Vous êtes sur la page 1sur 14

Increased Mitochondrial Biogenesis and

Reactive Oxygen Species Production


Accompany Prolonged CD4+ T Cell
Activation
This information is current as
of October 29, 2018. Billur Akkaya, Alexander S. Roesler, Pietro Miozzo,
Brandon P. Theall, Jafar Al Souz, Margery G. Smelkinson,
Juraj Kabat, Javier Traba, Michael N. Sack, Joseph A.
Brzostowski, Mirna Pena, David W. Dorward, Susan K.
Pierce and Munir Akkaya

Downloaded from http://www.jimmunol.org/ by guest on October 29, 2018


J Immunol published online 29 October 2018
http://www.jimmunol.org/content/early/2018/10/28/jimmun
ol.1800753

Supplementary http://www.jimmunol.org/content/suppl/2018/10/29/jimmunol.180075
Material 3.DCSupplemental

Why The JI? Submit online.


• Rapid Reviews! 30 days* from submission to initial decision
• No Triage! Every submission reviewed by practicing scientists
• Fast Publication! 4 weeks from acceptance to publication
*average

Subscription Information about subscribing to The Journal of Immunology is online at:


http://jimmunol.org/subscription
Permissions Submit copyright permission requests at:
http://www.aai.org/About/Publications/JI/copyright.html
Email Alerts Receive free email-alerts when new articles cite this article. Sign up at:
http://jimmunol.org/alerts

The Journal of Immunology is published twice each month by


The American Association of Immunologists, Inc.,
1451 Rockville Pike, Suite 650, Rockville, MD 20852
All rights reserved.
Print ISSN: 0022-1767 Online ISSN: 1550-6606.
Published October 29, 2018, doi:10.4049/jimmunol.1800753
The Journal of Immunology

Increased Mitochondrial Biogenesis and Reactive Oxygen


Species Production Accompany Prolonged CD4+ T
Cell Activation

Billur Akkaya,*,1 Alexander S. Roesler,†,1,2 Pietro Miozzo,†,3 Brandon P. Theall,†


Jafar Al Souz,* Margery G. Smelkinson,‡ Juraj Kabat,‡ Javier Traba,x Michael N. Sack,x
Joseph A. Brzostowski,† Mirna Pena,† David W. Dorward,{ Susan K. Pierce,† and
Munir Akkaya†
Activation of CD4+ T cells to proliferate drives cells toward aerobic glycolysis for energy production while using mitochondria
primarily for macromolecular synthesis. In addition, the mitochondria of activated T cells increase production of reactive oxygen
species, providing an important second messenger for intracellular signaling pathways. To better understand the critical changes

Downloaded from http://www.jimmunol.org/ by guest on October 29, 2018


in mitochondria that accompany prolonged T cell activation, we carried out an extensive analysis of mitochondrial remodeling
using a combination of conventional strategies and a novel high-resolution imaging method. We show that for 4 d following
activation, mouse CD4+ T cells sustained their commitment to glycolysis facilitated by increased glucose uptake through increased
expression of GLUT transporters. Despite their limited contribution to energy production, mitochondria were active and showed
increased reactive oxygen species production. Moreover, prolonged activation of CD4+ T cells led to increases in mitochondrial
content and volume, in the number of mitochondria per cell and in mitochondrial biogenesis. Thus, during prolonged activation,
CD4+ T cells continue to obtain energy predominantly from glycolysis but also undergo extensive mitochondrial remodeling,
resulting in increased mitochondrial activity. The Journal of Immunology, 2018, 201: 000–000.

A
ctivation in response to foreign molecules results in a T cells, which, upon activation, remodel their energy production
cascade of changes in immune cells. These changes machinery predominantly toward glycolysis with limited increase
include not only rapid proliferation and differentiation in oxidative phosphorylation (OXPHOS) (6), activated B cells
but also the extensive cellular remodeling that accompanies them. show a more proportional increase in both glycolysis and
The cellular changes that immune cells go through to meet the OXPHOS, and they maintain part of their OXPHOS capacity by
needs of activated cells have gained considerable interest in recent diverting a portion of glucose toward oxidation in mitochondria
years (1, 2). Most of the advances in the area focus on how im- through increased pyruvate dehydrogenase activity (7). So, as
mune activation is coupled to changes in cellular metabolism and compared with activated T cells, mitochondria in activated B cells
how these metabolic changes are maintained. Although cellular contribute more to overall energy production (3).
activation and differentiation is a process that demands high en- The process through which activated T cells use their mito-
ergy production, the way cells meet this demand is not uniform. chondria for generation of macromolecular intermediates, such as
For instance, recent studies showed that both B and T lymphocytes lipid biosynthesis from citrate and nucleic acids through 1-carbon
rely on aerobic respiration and fatty acid oxidation for their qui- metabolism, rather than energy production, resembles a similar
escent state energy needs, and upon activation, they shift toward choice that exists in rapidly proliferating tumor cells (8–11). This
glucose as the main energy source (3–5). However, in contrast to phenomenon, termed the Warburg Effect, represents a shift in

*Laboratory of Immunology, National Institute of Allergy and Infectious Diseases, This work was supported by the Intramural Research Programs of the National
National Institutes of Health, Bethesda, MD 20892; †Laboratory of Immunogenetics, Institute of Allergy and Infectious Diseases and the National Heart, Lung, and Blood
National Institute of Allergy and Infectious Diseases, National Institutes of Health, Institute, National Institutes of Health.
Rockville, MD 20852; ‡Research Technologies Branch, National Institute of Allergy
M.A. conceived and supervised the project; M.A., B.A., A.S.R., P.M., B.P.T., and J.T.
and Infectious Diseases, National Institutes of Health, Bethesda, MD 20892;
x designed experiments; M.A., B.A., A.S.R., P.M., B.P.T., J.A.S., M.G.S., M.P., J.T., and
Cardiovascular Branch, National Heart, Lung, and Blood Institute, National Insti-
D.W.D. carried out experiments; M.A., B.A., B.P.T., J.K., and J.T. analyzed data; M.A.
tutes of Health, Bethesda, MD 20892; and {Research Technologies Branch, National
and B.A. wrote the manuscript; S.K.P., M.N.S., and J.A.B. edited the manuscript.
Institute of Allergy and Infectious Diseases, National Institutes of Health, Hamilton,
MT 59840 Address correspondence and reprint requests to Dr. Munir Akkaya, Laboratory of
1 Immunogenetics, 5625 Fishers Lane, Room 4S04B, Rockville, MD 20852. E-mail
B.A. and A.S.R. contributed equally to this work.
address: munir.akkaya@nih.gov
2
Current address: Duke University School of Medicine, Durham, NC.
The online version of this article contains supplemental material.
3
Current address: University of Massachusetts Medical School, Worcester, MA.
Abbreviations used in this article: A/R, antimycin/rotenone; COXIV, cytochrome c
ORCIDs: 0000-0002-6808-3776 (B.A.); 0000-0001-5768-1589 (A.S.R.); 0000-0003- oxidase subunit IV; DC, dendritic cell; 2-DG, 2-deoxy-D-glucose; DNP, dinitrophe-
3486-4061 (P.M.); 0000-0003-3852-5139 (B.P.T.); 0000-0003-2345-2803 (J.A.S.); nol; ECAR, extracellular acidification rate; MFI, mean fluorescence intensity;
0000-0001-7777-5574 (M.G.S.); 0000-0002-3411-0000 (M.N.S.); 0000-0003-0257- mTORC1, mTOR complex 1; 2-NBDG, 2-(N-(7-nitrobenz-2-oxa-1,3-diazol-4-yl)
3905 (J.A.B.); 0000-0002-1710-731X (D.W.D.); 0000-0001-7261-3437 (S.K.P.); amino)-2-deoxyglucose; OCR, oxygen consumption rate; OXPHOS, oxidative phos-
0000-0002-9949-9424 (M.A.). phorylation; qPCR, quantitative PCR; ROS, reactive oxygen species; RT, room tem-
perature; STED, stimulated emission depletion; TMRM, tetramethylrhodamine,
Received for publication May 29, 2018. Accepted for publication September 25,
methyl ester, perchlorate; TOM20, translocase of the outer membrane 20; VDAC1,
2018.
voltage-dependent anion channel 1; WT, wild-type.

