MICU1 Is an Essential Gatekeeper for MCU-Mediated Mitochondrial Ca2+ Uptake that Regulates Cell...

26
MICU1 Is an Essential Gatekeeper for MCU-Mediated Mitochondrial Ca 2+ Uptake that Regulates Cell Survival Karthik Mallilankaraman, 1,2,8 Patrick Doonan, 1,2,8 Ce ´ sar Ca ´ rdenas, 3,6,8 Harish C. Chandramoorthy, 1,2,8 Marioly Mu ¨ ller, 3 Russell Miller, 4 Nicholas E. Hoffman, 1,2 Rajesh Kumar Gandhirajan, 1,2 Jordi Molgo ´, 7 Morris J. Birnbaum, 4 Brad S. Rothberg, 1 Don-On Daniel Mak, 3 J. Kevin Foskett, 3,5, * and Muniswamy Madesh 1,2, * 1 Department of Biochemistry 2 Center for Translational Medicine Temple University, Philadelphia, PA 19140, USA 3 Department of Physiology 4 Institute for Diabetes, Obesity, and Metabolism 5 Department of Cell and Developmental Biology University of Pennsylvania, Philadelphia, PA 19104, USA 6 Anatomy and Developmental Biology Program, Institute of Biomedical Sciences, University of Chile, Independencia 7024, Santiago 8380453, Chile 7 CNRS, Institut de Neurobiologie Alfred Fessard, FRC2118, Laboratoire de Neurobiologie et De ´ veloppement, UPR 3294, 91198 Gif-sur-Yvette Cedex, France 8 These authors contributed equally to this work *Correspondence: [email protected] (J.K.F.), [email protected] (M.M.) http://dx.doi.org/10.1016/j.cell.2012.10.011 SUMMARY Mitochondrial Ca 2+ (Ca 2+ m ) uptake is mediated by an inner membrane Ca 2+ channel called the uniporter. Ca 2+ uptake is driven by the considerable voltage present across the inner membrane (DJ m ) gener- ated by proton pumping by the respiratory chain. Mitochondrial matrix Ca 2+ concentration is main- tained five to six orders of magnitude lower than its equilibrium level, but the molecular mechanisms for how this is achieved are not clear. Here, we demon- strate that the mitochondrial protein MICU1 is required to preserve normal [Ca 2+ ] m under basal conditions. In its absence, mitochondria become constitutively loaded with Ca 2+ , triggering excessive reactive oxygen species generation and sensitivity to apoptotic stress. MICU1 interacts with the uniporter pore-forming subunit MCU and sets a Ca 2+ threshold for Ca 2+ m uptake without affecting the kinetic proper- ties of MCU-mediated Ca 2+ uptake. Thus, MICU1 is a gatekeeper of MCU-mediated Ca 2+ m uptake that is essential to prevent [Ca 2+ ] m overload and associ- ated stress. INTRODUCTION Mitochondrial Ca 2+ homeostasis plays important roles in cellular physiology. Ca 2+ flux across the inner mitochondrial membrane (IMM) regulates cell bioenergetics, cytoplasmic Ca 2+ ([Ca 2+ ] i ) signals, and activation of cell death pathways (Balaban, 2009; Denton and McCormack, 1980; Duchen et al., 2008; Gunter and Gunter, 1994; Hajno ´ czky et al., 1995; Hansford, 1994; Her- rington et al., 1996; Lemasters et al., 2009; McCormack et al., 1990; Orrenius et al., 2003; Szalai et al., 1999). Mitochondrial Ca 2+ (Ca 2+ m ) uptake has been studied for more than five decades, with crucial insights into the underlying mechanisms enabled by development of the chemiosmotic hypothesis and appreciation of the considerable voltage across the IMM (DJ m ) generated by proton pumping in the respiratory chain (Carafoli, 1987; Drago et al., 2011; Nicholls, 2005; O’Rourke, 2007; Rottenberg and Scarpa, 1974). Ca 2+ uptake is an electro- genic process driven by DJ m and mediated by a Ca 2+ -selective ion channel (MiCa; Kirichok et al., 2004) called the uniporter (Bernardi, 1999; Igbavboa and Pfeiffer, 1988; O’Rourke, 2007; Santo-Domingo and Demaurex, 2010). Properties of the uni- porter have been derived primarily from studies of isolated mito- chondria, where it was generally found to have a low apparent Ca 2+ affinity (10–70 mM) with variable cooperativity (Bragadin et al., 1979; Gunter et al., 1994). Agonist-induced [Ca 2+ ] i signals can be rapidly transduced to the mitochondrial matrix despite this apparent low affinity because mitochondria can exist in close apposition to sites of Ca 2+ release where local [Ca 2+ ] i can be higher than in the bulk cytoplasm (Carafoli and Lehninger, 1971; Collins et al., 2001; Filippin et al., 2003; Nicholls, 2008; Palmer et al., 2006; Rizzuto et al., 1998, 2004). Nevertheless, higher-affinity mitochondrial Ca 2+ uptake has been observed in many studies (Santo-Domingo and Demaurex, 2010; Spa ¨t et al., 2008). Furthermore, patch-clamp electrophysiology sug- gests that the uniporter pore has high Ca 2+ affinity (dissociation constant <2 nM) that enables it to have high Ca 2+ selectivity (Kirichok et al., 2004). In addition, the open probability of the uniporter channel is voltage dependent, reaching nearly unity 630 Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc.

Transcript of MICU1 Is an Essential Gatekeeper for MCU-Mediated Mitochondrial Ca2+ Uptake that Regulates Cell...

MICU1 Is an Essential Gatekeeperfor MCU-Mediated Mitochondrial Ca2+

Uptake that Regulates Cell SurvivalKarthik Mallilankaraman,1,2,8 Patrick Doonan,1,2,8 Cesar Cardenas,3,6,8 Harish C. Chandramoorthy,1,2,8 Marioly Muller,3

Russell Miller,4 Nicholas E. Hoffman,1,2 Rajesh Kumar Gandhirajan,1,2 Jordi Molgo,7 Morris J. Birnbaum,4

Brad S. Rothberg,1 Don-On Daniel Mak,3 J. Kevin Foskett,3,5,* and Muniswamy Madesh1,2,*1Department of Biochemistry2Center for Translational MedicineTemple University, Philadelphia, PA 19140, USA3Department of Physiology4Institute for Diabetes, Obesity, and Metabolism5Department of Cell and Developmental Biology

University of Pennsylvania, Philadelphia, PA 19104, USA6Anatomy and Developmental Biology Program, Institute of Biomedical Sciences, University of Chile, Independencia 7024,

Santiago 8380453, Chile7CNRS, Institut de Neurobiologie Alfred Fessard, FRC2118, Laboratoire de Neurobiologie et Developpement, UPR 3294,

91198 Gif-sur-Yvette Cedex, France8These authors contributed equally to this work

*Correspondence: [email protected] (J.K.F.), [email protected] (M.M.)http://dx.doi.org/10.1016/j.cell.2012.10.011

SUMMARY

Mitochondrial Ca2+ (Ca2+m) uptake is mediated by aninner membrane Ca2+ channel called the uniporter.Ca2+ uptake is driven by the considerable voltagepresent across the inner membrane (DJm) gener-ated by proton pumping by the respiratory chain.Mitochondrial matrix Ca2+ concentration is main-tained five to six orders of magnitude lower than itsequilibrium level, but the molecular mechanisms forhow this is achieved are not clear. Here, we demon-strate that the mitochondrial protein MICU1 isrequired to preserve normal [Ca2+]m under basalconditions. In its absence, mitochondria becomeconstitutively loaded with Ca2+, triggering excessivereactive oxygen species generation and sensitivity toapoptotic stress. MICU1 interacts with the uniporterpore-forming subunit MCU and sets a Ca2+ thresholdfor Ca2+m uptake without affecting the kinetic proper-ties of MCU-mediated Ca2+ uptake. Thus, MICU1 isa gatekeeper of MCU-mediated Ca2+m uptake thatis essential to prevent [Ca2+]m overload and associ-ated stress.

INTRODUCTION

Mitochondrial Ca2+ homeostasis plays important roles in cellular

physiology. Ca2+ flux across the inner mitochondrial membrane

(IMM) regulates cell bioenergetics, cytoplasmic Ca2+ ([Ca2+]i)

signals, and activation of cell death pathways (Balaban, 2009;

630 Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc.

Denton and McCormack, 1980; Duchen et al., 2008; Gunter

and Gunter, 1994; Hajnoczky et al., 1995; Hansford, 1994; Her-

rington et al., 1996; Lemasters et al., 2009; McCormack et al.,

1990; Orrenius et al., 2003; Szalai et al., 1999). Mitochondrial

Ca2+ (Ca2+m) uptake has been studied for more than five

decades, with crucial insights into the underlying mechanisms

enabled by development of the chemiosmotic hypothesis and

appreciation of the considerable voltage across the IMM

(DJm) generated by proton pumping in the respiratory chain

(Carafoli, 1987; Drago et al., 2011; Nicholls, 2005; O’Rourke,

2007; Rottenberg and Scarpa, 1974). Ca2+ uptake is an electro-

genic process driven by DJm and mediated by a Ca2+-selective

ion channel (MiCa; Kirichok et al., 2004) called the uniporter

(Bernardi, 1999; Igbavboa and Pfeiffer, 1988; O’Rourke, 2007;

Santo-Domingo and Demaurex, 2010). Properties of the uni-

porter have been derived primarily from studies of isolated mito-

chondria, where it was generally found to have a low apparent

Ca2+ affinity (10–70 mM) with variable cooperativity (Bragadin

et al., 1979; Gunter et al., 1994). Agonist-induced [Ca2+]i signals

can be rapidly transduced to the mitochondrial matrix despite

this apparent low affinity because mitochondria can exist in

close apposition to sites of Ca2+ release where local [Ca2+]ican be higher than in the bulk cytoplasm (Carafoli and Lehninger,

1971; Collins et al., 2001; Filippin et al., 2003; Nicholls, 2008;

Palmer et al., 2006; Rizzuto et al., 1998, 2004). Nevertheless,

higher-affinity mitochondrial Ca2+ uptake has been observed in

many studies (Santo-Domingo and Demaurex, 2010; Spat

et al., 2008). Furthermore, patch-clamp electrophysiology sug-

gests that the uniporter pore has high Ca2+ affinity (dissociation

constant <2 nM) that enables it to have high Ca2+ selectivity

(Kirichok et al., 2004). In addition, the open probability of the

uniporter channel is voltage dependent, reaching nearly unity

at normalDJm (��180mV) (Kirichok et al., 2004). Thus, whereas

rapid and substantial Ca2+m uptake can take place in regions

of high [Ca2+]i microdomains, the high open probability and

Ca2+ affinity of the uniporter pore suggest that the large thermo-

dynamic driving force for Ca2+ uptake would result in Ca2+moverload in the absence of regulatory mechanisms to limit the

activity of the channel. However, the identity of such mecha-

nisms remains elusive.

Until recently, the molecular identity of the uniporter was

unknown. MICU1 was identified as a protein that localized to

the IMM and was suggested to be required for uniporter-

mediated Ca2+ uptake (Perocchi et al., 2010). Subsequently,

MCU was identified as the likely ion-conducting pore of the uni-

porter (Baughman et al., 2011; De Stefani et al., 2011). MICU1

and MCU biochemically interact, and their expression patterns

are tightly coupled across tissues and species (Baughman

et al., 2011). Nevertheless, the functional relationship between

these two uniporter components is unknown.

Here, we report that loss of MICU1 leads to constitutive Ca2+maccumulation through MCU. MICU1 is not required for MCU-

mediated Ca2+m uptake. Instead, MICU1 is a gatekeeper for

MCU-mediated Ca2+ uptake, establishing a threshold (the set

point) that prevents Ca2+ uptake in low [Ca2+]i (<3 mM) but

does not confer low apparent Ca2+ affinity or cooperativity of

MCU-mediated Ca2+ uptake at higher [Ca2+]i. MICU1 regulation

of MCU requires each of its Ca2+-binding EF hands, suggesting

that they provide the high-affinity [Ca2+] sensing mechanism that

enables MICU1 to exert its regulation. MICU1 senses matrix

[Ca2+] because it inhibits MCU-mediated Ca2+ uptake only

when [Ca2+]m is low. These findings reveal a previously unknown

role ofMICU1 as a gatekeeper to limit MCU-mediatedCa2+ influx

to prevent Ca2+m overload and its associated stress under

resting conditions.

RESULTS

Knockdown of MICU1 Causes Basal Ca2+mAccumulationHeLa cells were transfected with MICU1 small interfering RNA

(siRNA) (Table S1 available online) to silence its expression or

with a nontargeting scrambled siRNA as control (Figure S1A).

