Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

The FASEB Journal • Research Communication

Crucial yet divergent roles of mitochondrial redox state


in skeletal muscle vs. brown adipose tissue energetics
Ryan J. Mailloux, Cyril Nii-Klu Adjeitey, Jian Ying Xuan, and Mary-Ellen Harper1
Department of Biochemistry, Microbiology, and Immunology, Faculty of Medicine, University of
Ottawa, Ottawa, Ontario, Canada

ABSTRACT Reduced glutathione (GSH) is the major protein antioxidant species in cells (3). Reactive oxygen
determinant of redox balance in mitochondria and as species (ROS) are a normal byproduct of fuel metabo-
such is fundamental in the control of cellular bioener- lism in cells and are generated predominantly in mito-
getics. GSH is also the most important nonprotein chondria. As the singlet electrons of ROS can damage
antioxidant molecule in cells. Surprisingly, the effect of cellular constituents and functions, there are many
redox environment has never been examined in skele- enzymatic and nonenzymatic mechanisms to keep cel-
tal muscle and brown adipose tissue (BAT), two tissues lular ROS in check. The GSH-mediated scavenging of
that have exceptional dynamic range and that are ROS generates inactive glutathione disulfide (GSSG),
relevant to the development of obesity and related which, in turn, is rapidly converted back to GSH by
diseases. Here, we show that the redox environment NADPH and glutathione reductase (3). This antioxida-
plays crucial, yet divergent, roles in modulating mito- tive system allows mitochondria to maintain a high
chondrial bioenergetics in skeletal muscle and BAT. GSH/GSSG ratio and buffer locally generated ROS
Skeletal muscle mitochondria were found to naturally during metabolic processes. Recent work has also
have a highly reduced environment (GSH/GSSG⬇46), shown that GSH can regulate the redox environment of
and this was associated with fairly high (⬃40%) rates of mitochondria by forming disulfide bridges with ex-
state 4 (nonphosphorylating) respiration and decreased posed thiols on proteins, a process referred to as
reactive oxygen species (ROS) emission. The degluta- glutathionylation (4 –7). Very recent studies reveal that
thionylation of uncoupling protein 3 (UCP3) following this type of reversible reaction is a key post-translational
an increase in the reductive potential of mitochondria control mechanism required to modulate protein func-
results in a further increase in nonphosphorylating tion in response to changes in redox status (8, 9).
respiration (⬃20% in situ). BAT mitochondria were Fluctuations in ROS induce the glutathionylation and
found to have a much more oxidized status (GSH/ deglutathionylation of proteins, and the processes are
GSSG⬇13) and had basal reactive oxygen species emis- highly specific (5, 10). For example, certain proteins,
sion that was higher (⬃250% increase in ROS genera- such as actin, SDH, uncoupling protein 2 (UCP2), and
tion) than that in skeletal muscle mitochondria. When UCP3, are deglutathionylated by ROS (5, 11, 12). In
redox status was subsequently increased (i.e., more contrast, other proteins, including some electron trans-
reduced), UCP1-mediated uncoupling was more sensi- port chain respiratory complexes and tricarboxylic acid
tive to GDP inhibition. Surprisingly, BAT was found to cycle enzymes, are controlled by glutathionylation in
be devoid of glutaredoxin-2 (Grx2) expression, while response to ROS (13–16). More overt oxidative stres-
there was abundant expression in skeletal muscle. sors that induce a great decrease in GSH/GSSG can
Taken together, these findings reveal the importance of
prompt the spontaneous hyperglutathionylation of pro-
redox environment in controlling bioenergetic func-
teins, resulting in metabolic dysfunction (17).
tions in both tissues, and the highly unique character-
Two tissues that have remarkable degrees of meta-
istics of BAT in this regard.—Mailloux, R. J., Adjeitey,
bolic activity and flexibility are skeletal muscle and
C. N.-K., Xuan, J. Y., and Harper, M.-E. Crucial yet
brown adipose tissue (BAT). Both have the fairly
divergent roles of mitochondrial redox state in skeletal
unique ability to oxidize a range of carbon sources to
muscle vs. brown adipose tissue energetics. FASEB J. 26,
satisfy energy demands. Regardless of energy source,
363–375 (2012). www.fasebj.org
the graded transfer of electrons from NADH and
FADH2 to diatomic oxygen (O2) is used to form the
Key Words: glutathione 䡠 thermogenesis 䡠 uncoupling proteins
1
Correspondence: University of Ottawa, Faculty of Medi-
By virtue of its high concentration (⬃10 mM) and cine, Department of Biochemistry, Microbiology, and Immu-
nology, 451 Smyth Road, Ottawa, ON, Canada, K1H 8M5.
powerful reductive potential (E0⫽⫺240 mV), reduced E-mail: maryellen.harper@uottawa.ca
glutathione (GSH) is the major redox buffer in mito- doi: 10.1096/fj.11-189639
chondria (1, 2). The redox buffering capacity of GSH This article includes supplemental data. Please visit http://
allows this molecule to function as the principal non- www.fasebj.org to obtain this information.

0892-6638/12/0026-0363 © FASEB 363


protonmotive force across the mitochondrial inner determinations. Mice were housed either at room tempera-
membrane (MIM) (18). Protonmotive force can then ture (RT; 23°C) or 4°C for 24 h prior to experimentation. All
be used to drive ATP production (coupled respiration) experiments on skeletal muscle mitochondria were per-
formed on mice housed at RT. During cold exposure, mice
or can be dissipated by proton-leak reactions (nonphos- were given free access to food and water but no housing
phorylating, uncoupled respiration) (18). Proton leak- materials. On the day of experiments, mice were quickly
age is mediated through uncoupling protein (UCP) euthanized (decapitation), and the skeletal muscle and BAT
and non-UCP mechanisms (19). Recent work has indi- were immediately extracted and placed in ice-cold buffer with
cated that brown fat and skeletal muscle cells can share or without 2 mM DTT. All experiments were performed
the same ancestral cellular lineage (20). In fact, Sca1⫹ according to the principles and guidelines of the Canadian
Council of Animal Care and the study was approved by the
progenitor cells found in white fat and skeletal muscle
Animal Care Committee of the University of Ottawa.
can easily be induced to differentiate into brown adi-
pocytes, while those in brown fat are constitutively
Mitochondrial isolation
committed as brown adipocyte precursors (21). Despite
their similarity in fuel oxidation and common cell
Buffers were ice-cold, and procedures were performed on ice
lineages, substrate oxidation in skeletal muscle and
or at 4°C to maintain mitochondrial redox status and func-
BAT serves very different purposes. In skeletal muscle, tional characteristics. To test the effect of redox state on
substrate oxidation is mostly used to support coupled skeletal muscle and BAT mitochondria, isolations were per-
respiration to generate ATP for contractile functions formed in the presence or absence of 2 mM DTT. DTT is a
(22). However, it must be kept in mind that uncoupled strong reducing agent that maintains a more reduced GSH
respiration in skeletal muscle can account for ⬃20 – pool in mitochondria (7). Notably, however, DTT was not
40% of energy expenditure in this tissue, and its included in any of the mitochondrial resuspension and assay
buffers. Muscle mitochondria were isolated as described by
quantitative importance is greatest when ATP demands Mailloux et al. (11). Skeletal muscle from forelimbs and
of muscle are low (e.g., when muscle is at rest; refs. 22, hindlimbs was dissected, cleaned of connective tissue and fat,
23). In BAT, substrate oxidation rates are very high and placed in buffer A (140 mM KCl, 20 mM HEPES, 5 mM
when UCP1 is activated to fulfill this tissue’s essential MgCl2, and 1 mM EGTA; pH 7.0). Muscle was weighed and
role in thermoregulation (24, 25). Under conditions of minced on a Teflon board and then placed in 30 ml of
thermoneutrality, BAT is inactive, and only basal cellu- homogenizing medium (buffer A supplemented with 1 mM
lar reactions occur (26). However, when an animal is ATP and 1% BSA, w/v) containing 1 U of protease (subtilisin
A). Tissue was homogenized using a glass/Teflon Potter-
exposed to the cold, BAT metabolic activity increases Elvehjem tissue grinder and then centrifuged at 800 g (9 min,
substantially to support thermogenesis (27, 28). While 4°C) to remove undisrupted tissue and whole cells. The
UCP3 is also expressed in BAT, UCP3 is the only UCP supernatant was centrifuged at 12,000 g (9 min, 4°C), and the
protein expressed in muscle. UCP3-mediated proton resulting pellet was resuspended in 1 ml of buffer A and
leakage is thought to be activated by ROS to provide a incubated on ice for 5 min to allow myofibrillar repolymer-
negative feedback loop mitigating further ROS produc- ization. Samples were then diluted to 30 ml with buffer A and
centrifuged at 800 g (9 min, 4°C) to remove myofibrils.
tion by mitochondria (29, 30). We recently demon-
Finally, the supernatant was centrifuged at 12,000 g (9 min,
strated that UCP3, unlike UCP1, is controlled by revers- 4°C) to pellet the mitochondria. The final pellet was resus-
ible glutathionylation/deglutathionylation, a process pended in buffer A devoid of DTT.
induced by ROS (11). BAT mitochondria were isolated as described by Mailloux
Although mitochondrial metabolism in both tissues et al. (11). Briefly, interscapular BAT was removed, cleaned of
has been investigated extensively, very little informa- white adipose tissue, and placed in 30 ml of 250 mM sucrose
tion exists on the influence of redox state. This is solution. The tissue was minced and homogenized using a
glass/Teflon Potter-Elvehjem tissue grinder in 30 ml of 250
particularly true for BAT, a tissue that has recently mM sucrose solution. The homogenate was centrifuged at
attracted a great deal of attention since it has been 8500 g (9 min, 4°C), and the pellet was resuspended in 250
unequivocally demonstrated that BAT exists in adult mM sucrose plus 0.2% BSA and spun at 800 g (9 min, 4°C).
humans (31–33). Here, we examined the effect of the The supernatant was centrifuged at 8500 g (9 min, 4°C), and
redox environment on bioenergetics and ROS in skel- the mitochondrial pellet was resuspended in buffer B (125
etal muscle and BAT mitochondria. We report differ- mM sucrose, 25 mM Trizma base, 5 mM MgCl2, and 1 mM
ences in redox environment and ROS production and EGTA, pH 7.4) devoid of DTT. Protein concentration was
determined by Bradford assay using BSA as the standard.
their control in these two tissues. We further demon- Mitochondrial preparations were kept on ice, and all exper-
strate that these differences were matched by variations iments were conducted within 2 h following isolation. Exper-
in GSH and GSSG levels. iments using skeletal muscle mitochondria were conducted in
buffer A. All BAT mitochondria experiments were performed
using buffer B and performed under state 4 respiratory
conditions. Aliquots of mitochondria were also used for
MATERIALS AND METHODS HPLC analyses or stored at ⫺80°C for later immunoblot
determinations.
Mice and tissue extraction
Mitochondrial membrane potential (⌬⌿m) determinations
Male C57BL6J mice (9 –12 wk of age) fed a standard ad libitum
diet (44.2% carbohydrate, 6.2% fat, 18.6% crude protein; diet Kinetic determinations were conducted in isolated mitochon-
T.2018; Harlan, Indianapolis, IN, USA) were used for all dria and the fluorophore, safranin O (5 ␮M; excitation 485