www.jimmunol.org/cgi/doi/10.4049/jimmunol.1800753
2 T CELL ACTIVATION INDUCES MITOCHONDRIAL REMODELING

rapidly proliferating cells toward glycolysis and lactate production (Corning/Life Sciences, Tewksbury, MA) previously coated with 4 mg/ml
even in the presence of oxygen (12). Despite the inefficiency of anti-mouse CD3ε (Clone 145-2C11; Biolegend, San Diego, CA) and
4 mg/ml anti-mouse CD28 (Clone 37.12; Biolegend). Cells were incubated
glycolysis as a source of energy compared with OXPHOS, this for up to 96 h in a 5% CO2 humidified tissue culture incubator at 37˚C.
commitment allows for the use of the mitochondrial TCA cycle
for macromolecular synthesis in proliferating T cells (9, 10, Adoptive transfer and in vivo CD4+ T cell activation
13–15). Furthermore, recent studies showed that mitochondrial For adoptive transfer experiments, naive CD4+ T cells purified from spleens
reactive oxygen species (ROS) production increases upon T cell of CD45.1+CD45.2+ wild-type (WT) mice and CD45.2+ OT-II transgenic
activation, and this acts as a second signal in regulating multiple mice were mixed at a 1:1 ratio, stained with e450 cell proliferation dye
downstream elements (15, 16). (Thermo Fisher, Waltham, MA) according to the manufacturer’s recom-
mendations, and resuspended in PBS. Two hundred microliters of PBS
However, despite the advances in our understanding of T cell containing 2 3 106 e450-stained CD4+ T cells were transferred into a
immunometabolism, key questions remain unanswered. Because CD45.2 WT mouse i.v. via tail vein injection. Twenty-four hours after
most studies focus on early time points after T cell activation, we do T cell transfer, mice were injected i.v. with 100 ml of PBS containing 7.5 3
not clearly know whether the shift toward glycolysis is sustained 105 DCs previously loaded with either OVA(323–339) or lymphocytic cho-
riomeningitis virus gp(61–80) peptides. Four days after DC transfer, mice
after prolonged activation. In addition, we do not know what type of were euthanized, and spleens were harvested.
structural mitochondrial remodeling, if any, accompanies sustained
T cell activation. Preparation of samples of light microscopy
In this study, we addressed these key questions by applying a Eighteen-millimeter no. 1.5 circular glass coverslips (catalog no. 64-0714;
range of metabolic and cellular analyses to naive and CD4+ T cells Warner Instruments, Hamden, CT) were placed inside wells of 12-well
activated both in vitro and in vivo in a comparative fashion. Our tissue culture plates (Corning), and 600 ml of 0.01% poly-L-lysine solution
(Sigma-Aldrich) was carefully applied on top of the coverslip, avoiding