Strikingly, [Ca2+]m was constitutively elevated under resting

conditions in MICU1 knockdown (KD) cells (Figures S1B and

S1C). It was previously reported that MICU1 expression was

required for uniporter-mediated Ca2+ uptake (Perocchi et al.,

2010). Nevertheless, MICU1 KD did not affect the ability of mito-

chondria to accumulate Ca2+ in response to an agonist-induced

[Ca2+]i rise (Figures S1D, S1E, and S1G) nor did it affect the

magnitude of a histamine-induced [Ca2+]i rise (Figure S1F) or

DJm (Figure S1H).

MCU was identified as the pore-forming subunit of the mito-

chondrial uniporter (Baughman et al., 2011; De Stefani et al.,

2011). To confirm that MICU1 and MCU KD are not functionally

equivalent, we compared effects of knocking down expression

of either protein in independent assays. We previously demon-

strated that inhibition of basal inositol trisphosphate receptor

(InsP3R)-mediated Ca2+ release resulted in insufficient Ca2+

transfer from endoplasmic reticulum (ER) to mitochondria to

support optimal bioenergetics. As a consequence, cellular

[AMP]:[ATP] ratio is elevated, and autophagy is activated as

a prosurvival mechanism (Cardenas et al., 2010). If MCU and

MICU1 are each required for Ca2+m uptake, we speculated that

KD of either protein would elevate [AMP]:[ATP] and induce auto-

phagy. Nevertheless, whereas [AMP]:[ATP] was increased in

MCU KD cells, it was unchanged in MICU1 KD cells (Figure S1I).

Elevated autophagy was observed (Cardenas et al., 2010) in

MCU KD, but it was unchanged in MICU1 KD cells (Figures

S1J and S1K). Whereas the effects of MCU KD are as expected

if MCU plays an essential role in Ca2+m uptake, those of MICU1

are not. These results strengthen the conclusion that MICU1 is

not essential for Ca2+m uptake and support the notion that it

plays a distinct role.

We generated stable clones by using lentiviral short hairpin

RNAs (shRNAs) targeting different regions of the MICU1 gene

(Table S2 and Figure S1L). Two, shHe#B4 and shHe#B6, had

80.7% and 82.6% messenger RNA (mRNA) KD, respectively

(Figure S1L). We used clone shHe#B6 for subsequent experi-

ments. As in cells treated withMICU1 siRNA, [Ca2+]m was consti-

tutively elevated in stable MICU1 KD cells (Figures 1A and 1B),

with elevated basal [Ca2+]m correlated with the degree of KD

(Figures 1B and S1L). Also, as in siRNA-treated cells, hista-

mine-induced Ca2+m uptake (Figures 1C and 1E) and rise of

[Ca2+]i (Figure 1D) were unaltered in MICU1 KD cells. DJm (Fig-

ure 1F) and [AMP]:[ATP] (Figure S1M) were not affected either,

and autophagy was not activated (Figure S1N). Lack of auto-

phagy induction in MICU1 KD cells was not due to intrinsic

defects in autophagy because Xestospongin B (XeB), a specific

InsP3R inhibitor, induced autophagy similarly in control

(shHe#B8) and KD cells (Figure S1N).We extended these studies

to human endothelial cells (EC) with MICU1 stably knocked

down. Similar to HeLa cells, [Ca2+]m was constitutively elevated

(Figures 1G and 1H), whereas histamine-induced Ca2+m uptake

(Figures 1I and 1K) and rise of [Ca2+]i (Figure 1J) were unaffected,

and DJm was normal (Figure 1L).

Defective Ca2+m uptake can result in enhanced agonist-

induced [Ca2+]i responses due to diminished cytoplasmic

buffering (Hoth et al., 2000; SimpsonandRussell, 1998). In agree-

ment, [Ca2+]i was elevated for a prolonged period in response to

histamine in MCU KD cells, whereas a transient response was

observed in MICU1 KD cells (Figure S1O). The Ca2+-dependent

transcription factor NFAT was strongly activated in MCU KD

cells, which is consistent with the prolonged [Ca2+]i signal,

whereas only weak activation was observed in MICU1 KD

cells (Figure S1P). Further, nuclear NFATC3-GFP-positive cells

showed increased nuclear translocation in MCU-silenced cells

compared with MICU1 KD cells (Figures S1Q and S1R).

Together, these results suggest that MICU1 is not required

for uniporter-dependent Ca2+ uptake, nor does it function

similarly to MCU. Instead, they suggest that MICU1 may play

a role in constitutively suppressing Ca2+m uptake under resting

conditions.

Constitutive Ca2+m Accumulation in MICU1 KD CellsOccurs through MCU-Mediated Ca2+ UptakeWe confirmed (Perocchi et al., 2010) that MICU1 biochemically

interacts with MCU (Figure S2A). MICU1 KD did not alter MCU

Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc. 631

Figure 1. MICU1 Silencing Promotes Basal Ca2+m Accumulation

(A) Representative confocal images of rhod-2 fluorescence in HeLa Neg shRNA (control) and MICU1 shRNA clone KD B6 cells.

(B) Quantification of basal rhod-2 fluorescence in HeLa Neg shRNA (control) and MICU1 shRNA KD clones. Mean ±SEM; ns, not significant; **p < 0.01, n = 3.

(C) Kinetics of [Ca2+]m in HeLa Neg shRNA (control) and MICU1 shRNA KD clone B6 cells in response to histamine (100 mM). Solid lines, mean; shaded

regions, ±SEM; n = 3.

(D and E) Quantification of cytosolic (purple) andmitochondrial (green) [Ca2+] peak amplitudes in HeLa Neg shRNA (control) andMICU1 shRNA KD clone B6 cells

after stimulation with 100 mM histamine. Mean ±SEM; n = 3.

632 Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc.

Figure 2. MICU1Knockdown-InducedBasal

Ca2+m Accumulation Is Mediated by MCU-

Dependent Ca2+ Uptake

(A) Representative confocal images of basal rhod-

2 fluorescence in HeLa cells stably expressing

Neg shRNA, MICU1 KD shRNA, and MICU1

Rescue andWT cells treated with MCU siRNA and

MICU1 KD cells treated with MCU siRNA.

(B) Quantification of basal rhod-2 fluorescence of

cells in (A). Mean ±SEM; ***p < 0.001, n = 3.

(C) Responses of [Ca2+]m in cells in (A) to

100 mM histamine. Solid lines, mean; shaded

regions, ±SEM; n = 3.

(D) Quantification of basal rhod-2 fluorescence in

ECs stably expressing Neg shRNA, MICU1 KD

shRNA, and MICU1 Rescue and WT cells treated

with MCU siRNA and MICU1 KD cells treated with

MCU siRNA. Mean ±SEM; ***p < 0.001, n = 3.

(E) Responses of [Ca2+]m in ECs in (D) to

100 mM histamine. Solid lines, mean; shaded

regions, ±SEM; n = 3.

(F and G) Representative traces and quantification

of basal bath [Ca2+] (FuraFF fluorescence) in

Neg shRNA (black) and MICU1KD (red) per-

meabilized HEK293 cells before and after Ru360.

Mean ±SEM; ns, not different; **p < 0.01, n = 3.

See also Figure S2.

expression (Figure S2B). We therefore considered that MICU1

may function to inhibit MCU-mediated Ca2+ uptake and that

the observed elevated basal [Ca2+]m in MICU1 ablated cells

occurs through uncontrolled MCU-dependent Ca2+ uptake. To

(F) Quantification of TMRE fluorescence to assess DJm in HeLa Neg shRNA (control) and MICU1 shRNA K

(G) Representative confocal images of rhod-2 fluorescence in Neg shRNA (control) and MICU1 shRNA clon

(H) Quantification of basal rhod-2 fluorescence in Neg shRNA (control) and MICU1 KD ECs. Mean ±SEM; **

(I) Kinetics of [Ca2+]m in Neg shRNA (control) and MICU1 KD ECs in response to histamine (100 mM). Solid li

(J and K) Quantification of cytosolic (purple) and mitochondrial (green) [Ca2+] peak amplitudes in Neg shRNA

histamine (100 mM). Mean ±SEM; n = 3.

(L) Quantification of TMRE fluorescence to assess DJm in Neg shRNA (control) and MICU1 shRNA KD B6 E

Cell 151, 630–644,

test this, we silenced MCU expression

by >85% (Figure S2C) in control negative

shRNA and MICU1 KD HeLa and

ECs (Figure S2D). Histamine-induced

Ca2+m uptake was effectively abrogated,

demonstrating functional efficacy of

MCU KD (Figures 2C and 2E). Impor-

tantly, silencing MCU abolished elevated

basal [Ca2+]m in MICU1 KD cells (Fig-

ures 2A–2E and S2H). This indicates

that Ca2+m accumulation induced by

loss of MICU1 is mediated by MCU-

dependent Ca2+ uptake. To further verify

this, we assessed Ca2+m uptake in digi-

tonin-permeabilized stable MICU1 KD

HEK293 cells (Figure S2E) bathed in intra-

cellular-like medium (ICM) containing

thapsigargin (Tg) to prevent ER Ca2+

uptake and FuraFF to monitor [Ca2+] in

the medium. Following exposure of control cells to Tg, Ca2+

leak from the ER caused ICM [Ca2+] to increase progressively

over �5 min to �1.8 mM (Figure 2F). Subsequent exposure to

the MCU inhibitor Ru360 had little effect (Figure 2F), indicating

D clones. Mean ±SEM; n = 3. See also Figure S1.

e KD B6 ECs.

p < 0.01, n = 3.

nes, mean; shaded regions, ±SEM; n = 3.

(control) and MICU1 KD ECs after stimulation with

Cs. Mean ±SEM; n = 3.

October 26, 2012 ª2012 Elsevier Inc. 633

634 Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc.

that mitochondria had not taken up Ca2+. In contrast, ICM [Ca2+]

did not increase during ER Ca2+ leakage from MICU1 KD cells

(Figure 2F). Subsequent exposure to Ru360 caused a rapid

rise of ICM [Ca2+] to levels observed in control cells (Figures 2F

and 2G). These results indicate that Ru360-sensitive MCU-

dependent Ca2+ uptake buffered Ca2+ released from the ER in

the MICU1 KD cells. Further, the observation that Ca2+ efflux

was active in MICU1 KD cells, evident upon addition of Ru360,

indicates that MICU1 KD does not alter Ca2+m efflux (Figures

2F and S5). Importantly, this buffering occurred in MICU1 KD

cells at [Ca2+] under which MCU-dependent Ca2+ uptake is nor-

mally not active. Accordingly, these results suggest that MICU1

normally acts as a brake to prevent MCU-dependent Ca2+muptake at low [Ca2+]i.

To test this further, we examined phosphorylation of the mito-

chondrial protein pyruvate dehydrogenase (PDH). PDH phos-

phorylation by pyruvate kinase suppresses its activity, whereas

dephosphorylation by Ca2+-dependent pyruvate dehydroge-

nase phosphatase (PDP) enhances it (Cardenas et al., 2010). If

MICU1 normally acts to prevent MCU-mediated Ca2+ uptake

under basal conditions, we speculated that PDH would be

constitutively hypophosphorylated in MICU1 KD cells by

enhanced Ca2+-activated PDP activity. In agreement, PDH was

hypophosphorylated inMICU1KDcells (FigureS2F). As a control,

XeB, which blocks ER Ca2+ release (Cardenas et al., 2010),

increased PDH phosphorylation in the MICU1 KD cells to levels

comparable to those in the control cells (Figure S2G).

MICU1 Inhibits MCU-Mediated Ca2+m Influx

in Low [Ca2+]Our results suggest that MICU1 may regulate the apparent Ca2+

affinity or threshold of MCU-dependent Ca2+ uptake. To

examine this in more detail, we used the permeabilized cell

protocol and simultaneously monitored the rate of Ca2+m uptake

and DJm in response to boluses of Ca2+ added to ICM. Cells

(�6 million) were permeabilized in ICM containing Tg. After

450 s, a single bolus of Ca2+ was added, and the Fura FF (Fig-

ure 3A) and JC-1 (Figure 3B) signals were monitored for 300 s

before an uncoupler (CCCP) was added to depolarize the IMM

and to release accumulated Ca2+ (Figures 3A and 3F). Ca2+

uptake in these conditions, monitored as a decrease in ICM

[Ca2+], is completely dependent on MCU (Figure 3G). In

response to a 0.5 mM Ca2+ bolus, ICM [Ca2+] rapidly rose to

a stable plateau in control cells, indicating lack of Ca2+m uptake

at this [Ca2+], as expected. In contrast, ICM [Ca2+] decreased to

Figure 3. MICU1 Controls MCU-Mediated Ca2+m Uptake in Low [Ca2+]

(A) Permeabilized stable Neg shRNA (black) andMICU1 KD (red) HEK293 cells we

lines, mean; shaded regions, ±SEM; n = 3.