364 Vol. 26 January 2012 The FASEB Journal 䡠 www.fasebj.org MAILLOUX ET AL.
nm, emission 580 nm) using a BioTEK FLx800 reader and addition of substrates or allosteric regulators. Data were
Gen5 software (BioTek Instruments, Inc., Winooski, VT, normalized to background fluorescence and mitochondrial
USA). Mitochondria (0.5 mg/ml) were preincubated in ei- concentration.
ther buffer A or buffer B at 37°C prior to initiating experi-
ments. For skeletal muscle, mitochondria were assayed under Western blot analysis
state 3 and state 4 conditions. Mitochondria were energized
for 1.5 min with 5 mM succinate and then deenergized with
Levels of glutaredoxin-1 (Grx1), glutaredoxin-2 (Grx2), and
0.5 mM ADP for 4 min. Mitochondria were then treated with
thioredoxin-2 (Trx2) were tested by immunoblot. Immedi-
1.3 ␮g/ml of oligomycin to assess the effect of redox environ-
ately following isolation, mitochondria were diluted to 1
ment on state 4 respiration. Rates of change in fluorescence
mg/ml in Laemmli buffer and then electrophoresed on a
(AU/min/mg protein) were then calculated for state 3 and
10% isocratic SDS gel. On completion, gel slabs were equili-
state 4 conditions, respectively. For BAT mitochondria, the
brated in transfer buffer and then electroblotted onto nitro-
effect of redox environment on UCP1 function was tested by
cellulose membranes at RT for 1 h at 100 V with a maximum
GDP and oleic acid titrations. Immediately prior to the assay,
amperage of 400 mA. The membranes were blocked with
mitochondria were preincubated in 1.3 ␮g/ml oligomycin to
Tris-buffered saline containing 1% (v/v) Tween-20 (TBS-T).
induce state 4 (nonphosphorylating) respiratory conditions;
Membranes were then probed for 24 h at 4°C with primary
subsequent changes in membrane potential were probed
antibodies directed against Grx1 (1:500; Abcam, Cambridge,
using UCP1 inhibitors (GDP) and activators (oleate). Mito-
MA, USA), Grx2 (1:500; Abcam), Trx2 (1:500; Abcam), UCP1
chondria were treated with 5 mM pyruvate and 3 mM malate
(1:10,000; Sigma, St. Louis, MO, USA), and UCP3 (Abcam;
for 1 min, and then GDP was added incrementally over time.
1:1000). For UCP1, membranes were probed for only 1 h at
Oleic acid was then titrated to test UCP1 reactivation (incre-
RT. SDH served as the loading standard (1:2000, SDHA
ments of 0.1 mM). Fresh oleate was used in every experiment.
antibody; Santa Cruz Biotechnology, Santa Cruz, CA, USA).
The time of addition for each chemical is indicated in the
Membranes were then incubated for 1 h at RT with the
figures. As a control, safranin O fluorescence in the presence
requisite horseradish peroxidase-conjugated secondary anti-
of mitochondria and in the absence of substrate was per-
bodies (anti-rabbit, 1:2000; Santa Cruz Biotechnology). Blots
formed immediately before each experiment.
were incubated in enhanced chemiluminescent substrate for
visualization (ECL kit; Thermo Scientific, Waltham, MA,
Oxygen consumption measurements USA).

BAT mitochondria (0.5 mg/ml) isolated in the presence or Protein-glutathione (PSSG) adduct detection
absence of DTT at 37°C were studied in a Clark-type oxygen
electrode (Hansatech, Norfolk, UK). Prior to determinations, Following mitochondrial isolation and determination of pro-
mitochondria were warmed for 10 min in the electrode tein content, mitochondria were washed once with either
chamber in buffer B at 37°C. Buffer B was assumed to contain buffer A (skeletal muscle) or buffer B (BAT) and then treated
406 nmol O/ml. State 4 respiratory conditions were then with Laemmli buffer containing 25 mM N-ethylmaleimide
induced by adding oligomycin (1.3 ␮g/ml) to the chamber. (NEM). The NEM was included to prevent spontaneous
State 4 respiration was then tested following the addition of 5 glutathionylation events and to quench any residual DTT.
mM pyruvate and 3 mM malate. Experiments were performed Protein (30 ␮g) was electrophoresed in a 10% linear gradient
under state 4 conditions to assess UCP1-mediated respiration SDS gel under standard conditions. On completion, gel slabs
only. Mitochondria were then treated acutely with GDP (final were incubated in transfer buffer and then used for electro-
concentration of 1 mM), followed by a treatment with oleate blotting, as described above. Membranes were blocked for 1 h
(final concentration of 0.1 mM). Measurements of oxygen at RT in TBS-T containing 5% (w/v) nonfat skim milk. PSSG
consumption rate (OCR) were completed within an 8- to adducts were detected using the anti-PSSG (1:500; Virogen,
12-min period. Boston, MA, USA). Following a 24-h incubation in primary
antibody at 4°C, membranes were probed for 1 h at RT with
Mitochondrial ROS emission determinations anti-mouse horseradish peroxidase-conjugated secondary an-
tibody (1:2000, Santa Cruz Biotechnology). Immunoreactive
bands were detected as described above. Blots were quanti-
ROS emission experiments were conducted using the BioTEK
fied using ImageJ software (U.S. National Institutes of Health,
SynergyMX microplate reader and Gen5 software. Mitochon-
Bethesda, MD, USA).
dria isolated in the presence or absence of DTT were tested
kinetically for ROS emission in the presence of different
substrates and allosteric regulators. ROS production was GSH/GSSG determinations
assessed fluorometrically using 20 ␮M DCF diacetate (DCFH-
DA). For all experiments, mitochondria were prewarmed in Mitochondrial GSH and GSSG levels were assessed by HPLC.
the requisite buffer for 10 min at 37°C. Baseline measure- Following one wash in either buffer A or buffer B, mitochon-
ments of DCFH-DA fluorescence were performed in the dria isolated from skeletal muscle and BAT were diluted
presence of mitochondria prior to addition of substrates. For to 0.5 mg/ml in ice-cold 1% (w/v) trichloroacetic acid
skeletal muscle mitochondria (0.5 mg/ml), ROS production (TCA)/1% (w/v) meta-phosphoric acid solution and incu-
was tested in the presence of succinate (5 mM) or pyruvate/ bated on ice for 10 min. Precipitate was removed by centrif-
malate (5 mM/3 mM). ROS production was then tested ugation for 10 min at 12,000 g at 4°C. The supernatant was
under oligomycin-induced state 4 conditions (1.3 ␮g/ml). then collected and stored at ⫺80°C. On the day of experi-
ROS emission in BAT mitochondria was tested in buffer B ments, samples were injected into an Agilent 1100 Series
under oligomycin-induced state 4 conditions. BAT mitochon- HPLC equipped with a Pursuit C18 column (150⫻4.6 mm, 5
dria (0.5 mg/ml) were treated with 5 mM pyruvate and 3 mM ␮m; Agilent Technologies, Santa Clara, CA, USA). For proper
malate, and then ROS emission was ascertained following the separation of GSH and GSSG, a flow rate of 1 ml/min was
addition of GDP (1 mM) and oleate (0.1 mM). Experiments used, with a mobile phase consisting of 0.1% (w/v) TCA and
were conducted for 20 –30 min. Rates of change in ROS HPLC-grade methanol in a 90:10 ratio. GSH and GSSG were
production (AU/min/mg protein) were calculated after the detected using an Agilent UV-Vis variable-wavelength detec-