Downloaded from http://www.jimmunol.org/ by guest on October 29, 2018


data showed a continued dominance of glycolysis over OXPHOS
spillover to the underside. After 5 min of incubation at room temperature
as the source of energy in activated T cells even 4 d after acti- (RT), liquid was aspired, and coverslip was dried for 2 h to overnight in a
vation. This was accompanied by a gradual increase in the ex- ventilating tissue culture hood.
pression of GLUT transporters, which allowed for increased Cells harvested upon activation or purification were stained with RPMI
glucose uptake. Using a novel mitochondrial imaging strategy 1640 without phenol red supplemented with 2% FCS, 1% HEPES, and 1/250
dilution of LIVE/DEAD Fixable Green Dead Cell Stain (Thermo Fisher
optimized for lymphocytes, we showed that mitochondrial
Scientific) for 20 min at 4˚C protected from light. Cells were then washed
remodeling goes in favor of increasing mitochondrial size, vol- and resuspended with prewarmed RPMI 1640 without phenol red supple-
ume, and number in activated cells, which highlights the impor- mented with 2% FCS, 1% HEPES and 150 nM MitoTracker Red CMXROS
tance of mitochondria in rapidly proliferating activated T cells (Thermo Fisher Scientific). Cells were incubated at 37˚C for 30 min, fol-
despite their limited role in energy production. lowed by two washes in a protein-free buffer (RPMI 1640 without phenol
red, PBS, or HBSS). After the final wash, cells were resuspended in protein-
free buffer at ∼3 3 106 cells/ml, and 500 ml of the cell suspension was
Materials and Methods transferred onto the poly-L-lysine–coated coverslip (note that although these
Animals, cells, and reagents numbers yielded optimal coating density for lymphocytes tested here, ad-
justments might be required if dramatic alterations in the size of cells are
Eight- to twelve-week-old C57BL/6 female mice purchased from anticipated because of specific experimental conditions). After 10 min of
The Jackson Laboratory (Bar Harbor, ME) were used for isolation incubation at RT, suspension was aspirated and washed using FACS buffer
of lymphocyte subsets. CD45.1+ CD45.2+ mice were generated by (HBSS containing 2% FCS and sodium azide). This was followed by a
breeding C57BL/6 mice with B6.SJL-Ptprca Pepcb/BoyJ mice (purchased 10-min incubation with anti-CD16/32 Ab (Biolegend). Cells were then
from The Jackson Laboratory). OT-II TCR transgenic mice were obtained washed with FACS buffer, fixed, and permeabilized using the Cytofix/
from Taconic Farms (Hudson, NY). Mice were maintained at National
Cytoperm kit (BD Biosciences, San Jose, CA). Rabbit anti-mouse translo-
Institute of Allergy and Infectious Diseases animal facilities, according to
case of the outer membrane 20 (TOM20) (2 mg/ml) (Abcam, Cambridge,
Animal Care and Use Committee standards.
MA) followed by ATTO-647N–conjugated anti-rabbit secondary (2 mg/ml)
For lymphocyte isolation, mice were euthanized by CO2 asphyxiation
(Sigma-Aldrich) was used to stain TOM20 intracellularly. Fixation, per-
followed by cervical dislocation, and spleens were harvested. Naive CD4+
meabilization, and intracellular staining steps were carried out by following
T cells were isolated using an isolation strategy described elsewhere (17).
the BD Cytofix/Cytoperm protocol. DAPI staining of nuclei was carried out
Cell purities were checked by staining with phenotyping Abs and found to
next by incubating the cells with PBS supplemented with 300 nM DAPI
be over 95% as measured by flow cytometry. For the isolation of dendritic
for 5 min at RT. Coverslips were washed once and mounted on slides by
cells (DCs), spleens were removed and flushed by complete RPMI 1640
using ∼10–15 ml of mounting media prepared by mixing dissolved
containing Liberase Blendzyme II and 2 mg/ml DNase, both purchased Mowiol 4-88 (Polysciences, Warrington, PA) with 0.1% aqueous solution
from Roche (Indianapolis, IN). Spleens were then fragmented and incu- of p-phenylenediamine at a 9:1 ratio, according to the manufacturer’s guide-
bated at 37˚C for 30 min. After incubation, RBCs were lysed with lines. Slides were left at RT in the dark overnight to ensure proper hard-
ammonium-chloride-potassium lysing buffer. DCs were isolated using ening before imaging. For long-term storage, samples were kept at 220˚C.
CD11c Microbeads (Miltenyi Biotec, Auburn, CA) and autoMACS
(Miltenyi Biotec) according to the manufacturer’s protocol. For Light microscopy and image analysis
experiments involving cell culture, cells were maintained in complete
media (RPMI 1640 media containing 50 U/ml penicillin, 50 mM 2-ME, Images were collected on a Leica TCS SP8 STED 33 system equipped with
white light and UV excitation lasers, a pulsed 775-nm depletion laser, an
50 mM streptomycin, 2 mM L-glutamine, 0.1 mM nonessential amino
HC PL APO 1003/1.40 oil STED White objective, and gated HyD de-
acids, 10% FCS, 1 mM sodium pyruvate, and10 mM HEPES.) tectors. To exclude dead cells from images, an initial single-slide image
To test the involvement of NO and mTOR in the regulation of was taken from the area of interest using all four lasers, and upon con-
metabolic remodeling, NO scavenger 2-(4-Carboxyphenyl)-4,4,5,5-
firmation of cell viability, a 488-nm laser used to detect the LIVE/DEAD
tetramethylimidazoline-1-oxyl-3-oxide potassium (Carboxy-PTIO) and stain was switched off to decrease sampling time and photobleaching.
mTOR inhibitor rapamycin (both purchased from Sigma-Aldrich, Image analysis was carried out using Imaris software version 9.1
St. Louis, MO) were added to cultures at 500 mM and 200 nM final (Bitplane, Zurich, Switzerland). Channel arithmetic extension in Imaris
concentrations, respectively. was used to sum the TOM20 and MitoTracker fluorescence intensities
In vitro CD4+ T cell activation together, and the resulting channel was selected to create a three-
dimensional surface rendering of the mitochondrion via surface crea-
For nonpolarizing CD4+ T cell (Th0) activation, a protocol previously tion and watershed splitting algorithm. The seed points were determined
described elsewhere was used (18). Briefly, aseptically purified naive CD4+ with a region growing estimated diameter of 0.5 mm. Other parameters
T cells were resuspended in complete media supplemented with 100 U/ml used in the three-dimensional surface creation, such as surface grain size
IL-2 (Peprotech, Rocky Hill, NJ) and 10 mg/ml anti-TGF-b (Bioxcell, and diameter of the largest sphere, were adjusted individually, according
West Lebanon, NH) and plated on sterile 24-well tissue culture plates to the fluorescence intensities.
The Journal of Immunology 3

Transmission electron microscopy used: Histon 3 (9715S; Cell Signaling Technology), HSP60 (4870S; Cell
Signaling Technology), SIRT3 (5490S; Cell Signaling Technology), and
For transmission electron microscope imaging, activated and naive CD4+ TOM20 (sc-11415; Santa Cruz Biotechnology). Images were captured
T cells were harvested and fixed by directly mixing the cell culture sus- using the Odyssey system (LI-COR).
pensions with excess amounts of the modified Karnovsky fixative (0.1 M
sodium phosphate buffer containing 4% paraformaldehyde and 2.5% Quantitative PCR
glutheraldehyde) (Electron Microscopy Sciences, Hatfield, PA). Samples
were processed following a previously published microwave irradiation The ratio of mitochondrial to genomic DNA was used to assess the mi-
strategy (19) with the following modifications: centrifugations were carried tochondrial biogenesis potential of T cells. For this purpose, viable cells
out at 800 3 g, cells were infiltrated in Araldite resin (SPI, West Chester, were FACS sorted, and DNA was isolated using DNeasy Blood and Tissue
PA), sections were cut at 200-nm thickness, and no additional staining was Kit (Qiagen, Hilden, Germany) and quantitative PCR (qPCR) was carried
performed on sections. UltraScan 4000 camera (Gatan, Pleasanton, CA) out with different dilutions of DNA using iQ SYBR Green Supermix (Bio-
was used for image collection. Rad, Hercules, CA). To represent genomic DNA, Mouse 18S ribosomal
DNA was amplified using 59-TAGAGGGACAAGTGGCGTTC-39 and 59-
Extracellular flux assay CGCTGAGCCAGTCAGTGT-39. To represent mitochondrial DNA, mouse
cytochrome oxidase subunit I was amplified using 59-GCCCCAGATA-
Oxygen consumption rate (OCR) and extracellular acidification rate TAGCATTCCC-39 and 59-GTTCATCCTGTTCCTGCTCC-39, and 12S
(ECAR) measurements were performed using Seahorse XF96 analyzer ribosomal DNA was amplified using 59-ACCGCGGTCATACGATTAAC-
(Agilent Technologies, Santa Clara, CA) following a previously published 39 and 59-CCCAGTTTGGGTCTTAGCTG-39. The transcriptional activity
protocol (20). For each experiment, activated and naive CD4+ T cells were of various genes encoding antioxidant enzymes was measured using RNA
FACS sorted prior to assay to exclude nonviable cells. Chemicals used for isolated from freshly isolated naive CD4+ T cells and viable FACS-sorted
glycolysis and mitochondrial stress tests were administered at the fol- 4-d activated CD4+ T cells following a protocol and primer sets published
lowing final concentrations: oligomycin (1 mM), 2,4 dinitrophenol (DNP) elsewhere (7).
(0.1 mM), antimycin A (1 mM), rotenone (1 mM), glucose (10 mM), and