(B) Representative traces indicate DJm (JC-1 ratio) in permeabilized Neg shRNA (

(C) Change in bath [Ca2+] due to Ca2+m uptake in response to various bath [C

exponential fit of bath [Ca2+] kinetics following Ca2+ pulses. Inset shows respons

(D) Rate of [Ca2+]m as function of bath [Ca2+] derived from single exponential fits

data. Reduced uptake rate observed at [Ca2+] > 20 mM is due to DJm depolariza

(E) Kinetic parameters derived from Hill equation fits of data in (C).

(F) Total Ca2+m uptake after each Ca2+ pulse was determined in neg shRNA (black

obtained from cells with no added Ca2+ pulse. Solid lines, mean; shaded regions

(G) Ru360 blocks basal Ca2+m uptake in MICU1 KD cells as well as uptake in KD

experiments. See also Figure S3.

nearly the prepulse level in MICU1 KD cells (Figure 3A). Subse-

quent CCCP addition released significantly more Ca2+ from

MICU1 KD mitochondria, indicating that the decrease in ICM

[Ca2+] following the initial pulse was due to Ca2+m uptake (Fig-

ure 3F, see inset). Similarly, in response to a 1 mM Ca2+ pulse,

Ca2+ uptake in control cells was absent, whereas uptake in

MICU1 KD cells was substantial (Figure 3A). The difference in

Ca2+m uptake between control and MICU1 KD cells was not

due to a greater driving force for Ca2+ uptake in MICU1 KD cells

because DJm both before and during the Ca2+ pulses were

similar in control and KD cells (Figure 3B), and MICU1 KD mito-

chondria have higher basal [Ca2+]m (Figures 1B, 1H, 2B, and 2D).

In addition, it cannot be accounted for by different expression

levels of MCU (Figure S2B). In response to a 2 mM Ca2+ pulse,

control mitochondria buffered the [Ca2+] but to a level (the set

point; Nicholls, 2005) that was significantly higher than that

achieved by MICU1 KD mitochondria (Figure 3A). At higher

concentrations (5–50 mM), both control and MICU1 KD cells ex-

hibited robust Ca2+m uptake (Figure 3A) with similar transient

DJm depolarizations at 20 and 50 mMpulses (Figure 3B). Recon-

stitution of MICU1 (Figures S2E and S2G) in MICU1 KD HeLa

cells prevented elevated basal [Ca2+]m and reverted Ca2+muptake (Figures 2A–2E). Similarly, rescue of MICU1 in permeabi-

lized HEK293 cells prevented elevated basal [Ca2+]m and re-

verted Ca2+m uptake in response to low [Ca2+] bolus additions

(1–5 mM) to control levels (Figures S3A–S3D).

The ICM [Ca2+] decay kinetics were all well fitted assuming

a single exponential process. The fits were used to calculate

the change in ICM [Ca2+] due to Ca2+m uptake following the

pulse, as well as the rate of Ca2+m uptake. Mitochondria in

MICU1 KD cells took up more Ca2+ than control cells in the low

[Ca2+] regime, from 0.5 to �5 mM (Figure 3C, inset). In contrast,

MICU1KDwaswithout consequence at higher [Ca2+] (Figure 3C).

Surprisingly, Hill equation fits of the uptake rates yielded similar

kinetic parameters in control and MICU1 KD cells (Figure 3D).

The Kd, or apparent Ca2+ affinity of Ca2+m uptake, was �11 mM

in control and MICU1 KD cells, with Hill coefficients of 3.2 and

3.4, respectively (Figure 3E). These results suggest that MICU1

plays an important role in limiting Ca2+m uptake in the low

[Ca2+] regime without affecting the overall kinetic behavior of

MCU-mediated Ca2+ uptake. The total amount of Ca2+ buffered

by control and MICU1 KD mitochondria in response to the Ca2+

pulses was evaluated bymeasuring the amount of Ca2+ released

by CCCP (Figure 3F). Mitochondria with MICU1 knocked down

took up more Ca2+ than control mitochondria in the low [Ca2+]

i

re pulsed with different [Ca2+] as indicated. Traces show bath [Ca2+] (mM). Solid

black) andMICU1 KD (red) HEK293 cells in response to Ca2+ pulses and CCCP.

a2+] in control (black) and MICU1 KD (red) cells. Uptake derived from single

es up to 5 mM bath [Ca2+].

of bath [Ca2+] responses following Ca2+ pulses. Solid line is Hill equation fit of

tion and was not used in fitting.

) and MICU1 KD (red) cells by recording Ca2+ released by CCCP. Inset, traces

, mean ±SEM; n = 3. **p < 0.01 and *p < 0.05, respectively.

and control cells in response to Ca2+ pulses. Representative traces from three

Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc. 635

Figure 4. MICU1 Silencing Lowers the Set Point for Basal MCU-Mediated Ca2+ Uptake(A) Overlay of bath [Ca2+] responses to 0, 0.5, 1, 2, and 5 mM Ca2+ pulses in permeabilized Neg shRNA (black; left) and MICU1 KD cells (red; right). Solid lines,

means; shaded areas, mean ±SEM; n = 3 at each [Ca2+].

636 Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc.

regime, up to 5 mMCa2+. Notably, however, the amount taken up

versus [Ca2+] curve for the two groupswas parallel over the entire

range of [Ca2+] (Figure 3F), with the offset between them ac-

counted for by Ca2+ accumulated by MICU1 KD mitochondria

before the pulses (the offset in Figure 3A). These results demon-

strate that MICU1 KD does not change Ca2+m buffering capacity.

Taken together, these data suggest that MICU1 acts as a ‘‘gate-

keeper’’ that sets a [Ca2+] threshold to prevent Ca2+m uptake in

low [Ca2+]i, but it does not play a role in conferring thewell-estab-

lished cooperativity of Ca2+m uptake.

MICU1 Is a Gatekeeper of MCU-Mediated Ca2+m InfluxOur results suggest that MICU1 regulates MCU-dependent

Ca2+m uptake in low [Ca2+]. First, mitochondria in MICU1 KD

cells have constitutively elevated [Ca2+]m (Figures 1A–1C, 1G–

1I, and 2A–2E). This suggests that, in the absence of MICU1,

MCU-mediated Ca2+ uptake is active at [Ca2+]i that exists in

resting cells. Second, mitochondria in permeabilized MICU1

KD cells take up Ca2+ from boluses added to the ICM at concen-

trations (up to �3 mM) where control mitochondria do not

(Figures 3A and 3C). Third, in response to slow addition of

Ca2+ to the medium (in permeabilized cells caused by ER Ca2+

leak), ICM Ca2+ is unbuffered and consequently rises in control

cells but is buffered by Ca2+m uptake in MICU1 KD cells (Figures

3A and 4A). During the 450 s following Tg-induced ER leak, ICM

[Ca2+] rose by nearly 1 mM to 1.8 ± 0.06 mM in control cells but

rose by only �160 nM to 0.9 ± 0.02 mM in MICU1 KD cells

(Figures 4B and 4C), which is similar to the set point value re-

ported previously (Nicholls, 1978). MICU1 KD mitochondria

also reduced ICM [Ca2+] to lower levels than controls following

acute Ca2+ pulses. Following Ca2+ pulses in the 0.5–5 mM

regime, MICU1 KD mitochondria reduced ICM [Ca2+] during

the subsequent 300 s to 1.3 mM, whereas control mitochondria

reduced it to 2.7 mM (Figures 4D and 4E), similar to the �3 mM

threshold determined (Figures 3C and S3E). In response to

higher [Ca2+] pulses (Figure 4F), mitochondria in both cells

rapidly buffered Ca2+, but MICU1 KD mitochondria reduced

ICM [Ca2+] to �1.9 mM, whereas control mitochondria buffered

it less effectively to a higher level, again to �3 mM (Figures 4G

and 4H).

Elevated [Ca2+]m stimulates oxidative phosphorylation (OX-

PHOS) by activating mitochondrial enzymes (Denton and

McCormack, 1980; Hajnoczky et al., 1995; Robb-Gaspers

et al., 1998). InsP3-linked agonists trigger large increases in

[Ca2+]m because close appositions between mitochondria and

ER enable released Ca2+ to reach levels sufficient for rapid and

(B) Overlay of bath [Ca2+] kinetics during 450 s following plasmamembrane perme

cells. Solid lines, means; shaded areas, mean ±SEM; n = 3 at each [Ca2+].

(C) Scatterplot of bath [Ca2+] at 450 s from experiments shown in (B). Mean ±SE

(D) Overlay of bath [Ca2+] kinetics during 300 s following 0.5, 1, 2, and 5 mM ba

(red; right). Solid lines, means; shaded areas, mean ±SEM; n = 3 at each [Ca2+].

(E) Scatterplot of bath [Ca2+] at 300 s from experiments shown in (D). Mean ±SE

(F) Overlay of bath [Ca2+] responses to 10, 20, and 50 mMCa2+ pulses in permeabil

shaded areas, mean ±SEM; n = 3 at each [Ca2+].

(G) Overlay of bath [Ca2+] kinetics during 300 s following 10, 20, and 50 mM bath C

right). Solid lines, means; shaded areas, mean ±SEM; n = 3 at each [Ca2+].

(H) Scatterplot of bath [Ca2+] at 300 s from experiments in (G). ***p < 0.001. See

substantial permeation through MCU. If MICU1 acts as a gate-

keeper to set the [Ca2+]i threshold for Ca2+m uptake, we specu-

lated that its absence would enable agonist-induced [Ca2+]isignals tomore efficiently raise [Ca2+]m within the total mitochon-

drial population by enabling those that lack close appositions

to ER to take up released Ca2+. We hypothesized that such

recruitment would enable weak agonist stimulation to more

efficaciously enhance OX-PHOS. In control cells, low (10 mM)

histamine was without effect on O2 consumption rate (OCR)

(Figures S4A and S4B), whereas it was enhanced more than

3-fold in response to 100 mM histamine (Figures S4C and S4D).

In contrast, OCR was enhanced by more than 2-fold in response

to 10 mM histamine in MICU1 KD cells, nearly to the level

achieved by 100 mM histamine (Figures S4A–S4D). These results

are consistent with the hypothesis that MICU1 regulates the

threshold [Ca2+]i for MCU-dependent Ca2+m uptake.

EFHandDomains ofMICU1Regulate theCa2+ Thresholdfor MCU-Mediated Ca2+ UptakeOur results indicate that MICU1 regulates the threshold for

MCU-mediated Ca2+ uptake in the low Ca2+ regime extending

down to resting [Ca2+]i, suggesting that it senses [Ca2+] with

high affinity. MICU1 contains two highly conserved Ca2+-binding

EF hands. We speculated that high-affinity Ca2+ binding by the

EF hands enables MICU1 to sense low [Ca2+] to inhibit MCU.

To test this, we stably re-expressed in MICU1 KD cells shRNA-

insensitive MICU1 with either EF1 or EF2 hands disabled by

introduction of two point mutations of critical acidic residues

that disable Ca2+ binding (Figure 5A). As shown previously in

Figures S3A–S3D, re-expression of shRNA-insensitive MICU1

rescued the MICU1 KD cells, preventing MCU-dependent

Ca2+m uptake in permeabilized cells in response to elevated

ICM [Ca2+] induced by release of ER stores or in response to

an acute 1 mM Ca2+ pulse. In contrast, re-expression of MICU1

with either EF1 or EF2 mutated (Figure 5A) failed to rescue

MICU1 KD cells (Figures 5B–5E). Interestingly, Hill equation

fits of the uptake rates of cells stably expressing MICU1 with

either EF hand mutated yielded kinetic parameters similar to

those of control and MICU1 KD cells (Figures 5F and 5G).

Thus, MICU1 regulation of the [Ca2+] threshold for MCU-medi-

ated Ca2+ uptake is mediated by high-affinity Ca2+ binding by

its two EF hands.

MICU1 EF hands are located in the mitochondrial matrix

(Perocchi et al., 2010). Accordingly, MICU1 should respond to

[Ca2+]m. To test whether MICU1 senses [Ca2+]m specifically, per-

meabilized cells were exposed to a single 10 mM Ca2+ pulse.

abilization and exposure to Tg in Neg shRNA control (black) andMICU1KD (red)

M shown. ***p < 0.001.

th Ca2+ pulses in permeabilized Neg shRNA (black; left) and MICU1 KD cells

M shown. ***p < 0.001.

ized Neg shRNA (black; left) andMICU1 KD cells (red; right). Solid lines, means;

a2+ pulses in permeabilized Neg shRNA (black; left) and MICU1 KD cells (red;

also Figure S4.

Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc. 637

Figure 5. EF HandDomains ofMICU1Regu-

late the Ca2+ Threshold for MCU-Mediated

Ca2+ Uptake

(A) Knockdown of MICU1 mRNA levels in HEK293

Neg shRNA, MICU1 KD, MICU1 rescue, and

MICU1 KD stably expressing shRNA-insensitive

MICU1 cDNA harboring either D231A and E242K

mutation (EF1 mutant) or D421A and E432K

mutation (EF2 mutant) cells (mean ±SEM; n = 3).

(B) Permeabilized HEK293 Neg shRNA, MICU1

KD, MICU1 rescue, MICU1 KD stably expressing

shRNA-insensitive MICU1 cDNA harboring either

EF1 or EF2 mutations; cells were pulsed with 1 mM

Ca2+ as indicated. Representative traces show

bath [Ca2+] (mM).

(C) Quantification of basal bath [Ca2+] at 450 s

after permeabilization. *p < 0.05, **p <0.01, and

***p < 0.001. Mean ±SEM; n = 3.

(D) Change in bath [Ca2+] due to Ca2+m up-

take in response to 1 mM Ca2+ pulse. Uptake

derived from single exponential fit of bath

[Ca2+] kinetics following Ca2+ pulse. **p < 0.01

and ***p < 0.001. n.s., not significant. Mean ±SEM;

n = 3.

(E) Total Ca2+m uptake after basal accumulation

(no Ca2+ pulse) or 1 mM Ca2+ pulse determined

by recording Ca2+ released by CCCP addition at

t = 750 s. Mean ± SE; *p < 0.05 and **p < 0.01.

Mean ±SEM; n = 3.

(F) Rate of Ca2+m uptake as function of bath [Ca2+]

derived from single exponential fits of bath [Ca2+]

responses following Ca2+ pulses in permeabilized

EF1 or EF2 mutant cells. Solid line is Hill equation

fit of the data. Reduced uptake rate observed at

[Ca2+] > 20 mM is due to DJm depolarization and

was not used in the fitting.

(G) Kinetic parameters derived from Hill equation

fits of data in (F). See also Figure S5.

Both control and MICU1 KD cells rapidly take up Ca2+ and

reduce ICM [Ca2+] to a new steady state level similar to that

present before the pulse (Figures 3A and S5). Before the pulse,

MCU-mediated Ca2+ uptake at this ICM [Ca2+] was minimal in

control cells, as evidenced by lack of significant Ca2+ extrusion

observed upon block of Ca2+ uptake by Ru360, in contrast to

the behavior of MICU1 KD cells (Figure 2F). However, after the

pulse, both control andMICU1KD cells behaved similarly, exhib-

iting robust MCU-mediated Ca2+ uptake, as evidenced by the

high rate of Ca2+ extrusion mediated by the Na+/Ca2+ exchanger

upon Ru360 addition (Figure S5). Thus, the elevation of [Ca2+]mafter the 10 mM pulse relieved MICU1 inhibition of MCU-medi-

ated Ca2+ uptake, indicating that MICU1 senses resting matrix

[Ca2+] specifically.

638 Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc.

MICU1 Deficiency Elevates BasalMitochondrial ROS Generation andSensitizes Cells to Apoptotic CellDeathMitochondrial OX-PHOS and NADPH

oxidases are the major reactive oxygen

species (ROS) sources in most cells (Ha-

manaka and Chandel, 2010; Lambeth,

2004). Ca2+m overload is associated with excessive mitochon-

drial ROS (mROS) generation (Jacobson and Duchen, 2002).

Basal mROS levels were elevated by several-fold in MICU1 KD

cells (Figures 6A–6C). Elevated mROS in MICU1 KD cells was

unaltered by the NADPH oxidase inhibitor DP1, whereas the

mitochondrial complex III blocker antimycin A elevated mROS

in both cell types (Figure 6C). This indicates that NADPH

oxidases are not responsible for mROS elevation in MICU1 KD

cells. Importantly, either reconstitution of MICU1 or siRNA

silencing of MCU in MICU1 KD cells abrogated mROS elevation

(Figure 6D). Similar results were obtained in ECs (Figures 6E and

6F). Thus, failure of MICU1 to inhibit MCU-mediated Ca2+ uptake

under low [Ca2+]i conditions results in [Ca2+]m-dependent mROS

overproduction. Whereas constitutively elevated [Ca2+]m in

Figure 6. MICU1 Deficiency Elevates Basal

Mitochondrial ROS

(A and B) Representative confocal images show

MitoSOX Red fluorescence in Neg shRNA (A) and

MICU1 KD (B) HeLa cells. Bottom, enlarged

portions of fields showing single cells depicting

mitochondrial localization of indicator.

(C) Quantification of MitoSOX Red fluorescence

in Neg shRNA and MICU1 KD cells, with or with-

out NADPH oxidase inhibitor DPI (10 mM).

Mitochondrial respiratory complex III blocker

antimycin A (AA) used as positive control.

Mean ±SEM; **p < 0.01, ns, not different; n = 3.

(D) Quantification of MitoSOX Red fluorescence

in MICU1 KD, MCU siRNA-treated MICU1 KD,

and MICU1 rescue HeLa cells. Mean ±SEM; **p <

0.01; n = 3.

(E) Representative confocal images of MitoSOX

Red fluorescence in Neg shRNA, MICU1 KD,

MICU1 KD +MCU siRNA, and MICU1 rescue ECs.

(F) Quantification of MitoSOX Red fluorescence in

MICU1 KD, MCU siRNA-treated MICU1 KD, and

MICU1 rescue ECs. Mean ±SEM; **p < 0.01; n = 3.

See also Figure S6.

MICU1 KD cells was expected to enhance basal bioenergetics,

this was not observed (Figure S4). We asked whether elevated

basal mROS inhibited mitochondrial metabolism. MICU1 KD

cells were transduced for 36 hr with mitochondrial (manganese

superoxide dismutase; AdMnSOD) and cytosolic (glutathione

peroxidase 1; AdGPX1) antioxidants. Both markedly decreased

[AMP]:[ATP] (Figure S6A) and enhanced basal OCR (Figure S6B),

suggesting that chronic [Ca2+]m elevation, by generating exces-

sive mROS, inhibited Ca2+-activated OX-PHOS.

Physiological mROS is implicated in many processes, includ-

ing gene expression and cell growth, proliferation, and differenti-

ation, whereas excessive mROS promotes cell death (Balaban

et al., 2005; Hamanaka and Chandel, 2010). Chronic mROS

elevation in MICU1 KD HeLa cells did not alter proliferation (Fig-

ure 7A). However, MICU1 KD ECs had defective migration in

a wound-healing assay (Figure 7B). Strikingly, ceramide-induced

HeLa cell death (Figures S7A and S7B) was enhanced by nearly

100% in MICU1 KD cells (Figure 7C), and lipopolysaccharide +

cycloheximide (LPS/CHX)-induced EC death was enhanced by

MICU1 KD (Figure 7D). Importantly, cell death was prevented

in MICU1 rescue cells (Figures 7C and 7D). Overexpression of

antioxidants, mitochondrial MnSOD, and cytoplasmic GPX1

also strongly protected MICU1 KD cells (Figures 7C and 7D).

Thus, MICU1 inhibition of MCU-dependent Ca2+ uptake under

basal conditions is necessary to suppress Ca2+m accumulation,

excessive mROS production, and sensitivity to apoptotic stress.

Cell 151, 630–644,

DISCUSSION

Themost significant finding of our studies

is that MICU1 plays an essential role in

limiting Ca2+m uptake when [Ca2+]i is

low, within the range found in cells at

rest or during weak agonist stimulation.

In absence of this regulation, mitochon-

dria become constitutively loadedwith Ca2+ under resting condi-

tions, and they more effectively take up Ca2+ in low [Ca2+]i.

Whereas the latter effect enables cells to respond more sensi-

tively to weak agonist stimulation, the former has detrimental

consequences, including enhanced mROS generation and

susceptibility to cell stresses. MICU1 provides an essential

protective mechanism by limiting MCU-mediated Ca2+ uptake

in the face of a tremendous thermodynamic driving force. This

essential role likely accounts for the tightly correlated expression

and physical interaction of MICU1 and MCU.

MICU1 Is a Gatekeeper of MCU-Mediated MitochondrialCa2+ InfluxOur results suggest that MICU1 inhibits MCU-dependent Ca2+

uptake in low [Ca2+]. Mitochondria in MICU1 KD cells have

constitutively elevated [Ca2+]m, suggesting that MICU1 limits

MCU-mediated Ca2+ uptake at [Ca2+]i that exists in resting cells,

�50–100 nM. Even at such low [Ca2+]i, the thermodynamic

driving force for Ca2+ uptake across the IMM is prodigious

(Azzone et al., 1977; Bernardi, 1999). Furthermore, the open

probability of the uniporter Ca2+ channel is nearly unity at normal

DJm (Kirichok et al., 2004). It has been assumed that continuous

Ca2+ extrusion compensates for influx (Nicholls, 2005) or, more

commonly, that the low apparent Ca2+ affinity of the uniporter

Ca2+ uptake mechanism effectively offsets this driving force

(Carafoli, 1987; Drago et al., 2011; Nicholls, 2005; Rizzuto

October 26, 2012 ª2012 Elsevier Inc. 639

Figure 7. MICU1 Deficiency Impairs Cell Migration and Sensitizes Cells to Apoptotic Cell Death

(A) HeLa cells were labeled with CFSE, and cell proliferation was determined by flow cytometry. Representative histograms (left) and quantification (right) of CFSE

distribution after 72 hr. Mean ±SEM; n = 3.

(B) Representative images and quantification of cell migration in Neg shRNA, MICU1KD, and MICU1 rescue ECs 8 hr postscratch. Mean ±SEM; n = 3.

(C) Quantification of cell death after 20 hr of C2-Ceramide (40 mM) treatment. Neg shRNA and MICU1 KD cells treated with adenoviral manganese superoxide

dismutase (MnSOD) and glutathione peroxidase 1 (GPX1) for 36 hr before challenged with C2-Ceramide for 20 hr; mean ±SEM; * p < 0.05 and ***p < 0.001; n = 3.

(D) Quantification of cell death in endothelial cells after 12 hr of LPS and cycloheximide treatment. Neg shRNA and MICU1 KD endothelial cells treated with

adenoviral MnSOD and GPX1 for 36 hr before challenge with LPS and cycloheximide for 12 hr; mean ±SEM; ***p < 0.001; n = 3.

et al., 2004), but our results suggest that neither is the case.

Addition of Ru360 in control cells did not unmask ongoing

Ca2+ extrusion (Figure 2F), suggesting that Ca2+ influx is not

constitutive under basal conditions. In contrast, Ru360 addition

revealed ongoing Ca2+ extrusion in MICU1 KD cells, indicating

that MCU-mediated Ca2+ uptake is activated and compensated

for at a new higher steady-state [Ca2+]m by constitutive Ca2+

extrusion that remains intact in MICU1 KD cells (Figures 2F

and 2G). Studies of isolated mitochondria have suggested that

[Ca2+] required for half-maximal uptake rate is up to 70 mM,

with most estimates suggesting an apparent Ca2+ affinity of

�10 mM, or 100-fold higher than resting [Ca2+]i (Bernardi, 1999;

Gunter and Pfeiffer, 1990). We confirmed the low apparent

affinity of MCU-mediated Ca2+ uptake—in the 10 mM range—

although IMM depolarization by [Ca2+] >20 mM limits the

ability to estimate this well, suggesting that 10 mM is a lower

limit. Nevertheless, mitochondria in MICU1 KD cells accumu-

640 Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc.

lated Ca2+ at normal DJm from the cytoplasm where [Ca2+] =

70 nM. Importantly, therefore, the low apparent Ca2+ affinity of

the uniporter does not effectively counteract the thermodynamic

driving force. Accordingly, cells require regulatory mechanisms

to limit MCU-mediated Ca2+ uptake under normal resting condi-

tions. The results here suggest that MICU1 is an essential

component of this mechanism.