REDOX MODULATION OF MITOCHONDRIAL METABOLISM 365


tor operating at 215 nm. Retention times were confirmed by rates were tested as described above. Data were normalized to
injecting GSH and GSSG standards. GSH and GSSG were protein content/well. State 4 respiration was expressed as a
quantified using Agilent Chemstation software. percent of the basal respiration.
For diamide titrations, C2C12 cells were grown, differenti-
In situ measurement of bioenergetics ated, transfected, and then treated with or without 2 mM
DTT, as described above. Cells were transfected with either
pCMV-UCP3 or empty pCMV vector. Cells were tested for the
The XF24 Extracellular Flux Analyzer (Seahorse Bioscience, presence or absence of UCP3 by immunoblot (Supplemental
North Billerica, MA, USA) was employed to determine the Fig. S1). Immediately prior to loading the cell plate into the
mitochondrial bioenergetic characteristics in primary myo- Seahorse XF-24, medium was removed and replaced with
tubes incubated in the presence or absence of 2 mM DTT. fresh prewarmed DTT-free Seahorse medium containing 1.3
Primary myoblasts were isolated and purified from wild-type ␮g/ml of oligomycin to induce state 4 respiratory conditions.
C57BL6J mice, as described by Rando et al. (34). Myoblasts Following assessment of state 4 respiration, diamide was
were subcultured and maintained in DMEM containing 10% injected sequential to final concentrations of 50, 100, 200,
(v/v) FBS, 5 mM glucose, 1 mM pyruvate, 2% (v/v) antimy- and 300 ␮M, respectively. OCR was tested for two intervals
cotics-antibiotics (AA), and 2.5 ng/ml fibroblast growth fac- following each injection, each including 2 min of mixing, 2
tor (FGF). For experiments, primary myoblasts were seeded at min of incubation, and 2 min of measurement. Data were
50,000 cells/ml and grown in Matrigel-coated 24-well Sea- normalized to protein content/well.
horse XF24 culture plates. On reaching ⬃80% confluency,
cells were differentiated for up to 5 d in DMEM containing
2% (v/v) FBS, 5 mm glucose, 1 mM pyruvate, and 2% AA. On Statistical analysis
the day of experiments, differentiated cells were incubated
for 45 min at 37°C in HCO3-free DMEM containing 5 mM Student’s t tests were performed using Excel software
glucose, 4 mM l-glutamine, and 1 mM pyruvate (Seahorse (Microsoft, Redmond, WA, USA). Results are expressed as
medium) in the absence or presence of 2 mM DTT at means ⫾ se unless indicated otherwise in the figure
ambient CO2. The medium was then aspirated, cell monolay- captions.
ers were washed once with warmed PBS, and then fresh
prewarmed DTT-free Seahorse medium was added. Fluori-
metric sensors enabled the sensitive in situ measurement of
OCR and extracellular acidification rate (ECAR) in the
RESULTS
microenvironment surrounding the cells. Measurements of
OCR and ECAR were performed for 3 measurement intervals Glutathione levels are drastically different between
to assess basal rates (1 measurement interval consists of a skeletal muscle and BAT mitochondria and are
2-min mixing, 2-min incubation, and 2-min measurement differentially affected by the reducing agent, DTT
step). The effect of DTT on state 4 respiration was ascertained
by injecting oligomycin (1.3 ␮g/ml). State 4 respiration was
The absolute GSH and GSSG levels in skeletal muscle
tested over 3 measurement intervals. OCR and ECAR were
normalized to total cellular protein/well using the Bradford and BAT mitochondria (from RT mice) treated or not
assay. treated with DTT are presented in Table 1. Comparing
To assess whether DTT activates UCP3 in situ, C2C12 cells the GSH/GSSG of skeletal muscle and BAT revealed
were grown [DMEM, 5 mM glucose, 1 mM pyruvate, 10% that the GSH pool in BAT mitochondria was far more
(v/v) FBS, and 2% (v/v) AA] and differentiated [DMEM, 5 oxidized than skeletal muscle. This would indicate that,
mM glucose, 1 mM pyruvate 2% (v/v) FBS, and 2% (v/v) AA] in contrast to skeletal muscle, the BAT mitochondrial
on Matrigel-coated Seahorse plates (35). On full differentia-
tion, cells were transiently transfected with pCMV-UCP3
environment is far more oxidizing. Indeed, isolation of
plasmid (Origene, Rockville, MD, USA) using Lipofectamine skeletal muscle mitochondria in DTT increased the
2000 (Invitrogen, Carlsbad, CA, USA), according to the GSH/GSSG ratio substantially (Fig. 1A). BAT mito-
manufacturer’s instructions. Transfection protocols were op- chondria isolated from mice exposed to RT conditions
timized to avoid UCP3 overexpression and resultant uncon- and treated with DTT also contained a more reduced
trolled proton leakage. Following a 36-h transfection, cells GSH pool in comparison to mitochondria isolated in
were washed twice with PBS and incubated in Seahorse the absence of DTT (Fig. 1A).
medium containing 2 mM DTT, as described above. Cells
incubated in the absence of DTT served as the control. Changes in the GSH/GSSG ratio are often accompa-
Immediately prior to loading the plate in the Seahorse XF-24, nied by alterations in the levels of PSSG adducts.
the medium was removed and replaced with fresh prewarmed Skeletal muscle mitochondria isolated in the presence
DTT-free Seahorse medium. Basal and state 4 respiration of DTT displayed a significant 36% decrease in PSSG

TABLE 1. Absolute levels of GSH and GSSG in skeletal muscle and BAT mitochondria isolated in the presence or absence of DTT

Skeletal muscle BAT, RT BAT, 4°C

Parameter ⫺DTT ⫹DTT ⫺DTT ⫹DTT ⫺ DTT ⫹DTT

GSH (nmol/mg protein) 31.50 ⫾ 0.021 34.50 ⫾ 0.028* 30.73 ⫾ 0.002 34.56 ⫾ 0.003** 9.43 ⫾ 0.40 9.56 ⫾ 0.79
GSSG (nmol/mg protein) 0.709 ⫾ 0.014 0.438 ⫾ 0.004* 2.417 ⫾ 0.001 1.242 ⫾ 0.006** 1.66 ⫾ 0.36 1.41 ⫾ 0.34

BAT mitochondria were collected from mice exposed to RT and 4°C conditions. Absolute GSH and GSSG levels were determined by HPLC,
as described in Materials and Methods. Statistical significance was determined by comparing ⫺DTT and ⫹DTT; n ⫽ 4. *P ⬍ 0.05, **P ⬍ 0.005;
Student’s t test.

366 Vol. 26 January 2012 The FASEB Journal 䡠 www.fasebj.org MAILLOUX ET AL.
Figure 1. Effects of DTT pretreatment on redox environment in skeletal muscle and BAT mitochondria. BAT mitochondria were
isolated from mice exposed to RT conditions. Mitochondria from skeletal muscle and BAT were isolated in the presence of 2
mM DTT, as described in Materials and Methods. Mitochondria isolated in the absence of DTT served as the control. A) Effect
of DTT pretreatment on the mitochondrial GSH/GSSG ratio. Immediately following isolation, mitochondria were diluted to 0.5
mg/ml in a solution containing 1% (w/v) TCA/1% (w/v) meta-phosphoric acid and then clarified for HPLC analysis of GSH
and GSSG levels. GSH and GSSG retention times were confirmed, and levels were quantified by injecting varying amounts of
standard solutions. Following quantification, the GSH/GSSG ratio for skeletal muscle and BAT pretreated with DTT was
determined. Absolute amounts of mitochondrial GSH and GSSG are shown in Table 1; n ⫽ 5. B) Assessment of PSSG levels in
skeletal muscle and BAT mitochondria. Following isolation, 30 ␮g of mitochondrial protein was electrophoresed and tested for
PSSG levels using an anti-PSSG antibody (virogen). Blots were subsequently quantified using ImageJ software; n ⫽ 4.
C) Immunoblot analysis for enzymes involved in reversible glutathionylation. Mitochondrial protein (60 ␮g) was electropho-
resed and tested for Grx1, Grx2, and Trx2 levels, respectively. SDH served as the loading control. Values are expressed as
means ⫾ sd, using Student’s t test.