Downloaded from http://www.jimmunol.org/ by guest on October 29, 2018


2-deoxy-D-glucose (2-DG) (50 mM). Statistical significance analysis
Flow cytometry Methods used to analyze statistical significance between experimental
groups and the p value ranges are described in respective figure
For measurement of mitochondrial membrane potential, a final concen-
legends. Statistical analysis was calculated using Graph Pad Prism
tration of 5 mM FCCP (depolarizes the membrane to show the back-
Version 7.
ground staining), 6 mM oligomycin (hyperpolarizes the membrane to
show maximum achievable potential), or plain media (to show current
potential) were added directly to the cell cultures. After 10 min incu-
bation at 37˚C, staining mixture containing tetramethylrhodamine, Results
methyl ester, perchlorate (TMRM) (Thermo Fisher Scientific) (30 nM Prolonged activation of CD4+T cells results in cellular
final concentration) and SYTOX Blue Dead Cell Stain (Thermo Fisher changes to facilitate glycolysis
Scientific) (1 mM final concentration) were added to the cells. Cultures
were incubated for an additional 30 min and then analyzed in flow Naive CD4+ T cells harvested from mouse spleens were acti-
cytometry. The formula 100 3 (mean fluorescence intensity [MFI](TMRM alone) 2 vated in vitro using plate-bound Abs specific for CD3 and CD28
MFI(TMRM + FCCP))/(MFI(TMRM + oligomycin) 2 MFI(TMRM + FCCP)) was used to in media containing soluble IL-2 and anti–TGF-b Ab for 4 d to
standardize the measured TMRM MFI levels in the percentage of the maximum
potential format.
promote nonpolarized activation and proliferation. As expected,
For flow cytometric detection of ROS production, 5 mM MitoSOX Red T cells proliferated and upregulated the activation marker CD44
(Thermo Fisher Scientific) or 500 nM CellROX (Thermo Fisher Scientific) as assessed by flow cytometry (Supplemental Fig. 1A). We
were used together with 1 mM SYTOX Blue Dead Cell Stain. initially compared freshly isolated naive and activated CD4+
For intracellular staining of GLUT transporters and mitochondrial T cells in a Seahorse analyzer using a glucose stress test as
markers, harvested cells were resuspended in FACS buffer supplemented
with 1/250 dilution of LIVE/DEAD Near-IR Dead Cell Stain (Thermo described previously (20). For this assay, both cell types were
Fisher Scientific) and any relevant surface-staining Abs. Upon 30 min immobilized in chambers and starved of glucose for 30 min.
incubation at 4˚C, cells were washed, fixed, and permeabilized by either Following three baseline recordings of ECAR, which correlates
BD Biosciences CytoFix/Cytoperm kit (for GLUT stains) or Biolegend to lactate production by glycolysis, glucose was added to the
Foxp3 stain kit (for mitochondrial stains). Fluorochrome-conjugated mAbs
against GLUT-1, GLUT-3, voltage-dependent anion channel 1 (VDAC1),
wells to stimulate glycolysis. Next antimycin/rotenone (A/R)
TOM 20, and cytochrome c oxidase subunit IV (COXIV) used in these was added to block complex I and III of the respiratory chain,
experiments were purchased from Abcam. which is expected to increase glycolytic capacity to its maxi-
For identification of the origin of adoptively transferred cells, Abs against mum (24). Finally, 2-DG was added to the wells to stop gly-
mouse CD4, CD45.1, and CD45.2 were used. Activation states of cells were colysis to demonstrate the direct link between the ECAR and the
detected by Abs against CD44. mTOR complex 1 (mTORC1) activity was
evaluated by measuring surface expressions of mTOR-dependent markers glycolytic performance. We observed that activated T cells
CD71 and CD98 (21–23), using fluorescently labeled mAbs. These Abs rapidly responded to glucose by increasing their ECAR val-
were purchased from Biolegend. ues, which was further enhanced by adding A/R. In contrast,
Real-time glucose uptake of the cells was monitored by adding 2-(N- naive T cells did not show any significant response to stimula-
(7-nitrobenz-2-oxa-1,3-diazol-4-yl)amino)-2-deoxyglucose (2-NBDG) (Thermo
Fisher Scientific) to cultured cells. For this purpose, activated and naive
tion, showing the relative insignificance of glycolysis in resting
T cells were stained with a viability marker and then incubated with 2- T cell energy production (Fig. 1A, 1B). These observations were
NBDG for various durations. Cells remained in 37˚C culture until being in line with the findings of a previous study that showed gradual
harvested at various time points and were then rapidly analyzed in a flow increases in ECAR values and lactate production for both hu-
cytometer. man CD4+ and CD8+ T cells during the course of a 3-d in vitro
Data acquisition was done using BD LSR II or BD 320 cytometers,
and data were analyzed using FlowJo software (version 10) (FlowJo, activation (25).
Ashland, OR). Next, we ruled out the possibility that confounding effects
resulting from glucose starvation might be accounting for these
Western blot
findings by repeating the experiment in the absence of glucose
For mitochondrial content detection by Western blot, freshly isolated naive deprivation, which showed similar results (Supplemental Fig.
CD4+ T cells and in vitro–activated CD4+ T cells were stained with a 1B). Consistent with these observations, as compared with
viability dye, and live cells were FACS sorted into protein-free buffer.
Equal numbers of viable cells from both conditions were lysed using
naive T cells, activated T cells increased GLUT-1 and GLUT-3
radioimmunoprecipitation assay buffer. Lysates were run on SDS-PAGE expressions (Fig. 1C–F) that corresponded to their increased
and transferred to nitrocellulose membranes. The following Abs were ability to take up glucose as measured by 2-NBDG (Fig. 1G).
4 T CELL ACTIVATION INDUCES MITOCHONDRIAL REMODELING

