However, glutaredoxins were found to be inhibited by the purified glutathione adduct of cisplatin, bis-(glutathionato)platinum(II) (GS-Platinum complex, GS-Pt), with an IC50 = 350 μM in the standard β-hydroxyethyl disulfide-coupled assay for human Grx. Also the mammalian Trx system was inhibited by GS-Pt with similar efficiency (IC50 = 325 μM), whereas neither the E. coli Trx system nor glutathione reductase were inhibited. Formation of GS-Pt is a major route for cellular elimination of CDDP. The fact that GS-Pt inhibits the mammalian Trx as well as Grx systems shows that CDDP may exert effects at several stages of its metabolism, including after conjugation with GSH, which are intimately linked with the cellular disulfide/dithiol redox regulatory systems.