Such mechanisms could operate by controlling kinetic pro-

perties of MCU-mediated Ca2+ uptake. However, MICU1 does

not appear to utilize this strategy. The apparent Vmax, Kd, and

Hill coefficient that describe the [Ca2+] dependence of MCU-

mediated Ca2+ uptake are essentially identical in control and

MICU1 KD cells. It was previously speculated that MICU1 might

contribute to Ca2+ cooperativity of uniporter-mediated Ca2+

uptake (Perocchi et al., 2010), but our results suggest that the

cooperativity is not mediated by MICU1 and resides elsewhere,

possibly the MCU channel itself. Our results suggest instead that

MICU1 acts as a Ca2+ sensor that sets a threshold [Ca2+]

for Ca2+m uptake. The [Ca2+]i threshold for mitochondria that

had not been preloaded with Ca2+ appears to be �3 mM. This

value was suggested by the steady-state bath [Ca2+] achieved

by [Ca2+]m uptake following exposure to boluses of Ca2+ or

to slow rises in Ca2+ caused by ER Ca2+ leak. Below 3–4 mM

Ca2+, MCU-mediated Ca2+ uptake is strongly inhibited by

MICU1.

The mechanism by which MICU1 inhibits MCU activity at

[Ca2+]i <�3 mM remains to be determined. MICU1 biochemically

interacts with MCU (Figure S2A; Baughman et al., 2011),

suggesting that MICU1 might regulate MCU activity directly,

perhaps by acting as an endogenous ligand that blocks the

permeation pathway or by altering MCU channel gating. Failure

to rescue enhanced Ca2+ uptake in MICU1 KD cells with MICU1

containing disabling mutations in either Ca2+-binding EF hand

suggests that MICU1 plays an important role as a Ca2+ sensor

in the mechanism. The requirement for two functional EF hands

is consistent with MICU1 binding Ca2+ with high affinity (Gifford

et al., 2007), as required for it to operate in the low [Ca2+] regime

as observed. Our results suggest that the EF hands of MICU1 are

bound to Ca2+ under both [Ca2+]i and [Ca2+]m resting conditions,

and this enables MICU1 to function as a brake to limit Ca2+

permeation through MCU. However, the lack of effect of

MICU1, with or without functional EF hands, on the kinetic

properties of MCU suggests that high [Ca2+]m can overcome

this block. Whether high [Ca2+]m suppresses MICU1 inhibition

by disrupting the physical interaction between MICU1 and

MCU (not supported by preliminary data) or by other MCU-

intrinsic or -extrinsic Ca2+ regulatory mechanisms remains to

be determined.

A ruthenium-red-sensitive [Ca2+]m uptake mode referred to as

the ‘‘rapidmode,’’ or RaM, describes a phenomenon observed in

isolated mitochondria in which Ca2+ uptake during the initial

300 ms of a Ca2+ pulse is much faster than the subsequent

rate (Gunter et al., 2004; Sparagna et al., 1995). Of note, RaM

was not observed in cases in which mitochondria had been pre-

incubated with�150 nM bath Ca2+ (Buntinas et al., 2001; Gunter

et al., 2004). The identification here of MICU1 as a high-affinity

Ca2+ brake on MCU-mediated Ca2+ uptake may account

for these observations. Thus, in mitochondria incubated in

Ca2+-depleted media, Ca2+ likely unbinds from MICU1, relieving

inhibition of MCU. Upon addition of Ca2+, uninhibited MCU

would initially take up Ca2+ at a rapid rate until Ca2+ permeation

into the matrix, shown to occur <100 ms after a Ca2+ pulse

(Gunter et al., 2004; Sparagna et al., 1995; Territo et al., 2001),

enables [Ca2+]m to reach levels for Ca2+ binding to the EF hands,

triggering MICU1-mediated rapid inhibition of influx. In this

model, the inability to observe RaMwhenmitochondria are incu-

bated in 150 nM Ca2+ is due to matrix [Ca2+] being sufficiently

high to ensure Ca2+ occupancy of the EF hands, enabling

MICU1 to exert its normal inhibition of Ca2+ uptake.

Functional Implications of the Role of MICU1Ca2+m uptake regulates cell bioenergetics, physiological ROS

signaling, and [Ca2+]i signals. Excessive Ca2+m uptake can

have deleterious effects, including inner membrane depolariza-

tion, mROS overproduction, sensitization to apoptotic and

necrotic stimuli, and activation of the permeability transition

pore and subsequent cell death pathways (see Introduction for

references). Our results here demonstrate an essential role of

MICU1 to control Ca2+m uptake and protect cells against delete-

rious consequences associated with Ca2+m overload. In the

absence of MICU1, excessive MCU-mediated Ca2+ uptake led

to elevated [Ca2+]m under basal conditions and mROS overpro-

duction and sensitivity to apoptotic stimuli. Physiologically

controlled delivery of Ca2+ to mitochondria can stimulate cellular

bioenergetics (Balaban, 2009; Denton and McCormack, 1980;

Robb-Gaspers et al., 1998; Territo et al., 2000). Lack of this

delivery (Cardenas et al., 2010), as shown here in response to

knockdown of MCU, causes bioenergetic stress. Conversely,

overexpression of MCU to promote excessive Ca2+m uptake

can potentiate ceramide-induced cell death (De Stefani et al.,

2011).MICU1KD similarly potentiated stress-induced cell death.

Of note, antioxidants mitigated these effects. Low mROS levels

are required to maintain normal cellular functions, whereas

aberrant mROS production leads to oxidative stress that is asso-

ciated with cellular damage that ultimately leads to cell loss and

organ failure (Hamanaka and Chandel, 2010). Our results

emphasize the critical role played by MICU1 in regulating the

nature of Ca2+-dependent mitochondrial oxidant signaling.

In summary, our study has identified an essential role of

MICU1 as a gatekeeper that sets the [Ca2+]i threshold without

affecting the kinetic properties of MCU-mediated Ca2+m uptake.

This essential regulation protects cells from Ca2+m overload and

consequent deleterious stress. The interaction between MICU1

and MCU may be an important target for cellular regulation

that tunes Ca2+m uptake to regulate bioenergetics and [Ca2+]iand oxidant signaling in physiological and pathophysiological

situations.

EXPERIMENTAL PROCEDURES

Cell Lines

Cells were grown in Dulbecco’s modified Eagle’s medium (DMEM) supple-

mented with 10% FBS, 100 U/ml penicillin, and 100 mg/ml streptomycin at

37�C, 5% CO2. Details can be found in Extended Experimental Procedures.

Endothelial cells were grown in endothelial cell growth supplement (ECGS)

supplemented condition.

RNA Interference

HeLa, HEK293, or endothelial cells were transfected with pools of distinct

proprietary siRNAs and used 72 hr posttransfection unless otherwise stated.

Details can be found in Extended Experimental Procedures.

Generation of Stable MICU1 Knockdown Cell Lines

Cells were transduced with MICU1 shRNA lentiviruses. Knockdown was

assessed by qRT-PCR. Details can be found in Extended Experimental

Procedures.

MICU1 Rescue Cells

MICU1 rescues were created by expressing constructs resistant to MICU1

shRNA knockdown. Details can be found in Extended Experimental

Procedures.

Quantitative Real-Time PCR Analysis

Total RNAwas isolated from cells by using RNAesyMinikit (QIAGEN), and total

RNA was reverse transcribed. Quantitative real-time PCR reactions were

Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc. 641

performed with gene-specific primers. Details can be found in Extended

Experimental Procedures.

Simultaneous Measurement of Cytoplasmic and Mitochondrial Ca2+

Concentrations

Cells grown on 25 mm glass coverslips were loaded with 2 mM rhod-2/AM

(50 min) and 5 mM Fluo-4/AM (30 min) in extracellular medium as previously

described (Madesh et al., 2005). Coverslips were mounted in an open perfu-

sion microincubator (PDMI-2; Harvard Apparatus) at 37�C and imaged. After

1 min of baseline recording, histamine (100 mM) was added, and confocal

images were recorded every 3 s (510 Meta; Carl Zeiss, Inc.) at 488 and

561 nm excitation using a 633 oil objective. Images were analyzed and quan-

titated by using ImageJ (NIH) (Hawkins et al., 2010; Madesh et al., 2005; Mal-

lilankaraman et al., 2011).

Simultaneous Measurement of Ca2+ Uptake and Mitochondrial

Membrane Potential in Permeabilized Cells

Negative shRNA, MICU1 KD, MICU1 rescue, MICU1 EF1 mutant, and MICU1

EF2mutant cellswere trypsinized, counted (63106), andwashed in anextracel-

lular-like Ca2+-free buffer (in mM: 120 NaCl, 5 KCl, 1 KH2PO4, 0.2 MgCl2, 0.1

EGTA, and 20 HEPES-NaOH [pH 7.4]). Following centrifugation, cells were

transferred to an intracellular-like medium (permeabilization buffer, in mM: 120

KCl, 10 NaCl, 1 KH2PO4, 20 HEPES-Tris [pH 7.2], protease inhibitors [EDTA-

free complete tablets, Roche Applied Science], and 2 mM thapsigargin and digi-

tonin [40 mg/ml]). The cell suspension supplemented with succinate (2 mM) was

placed in a fluorimeter andpermeabilizedby gentle stirring. FuraFF (0.5 mM)was

added at 0 s and JC-1 (800 nm) at 20 s to simultaneously measure extrami-

tochondrial Ca2+ and DJm. Fluorescence was monitored in a temperature-

controlled (37�C) multiwavelength-excitation dual wavelength-emission spec-

trofluorometer (Delta RAM, Photon Technology International) using 490 nm

excitation/535 nm emission for the monomer, 570 nm excitation/595 nm emis-

sion for the J-aggregate of JC1, and 340 nm/380 nm for FuraFF. The ratiometric

dye FuraFF was calibrated as previously described. At 450 s, Ca2+ pulses were

added, and DJm and extramitochondrial Ca2+ were monitored simultaneously.

CCCP was added at 750 s to determine mitochondrial Ca2+ content.

Mitochondrial Ca2+ uptake rate was derived by fitting the decay of bath

[Ca2+] after a Ca2+ pulse to the equation:

y = y0 +A exp

�� ½t � t0�

t

�;

where A is the change in bath [Ca2+] between the peak and the decay asymp-

tote and t is the decay time constant. Ca2+ uptake rate (A/t) as a function of

extramitochondrial calcium, [Ca2+]out, was fitted (Igor Pro 6.2) to a Hill equation

to derive kinetic parameters of MCU-mediated Ca2+ uptake:

Ca2+ uptake rate=Vmaxh

1+ ðKact=½Ca2+ �ÞHi;

where Vmax is the maximum rate, Kact is the half-maximal bath [Ca2+], and H is

the Hill coefficient. The total amount of Ca2+ accumulated by MCU-mediated

Ca2+ uptake (mitochondrial Ca2+ buffering capacity) was evaluated by

measuring the change in bath [Ca2+] in response to the uncoupler CCCP.

MCU-mediated Ca2+ accumulation as a function of extramitochondrial

[Ca2+] was fitted (Origin 7.0) with a 3-parameter model: y = a � b ln(x + c).

Immunoprecipitation

Transfected cells were harvested and lysed. Anti-Flag agarose beads

(Novus) were used for immunoprecipitation. Anti-GFP (Cell Signaling) was

used to detect MCU. Details can be found in Extended Experimental

Procedures.

Western Blotting and Treatments

Standard techniques were used. Chemiluminescence detection was per-

formed at exposures within the linear range. Details can be found in Extended

Experimental Procedures.

642 Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc.

Oxygen Consumption

OCR was measured at 37�C by using XF24 extracellular analyzer (Seahorse

Bioscience) as described (Cardenas et al., 2010). Details can be found in

Extended Experimental Procedures.

Measurement of Mitochondrial Superoxide

Mitochondrial superoxide was measured by using the mitochondrial O2d-

indicator MitoSOX Red (molecular probes; Invitrogen) as described (Mallilan-

karaman et al., 2011; Mukhopadhyay et al., 2007). Briefly, cells grown

on coated glass coverslips were loaded with 5 mM MitoSOX Red for

30 min, and coverslips were mounted in an open perfusion microincubator

(PDMI-2; Harvard Apparatus) at 37�C and imaged. Confocal (510 Meta; Carl

Zeiss, Inc.) images were obtained at 561 nm excitation by using a 633 oil

objective. For DPI treatment, cells were incubated with 30 mM DPI for 30 min

prior to dye loading. For antimycin A treatment, cells were incubated with

2 mM antimycin A for 30 min prior to dye loading. Images were analyzed,

and the mean MitoSOX Red fluorescence was quantified by using Image J

software (NIH).

Cell Proliferation Assay

Cell proliferation was measured by using the CellTrace CFSE Cell Proliferation

Kit (Molecular Probes, Invitrogen). Details can be found in Extended Experi-

mental Procedures.

Migration Assay

To assess endothelial cell migration, a scratch assay was employed as

described in Extended Experimental Procedures.

Cell Death Assay

Cell death was evaluated with the Aqua LIVE/DEAD Fixable Dead Cell Stain Kit

(Molecular Probes, Invitrogen). Details can be found in Extended Experimental

Procedures.