adducts (P⬍0.01; Fig. 1B). Immunoreactive bands in that UCP3-mediated proton leakage is activated by the
the low-molecular-weight range (⬃34 kDa) displayed deglutathionylation of UCP3 (9, 11). Hence, we exam-
the most dramatic decreases in intensity when skeletal ined the effect of the presence and absence of DTT on
muscle mitochondria were isolated in DTT. In contrast, mitochondrial bioenergetics. First, we tested the effect
DTT decreased PSSG levels by 27% in BAT mitochon- of DTT on skeletal muscle mitochondrial membrane
dria, but this did not reach statistical significance (Fig. potential using safranin O. Mitochondria were initially
1B). There were apparently a number of proteins that energized for ⬃1.5 min with 5 mM succinate, and then
were glutathionylated in BAT mitochondria; however, ADP and oligomycin were added sequentially to induce
DTT pretreatment had less of an effect on decreasing state 3 and 4 respiratory conditions, respectively. While
the amount of PSSG adducts when compared to effects there was a trend for an increase, the pretreatment
in skeletal muscle. We next tested the levels of putative of mitochondria with DTT had no significant effect
glutathionylating enzymes in mitochondria from both on rate of mitochondrial membrane depolarization
tissues. In Fig. 1C, skeletal muscle mitochondria con- under state 3 respiratory conditions (ADP⫹succinate;
tained Grx1, Grx2, and Trx2 proteins. In contrast, only Fig. 2A). Examination of membrane potentials under
Grx1 and Trx2 were detected in BAT mitochondria. oligomycin-induced state 4 respiratory conditions did,
There also appeared to be a slight shift in the mobility however, reveal that DTT pretreatment diminished the
of Grx1 between the skeletal muscle and BAT mito- rate of membrane potential repolarization, consistent
chondria. BAT mitochondria contained more Trx2, with the induction of leak-dependent processes (Fig.
which is thought to play a greater role in limiting 2B). The increase in state 4 respiration was confirmed
oxidative damage than in glutathionylation (36). Over- in situ using primary myotubes from wild-type mice.
all, these findings reveal that the mitochondrial redox Indeed, preincubation in DTT led to a significant
state and key redox enzymes are substantially different increase in state 4 respiration in cells (Fig. 2C). Assess-
between skeletal muscle and BAT. ment of the cellular ECAR revealed that DTT treatment
diminished glycolytic metabolism (Fig. 2C). We then
A more reductive mitochondrial environment determined whether DTT pretreatment diminished ROS
increases UCP3-mediated proton leakage in skeletal production from the skeletal muscle mitochondria (Fig.
muscle 2D). No significant changes in the rate of ROS production
were observed following succinate addition (slope⫺DTT⫽
Protein glutathionylation is often associated with 36.98⫾14.05 vs. slope⫹DTT⫽25.83⫾14.79, P⫽0.1, Stu-
changes in redox state, and we recently demonstrated dent’s t test). However, DTT pretreatment significantly

REDOX MODULATION OF MITOCHONDRIAL METABOLISM 367


Figure 2. A more reducing redox environment increases state 4 respiration and
decreases ROS production in skeletal muscle mitochondria. Mitochondria from
skeletal muscle were isolated in the presence of 2 mM DTT, as described in
Materials and Methods. Mitochondria isolated in the absence of DTT served as
the control. A) Effect of alterations in redox environment on mitochondrial
membrane potential under state 3 conditions. Mitochondrial membrane poten-
tial was determined using safranin O. Mitochondria (0.5 mg/ml) from skeletal
muscle were energized with 5 mM succinate and then treated with ADP (0.5
mM). Rate of change in fluorescence following ADP addition was then calcu-
lated; n ⫽ 5. B) Effect of alterations in redox environment on mitochondrial
membrane potential under state 4 conditions. Mitochondria (0.5 mg/ml) from
skeletal muscle that were energized with succinate (5 mM) and treated with ADP
(0.5 mM) were then treated with oligomycin (1.3 ␮g/ml). Rate of change in
fluorescence following oligomycin addition was then calculated; n ⫽ 5. C) Pretreatment of primary myotubes isolated
from wild-type (WT) mice with DTT increases oligomycin-induced state 4 respiration. Primary myotubes grown and
differentiated in Seahorse tissue culture plates were treated with 2 mM DTT for 45 min, and then OCR and ECAR were
tested in situ using the Seahorse Extracellular Flux Analyzer. Following the assessment of OCR and ECAR under basal
conditions, OCR and ECAR were tested under oligomycin-induced state 4 conditions. Data are expressed as a
percentage of basal conditions. OCR was normalized to protein content/well; n ⫽ 4. D) A more reductive environment
diminishes state 4 ROS generation. Mitochondria (0.5 mg/ml) were tested for ROS emission using DCFH-DA (20 ␮M).
ROS emission was tested kinetically with succinate (5 mM) and oligomycin (1.3 ␮g/ml). Rate of ROS production
following the addition of substrate or oligomycin was then calculated; n ⫽ 5. Values are expressed as means ⫾ se, using
Student’s t test.

diminished the rate of ROS production during oligomy- thionylation catalyst. We first determined whether DTT
cin-induced state 4 conditions (after oligomycin treat- could enhance leakage through UCP3 using C2C12
ment, slope⫺DTT⫽71.41⫾18.22 vs. slope⫹DTT⫽38.84⫾ cells transiently transfected with pCMV-UCP3. It is
18.19, Pⱕ0.01, Student’s t test). Rotenone treatment important to note that C2C12 cells do not express
diminished ROS production in mitochondria energized endogenous UCP3 (29). C2C12 cells preincubated in
by succinate and treated with oligomycin, confirming that the presence or absence of DTT did not display any
complex I is a major source for succinate-mediated ROS changes in the basal respiration and oligomycin-in-
production (Supplemental Fig. S2). However, mitochon- duced state 4 respiration (Fig. 3A). However, DTT
dria not exposed to DTT produced more ROS following pretreatment significantly increased the contribution
the oligomycin and rotenone treatments. We also tested of nonphosphorylating respiration to the total respira-
the effect of complex I substrates (pyruvate/malate) tion in the UCP3-overexpressing cells (Fig. 3B). Di-
on the rate of ROS production from mitochondria iso- amide titration experiments revealed that this DTT-
lated in the presence or absence of DTT (Supplemental mediated activation of UCP3 could be reversed by
Fig. S2). Again, ROS production following induction of glutathionylation. For this set of experiments, C2C12
state 4 respiration with oligomycin was found to be lower cells transiently transfected with either pCMV-UCP3 or
in mitochondria pretreated with DTT. empty pCMV vector were preincubated in the presence
or absence of DTT, and then state 4 respiratory condi-
Degree of glutathionylation controls UCP3-mediated tions were induced with oligomycin. Titration of di-
proton leakage amide up to a concentration of 200 ␮M resulted in a
progressive decrease in state 4 respiration (up to 40%)
The above results indicate that increased leak-depen- in UCP3-overexpressing cells treated with DTT (Fig.
dent respiration accounts, in part, for the decrease in 3C). Raising the diamide concentration to 300 ␮M had
mitochondrial ROS generation. Hence, we decided to no further effect on state 4 respiration. Diamide treat-
test whether DTT-induced increase in leak-dependent ment also diminished state 4 respiration in the cells
respiration was due to UCP3 deglutathionylation, and transfected with empty pCMV vector, but not to the
whether this DTT-mediated activation of UCP3 could same extent as the UCP3-expressing cells, indicating
be reversed by acute treatment with diamide, a gluta- that other proteins involved in catalyzing proton

368 Vol. 26 January 2012 The FASEB Journal 䡠 www.fasebj.org MAILLOUX ET AL.
Figure 3. UCP3 is required for the DTT-mediated increases in state 4 respiration. For all experiments, respiratory rates were
tested in situ with the Seahorse Extracellular Flux Analyzer. A) Absolute basal and oligomycin-induced respiratory rates in C2C12
cells transiently transfected with pCMV-UCP3 treated with or without 2 mM DTT. Cells were treated with or without 2 mM DTT
45 min prior to the assay. Following assessment of basal OCR, oligomycin-induced state 4 was tested; n ⫽ 4. B) Contribution of
leak-dependent respiration to total respiration. Oligomycin-induced state 4 respiration is expressed as a percentage of basal
OCR; n ⫽ 4. C, D) Deglutathionylation of UCP3 is responsible for the DTT-mediated increases in state 4 respiration. C2C12
myotubes transiently transfected with empty pCMV vector (⫺UCP3) or pCMV-UCP3 (⫹UCP3) were treated with (C) or without
(D) 2 mM DTT for 45 min, and then state 4 conditions were induced immediately with oligomycin (1.3 ␮g/ml). Effect of
glutathionylation on state 4 respiration was then tested by diamide titration. Following assessment of state 4 respiration in
UCP3⫺/⫹ cells treated with or without DTT, cells were acutely treated with increasing amounts of diamide (50, 100, 200, and
300 ␮M, respectively). Data are expressed as a percentage of the state 4 OCR induced with oligomycin prior to diamide
treatment (0 ␮M). State 4 OCR was normalized to protein content/well; n ⫽ 6. Values are expressed as means ⫾ sd, using
Student’s t test.

leakage are also inhibited by glutathionylation (Fig. Effects of a reductive environment on BAT
3C). Cells not treated with DTT were also tested for mitochondrial bioenergetics
UCP3 inhibition by diamide titration. Small, but
significant, decreases in state 4 respiration in UCP3-
Although DTT treatment increased the GSH/GSSG
expressing cells were observed when treated with 100 ratio in BAT mitochondria, it did not significantly
␮M and 200 ␮M diamide (Fig. 3D). Intriguingly, at decrease the amount of PSSG adducts. We previously
200 ␮M diamide, only a 20% decrease in state 4 demonstrated that in BAT mitochondria, unlike skel-
respiration was observed, indicating that pretreat- etal muscle mitochondria, glutathionylation appears
ment with DTT may liberate cysteine residues from to play a much less important role in modulating
glutathione in situ, making these residues available UCP1 (11). Hence, we tested the effect of redox
for reglutathionylation by diamide. Moreover, bring- environment on the bioenergetics of BAT mitochon-
ing the concentration of diamide up to 300 ␮M dria. Membrane potential determinations revealed
actually restored oligomycin-induced state 4 respira- that DTT pretreatment sensitized UCP1 to inhibition
tion to control levels (Fig. 3D). Hence, our results by GDP (Fig. 4A). In the mitochondria treated with
indicate that, at least in C2C12 cells, UCP3 is partially DTT, GDP induced a 2.5-fold increase in safranin O
glutathionylated under normal conditions and that fluorescence (indicative of large increases in mem-
full deglutathionylation is required for the full acti- brane potential), whereas GDP only increased the
vation of UCP3-mediated proton leakage. fluorescence by 1.6-fold in mitochondria not treated