Downloaded from http://www.jimmunol.org/ by guest on October 29, 2018


FIGURE 1. Activated CD4+ T cells remodel their glucose uptake machinery to meet increased energy demand. For all experiments, purified naive
and 96-h ex vivo–activated CD4+ T cells were used. (A and B) Cells were FACS sorted for viability, and 5 3 105 viable cells were immobilized onto
each well of a 96-well Seahorse analyzer in which they were starved of glucose for 30 min, and basal ECAR levels were measured. Glucose, A/R, and
2-DG were consequently added after this period, and changes in ECAR values were recorded. (A) Arrowheads indicate the time points treatments were
added. Symbols and error bars refer to mean and SD of triplicates. (B) ECAR values pooled from four independent glycolysis stress tests (glycolysis =
ECARpostglucose 2 ECARbasal; glycolytic capacity = ECARpost–A/R 2 ECARpost–2-DG). Symbols demonstrate the means of individual experiments, and
lines mark the mean of the pooled data. (C–F) The expression levels of GLUT-1 (C and D) and GLUT-3 (E and F) as measured by flow cytometry.
Representative histograms (C and E) and bar graphs (D and F) demonstrating the MFI values are shown. Bars and error bars represent mean and SD of
triplicates. (G) Cells were stained with LIVE/DEAD and incubated in the presence of 10 mM 2-NBDG for up to 90 min. At each time point, aliquots
were harvested and run on a flow cytometer. MFI values were then normalized by subtracting the background MFI. Symbols and error bars represent
mean and SD of quadruplicates, respectively. (H) Expressions of Slc2a1 and Slc2a3 genes encoding GLUT-1 and GLUT-3, respectively, are quantified
using RNA isolated from naive and activated T cells. Bars and error bars represent mean and SEM of five independent experiments, respectively. Data
in (A) and (C)–(G) represent four independent experiments. Statistical significance was measured with Welch t test (B, D, F, and H) or two-way ANOVA
with Sidak multiple comparisons analysis (G). *0.01 , p # 0.05, **0.001 , p # 0.01, ****p # 0.0001.
The Journal of Immunology 5

Downloaded from http://www.jimmunol.org/ by guest on October 29, 2018

FIGURE 2. T cells continuously activated in vitro also increase their mitochondrial respiration despite greatly accelerated glycolysis. For all
experiments, purified naive and 96-h ex vivo–activated CD4+ T cells were used. (A–C) Cells were FACS sorted for viability, and 5 3 105 viable cells
were immobilized onto each well of a 96-well Seahorse analyzer. Upon basal OCR measurements, cells were sequentially treated with oligomycin;
2,4 DNP; and A/R on time points indicated with arrowheads. (A) Graph represents the changes in OCR values. Symbols and error bars refer to mean
and SD of each recording. (B) OCR values pooled from four independent mitochondrial stress tests (basal: OCRinitial 2 OCRpost–A/R; maximal
respiration = OCRpost-DNP 2 OCRpost–A/R; ATP-coupled respiration = OCRbasal 2 OCR postoligomycin). Symbols demonstrate means of individual
experiments, and lines mark the mean of the pooled data. (C) Ratio of the basal OCR to basal ECAR values, pooled from the mitochondrial stress
tests above. (D–F) Mitochondrial membrane potentials were measured using TMRM either alone or in combination with oligomycin or FCCP.
Representative histograms (D), MFI bar graphs (E), and percentage of the maximum potential graphs (F) are shown. Data represent four independent
experiments each carried out with triplicates. Statistical significance was calculated using Welch t test. p . 0.05 = ns; *0.01 , p # 0.05, ***0.0001
, p # 0.001. ns, not significant.

Activated T cells were actively transcribing Slc2a1 and Slc2a3, Thus, upon prolonged activation, T cells sustained increased
the genes encoding GLUT1 and GLUT3, respectively. These glycolytic activity, and the high expression levels of GLUT1
genes were up to 40-fold more transcriptionally active in and GLUT3 facilitated glucose uptake to fuel the glycolytic
activated T cells compared with naive T cells (Fig. 1H). activity.
6 T CELL ACTIVATION INDUCES MITOCHONDRIAL REMODELING

FIGURE 3. Increased ROS production following


T cell stimulation persists during prolonged acti-
vation without causing oxidative stress-induced
mitochondrial dysfunction. For all experiments,
purified naive and 96-h ex vivo–activated CD4+
T cells were used. (A and B) Cells were stained

Downloaded from http://www.jimmunol.org/ by guest on October 29, 2018


with SYTOX Blue viability dye together with
CellROX (A) or MitoSOX (B). Representative his-
togram overlays (top panels) and graphs of MFI
values (bottom panels) are shown. Bars and error
bars represent mean and SD, respectively. Data
represent three independent experiments, each
carried out with quadruplicates. (C) Transcription
of genes coding for antioxidant enzymes gluta-
thione reductase (Gsr), glutathione peroxidase 1
(Gpx1), superoxide dismutase (Sod1 and Sod2), and
catalase (Cat), as measured by qPCR, are shown for
both activated and naive T cells. Bars and error bars
refer to the mean of individual samples obtained
from five independent experiments and SEM, re-
spectively. (D) Transmission electron microscopy
electron micrographs of naive and activated T cells.
Images represent sections obtained from 10 indi-
vidual cells. N refers to nucleus (scale bar, 400
nm). Statistical significance was calculated using
Welch t test. p . 0.05 = ns, **0.001 , p # 0.01,
***0.0001 , p # 0.001. ns, not significant.

T cells preserve their mitochondrial reserves upon electron transport chain through complex I and III was blocked
prolonged activation using A/R, which dropped the oxygen consumption to the basal
We performed a mitochondrial stress test to determine the con- level, confirming the link between the oxygen consumption
tribution of mitochondria to cellular energy production in ac- measurements and mitochondrial activity. We showed that,
tivated and naive CD4+ T cells using a Seahorse extracellular compared with naive T cells, activated T cells used their mi-
flux analyzer. Cells were immobilized, and their basal oxygen tochondria more actively for energy production as shown by
consumption was recorded. This was followed by oligomycin significantly higher oxygen consumption levels (Fig. 2A, 2B).
treatment to inhibit ATP production, which leads to a decrease However, the OCR/ECAR ratio was significantly lower in ac-
in oxygen consumption. 2,4 DNP was added to uncouple tivated cells, indicating that the prolonged activation-induced
OXPHOS from electron transport, which increases oxygen increase in glycolysis surpasses the increase in mitochondrial
consumption to its maximum possible level. Finally, the OXPHOS (Fig. 2C).
The Journal of Immunology 7

FIGURE 4. Prolonged T cell ac-


tivation leads to an increase in total
mitochondrial content. For all ex-
periments, purified naive and 96-h
ex vivo–activated CD4+ T cells were
used. (A and B) Cells were stained
with a viability dye together with
various mitochondria-specific markers
and analyzed in flow cytometry. Rep-
resentative histogram overlays (A) and
MFI graphs (B) are shown. Bars and
error bars represent mean of triplicates
and SD, respectively. (C) Live cells
were FACS sorted, and equal num-
bers of naive and activated cells
were lysed, separated in protein gel,
and immunoblotted for mitochon-
drial markers as well as H3 for load-
ing control. Representative Western

Downloaded from http://www.jimmunol.org/ by guest on October 29, 2018


blot image is shown. Data represent
two independent experiments. (D)
Total DNA was isolated from FACS-
sorted live naive and activated CD4+
T cells. Graphs show mitochondrial
DNA copy numbers relative to ge-
nomic DNA that were measured
using qPCR. Bars and error bars
refer to the mean of individual
samples obtained from five inde-
pendent experiments and SEM, re-
spectively. Statistical significance
was calculated using Welch t test.
*0.01 , p # 0.05; ***0.0001 ,
p # 0.001; ****p # 0.0001.