Flow Cytometry

Details can be found in Extended Experimental Procedures.

Luciferase Assay

Cells were transfected with NFAT-luciferase reporter plasmids, and luciferase

activity was detected by using Bright-Glo Luciferase Assay System (Promega).

Details can be found in Extended Experimental Procedures.

NFAT Nuclear Translocation

Cells were transduced with Ad5-NFATC3-GFP. Confocal images were

acquired, and nuclear translocation was assessed by counting the number

of nuclear-GFP-positive cells. Details can be found in Extended Experimental

Procedures.

Statistical Analyses

Unless indicated, all experiments were repeated thrice, and recordings are

representative of the mean fluorescence value of all cells/field or parameter.

Data from multiple experiments were quantified to determine peak n-fold or

percent change, expressed as mean ±SEM, and differences between groups

were analyzed by using two-tailed Student’s t test or, when not normally

distributed, a Mann-Whitney U test. Differences in means among multiple

data sets were analyzed by using one-way analysis of variance (ANOVA)

with the Kruskal-Wallis test, followed by pairwise comparison using the

Dunn test. p value < 0.05 was considered significant in all analyses. Data

were computed either with Graphpad Prism version 5.0 or with SigmaPlot

Software.

SUPPLEMENTAL INFORMATION

Supplemental Information includes Extended Experimental Procedures, seven

figures, and two tables and can be foundwith this article online at http://dx.doi.

org/10.1016/j.cell.2012.10.011.

ACKNOWLEDGMENTS

This work was supported by NIH grants HL086699, HL086699-01A2S1,

and 1S10RR027327-01 to M. Madesh and by GM56328 and MH059937

to J.K.F. We thank Jun Yang for biochemistry assistance and the University

of Pennsylvania IDOM (P30-DK19535) for assistance. C.C. was supported

by the American Heart Association. R.M. was supported by an NRSA

Fellowship.

Received: March 26, 2012

Revised: July 30, 2012

Accepted: October 5, 2012

Published: October 25, 2012

REFERENCES

Azzone, G.F., Bragadin, M., Pozzan, T., and Antone, P.D. (1977). Proton

electrochemical potential in steady state rat liver mitochondria. Biochim.

Biophys. Acta 459, 96–109.

Balaban, R.S. (2009). The role of Ca(2+) signaling in the coordination of

mitochondrial ATP production with cardiac work. Biochim. Biophys. Acta

1787, 1334–1341.

Balaban, R.S., Nemoto, S., and Finkel, T. (2005). Mitochondria, oxidants, and

aging. Cell 120, 483–495.

Baughman, J.M., Perocchi, F., Girgis, H.S., Plovanich, M., Belcher-Timme,

C.A., Sancak, Y., Bao, X.R., Strittmatter, L., Goldberger, O., Bogorad, R.L.,

et al. (2011). Integrative genomics identifies MCU as an essential component

of the mitochondrial calcium uniporter. Nature 476, 341–345.

Bernardi, P. (1999). Mitochondrial transport of cations: channels, exchangers,

and permeability transition. Physiol. Rev. 79, 1127–1155.

Bragadin, M., Pozzan, T., and Azzone, G.F. (1979). Kinetics of Ca2+ carrier in

rat liver mitochondria. Biochemistry 18, 5972–5978.

Buntinas, L., Gunter, K.K., Sparagna, G.C., and Gunter, T.E. (2001). The rapid

mode of calcium uptake into heart mitochondria (RaM): comparison to RaM in

liver mitochondria. Biochim. Biophys. Acta 1504, 248–261.

Carafoli, E. (1987). Intracellular calcium homeostasis. Annu. Rev. Biochem. 56,

395–433.

Carafoli, E., and Lehninger, A.L. (1971). A survey of the interaction of calcium

ions with mitochondria from different tissues and species. Biochem. J. 122,

681–690.

Cardenas, C., Miller, R.A., Smith, I., Bui, T., Molgo, J., Muller, M., Vais, H.,

Cheung, K.H., Yang, J., Parker, I., et al. (2010). Essential regulation of cell

bioenergetics by constitutive InsP3 receptor Ca2+ transfer to mitochondria.

Cell 142, 270–283.

Collins, T.J., Lipp, P., Berridge, M.J., and Bootman, M.D. (2001). Mitochondrial

Ca(2+) uptake depends on the spatial and temporal profile of cytosolic Ca(2+)

signals. J. Biol. Chem. 276, 26411–26420.

De Stefani, D., Raffaello, A., Teardo, E., Szabo, I., and Rizzuto, R. (2011). A

forty-kilodalton protein of the inner membrane is the mitochondrial calcium

uniporter. Nature 476, 336–340.

Denton, R.M., and McCormack, J.G. (1980). The role of calcium in the regula-

tion of mitochondrial metabolism. Biochem. Soc. Trans. 8, 266–268.

Drago, I., Pizzo, P., and Pozzan, T. (2011). After half a century mitochondrial

calcium in- and efflux machineries reveal themselves. EMBO J. 30, 4119–

4125.

Duchen, M.R., Verkhratsky, A., and Muallem, S. (2008). Mitochondria and

calcium in health and disease. Cell Calcium 44, 1–5.

Filippin, L., Magalhaes, P.J., Di Benedetto, G., Colella, M., and Pozzan, T.

(2003). Stable interactions between mitochondria and endoplasmic reticulum

allow rapid accumulation of calcium in a subpopulation of mitochondria. J.

Biol. Chem. 278, 39224–39234.

Gifford, J.L., Walsh, M.P., and Vogel, H.J. (2007). Structures and metal-ion-

binding properties of the Ca2+-binding helix-loop-helix EF-hand motifs.

Biochem. J. 405, 199–221.

Gunter, T.E., and Pfeiffer, D.R. (1990). Mechanisms by which mitochondria

transport calcium. Am. J. Physiol. 258, C755–C786.

Gunter, K.K., and Gunter, T.E. (1994). Transport of calcium by mitochondria.

J. Bioenerg. Biomembr. 26, 471–485.

Gunter, T.E., Gunter, K.K., Sheu, S.S., and Gavin, C.E. (1994). Mitochondrial

calcium transport: physiological and pathological relevance. Am. J. Physiol.

267, C313–C339.

Gunter, T.E., Yule, D.I., Gunter, K.K., Eliseev, R.A., and Salter, J.D. (2004).

Calcium and mitochondria. FEBS Lett. 567, 96–102.

Hajnoczky, G., Robb-Gaspers, L.D., Seitz, M.B., and Thomas, A.P. (1995).

Decoding of cytosolic calcium oscillations in the mitochondria. Cell 82,

415–424.

Hamanaka, R.B., and Chandel, N.S. (2010). Mitochondrial reactive oxygen

species regulate cellular signaling and dictate biological outcomes. Trends

Biochem. Sci. 35, 505–513.

Hansford, R.G. (1994). Physiological role of mitochondrial Ca2+ transport. J.

Bioenerg. Biomembr. 26, 495–508.

Hawkins, B.J., Irrinki, K.M., Mallilankaraman, K., Lien, Y.C., Wang, Y., Bhanu-

mathy, C.D., Subbiah, R., Ritchie, M.F., Soboloff, J., Baba, Y., et al. (2010).

S-glutathionylation activates STIM1 and alters mitochondrial homeostasis.

J. Cell Biol. 190, 391–405.

Herrington, J., Park, Y.B., Babcock, D.F., and Hille, B. (1996). Dominant role of

mitochondria in clearance of large Ca2+ loads from rat adrenal chromaffin cells.

Neuron 16, 219–228.

Hoth, M., Button, D.C., and Lewis, R.S. (2000). Mitochondrial control of

calcium-channel gating: a mechanism for sustained signaling and transcrip-

tional activation in T lymphocytes. Proc. Natl. Acad. Sci. USA 97, 10607–

10612.

Igbavboa, U., and Pfeiffer, D.R. (1988). EGTA inhibits reverse uniport-depen-

dent Ca2+ release from uncoupled mitochondria. Possible regulation of the

Ca2+ uniporter by a Ca2+ binding site on the cytoplasmic side of the inner

membrane. J. Biol. Chem. 263, 1405–1412.

Jacobson, J., and Duchen, M.R. (2002). Mitochondrial oxidative stress and cell

death in astrocytes—requirement for stored Ca2+ and sustained opening of the

permeability transition pore. J. Cell Sci. 115, 1175–1188.

Kirichok, Y., Krapivinsky, G., and Clapham, D.E. (2004). The mitochondrial

calcium uniporter is a highly selective ion channel. Nature 427, 360–364.

Lambeth, J.D. (2004). NOX enzymes and the biology of reactive oxygen. Nat.

Rev. Immunol. 4, 181–189.

Lemasters, J.J., Theruvath, T.P., Zhong, Z., and Nieminen, A.L. (2009).

Mitochondrial calcium and the permeability transition in cell death. Biochim.

Biophys. Acta 1787, 1395–1401.

Madesh, M., Hawkins, B.J., Milovanova, T., Bhanumathy, C.D., Joseph, S.K.,

Ramachandrarao, S.P., Sharma, K., Kurosaki, T., and Fisher, A.B. (2005).

Selective role for superoxide in InsP3 receptor-mediated mitochondrial

dysfunction and endothelial apoptosis. J. Cell Biol. 170, 1079–1090.

Mallilankaraman, K., Gandhirajan, R.K., Hawkins, B.J., andMadesh,M. (2011).

Visualization of vascular Ca2+ signaling triggered by paracrine derived ROS. J.

Vis. Exp. 58, e3511. http://dx.doi.org/10.3791/3511.

McCormack, J.G., Halestrap, A.P., and Denton, R.M. (1990). Role of calcium

ions in regulation of mammalian intramitochondrial metabolism. Physiol.

Rev. 70, 391–425.

Mukhopadhyay, P., Rajesh, M., Hasko, G., Hawkins, B.J., Madesh, M., and

Pacher, P. (2007). Simultaneous detection of apoptosis and mitochondrial

superoxide production in live cells by flow cytometry and confocal micros-

copy. Nat. Protoc. 2, 2295–2301.

Nicholls, D.G. (1978). The regulation of extramitochondrial free calcium ion

concentration by rat liver mitochondria. Biochem. J. 176, 463–474.

Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc. 643

Nicholls, D.G. (2005). Mitochondria and calcium signaling. Cell Calcium 38,

311–317.

Nicholls, D.G. (2008). Forty years of Mitchell’s proton circuit: From little grey

books to little grey cells. Biochim. Biophys. Acta 1777, 550–556.

O’Rourke, B. (2007). Mitochondrial ion channels. Annu. Rev. Physiol. 69,

19–49.

Orrenius, S., Zhivotovsky, B., and Nicotera, P. (2003). Regulation of cell death:

the calcium-apoptosis link. Nat. Rev. Mol. Cell Biol. 4, 552–565.

Palmer, A.E., Giacomello, M., Kortemme, T., Hires, S.A., Lev-Ram, V., Baker,

D., and Tsien, R.Y. (2006). Ca2+ indicators based on computationally rede-

signed calmodulin-peptide pairs. Chem. Biol. 13, 521–530.

Perocchi, F., Gohil, V.M., Girgis, H.S., Bao, X.R., McCombs, J.E., Palmer, A.E.,

and Mootha, V.K. (2010). MICU1 encodes a mitochondrial EF hand protein

required for Ca(2+) uptake. Nature 467, 291–296.

Rizzuto, R., Pinton, P., Carrington, W., Fay, F.S., Fogarty, K.E., Lifshitz, L.M.,

Tuft, R.A., and Pozzan, T. (1998). Close contacts with the endoplasmic

reticulum as determinants of mitochondrial Ca2+ responses. Science 280,

1763–1766.

Rizzuto, R., Duchen, M.R., and Pozzan, T. (2004). Flirting in little space: the ER/

mitochondria Ca2+ liaison. Sci. STKE 2004, re1.

Robb-Gaspers, L.D., Burnett, P., Rutter, G.A., Denton, R.M., Rizzuto, R., and

Thomas, A.P. (1998). Integrating cytosolic calcium signals into mitochondrial

metabolic responses. EMBO J. 17, 4987–5000.

644 Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc.

Rottenberg, H., and Scarpa, A. (1974). Calcium uptake and membrane

potential in mitochondria. Biochemistry 13, 4811–4817.

Santo-Domingo, J., and Demaurex, N. (2010). Calcium uptake mechanisms of

mitochondria. Biochim. Biophys. Acta 1797, 907–912.