REDOX MODULATION OF MITOCHONDRIAL METABOLISM 369


Figure 4. Influence of a changing redox environment on UCP1 in BAT mitochondria from RT mice. Mitochondria from BAT
were isolated in the presence of 2 mM DTT as described in Materials and Methods. Mitochondria isolated in the absence of DTT
served as the control. Mitochondria were treated with oligomycin (1.3 ␮g/ml) to induce state 4 respiratory conditions. A) Effect
of alterations in redox environment on mitochondrial membrane potential. Mitochondrial energization was tested using
safranin O. Mitochondria (0.5 mg/ml) from BAT were energized at time 0 with 5 mM pyruvate and 3 mM malate, and then GDP
(G) was titrated into the reaction at 0.5 mM increments to test the effect of DTT on UCP1 inhibition. Effect of DTT on UCP1
reactivation was then tested by adding increasing amounts of oleate (OA) at 0.1 mM increments. The change in fluorescence
following the addition of GDP and oleate was calculated from the background fluorescence. *P ⱕ 0.05. B) DTT increases
sensitivity of UCP1 to GDP but not oleate. O2 consumption determinations were performed using a Clark-type electrode under
oligomycin-induced state 4 conditions. Respiration was initially tested with 5 mM pyruvate and 3 mM malate, followed by GDP
(1 mM) and oleate (0.1 mM) treatment. C) DTT pretreatment increases ROS emission in BAT mitochondria isolated from RT
mice. ROS emission was tested under oligomycin-induced state 4 respiratory conditions. ROS production was detected with
DCFH-DA (20 ␮M). Following the addition of 5 mM pyruvate and 3 mM malate, ROS production was determined following GDP
(1 mM) and oleate (0.1 mM) treatment. Data were normalized to background fluorescence and amount of protein. Values are
expressed as means ⫾ se, using Student’s t test; n ⫽ 5.

with DTT (fold change calculated from the total GDP BAT mitochondrial redox environment in
response). It is important to note that GDP was mitochondria from cold-exposed mice
titrated in 0.5 mM increments, reaching a final
concentration of 2 mM (Fig. 4A). Since UCP1 has Given that UCP1 expression and uncoupling increase
been reported to be fully inhibited by 1 mM GDP, substantially in BAT when animals are exposed to a
and GDP is also known to inhibit ANT, some of the cold environment and that the activation of mitochon-
GDP effects may be due to ANT inhibition (37). In drial uncoupling can drastically alter mitochondrial
contrast to GDP, no difference in oleate sensitivity redox poise, we measured GSH and GSSG pools in BAT
was observed between mitochondria isolated in the mitochondria isolated from cold-exposed mice. The
absence or presence of DTT (Fig. 4A). BAT mito- absolute amounts of GSH and GSSG are presented in
chondrial oxygen consumption determinations con- Table 1. The GSH/GSSG ratio was even more oxidized
firmed these observations. A more pronounced in the BAT mitochondria from cold-exposed mice
decrease in oxygen consumption was observed in (Fig. 5A). As expected, cold exposure altered the
DTT-exposed mitochondria acutely treated with 1 bioenegetics of BAT mitochondria. Specifically, mito-
mM GDP (Fig. 4B). Although a trend for increased chondrial membrane potential was more depolarized
respiration was observed following oleate addition to following cold exposure, and UCP1 was more sensitive
the DTT-treated mitochondria, this did not reach to the allosteric regulators GDP and oleate (Fig. 5B). In
significance. We also tested ROS emission in BAT addition, OCRs were double what was observed with
mitochondria isolated in the presence or absence of mitochondria from RT mitochondria (Fig. 5C). In-
DTT. The addition of GDP did not increase the rate creasing mitochondrial redox state by DTT pretreat-
of ROS production (Fig. 4C). However, the addition ment still elevated the sensitivity of UCP1 to GDP
of oleate increased ROS emission in mitochondria pre- inhibition (Fig. 5B). No differences in membrane po-
treated with DTT. A comparison of the rates of ROS forma- larity were detected prior to GDP addition. In addition,
tion following oleate addition indicates that DTT pretreat- there was no statistically significant difference in the
ment enhances mitochondrial ROS formation in BAT effect of oleate on UCP1 in mitochondria isolated in
(slope⫺DTT⫽236.94⫾63.63 vs. slope⫹DTT⫽396.71⫾67.52, the presence or absence of DTT. Measurements of
P⫽0.05, Student’s t test). Oleate did not increase in the rate OCR confirmed these results (Fig. 5C). Exposure of
of ROS emission from mitochondria not pretreated with DTT-treated mitochondria to GDP resulted in a more
DTT. Hence, unlike skeletal muscle, BAT mitochondria significant decrease in O2 consumption. No differences
treated with DTT display a metabolic shift that favors in- in oxygen consumption were recorded when GDP-
creased sensitivity to UCP1 inhibition and higher ROS inhibited UCP1 was reactivated with oleate. We also
generation. tested the effect of the presence or absence of DTT on

370 Vol. 26 January 2012 The FASEB Journal 䡠 www.fasebj.org MAILLOUX ET AL.
Figure 5. Influence of a changing redox environment on UCP1 in BAT
mitochondria from cold-exposed mice. Mice were exposed to cold (4°C) for
24 h, and then mitochondria from BAT were isolated in the presence of 2 mM
DTT, as described in Materials and Methods. Mitochondria isolated in the
absence of DTT served as the control. Mitochondria were treated with
oligomycin (1.3 ␮g/ml) to induce state 4 respiratory conditions. A) Effect of
DTT treatment on GSH/GSSG ratio. Immediately following isolation, mito-
chondria were diluted to 0.5 mg/ml in a solution containing 1% (w/v) TCA
and 1% (w/v) meta-phosphoric acid, and then clarified for HPLC analysis of
GSH and GSSG levels. GSH and GSSG retention times were confirmed, and
levels were quantified by injecting varying amounts of standard solutions.
Following quantification, GSH/GSSG ratio was determined. Absolute amounts
of mitochondrial GSH and GSSG are shown in Table 1. Values are expressed
as means ⫾ sd, using Student’s t test; n ⫽ 4. B) Effect of alterations in redox environment on mitochondrial membrane
potential. Mitochondrial energization was tested using safranin O. Mitochondria (0.5 mg/ml) were energized at time 0 with
5 mM pyruvate and 3 mM malate under oligomycin-induced state 4 conditions, and then GDP (G) was titrated into the
reaction at 0.5 mM increments to test the effect of DTT on UCP1 inhibition. Effect of DTT on UCP1 reactivation was then
tested by adding oleate (OA) at 0.1 mM increments. Change in fluorescence following the addition of GDP and oleate was
calculated from the background fluorescence. Values are expressed as means ⫾ se, using Student’s t test; n ⫽ 4. C) DTT
increases sensitivity of UCP1 to GDP but not oleate. O2 consumption determinations were performed using a Clark-type
electrode under oligomycin-induced state 4 conditions. Respiration was initially tested with 5 mM pyruvate and 3 mM
malate, followed by GDP (1 mM) and oleate (0.1 mM) treatment. Values are expressed as means ⫾ se, using Student’s t
test; n ⫽ 4. D) Assessment of ROS production rates. ROS emission was tested under oligomycin-induced state 4 respiratory
conditions. ROS production was detected with DCFH-DA (20 ␮M). Following the addition of 5 mM pyruvate and 3 mM
malate, ROS production was determined following GDP (1 mM) and oleate (0.1 mM) treatment. Data were normalized to
background fluorescence and amount of protein. Values are expressed as means ⫾ se; n ⫽ 4.