Next, we compared the mitochondrial performance between permeability (28). However, we observed no change in tran-
naive and activated CD4+ T cells by measuring their relative scriptional activity of the genes encoding antioxidant enzymes in
mitochondrial membrane potential using TMRM either alone or in activated T cells (Fig. 3C), suggesting either ROS levels were
combination with oligomycin or FCCP. We showed that activated insufficient to induce transcription or there are other mechanisms
T cells have higher total membrane potential (Fig. 2D, 2E) but use that prevent upregulation of antioxidant genes to preserve
a similar percentage of their maximum potential compared with beneficial increase in ROS levels. Consistent with a beneficial role
naive T cells (Fig. 2F). These findings suggest that the increase in of increased ROS production, we observed no increase in per-
TMRM staining may either represent a continuation of mito- centage of maximum mitochondrial membrane potential (Fig. 2),
chondrial membrane hyperpolarization that was shown to occur which would rise during mitochondrial dysfunction. Moreover,
shortly after activation (26, 27) or merely a consequence of a electron micrographs of both naive and activated T cells showed
greater total mitochondrial content in activated T cells. Because healthy cristae structures and absence of swollen or vesicular-
both activated and resting T cells use ,30% of their mitochondrial swollen mitochondrial matrix areas, similar to those exemplified
membrane potential, both can be considered efficient in terms of in Refs. 7, 29, and 30, all of which rule out any ROS-induced
mitochondrial ATP production with adequate reserves. gross pathology in mitochondrial structures (Fig. 3D). Therefore,
sustained increase in ROS can be considered as a regulatory
Activated T cells maintain their mitochondrial activity despite
component of activation rather than a signature of pathology.
prolonged ROS production
It was shown previously that T cell activation leads to rapid in- Prolonged T cell activation leads to increases in
creases in mitochondrial ROS production, which acts as a second mitochondrial content
messenger to induce downstream cellular changes (16). We asked We explored whether during prolonged activation, T cells increase
whether this early increase in ROS production is sustained upon their mitochondrial content. To do so, we measured the expression
prolonged activation. We observed higher ROS production in ac- levels of various proteins specifically localized in mitochondria and
tivated T cells, as measured by both CellROX (Fig. 3A), a marker widely used to assess mitochondrial content (7, 31–33). T cells
for total intracellular ROS production, and MitoSOX (Fig. 3B), a activated for 4 d in vitro as compared with naive T cells showed
specific dye indicating mitochondrial ROS production. The in- increased levels of mitofilin and COXIV, as measured by flow
creases in intracellular ROS often trigger transcriptional activation cytometry; sirtuin 3 (Sirt3) and heat shock protein 60 (Hsp60), as
of antioxidant enzymes such as SOD1 and SOD2 to prevent measured by western blot; and VDAC1 and TOM20 as measured
pathologic consequences of oxidative stress, such as induction of by both (Fig. 4A–C). Upregulation was observed for all these
mitochondrial swelling, by increasing mitochondrial membrane markers, demonstrating an increase in the total mitochondrial
8 T CELL ACTIVATION INDUCES MITOCHONDRIAL REMODELING

Downloaded from http://www.jimmunol.org/ by guest on October 29, 2018

FIGURE 5. Metabolic remodeling due to activation can be prevented by the inhibition of NO or mTOR pathways. Purified naive CD4+ T cells were
activated up to 96 h ex vivo. (A and B) Time-dependent changes in activation status (CD44), GLUT1 expression, and mitochondrial content (TOM20 and
COXIV). Representative histogram overlays (A) and MFI graphs (B) are shown. Symbols and error bars represent the mean of triplicates and SD, re-
spectively. Time points were compared against 0 h. (C–E) Naive CD4+ T cells were activated for 24 h ex vivo in the presence of rapamycin or Carboxy-PTIO where
indicated. Representative histogram overlays and MFI graphs for NO levels (C), mTOR-dependent surface markers (D), (Figure legend continues)
The Journal of Immunology 9

content in T cells upon prolonged activation, consistent with the activated T cells. Carboxy-PTIO–mediated inhibition of the in-
increased TMRM staining shown in Fig. 2D, 2E. To assess crease in mitochondrial content and glucose transporters was more
the mitochondrial biogenic potential of the cells, we compared the pronounced compared with the level of inhibition, observed in the
ratio of COI or 12S ribosomal DNA (genes encoded in the mi- presence of rapamycin (Fig. 5E). This suggests that NO may be
tochondrial genome) to 18S ribosomal DNA (a gene encoded in using both mTOR-dependent and -independent pathways during
the nuclear genome). This analysis revealed that 4 d poststimu- T cell activation-induced metabolic remodeling.
lation in vitro, T cells retain an increased mitochondrial DNA
mass, which further supports increased mitochondrial biogenesis Changes observed in in vitro–activated T cells can be
capacity (Fig. 4D). recapitulated in vivo
Thus far, we have shown that prolonged in vitro activation of
Metabolic remodeling following T cell activation is linked to isolated T cells results in increases in glucose transporters,
NO-mediated increases in mTOR activity total mitochondrial content, and mitochondrial ROS production
Our data, so far, showed that prolonged T cell activation leads to a (Figs. 1–4). To test whether these changes also occur upon
remodeling that involves both glucose metabolism and mito- in vivo activation of the cells, we carried out an adoptive transfer
chondria. To confirm that these changes directly stem from TCR experiment (Fig. 6A). For this purpose, naive TCR transgenic
stimulus, we repeated our 4-d culture experiments using different CD4+ T cells isolated from CD45.1- CD45.2+ OT-II mice were
combinations of activation conditions. This comparative analysis mixed with naive polyclonal CD4+ T cells isolated from WT
revealed that the observed increases in GLUT and mitochondrial CD45.1+ CD45.2+ congenic mice at a 1:1 ratio. Cells were la-
markers can be recapitulated almost perfectly by plate-bound anti- beled with e-450 cell proliferation dye and adoptively transferred