Simpson, P.B., and Russell, J.T. (1998). Mitochondrial Ca2+ uptake and

release influence metabotropic and ionotropic cytosolic Ca2+ responses in

rat oligodendrocyte progenitors. J. Physiol. 508, 413–426.

Sparagna, G.C., Gunter, K.K., Sheu, S.S., and Gunter, T.E. (1995). Mitochon-

drial calcium uptake from physiological-type pulses of calcium. A description

of the rapid uptake mode. J. Biol. Chem. 270, 27510–27515.

Spat, A., Szanda, G., Csordas, G., and Hajnoczky, G. (2008). High- and

low-calcium-dependent mechanisms of mitochondrial calcium signalling.

Cell Calcium 44, 51–63.

Szalai, G., Krishnamurthy, R., and Hajnoczky, G. (1999). Apoptosis driven by

IP(3)-linked mitochondrial calcium signals. EMBO J. 18, 6349–6361.

Territo, P.R., Mootha, V.K., French, S.A., and Balaban, R.S. (2000). Ca(2+)

activation of heart mitochondrial oxidative phosphorylation: role of the

F(0)/F(1)-ATPase. Am. J. Physiol. Cell Physiol. 278, C423–C435.

Territo, P.R., French, S.A., Dunleavy, M.C., Evans, F.J., and Balaban, R.S.

(2001). Calcium activation of heart mitochondrial oxidative phosphorylation:

rapid kinetics of mVO2, NADH, AND light scattering. J. Biol. Chem. 276,

2586–2599.

Supplemental Information

Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc. S1

EXTENDED EXPERIMENTAL PROCEDURES

Cell LinesHeLa (ATCC# CCL-2), HEK293 (ATCC# CRL-1573) and 293T/17 (ATCC# CRL-11268) cells were grown in Dulbecco’s modified

Eagle’s medium supplemented (DMEM) supplemented with 10% FBS, 100 U/ml penicillin, and 100 mg/ml streptomycin at 37�Cand 5% CO2. All knockdown cells were grown in complete DMEM supplemented with puromycin (2 mg/ml). All rescue cells were

grown in complete DMEM supplemented with G418 (500 mg/ml). Endothelial cells were grown in ECGS supplementation condition.

RNA InterferenceHeLa or HEK293 cells (3 3 105/well) grown on 0.2% gelatin coated coverslips in six-well dishes were transfected with pools of four

distinct proprietary siRNAs to MICU1 (ON-TARGETplus SMARTpool, Dharmacon, USA) (50 nM) using DharmaFECT 1 transfection

reagent as per manufacturer’s instructions (Dharmacon). As controls, non-targeting siRNA duplexes (Dharmacon) were employed.

For MCU siRNA transfection, HeLa or endothelial cells were transfected with pools of two distinct siRNAs (MCU-1 GCCAGAGACA

GACAAUACUUU, MCU-2 GGGAAUUGACAGAGUUGCUdTdT, Dharmacon, USA) (50 nM) using DharmaFECT 1. Cells were used for

experiments 72 hr post-transfection unless otherwise stated.

Generation of Stable MICU1 Knockdown Cell LinesFour different lentiviruses carrying shRNAs targeting different regions ofmicu1were produced by co-transfecting 293T/17 cells with

MICU1 lentiviral shRNA constructs (Openbiosystems), psPAX2 and pMD2.G (Addgene), as described (Naldini et al., 1996). HeLa or

HEK293 cells (5X105/well) grown in 6 well plates were transduced with MICU1 lentiviruses, selected with puromycin (2 mg/ml) 48 hr

posttransduction for 6-10 days and expanded. Knockdown was assessed by qRT-PCR.

MICU1 Rescue CellsA MICU1 rescue construct resistant to MICU1 shRNA knockdown was created with four silent point mutations in the MICU1 shRNA

targeting region (ORIGENE). MICU1 knockdown (HeLa, HEK293 or endothelial cells) cells were transfected with MICU1 rescue

construct and 48 hr posttransfection selected with 500 mg/ml G418 sulfate.

MICU1 EF Hand Mutant CellsThe shRNA-insensitive MICU1with either EF1 or EF2 hands disabled by introduction of two point mutations of critical acidic residues

(EF1 mutant D231A, E242K and EF2 mutant D421A, E432K; Origene technologies) were stably re-expressed in MICU1 knockdown

HEK293 cells under G418 selection to generate EF1 and EF2 mutant cells.

Quantitative Real-Time PCR AnalysisTotal RNA was isolated from cells using RNAesy Minikit (QIAGEN) and total RNA (1cmg) was reverse transcribed with the high

capacity cDNA reverse transcription kit (Applied Biosystems ABI 4368814) in accordance with the manufacturer’s instruction. Quan-

titative Real-time PCR reactions were performed with gene specific solaris qPCR gene expression assay kit (Abgene UK) (for human

MICU1 and MCU: forward primer: CTTTGACCGAGAGGCTGCT, reverse primer: GTGAGTTCAGACGAAAC and Probe: TGTTT

GGACGCGATGTT. Forward primer: GTCAGTTCACACTCAAGCCTAT, reverse primer: TTGAAGCAGCAACGCGAACA, and Probe:

TCTATTCACCAGATGGT respectively), as per the manufacturer’s protocol. The relative gene expression was calculated with Neg

shRNA in the case of stable knockdown cells or scrambled siRNA for transient knockdowns using 7300 Real Time PCR system

RQ study software 1.4 (Applied Biosystems). The results were expressed as percent mRNA expression and plotted using Prism 5

software.

ImmunoprecipitationMCU-GFP and MICU1-Flag cDNAs were transfected into HEK293 cells which were harvested 36 hr post-transfection and lysed in

2.8 ml of PBS containing 1% Triton x-100 and protease inhibitors (Sigma) per T75 flask with brief sonication. Lysed cells were

centrifuged at 13,000 rpm for 30 min, and the supernatant was transferred and aliquoted into new tubes. 10 ml anti-Flag agarose

beads (Novus) was used for immunoprecipitation at 40�C for 6 hr. Beads were collected by centrifugation, washed three times in

1.5 ml cold lysis buffer, and resuspended in 1X SDS-PAGE sample buffer. After electrophoresis, proteins were transferred onto

Nitro-cellulose (GE). anti-GFP (Cell Signaling Technologies) was used to detect MCU.

Western Blotting and TreatmentsXeB (5 mM) was added in fresh media as indicated in text or figure legends. Treatments were terminated by rapid removal of medium

with cells on ice, followed by cell lysis with Cytobuster protein extraction reagent (Novagen) supplemented with protease and phos-

phatase inhibitors (complete PhosSTOP, Roche). Protein extracts were separated in 8%, 10% or 15%SDS-polyacrylamide gels and

transferred to PDVF membranes (Millipore). Blocking was at room temperature for 1 hr in 5% fat-free milk, and membranes were

incubated overnight at 4�C with primary antibody, and then for 1 hr at room temperature with a secondary antibody conjugated

to horseradish peroxidase. Chemiluminescence detection used ECL-plus reagent (Pierce) and a series of timed exposures of HyBlot

S2 Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc.

CL film (Denville Scientific) to ensure densitometric analyses were performed at exposures within the linear range. To ensure

equal protein loading across gels, membranes were submerged in stripping buffer (Restore Western blot stripping buffer; Pierce),

incubated at 37�C for 15 min, and re-probed with a loading control antibody. Films were scanned, and Image J was employed for

densitometric analysis.

Oxygen ConsumptionOxygen consumption rate (OCR) was measured in at 37�C using an XF24 extracellular analyzer (Seahorse Bioscience). Cells were

seeded in 24-well plates treated with CELL-TAK (BD Bioscience). After 24 hr, cells were loaded into themachine for O2 concentration

determinations. Cells were sequentially exposed to histamine (10 or 100 mM), oligomycin (1 mM), carbonylcyanide p-trifluoromethox-

yphenylhydrazone (FCCP; 300 nM) and rotenone (100 nM). After each injection, OCRwasmeasured for 5min, themediumwasmixed

and again measured for 5 min. Representative traces shown in Figure S4. Every point represents average of 10 different wells. Basal

OCRwas calculated as difference betweenOCRmeasured before and after oligomycin. MaximumOCRwas calculated as difference

between OCRmeasured after FCCP and that measured after exposure to rotenone plus myxothiazol. After every experiment protein

concentration was determinate by lysing the samples (Cardenas et al., 2010).

Cell Proliferation AssayTo assess cellular proliferation of HeLaMICU1 knockdown cells, Cell Trace CFSE Cell proliferation Kit (Molecular Probes, Invitrogen)

was used according to manufacturer’s instructions. Briefly, cells were grown on 100 cm dishes to 90% confluence and labeled with

5 mMCFSE for 15min at 37�C and quenched with pre-warmed complete media for 30min at 37�C. An equal number of CFSE stained

cells were plated and incubated at 37�C, 5% CO2 for 72 hr. After 72 hr, the cells were washed with PBS, stained with aqua live dead

discrimination dye as per manufacturer instructions. The cells were then fixed with 3.7% paraformaldehyde and the fluorescence

intensities acquired using a flow cytometer (LSRII; Becton Dickinson, CA, USA) as described in the flow cytometry methods (see

below) or stored at 4�C until flow acquisition.

Endothelial Cell Migration AssayNegative shRNA, MICU1 KD and MICU1 Rescue endothelial cells were seeded at a density of 0.5 3 105 cells/well in 6-well plates

overnight for confluent monolayer. A uniform 1.8 mm scratch running the entire length of the well was created using a sterile

200 ml tip. The wells were washed thrice with PBS to remove the cell debris and then bathed in 2 ml complete endothelial medium.

After 8 hr the cells were washed and fixed with CAMCO Quick Stain� II as per manufacturer instructions. The wells were photo-

graphed at multiple locations using phase contrast microscopy with a 4 x objective. Migration was quantified using Image J software

(NIH); results are expressed as percent cell migration (Craige et al., 2011).

Cell Death AssayFor quantitative assessment of cell death, Aqua LIVE/DEAD Fixable Dead Cell Stain kit (Molecular Probes, Invitrogen) was used.

HeLa cells grown in six well plates were treated with C2-Ceramide (40 mM) for 20 hr. Similarly, for experiments using adenoviral

MnSOD and GPX1, the cells were infected with Ad5MnSOD and Ad5GPX1 (University of Iowa Gene Transfer Core) at an MOI of

2,000 particles/cell. All infections were done at least 36 hr before C2-Ceramide treatment. Similar methodology was followed for

endothelial cells, except endothelial cells were treated with Lipopolysacharide (LPS) + cycloheximide (CHX) for 12 hr. Untreated cells

served as controls. After treatment, cells were collected along with floating cells, washed once with PBS and stained with aqua Live/

Dead stain as per manufacturer instructions. The cells were then fixed with 3.7% paraformaldehyde and fluorescence intensities

were acquired using a flow cytometer (LSRII; Becton Dickinson, CA, USA) as described in flow cytometry methods (below) or stored

at 4� until flow acquisition.

Flow CytometryFor flow cytometric analysis a minimum of 100,000 events per sample were collected and the data analyzed using Flow-Jo software.

Standard gating procedures using FMO controls were used to identify positive cells/peak. Relative fluorescence intensities were

plotted on a logarithmic scale using Flow-Jo software and quantification was plotted using GraphPad PRISM software version

5.0 (Graphpad).

Luciferase AssayHEK293 Neg shRNA, MICU1 knockdown (KD) and MCU siRNA treated cells grown on 96 well plates were transfected with either

NFAT-luciferase reporter plasmid (a kind gift from Dr. Gary Koretzky, University of Pennsylvania) or control vector using Mirus

LT-1 transfection reagent (Mirus) as per manufacturer’s instructions. 48 hr post-transfection, cells were treated with histamine

(100 mM) for 4 hr at 37�C and lysed with 50 mL/well of 13 passive lysis buffer (Promega) and the luciferase activity was detected using

Bright-Glo Luciferase Assay System (Promega) as per manufacturer’s instruction. Values were normalized with untreated controls

and the relative luciferase activity (RLU) was plotted using GraphPad PRISM software version 5.0 (Graphpad).

Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc. S3

NFAT Nuclear TranslocationHeLa NegshRNA, MICU1 knockdown (KD) and MCU KD cells grown on gelatin coated coverslips were transduced with

Ad5-NFATC3-GFP (a kind gift from Dr. Steven R. Houser, Temple University) at a MOI of 2,000 particles/cell. 48 hr posttransduction

cells were treated with histamine (100 mM) for 4 hr at 37�C. Confocal images were acquired and nuclear translocation was assessed

by counting the number of nuclear-GFP positive cells. Data were expressed as percent NFAT-GFP nuclear translocation.