ROS emission from BAT mitochondria isolated from flexibility. BAT relies largely on glycolysis for ATP
cold-exposed mice. Specifically, GDP did not increase production, and when fully activated, uses fatty acids to
the rate of ROS production in mitochondria pretreated support uncoupled respiration, while glycolytic ATP
with or without DTT (Fig. 5D). However, the addition production continues. In the fed state, skeletal muscle
of oleate increased the rate of mitochondrial ROS is highly glycolytic, whereas in the fasting state and
production in mitochondria pretreated with DTT (Fig. during aerobic exercise, its reliance on fatty acid oxi-
5D). Indeed, the slope of the line following oleate dation is high. Such metabolic distinctions may account
addition for the DTT-treated mitochondria was much for the observed differences in mitochondrial redox
steeper, indicating faster kinetics of ROS production state between these two tissues. We have demonstrated
(slope⫺DTT⫽979⫾420 vs. slope⫹DTT⫽1868⫾608, P⫽ that the differences in mitochondrial redox environ-
0.05, Student’s t test). ment are required to maintain the differences between
mitochondrial bioenergetics in skeletal muscle and
BAT mitochondria. To manipulate the redox environ-
DISCUSSION
ment, mitochondria were isolated in the absence or
In the present study, we provide the first analysis of the presence of DTT, a strong reducing agent. DTT is
effect of redox environment on bioenergetics in skele- commonly used to maintain a reductive environment
tal muscle and BAT mitochondria. Although a more during mitochondrial isolation since its highly negative
reductive environment was beneficial to skeletal muscle redox potential maintains glutathione in a reduced
mitochondria, these conditions seemed to have an state (38). The process of mitochondrial isolation, due
opposite effect on BAT mitochondria. These differ- to the exposure of mitochondria to ambient oxygen
ences may be related to the vastly different physiologi- levels can severely diminish mitochondrial function
cal roles of skeletal muscle and BAT. However, both (39). The recent work of Garcia et al. (7) reveals that
tissues are characterized by a high degree of metabolic mitochondria isolated from other tissues using a Per-

REDOX MODULATION OF MITOCHONDRIAL METABOLISM 371


coll gradient protocol diminishes redox state and mi- ⬃30°C (43). Hence, the BAT from the RT mice used in
tochondrial function. Hence, the isolation of mito- our study is slightly activated for thermogenesis, but
chondria in the presence of reducing agents, such as clearly not to the same extent as that from mice
DTT, protects enzymes from oxidative deactivation. In exposed to 4°C. It would be interesting to perform
the presence of DTT, skeletal muscle mitochondria additional studies on BAT from mice housed at ther-
displayed an increase in UCP3-dependent state 4 respi- moneutrality. Cold exposure does increase SOD in BAT
ration, which coincided with a decrease in ROS emis- mitochondria (44). Immunoblots on mitochondria
sion. Although this is thought to be due to the deglu- from the BAT of mice exposed to RT conditions
tathionylation and activation of UCP3, the increased revealed higher amounts of Trx2 and Grx1 following
availability of GSH would also quench ROS. DTT treatment, which can aid in quenching ROS. But
Early studies of liver mitochondria established that the Grx2, the matrix-associated Grx isoform, was too low in
redox environment influences mitochondrial metabolism amount to detect. Hence, increased ROS generation
(40). Indeed, these initial observations showed that sup- from BAT mitochondria is most likely due to the highly
plementing mitochondria with DTT increases mitochon- oxidative nature of this tissue, a unique characteristic
drial oxygen consumption. However, one caveat to these required to drive its thermogenic function. In the
early studies was that DTT was present during the actual present study, we not only confirmed that the GSH pool
measurements of oxygen consumption, and reducing is far more oxidized in comparison to skeletal muscle,
agents (carrying very negative redox potentials) decrease but that an oxidized GSH pool may be required for the
oxygen concentrations in solution. In the recent study by proper functioning of BAT mitochondria. In mitochon-
Garcia et al. (7), the importance of isolating mitochondria dria from both RT and cold-acclimated mice, UCP1
from tissues other than BAT or skeletal muscle in a displayed increased sensitivity to GDP inhibition when
reductive environment was demonstrated. The research- the redox environment was altered with DTT. We also
ers showed that DTT treatment maintains the metabolic tested the effect of DTT on ROS emission from BAT
efficiency of neuronal mitochondria by preserving the mitochondria. DTT did not have any effect on the rate
GSH pool, and thus the reductive potential of the matrix of mitochondrial ROS production when malate and
environment (7). Hill et al. (41) made similar observations pyruvate were used as substrates or when GDP was
in situ in vascular smooth muscle cells. These seminal added to inhibit UCP1. However, the addition of oleate
findings illustrate that redox plays a central role in mod- substantially increased the rate of ROS production in
ulating mitochondrial metabolism. However, it should be mitochondria pretreated with DTT. It should be noted
pointed out that these more current studies also included that the increases in ROS production are likely to be
DTT in their reaction mixtures, which again can alter associated with oleate oxidation pathways. Indeed,
oxygen levels in solution, and thus the measured rates of acute increases in fatty acid oxidation can be a signifi-
oxygen consumption. In our study, we pretreated mito- cant source of ROS in mitochondria (45). Cold expo-
chondria or cells with DTT and immediately prior to sure increased mitochondrial ROS generation even
determinations replaced the DTT-containing solutions further in comparison to RT mitochondria. In addi-
with buffers devoid of this compound. Hence, the ob- tion, BAT mitochondria generated far more ROS than
served changes were due to alterations in mitochondrial skeletal muscle mitochondria. Although we do not have
redox environment unencumbered by any DTT-induced an explanation for this high level of ROS generation by
decreases in oxygen tension. BAT mitochondria, we can only hypothesize that this is
Earlier studies in BAT mitochondria from mice ex- directly or indirectly due to the presence of UCP1
posed to different temperatures showed that the GSH and/or the absence of Grx2 (46). Indeed, oleate, a
pools are highly oxidized (42). This initial observation well-known UCP1 activator, increased ROS generation.
is intriguing, since we know now that a large fraction of Moreover, cold exposure, which increases mitochondrial
the mitochondrial GSH pool in other tissues is main- UCP1 content and activity increased ROS generation
tained in a reduced state (2). A high GSH/GSSG ratio even more. The observed increases in ROS production
ensures that the matrix environment is highly reducing, following oleate treatment and cold acclimation indicate
a physical property required for the proper functioning that UCP1 does not participate in mitigating mitochon-
of mitochondria (e.g., catalytic thiols are maintained in drial ROS emission. However, more in-depth investiga-
a reduced and active state, and a reductive environ- tions are warranted. Also, the effects of redox environ-
ment aids in quenching ROS). However, in the case of ment on UCP1 function should be investigated further in
BAT mitochondria, the matrix environment is highly UCP1-null mice. Hence, BAT mitochondria generate
oxidizing, especially in mitochondria isolated from large amounts of ROS, especially after stimulation of
cold-exposed mice. On BAT activation, O2 consump- UCP1. To our knowledge, there is only one previously
tion increases severalfold and appears to be accompa- published study in which BAT mitochondrial ROS pro-
nied by substantial increases in ROS production, which duction was measured. Mracek et al. (47) measured the
further oxidizes the GSH pool. However, mitochondria effect of different substrates on ROS emission in BAT
from mice housed at RT conditions also displayed high mitochondria; however, the researchers did not assess the
rates of ROS production, especially after oleate addi- effects of GDP and oleate on ROS emission.
tion. In this study, RT was ⬃23°C, which presents a mild Recently, we corroborated and extended the original
thermal stress to mice. Thermoneutrality for mice is findings of Echtay et al. (30) by showing that proton

372 Vol. 26 January 2012 The FASEB Journal 䡠 www.fasebj.org MAILLOUX ET AL.
leakage through UCP3 is activated by a ROS-induced ment increased the sensitivity of these cells to diamide
deglutathionylation (11). Only small amounts of ROS treatment. Since the only protein expressed in C2C12
activated UCP3, since higher amounts actually deacti- cells that is known to be associated with proton leakage
vated the protein. Moreover, we demonstrated that the is ANT, and since ANT may be modulated by glutathio-
ROS-mediated activation of UCP3 required the deglu- nylation (16), our current findings indicate that further
tathionylation of Cys25 or Cys259 and that the subse- research into the control of ANT-mediated proton
quent reglutathionylation of these residues inhibits leakage by glutathionylation in muscle is warranted.
UCP3-mediated proton leakage (11). In the present Like UCP3, UCP1 has been suggested to play a role
study, we not only confirmed our previous observations, in curtailing mitochondrial ROS emission through an
but also show that the degree of UCP3 glutathionyla- inducible proton-leak mechanism (48). However, the
tion dictates the contribution of this protein to proton role of UCP1 in controlling ROS production from
leakage. C2C12 cells overexpressing UCP3 displayed a mitochondria is highly contentious (49). We have also
small, but significant, increase in state 4 respiration recently showed that, unlike UCP3, UCP1 is not mod-
when pretreated with DTT. However, DTT pretreat- ulated by diamide-induced glutathionylation in BAT
ment increased the sensitivity of UCP3 to subsequent mitochondria (11). However, in the same publication,
diamide-mediated inhibition. Indeed, in UCP3-express- we did observe that UCP3 was controlled by glutathio-
ing C2C12 cells pretreated with DTT, there was a more nylation in BAT mitochondria (11). Indeed, BAT mi-
pronounced decline in state 4 respiration with increas- tochondria do express UCP3, and on the basis of our
ing doses of diamide. In the DTT-pretreated cells, a observations, it would seem that UCP3 and UCP1 are
concentration of up to 200 ␮M diamide was needed to controlled differently, which is surprising since they
fully inhibit UCP3. In contrast, UCP3-expressing cells share high homology in cysteine residues. The differ-
not exposed to DTT were less sensitive to diamide ence in control of these two closely related proteins
treatment; UCP3 was fully inhibited by 100 ␮M di- may be due to divergences in their physiological func-
amide. This would indicate that UCP3 is partially tions. While UCP1 is used for thermogenesis, UCP3 is
glutathionylated (and thus inhibited) in cells not pre- required to guard cells from ROS and/or to control
treated with DTT. Furthermore, the inhibition of UCP3 ROS-mediated signaling processes in cells (19). In the
by diamide seemed to dissipate over time, since state 4 present study, we investigated the glutathionylated pro-
respiration gradually returned to normal, indicating teome and the redox poise of BAT mitochondria. How-
that glutathionylation/deglutathionylation of UCP3 is a ever, we found no significant DTT-induced changes in the
transient event. Hence, these findings indicate that glutathionylated proteome; however, a trend for de-
DTT pretreatment (inducing deglutathionylation) is creased protein glutathionylation following DTT pretreat-
required to assess the full leakage capacity of UCP3. ment was observed. Intriguingly, though, we found that
As shown in Fig. 6, our findings indicate that redox BAT mitochondria, unlike skeletal muscle, do not contain
state affects the control of UCP3 by modulating the Grx2 (or the levels are too low for detection by immuno-
extent of its glutathionylation. Of note, we also ob- blot). The main enzymes thought to be involved in
served that C2C12 cells transiently transfected with driving reversible glutathionylation in the mitochondria
empty pCMV vector were sensitive to diamide-mediated are Grx1, Grx2, and, to a limited extent, Trx2. While Grx1
inhibition of state 4 respiration, although not to the is positioned in the intermembrane space/cytosol, Grx2 is
same degree as cells expressing UCP3. DTT pretreat- located in the matrix, where most ROS are produced

Figure 6. Model for the control of


UCP3-mediated proton leakage by re-
versible glutathionylation. Partial gluta-
thionylation of UCP3 inhibits proton
leakage. Complete deglutathionylation
of UCP3 results in the activation of the
protein and increases proton leakage.
Full glutathionylation of available cys-
teines inhibits UCP3-mediated proton
leakage.