Downloaded from http://www.jimmunol.org/ by guest on October 29, 2018


CD3 stimulation alone and that addition of IL-2 in culture or to WT CD45.2+ animal. Twenty-four hours posttransfer, recipi-
costimulation with CD28 have no effects in the absence of TCR ent mice were injected with DCs pulsed with either OVA(323–339)
stimulation (Supplemental Fig. 2A, 2B). or control peptide, namely lymphocytic choriomeningitis virus
Next, to have a more in-depth understanding of the dynamics of gp(61–80). Four days poststimulation, spleens were harvested and
metabolic and cellular remodeling resulting from prolonged T cell analyzed in a flow cytometer using a gating strategy outlined in
activation, we measured the changes in the expression of CD44, Fig. 6B. As expected, OT-II transgenic T cells that were acti-
GLUT1, and mitochondrial markers on a daily basis during the course vated by OVA(323–339)-pulsed DCs in vivo had gone through a
of 4-d CD4+ T cell activation in vitro (Fig. 5A, 5B). This analysis few rounds of proliferation and induced the expression of acti-
showed gradual increases in the expression levels of the activation vation marker CD44 (Fig. 6C). In vivo activation of OT-II T cells
marker CD44 and glucose transporter GLUT1. In contrast, markers of also increased the fluorescent intensities of GLUT-1, COXIV,
mitochondrial content increased within the first 24 h and plateaued TOM20, and MitoSOX as compared with control groups, indi-
after that, showing that the need for increased glucose uptake is cating that changes accompanying in vitro activation can be
corelated with the duration of activation, whereas the majority of recapitulated in vivo (Fig. 6D–G).
mitochondrial remodeling occurs within the first day of activation.
These data also complemented the sustained high transcriptional ac- Multicolor imaging strategy offers a comprehensive analysis of
tivity we showed in Slc2a genes in Fig. 1H. activation-induced mitochondrial changes in T cells
Earlier studies on human samples showed a direct link between Having observed that both mitochondrial content and biogenesis
T cell stimulation-induced NO production and the increase in increase during prolonged activation, we wanted to visualize the
mitochondrial calcium levels and mitochondrial membrane hy- morphology of the resulting mitochondria. Visualization of mi-
perpolarization (34, 35). NO, through induction of mTORC1 ac- tochondria in live T cells is challenging given the small cytoplasmic
tivity, was shown to be responsible for prevention of mitophagy by volume and the relative fragility of T cells ex vivo.
Rab4A-mediated depletion of Drp1 (36–38). Furthermore, a re- To visualize T cell mitochondria, we developed a strategy using a
cent clinical trial in systemic lupus erythematosus patients showed combination of a MitoTracker dye (MitoTracker Red CMXROS)
that mTOR blockade leads to a reduction in mitochondrial mass that accumulates in the mitochondrial matrix (40) and a fluo-
(39). To address whether a similar link between T cell activation, rescently labeled Ab specific for a subunit of TOM20 (41). The
NO production, and mTOR activity might be responsible for the nucleus was stained with DAPI, and dead cells were marked by
changes we observed in the expression of GLUT transporters and/ a viability dye (Fig. 7A). We used high-resolution stimulated
or mitochondrial markers, we activated CD4+ T cells in vitro in emission depletion (STED) microscopy to distinguish individual
the presence or absence of NO scavenger (Carboxy-PTIO) or mitochondria inside the small cytoplasmic volume of T cells. To
mTOR inhibitor (rapamycin). We first showed that T cell reduce excessive photobleaching due to high laser power with
activation-induced NO can be efficiently scavenged in the pres- STED imaging, we used photostable ATTO-647 fluorochrome to
ence of Carboxy-PTIO (Fig. 5C) and that the blockage of NO label TOM20-specific Ab. This dual mitochondrial staining ap-
results in a level of reduction in mTORC1-dependent cell surface proach both eliminated staining artifacts and also provided a more
markers CD71 and CD98 similar to those observed in the presence precise image of mitochondrial morphology. We then applied
of rapamycin (Fig. 5D). Upon confirming the link between T cell computer algorithms to sum the fluorescence intensities of TOM20
activation-mediated NO production and the induction of mTOR and MitoTracker stains and to create artificial surfaces based on the
activity, we showed that both rapamycin and Carboxy-PTIO combined fluorescence intensity to predict the gross mitochon-
prevented the upregulation of GLUT1, VDAC1, and TOM20 in drial morphology and measure total mitochondrial volume. The

and metabolic markers (E) are shown. Bars and error bars represent the mean of triplicates and SD, respectively. Data are representative of two independent
experiments. Freshly isolated naive CD4+ T cells were used as control. Statistical significance was measured using one-way ANOVA with Dunnett (B)
or Tukey (C–E) multiple comparisons analysis. p . 0.05 = ns, *0.01 , p # 0.05, **0.001 , p # 0.01, ***0.0001 , p # 0.001, ****p # 0.0001.
ns, not significant.
10 T CELL ACTIVATION INDUCES MITOCHONDRIAL REMODELING

Downloaded from http://www.jimmunol.org/ by guest on October 29, 2018

FIGURE 6. In vivo activation of CD4+ T cells leads to changes that are similar to those observed in vitro. (A) Schematic outline of the adoptive transfer
experiment. (B) Gating strategy used to discriminate adoptively transferred OT-II and polyclonal T cells in recipient mouse spleens. (C) Histogram overlays
showing the levels of CD44 expression (left) and the extent of proliferation (right) in adoptively transferred polyclonal and OT-II T cells 4 d poststimulation
with DCs pulsed with control glycoprotein peptide (top) or OVA peptide (bottom). (D–G) T cells were activated in vivo for (Figure legend continues)
The Journal of Immunology 11

Downloaded from http://www.jimmunol.org/ by guest on October 29, 2018


FIGURE 7. High-resolution multicolor imaging of lymphocytes reveals mitochondrial changes in CD4+ T cells induced upon prolonged activation. Cells
activated for 4 d in vitro or freshly isolated from spleens of C57BL/6 mice were stained with LIVE/DEAD and MitoTracker CMXROS red, then mounted
on poly-L-lysine–coated coverslips. Cells were then fixed, permeabilized, and stained for TOM20 and DAPI as described in Materials and Methods. (A)
Microscope image demonstrating viable (white arrowheads) and nonviable (yellow arrowheads) cells in the same area of interest (scale bar, 4 mm). (B)
Outline of the sequential image processing strategy used for distinguishing mitochondrial boundaries in lymphocytes. Naive CD4+ T cells, stained as
outlined above, were imaged for TOM20, MitoTracker, and DAPI in STED system. Images left to right show microscope images of TOM20, MitoTracker,
their merged view with DAPI, combined channel showing the sum of both mitochondrial markers in one channel, computer-generated surface of the
combined channel, and individual mitochondria estimates based on the watershed splitting algorithm applied to the combined surface (scale bar, 1 mm).
(C–F) Naive and activated T cells were imaged and analyzed in groups of five cells per area of interest as described above. Representative STED microscope
images (C) and data derived from image analysis showing average total mitochondrial volumes per cell (D), average number of mitochondria per cell (E),
and volume distribution of individual mitochondria (F) are given. Lines represent mean values, and symbols represent averages of five cells (D and E) or
individual mitochondrion (F). Data represent two independent experiments, each with at least 25 cells analyzed per condition. Statistical significance was
calculated using Welch t test. Scale bars, 1 mm. ****p # 0.0001.