SUPPLEMENTAL REFERENCES

Craige, S.M., Chen, K., Pei, Y., Li, C., Huang, X., Chen, C., Shibata, R., Sato, K., Walsh, K., and Keaney, J.F., Jr. (2011). NADPH oxidase 4 promotes endothelial

angiogenesis through endothelial nitric oxide synthase activation. Circulation 124, 731–740.

Naldini, L., Blomer, U., Gallay, P., Ory, D., Mulligan, R., Gage, F.H., Verma, I.M., and Trono, D. (1996). In vivo gene delivery and stable transduction of nondividing

cells by a lentiviral vector. Science 272, 263–267.

Figure S1. Consequences of MICU1 and MCU Knockdown on Bioenergetics, Autophagy, and NFAT Activation, Related to Figure 1

(A) MICU1 mRNA (left) and protein (right) levels in HeLa cells transiently transfected with scrambled and MICU1 siRNAs. (Mean ± SEM; n = 3; ***p < 0.001).

(B) HeLa cells treated with indicated siRNAs were loaded with cytosolic (fluo-4 AM) and mitochondrial (rhod-2 AM) Ca2+ indicators. Fluorescence was monitored

simultaneously by confocal microscopy during stimulation with histamine (100 mM). Representative confocal images of scrambled siRNA (Scr siRNA) andMICU1

siRNA treated HeLa cells showing rhod-2 fluorescence.

(C) Quantification of basal rhod-2 fluorescence in Scr siRNA and MICU1 siRNA treated cells.

(D and E) Kinetic traces of cytosolic (black) and mitochondrial (red) [Ca2+] in Scr siRNA and MICU1 siRNA treated HeLa cells in response to histamine (100 mM)

(solid lines are mean; shaded regions are Mean ± SEM; n = 3).

(F and G) Quantification of cytosolic (gray) and mitochondrial (red) [Ca2+] peak amplitudes in Scr siRNA and MICU1 siRNA treated cells. Mean ± SEM; ns, not

significant; n = 3.

(H) Quantification of TMRE fluorescence to assess DJm in Scr siRNA and MICU1 siRNA treated cells. Mean ± SEM; ns, not significant; n = 3.

(I) Transient knockdown of MCU but not MICU1 increases AMP:ATP ratio in HeLa cells, expressed as fold increase over cells transfected with scrambled siRNA

(Mean ± SEM; n = 3; **p < 0.01, ns, not significant).

(J and K)Western blot of LC3 and tubulin in HeLa cells with either MCU (C) or MICU1 (D) transiently knocked down (top) and quantification (bottom) in presence or

absence of Xestospongin B (XeB) expressed as fold increase over scrambled siRNA as control (Mean ± SEM; n = 3; ***, **p < 0.001 and 0.01, respectively; ns, not

significant).

(L) MICU1 mRNA (left) and protein (right) in HeLa cells stably expressing MICU1 lentiviral shRNAs targeting different regions of MICU1 (Mean ± SEM; n = 3).

(M) Lack of effect of stable MICU1 knockdown on basal and oligomycin-induced AMP:ATP ratios expressed as fold increase over Neg shRNA cells (Mean ± SEM;

n = 3; ***, **p < 0.001 and 0.01, respectively; ns, not significant).

(N) Western blot (top) of LC3 and tubulin and quantification (bottom) of LC3 in stable HeLa MICU1 knockdown clones in presence or absence of Xestospongin B

(XeB) (Mean ± SEM; n = 3).

(O) Representative responses to histamine (100 mM) of cytosolic and mitochondrial [Ca2+] in HeLa cells treated with siRNA against MICU1 or MCU.

(P) Quantification of NFAT-dependent luciferase activity in NFAT-Luc expressing HeLa cells transfected with MICU1 or MCU siRNA in response to histamine

(100 mM) for 4 hr (Mean ± SEM; n = 3).

(Q) Representative confocal micrographs showing NFATC3-GFP nuclear translocation before and after histamine stimulation in MICU1 and MCU

knockdown cells.

(R) Quantification of NFATC3-GFP nuclear accumulation in MICU1 and MCU knockdown cells.

S4 Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc.

Figure S2. MICU1 Knockdown Is without Effect on MCU Expression and Enhances PDH Activity, Related to Figure 2

(A) Western blot of MCU-GFP and MICU1-Flag showing co-immunoprecipitation in transiently transfected HeLa cells. Representative of 4 experiments.

(B) MCU protein expression, normalized to tubulin, in different stable MICU1 knockdown cells. Mean ± SEM, n = 3, n.s, not significant.

(C) MCU mRNA in Neg shRNA and MICU1 knockdown (KD) HeLa cells after MCU siRNA transfection, normalized to levels in control cells. (Mean ± SEM; n = 3).

(D) MCU mRNA in Neg shRNA and MICU1 knockdown (KD) ECs after MCU siRNA transfection, normalized to levels in control cells. (Mean ± SEM; n = 3).

(E) MICU1mRNA in Neg shRNA, MICU1 KD andMICU1 Rescue HEK293 cells. HEK293 cells were transduced with #B6 lentiviral shRNAs to generate a stable line

with MICU1 reduced to levels (by 82%) similar to those in the HeLa KD cells. Mean ± SEM, n = 3.

(F) Pyruvate dehydrogenase (PDH) is hypo-phosphorylated in cells with MICU1 KD. MICU1 shHe#B5 and shHe#B6 HeLa clones show low levels of PDH

phosphorylation in basal conditions compared with the HeLa shHe#B8 (control) clone. Treatment with XeB (5 mM, 1 hr) increased phospho-PDH in all clones

(Mean ± SEM; n = 3; *, **, ***: p < 0.05, 0.01 and 0.001, respectively).

(G) MICU1 mRNA in Neg shRNA, MICU1 KD and MICU1 rescue HeLa cells. Mean ± SEM, n = 3.

(H) MICU1 mRNA in Neg shRNA, MICU1 KD and MICU1 rescue ECs. Mean ± SEM, n = 3.

Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc. S5

Figure S3. Reconstitution of MICU1 Reverses Mitochondrial Ca2+ Uptake Phenotypes in MICU1 Knockdown Cells, Related to Figure 3

(A) Rescue of mitochondrial [Ca2+] set-point and Ca2+ uptake in response to low [Ca2+] pulses in permeabilized MICU1 KD HEK293 cells (red) by shRNA

insensitive MICU1 cDNA re-expression (blue). Solid lines mean, shaded areas Mean ± SEM (n = 3).

(B) Scatterplot quantification of basal extra-mitochondrial [Ca2+] (450 s after permeabilization) in Neg shRNA (black), MICU1 KD (red) and MICU1 rescue (blue)

cells. Mean ± SEM shown.

(C) Change in bath [Ca2+] due to mitochondrial Ca2+ uptake in response to various [Ca2+] pulses in control (black), MICU1 KD (red) and MICU1 rescue (blue) cells.

Uptake derived from single exponential fit of bath [Ca2+] kinetics following Ca2+ pulses. Mean ± SEM.; n = 3; **: p < 0.01; ns, not significant.

(D) Quantification of total mitochondrial Ca2+ release after CCCP addition. Mean ± SEM.; n = 3; *, **: p < 0.05 and 0.01, respectively; ns, not significant.

(E) HEK293Neg shRNA cells were permeabilizedwith digitonin in intracellular-likemedium (ICM) containing thapsigargin (Tg),DJm indicator JC-1 and bath [Ca2+]

indicator Fura FF. Permeabilized Neg shRNA and MICU1 KD cells were pulsed with different [Ca2+] (1, 1.3 and 1.6 mM) as indicated. CCCP added as indicated.

Traces show bath [Ca2+] (mM). Solid lines are mean; shaded regions are SEM; n = 3.

S6 Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc.

Figure S4. MICU1 Knockdown Enhances Respiration in Response to Weak Agonist Stimulation, Related to Figure 4

(A) Oxygen consumption rates (OCR) in Neg shRNA (black) and MICU1 knockdown (KD) (red) HeLa cells exposed sequentially to (a) 10 mM histamine, (b)

oligomycin, (c) FCCP and (d) rotenone. Mean ± SEM, n = 3.

(B) Summary of basal, 10 mM histamine-induced and maximal OCR. Mean ± SEM., n = 3. **, *: p < 0.01 and 0.05, respectively; ns, not significant.

(C) Oxygen consumption rates (OCR) in Neg shRNA (black) andMICU1 KD (red) HeLa cells exposed sequentially to (a) 100 mMhistamine, (b) oligomycin, (c) FCCP

and (d) rotenone. Mean ± SEM, n = 3.

(D) Summary of basal, 100 mM histamine-induced and maximal OCR. Mean ± SEM, n = 3. *: p < 0.05; ns, not significant.

Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc. S7

Figure S5. Elevated [Ca2+]m Normalizes Ca2+ Uptake in Low [Ca2+] between MICU1 Knockdown and Control Cells, Related to Figure 5(A and B) Neg shRNA (A) and MICU1 KD (B) HEK293 cells were permeabilized with digitonin in intracellular-like medium containing thapsigargin (Tg) and bath

[Ca2+] indicator Fura FF, and then pulsed with 10 mM Ca2+. After mitochondrial clearance of bath Ca2+, Ru360 caused elevation of bath [Ca2+], indicating that

steady-state bath [Ca2+] after pulse was maintained by balance of MCU-mediated Ca2+ uptake and 10 mM CGP37157-sensitive Na+-Ca2+ exchanger-mediated

extrusion. CCCP added as indicated. Solid line is mean; shaded areas are ± SEM (n = 3).

(C) Mitochondrial Ca2+ efflux rate derived from (A) and (B) during initial 60 s following Ru360 addition. Because bath [Ca2+] was at steady-state before Ru360

addition, efflux rate equals MCU-mediated influx rate. Lack of difference between MICU1 knockdown and control cells after [Ca2+]m rise induced by exposure to

10 mM Ca2+ pulse indicates MICU1 is important in regulating Ca2+ influx only when [Ca2+]m is low.

S8 Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc.

Figure S6. Attenuation of mROS Relieves Constitutive [Ca2+]m Accumulation-Induced Metabolic Changes in MICU1 KD Cells, Related to

Figure 6(A) Basal AMP:ATP ratios in Neg shRNA and MICU1 KD cells expressing antioxidant genes (MnSOD/GPX1). Mean ± SEM; n = 3; ***, **p < 0.001 and 0.01,

respectively; ns, not significant.

(B) Basal oxygen consumption rates (OCR) in Neg shRNA (black) and MICU1 knockdown (KD) (red) HeLa cells with or without antioxidant genes (MnSOD/GPX1)

expression. Mean ± SEM, n = 3.; ***p < 0.00; 1 ns, not significant.

Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc. S9

Figure S7. MICU1 Deficiency Sensitizes Cells to mROS-Mediated Apoptotic Cell Death, Related to Figure 7

(A) Representative histograms showing cell death in HeLa Neg shRNA,MICU1 KD andMICU1 KDRescue cells after 20 hr of C2-Ceramide (40 mM) treatment. Neg

shRNA andMICU1 KD cells treated with adenoviral manganese superoxide dismutase (MnSOD) and glutathione peroxidase 1 (GPX1) for 36 hr before challenged

with C2-Ceramide for 20 hr. Mean ± SEM; n = 3.

(B) Representative histograms showing cell death in Neg shRNA, MICU1 KD and MICU1 KD Rescue ECs after 12 hr of LPS and cyclohexamide treatment. Neg

shRNA andMICU1 KD ECs treated with adenoviral manganese superoxide dismutase (MnSOD) and glutathione peroxidase 1 (GPX1) for 36 hr before challenged

with LPS and CHX for 12 hr. Mean ± SEM; n = 3.

S10 Cell 151, 630–644, October 26, 2012 ª2012 Elsevier Inc.

Table S1. ON-TARGETplus SMARTpool siRNA Sequences, Related to Figures 1 and S1

Human CBARA1

GACAAAAUCUUCCGAUAUU

UCCUCGAAUUUCAGCGUAA

GCAGUUUGGUGGCAUGCUA

GGGUCUGGAUCAAUAUAUA

List shows the siRNA sequences used in the study.

Table S2. Lentiviral shRNA Sequences, Related to Figure 1

Gene name

Ref seq Clone ID ShRNA sense sequence

CBARA1 NM_006077.1 TRCN0000053368 GCAGCTCAAGAAGCACTTCAA

CBARA1 NM_006077.1 TRCN0000053369 CGATCCATAACACCCAATGAA

CBARA1 NM_006077.1 TRCN0000053370 GCAATGGCGAACTGAGCAATA

CBARA1 NM_006077.1 TRCN0000053372 CTTCGCTTTACCCAAACAGTA

List of shRNAs and their associated TRC clone identification numbers (IDs) and pubmed IDs.