REDOX MODULATION OF MITOCHONDRIAL METABOLISM 373


(50). The absence of Grx2 in the matrix of BAT mito- 4. Taylor, E. R., Hurrell, F., Shannon, R. J., Lin, T. K., Hirst, J., and
chondria would indicate that this tissue lacks at least one Murphy, M. P. (2003) Reversible glutathionylation of complex I
increases mitochondrial superoxide formation. J. Biol. Chem.
enzyme involved in GSH conjugation reactions in the 278, 19603–19610
matrix. BAT mitochondria did, nevertheless, contain 5. Wang, J., Boja, E. S., Tan, W., Tekle, E., Fales, H. M., English, S.,
abundant Trx2, which can drive glutathionylation reac- Mieyal, J. J., and Chock, P. B. (2001) Reversible glutathionyla-
tion regulates actin polymerization in A431 cells. J. Biol. Chem.
tions. However, Trx2 is not as efficient as Grx in mediat- 276, 47763– 47766
ing glutathionylation reactions (51). BAT mitochondria 6. Naoi, M., Maruyama, W., Yi, H., Inaba, K., Akao, Y., and
were also found to have a far more oxidized GSH pool Shamoto-Nagai, M. (2009) Mitochondria in neurodegenerative
than skeletal muscle mitochondria, consistent with the disorders: regulation of the redox state and death signaling
leading to neuronal death and survival. J. Neural Transm. 116,
propensity of BAT mitochondria to generate much 1371–1381
higher amounts of ROS. These conditions in BAT mito- 7. Garcia, J., Han, D., Sancheti, H., Yap, L. P., Kaplowitz, N., and
chondria could prompt the nonenzymatic glutathionyla- Cadenas, E. (2010) Regulation of mitochondrial glutathione
redox status and protein glutathionylation by respiratory sub-
tion of proteins. Indeed, ROS are well known to drive strates. J. Biol. Chem. 285, 39646 –39654
spontaneous addition of GSH to exposed cysteinyl groups 8. Shelton, M. D., Chock, P. B., and Mieyal, J. J. (2005) Glutare-
by oxidizing sulfur residues to more reactive sulfinic acids doxin: role in reversible protein S-glutathionylation and regula-
(50). GSSG could also glutathionylate the proteome tion of redox signal transduction and protein translocation.
Antioxid. Redox Signal. 7, 348 –366
through a simple thiol disulfide exchange reaction with 9. Mieyal, J. J., Gallogly, M. M., Qanungo, S., Sabens, E. A., and
exposed cysteines (52, 53). Although we do not provide Shelton, M. D. (2008) Molecular mechanisms and clinical
any conclusive evidence regarding UCP1 glutathionyla- implications of reversible protein S-glutathionylation. Antioxid.
tion, we can conclude that the mechanisms governing Redox Signal. 10, 1941–1988
10. Fratelli, M., Demol, H., Puype, M., Casagrande, S., Eberini, I.,
glutathionylation in BAT mitochondria differ greatly Salmona, M., Bonetto, V., Mengozzi, M., Duffieux, F., Miclet, E.,
from those in skeletal muscle. Bachi, A., Vandekerckhove, J., Gianazza, E., and Ghezzi, P.
In summary, this is the first study to profile the effect (2002) Identification by redox proteomics of glutathionylated
proteins in oxidatively stressed human T lymphocytes. Proc. Natl.
of redox environment on skeletal muscle and BAT Acad. Sci. U. S. A. 99, 3505–3510
bioenergetics. Surprisingly, we found that BAT mito- 11. Mailloux, R. J., Seifert, E. L., Bouillaud, F., Aguer, C., Collins, S.,
chondria have a far more oxidizing redox environment, and Harper, M. E. (2011) Glutathionylation acts as a control
and our results indicate that this is required for UCP1 switch for uncoupling proteins UCP2 and UCP3. J. Biol. Chem.
286, 21865–21875
function. These fundamental observations indicate the 12. Chen, Y. R., Chen, C. L., Pfeiffer, D. R., and Zweier, J. L. (2007)
very unique nature of BAT, as our results herein and Mitochondrial complex II in the post-ischemic heart: oxidative
those of others (7, 41) demonstrate that all other injury and the role of protein S-glutathionylation. J. Biol. Chem.
tissues are reliant on a more reductive environment 282, 32640 –32654
13. Hurd, T. R., Requejo, R., Filipovska, A., Brown, S., Prime, T. A.,
for their function. Indeed, the major physiological Robinson, A. J., Fearnley, I. M., and Murphy, M. P. (2008)
function of BAT is nonshivering thermogenesis, and Complex I within oxidatively stressed bovine heart mitochon-
this is unequivocally dependent on UCP1. Because of dria is glutathionylated on Cys-531 and Cys-704 of the 75-kDa
subunit: potential role of CYS residues in decreasing oxidative
their significant contributions to whole body energy damage. J. Biol. Chem. 283, 24801–24815
homeostasis, skeletal muscle and BAT are studied ex- 14. Applegate, M. A., Humphries, K. M., and Szweda, L. I. (2008)
tensively in order to better understand the pathogene- Reversible inhibition of alpha-ketoglutarate dehydrogenase by
sis of obesity and type 2 diabetes mellitus (54, 55). The hydrogen peroxide: glutathionylation and protection of lipoic
acid. Biochemistry 47, 473– 478
observations, herein, provide further insights into the 15. Kil, I. S., and Park, J. W. (2005) Regulation of mitochondrial
unique importance of mitochondrial redox balance on NADP⫹-dependent isocitrate dehydrogenase activity by gluta-
metabolic processes in these two tissues. thionylation. J. Biol. Chem. 280, 10846 –10854
16. Queiroga, C. S., Almeida, A. S., Martel, C., Brenner, C., Alves,
P. M., and Vieira, H. L. (2010) Glutathionylation of adenine
The authors thank Mahmoud Salkhordeh for the isolation nucleotide translocase induced by carbon monoxide prevents
and purification of the primary myoblasts. R.J.M. was funded mitochondrial membrane permeabilization and apoptosis.
by a postdoctoral fellowship provided by the Canadian Insti- J. Biol. Chem. 285, 17077–17088
tutes of Health Research (CIHR). Operating funding for this 17. Brennan, J. P., Wait, R., Begum, S., Bell, J. R., Dunn, M. J., and
research was provided by CIHR (Institute of Nutrition, Me- Eaton, P. (2004) Detection and mapping of widespread inter-
tabolism, and Diabetes; to M.E.H.). molecular protein disulfide formation during cardiac oxidative
stress using proteomics with diagonal electrophoresis. J. Biol.
Chem. 279, 41352– 41360
18. Harper, M. E., Green, K., and Brand, M. D. (2008) The
efficiency of cellular energy transduction and its implications
REFERENCES for obesity. Annu. Rev. Nutr. 28, 13–33
19. Azzu, V., and Brand, M. D. (2010) The on-off switches of the
1. Schafer, F. Q., and Buettner, G. R. (2001) Redox environment mitochondrial uncoupling proteins. Trends Biochem. Sci. 35,
of the cell as viewed through the redox state of the glutathione 298 –307
disulfide/glutathione couple. Free Radic. Biol. Med. 30, 1191– 20. Seale, P., Bjork, B., Yang, W., Kajimura, S., Chin, S., Kuang, S.,
1212 Scime, A., Devarakonda, S., Conroe, H. M., Erdjument-Brom-
2. Hurd, T. R., Costa, N. J., Dahm, C. C., Beer, S. M., Brown, age, H., Tempst, P., Rudnicki, M. A., Beier, D. R., and Spiegel-
S. E., Filipovska, A., and Murphy, M. P. (2005) Glutathiony- man, B. M. (2008) PRDM16 controls a brown fat/skeletal
lation of mitochondrial proteins. Antioxid. Redox Signal. 7, muscle switch. Nature 454, 961–967
999 –1010 21. Schulz, T. J., Huang, T. L., Tran, T. T., Zhang, H., Townsend,
3. Mari, M., Morales, A., Colell, A., Garcia-Ruiz, C., and Fernandez- K. L., Shadrach, J. L., Cerletti, M., McDougall, L. E., Giorgadze,
Checa, J. C. (2009) Mitochondrial glutathione, a key survival N., Tchkonia, T., Schrier, D., Falb, D., Kirkland, J. L., Wagers,
antioxidant. Antioxid. Redox Signal. 11, 2685–2700 A. J., and Tseng, Y. H. (2011) Identification of inducible brown

374 Vol. 26 January 2012 The FASEB Journal 䡠 www.fasebj.org MAILLOUX ET AL.
adipocyte progenitors residing in skeletal muscle and white fat. 40. Haugaard, N., Lee, N. H., Kostrzewa, R., and Haugaard, E. S.
Proc. Natl. Acad. Sci. U. S. A. 108, 143–148 (1969) Effects of a disulfide (Ellman’s reagent) and thiols on
22. Rolfe, D. F., and Brand, M. D. (1996) Contribution of mitochon- oxidative phosphorylation and ion transport by rat liver mito-
drial proton leak to skeletal muscle respiration and to standard chondria. Biochem. Pharmacol. 18, 2385–2391
metabolic rate. Am. J. Physiol. 271, C1380 –C1389 41. Hill, B. G., Higdon, A. N., Dranka, B. P., and Darley-Usmar,
23. Rolfe, D. F., and Brand, M. D. (1996) Proton leak and control of V. M. (2010) Regulation of vascular smooth muscle cell bioen-
oxidative phosphorylation in perfused, resting rat skeletal mus- ergetic function by protein glutathiolation. Biochim. Biophys. Acta
cle. Biochim. Biophys. Acta 1276, 45–50 1797, 285–295
24. Cannon, B., and Nedergaard, J. (2004) Brown adipose tissue: 42. Barja de Quiroga, G., Lopez-Torres, M., Perez-Campo, R.,
function and physiological significance. Physiol. Rev. 84, 277–359 Abelenda, M., Paz Nava, M., and Puerta, M. L. (1991) Effect
25. Enerback, S., Jacobsson, A., Simpson, E. M., Guerra, C., Ya- of cold acclimation on GSH, antioxidant enzymes and lipid
mashita, H., Harper, M. E., and Kozak, L. P. (1997) Mice lacking peroxidation in brown adipose tissue. Biochem. J. 277, 289 –
mitochondrial uncoupling protein are cold-sensitive but not 292
obese. Nature 387, 90 –94 43. Feldmann, H. M., Golozoubova, V., Cannon, B., and Neder-
26. Goubern, M., Yazbeck, J., Chapey, M. F., Diolez, P., and Moreau, gaard, J. (2009) UCP1 ablation induces obesity and abolishes
F. (1990) Variations in energization parameters and proton diet-induced thermogenesis in mice exempt from thermal stress
conductance induced by cold adaptation and essential fatty acid by living at thermoneutrality. Cell Metab. 9, 203–209
deficiency in mitochondria of brown adipose tissue in the rat. 44. Petrovic, V., Buzadzic, B., Korac, A., and Korac, B. (2009)
Biochim. Biophys. Acta 1015, 334 –340 Antioxidative defense and mitochondrial thermogenic response
27. Nicholls, D. G., and Rial, E. (1999) A history of the first in brown adipose tissue. Genes. Nutr. 5, 225–235
uncoupling protein, UCP1. J. Bioenerg. Biomembr. 31, 399 – 406 45. Seifert, E. L., Estey, C., Xuan, J. Y., and Harper, M. E. (2009)
28. Himms-Hagen, J. (1990) Brown adipose tissue thermogenesis: Electron transport chain-dependent and -independent mecha-
interdisciplinary studies. FASEB J. 4, 2890 –2898 nisms of mitochondrial H2O2 emission during long-chain fatty
29. Mailloux, R. J., Dumouchel, T., Aguer, C., Dekemp, R., Beanlands, R., acid oxidation. J. Biol. Chem. 285, 5748 –5758
and Harper, M. E. (2011) Hexokinase II acts through UCP3 to 46. Wu, H., Xing, K., and Lou, M. F. (2010) Glutaredoxin 2 prevents
suppress mitochondrial reactive oxygen species production and main- H2O2-induced cell apoptosis by protecting complex I activity in
tain aerobic respiration. Biochem. J. 437, 21865–21875 the mitochondria. Biochim. Biophys. Acta 1797, 1705–1715
30. Echtay, K. S., Roussel, D., St-Pierre, J., Jekabsons, M. B., Cadenas, S., 47. Mracek, T., Pecinova, A., Vrbacky, M., Drahota, Z., and Houstek,
Stuart, J. A., Harper, J. A., Roebuck, S. J., Morrison, A., Pickering, S., J. (2009) High efficiency of ROS production by glycerophos-
Clapham, J. C., and Brand, M. D. (2002) Superoxide activates mito- phate dehydrogenase in mammalian mitochondria. Arch.
chondrial uncoupling proteins. Nature 415, 96–99 Biochem. Biophys. 481, 30 –36
31. Cypess, A. M., Lehman, S., Williams, G., Tal, I., Rodman, D., 48. Oelkrug, R., Kutschke, M., Meyer, C. W., Heldmaier, G., and
Goldfine, A. B., Kuo, F. C., Palmer, E. L., Tseng, Y. H., Doria, A., Jastroch, M. (2010) Uncoupling protein 1 decreases superoxide
Kolodny, G. M., and Kahn, C. R. (2009) Identification and production in brown adipose tissue mitochondria. J. Biol. Chem.
importance of brown adipose tissue in adult humans. N. Engl. 285, 21961–21968
J. Med. 360, 1509 –1517 49. Silva, J. P., Shabalina, I. G., Dufour, E., Petrovic, N., Backlund,
32. Virtanen, K. A., Lidell, M. E., Orava, J., Heglind, M., Westergren, E. C., Hultenby, K., Wibom, R., Nedergaard, J., Cannon, B., and
R., Niemi, T., Taittonen, M., Laine, J., Savisto, N. J., Enerback, Larsson, N. G. (2005) SOD2 overexpression: enhanced mito-
S., and Nuutila, P. (2009) Functional brown adipose tissue in chondrial tolerance but absence of effect on UCP activity.
healthy adults. N. Engl. J. Med. 360, 1518 –1525 EMBO J. 24, 4061– 4070
33. Nedergaard, J., Bengtsson, T., and Cannon, B. (2010) Three 50. Gallogly, M. M., and Mieyal, J. J. (2007) Mechanisms of revers-
years with adult human brown adipose tissue. Ann. N. Y. Acad. ible protein glutathionylation in redox signaling and oxidative
Sci. 1212, E20 –E36 stress. Curr. Opin. Pharmacol. 7, 381–391
34. Rando, T. A., and Blau, H. M. (1994) Primary mouse myoblast 51. Chrestensen, C. A., Starke, D. W., and Mieyal, J. J. (2000) Acute
purification, characterization, and transplantation for cell-me- cadmium exposure inactivates thioltransferase (glutaredoxin),
diated gene therapy. J. Cell Biol. 125, 1275–1287 inhibits intracellular reduction of protein-glutathionyl-mixed
35. Mailloux, R. J., and Harper, M. E. (2010) Glucose regulates disulfides, and initiates apoptosis. J. Biol. Chem. 275, 26556 –
enzymatic sources of mitochondrial NADPH in skeletal muscle 26565
cells; a novel role for glucose-6-phosphate dehydrogenase. 52. Gilbert, H. F. (1995) Thiol/disulfide exchange equilibria and
FASEB J. 24, 2495–2506 disulfide bond stability. Methods Enzymol. 251, 8 –28
36. Caprioli, J., Munemasa, Y., Kwong, J. M., and Piri, N. (2009) 53. Shaked, Z., Szajewski, R. P., and Whitesides, G. M. (1980) Rates
Overexpression of thioredoxins 1 and 2 increases retinal gan- of thiol-disulfide interchange reactions involving proteins and
glion cell survival after pharmacologically induced oxidative kinetic measurements of thiol pKa values. Biochemistry 19, 4156 –
stress, optic nerve transection, and in experimental glaucoma. 4166
Trans. Am. Ophthalmol. Soc. 107, 161–165 54. Sastre, J., Pallardo, F. V., Llopis, J., Furukawa, T., Vina, J. R., and
37. Parker, N., Crichton, P. G., Vidal-Puig, A. J., and Brand, M. D. Vina, J. (1989) Glutathione depletion by hyperphagia-induced
(2009) Uncoupling protein-1 (UCP1) contributes to the basal obesity. Life Sci. 45, 183–187
proton conductance of brown adipose tissue mitochondria. 55. Robertson, R. P., Harmon, J., Tran, P. O., and Poitout, V.
J. Bioenerg. Biomembr. 41, 335–342 (2004) Beta-cell glucose toxicity, lipotoxicity, and chronic
38. Cleland, W. W. (1964) Dithiothreitol, a new protective reagent oxidative stress in type 2 diabetes. Diabetes 53(Suppl. 1),
for Sh groups. Biochemistry 3, 480 – 482 S119 –S124
39. Nulton-Persson, A. C., and Szweda, L. I. (2001) Modulation of
mitochondrial function by hydrogen peroxide. J. Biol. Chem. Received for publication June 9, 2011.
276, 23357–23361 Accepted for publication September 1, 2011.

REDOX MODULATION OF MITOCHONDRIAL METABOLISM 375

You might also like