surfaces, consisting of both matrix and membrane compartments of The images showed fewer, sparsely distributed mitochondria in
the mitochondria, were also split via watershed splitting algorithm naive T cells as compared with activated T cells (Fig. 7C). Image
to estimate the size and number of individual mitochondria in each analysis showed that the average total mitochondrial volume per
cell (Fig. 7B) (Supplemental Video 1). cell is almost 3-fold higher in activated cells (Fig. 7D).
We used this imaging strategy to compare freshly isolated naive Furthermore, compared with naive cells, activated T cells had at
CD4+ T cells with T cells that were activated for 4 d in culture. least a 2-fold increase in the estimated number of mitochondria

4 d as illustrated in (A). Histogram overlays (left) and MFI graphs (right) show the levels of GLUT1 (D), TOM20 (E), COXIV (F), and MitoSOX (G). Each
circle is an individual mouse. Data represent two independent experiments. Statistical significance was calculated using two-way ANOVA with Sidak multiple
comparisons test. p . 0.05 = ns, ****p # 0.0001. ns, not significant.
12 T CELL ACTIVATION INDUCES MITOCHONDRIAL REMODELING

(Fig. 7E), and the average volume of individual mitochondria in T cell activation and highlight the functional importance of mito-
these cells was larger (Fig. 7F). These results not only comple- chondria in activated T cells despite its limited role in energy pro-
ment our observations on total mitochondrial content (Figs. 4–6) duction. The data presented in this article not only increase our
but also show that the increased total mitochondrial content in understanding of mitochondrial remodeling following prolonged
activated T cells is a direct consequence of both individual mi- T cell activation but also offer novel methodological approaches that
tochondrial enlargement and increased mitochondria numbers. can be applied to decipher changes in the mitochondrial dynamics in
Altogether, our data suggest that activation-induced mitochondrial other cell types under numerous experimental settings.
biogenesis serves the purpose of increasing mitochondrial content
in multiple ways. Disclosures
The authors have no financial conflicts of interest.
Discussion
In this study, we have addressed key questions related to metabolic
and cellular changes that accompany prolonged T cell activation References
and proliferation. Our data showed that the early increases in 1. Buck, M. D., D. O’Sullivan, R. I. Klein Geltink, J. D. Curtis, C. H. Chang,
D. E. Sanin, J. Qiu, O. Kretz, D. Braas, G. J. van der Windt, et al. 2016. Mi-
glycolysis and OXPHOS in activated T cells are maintained during tochondrial dynamics controls T cell fate through metabolic programming. Cell
prolonged activation. We have shown that after 4 d of both in vitro 166: 63–76.
and in vivo activation, rapidly proliferating CD4+ T cells have 2. O’Neill, L. A., R. J. Kishton, and J. Rathmell. 2016. A guide to immunome-
tabolism for immunologists. Nat. Rev. Immunol. 16: 553–565.
already increased their GLUT transporters and are more able to 3. Caro-Maldonado, A., R. Wang, A. G. Nichols, M. Kuraoka, S. Milasta,
import glucose into the cell and yet are still transcriptionally ac- L. D. Sun, A. L. Gavin, E. D. Abel, G. Kelsoe, D. R. Green, and J. C. Rathmell.

Downloaded from http://www.jimmunol.org/ by guest on October 29, 2018


2014. Metabolic reprogramming is required for antibody production that is
tive to induce further GLUT expression. suppressed in anergic but exaggerated in chronically BAFF-exposed B cells. J.
Our findings mainly focused on the changes in mitochondrial Immunol. 192: 3626–3636.
dynamics. This was an unexplored area because of the lack of high- 4. Caro-Maldonado, A., V. A. Gerriets, and J. C. Rathmell. 2012. Matched
and mismatched metabolic fuels in lymphocyte function. Semin. Immunol. 24:
resolution imaging strategies optimized for lymphocytes. These 405–413.
cells are not only difficult to handle and culture ex vivo (20), but 5. Frauwirth, K. A., J. L. Riley, M. H. Harris, R. V. Parry, J. C. Rathmell, D. R. Plas,
also their massive nucleus to cytoplasm ratio poses a great diffi- R. L. Elstrom, C. H. June, and C. B. Thompson. 2002. The CD28 signaling
pathway regulates glucose metabolism. Immunity 16: 769–777.
culty for any researcher aiming to visualize cytoplasmic com- 6. Chang, C. H., J. D. Curtis, L. B. Maggi, Jr., B. Faubert, A. V. Villarino,
partments. We addressed these shortcomings by introducing D. O’Sullivan, S. C. Huang, G. J. van der Windt, J. Blagih, J. Qiu, et al. 2013.
Posttranscriptional control of T cell effector function by aerobic glycolysis. Cell
viability staining to exclude nonviable cells as well as target the 153: 1239–1251.
mitochondria with two independent markers for added specificity. 7. Akkaya, M., J. Traba, A. S. Roesler, P. Miozzo, B. Akkaya, B. P. Theall,
By taking advantage of this novel multicolor mitochondria stain- H. Sohn, M. Pena, M. Smelkinson, J. Kabat, et al. 2018. Second signals rescue
B cells from activation-induced mitochondrial dysfunction and death. Nat.
ing protocol in a high-resolution STED microscope setting and Immunol. 19: 871–884.
processing the raw data further by using an optimized computer 8. van der Windt, G. J., and E. L. Pearce. 2012. Metabolic switching and fuel
algorithm, we have been able to visualize individual mitochondria choice during T-cell differentiation and memory development. Immunol. Rev.
249: 27–42.
and estimate their boundaries even at close proximity areas. 9. Ron-Harel, N., D. Santos, J. M. Ghergurovich, P. T. Sage, A. Reddy,
These analyses showed that upon continuous activation, CD4+ S. B. Lovitch, N. Dephoure, F. K. Satterstrom, M. Sheffer, J. B. Spinelli, et al.
2016. Mitochondrial biogenesis and proteome remodeling promote one-carbon
T cells upregulate their total mitochondrial volume and number as metabolism for T cell activation. Cell Metab. 24: 104–117.
well as the size of individual mitochondria. Our comparative anal- 10. Berod, L., C. Friedrich, A. Nandan, J. Freitag, S. Hagemann, K. Harmrolfs,
ysis on the ratio of mitochondrial DNA to genomic DNA revealed A. Sandouk, C. Hesse, C. N. Castro, H. Bähre, et al. 2014. De novo fatty acid
synthesis controls the fate between regulatory T and T helper 17 cells. [Published
that activated T cells have increased mitochondrial biogenesis. In erratum appears in 2015 Nat Med. 21: 414.] Nat. Med. 20: 1327–1333.
addition to the structural mitochondrial remodeling, we showed that 11. Wahl, D. R., C. A. Byersdorfer, J. L. Ferrara, A. W. Opipari, Jr., and G. D. Glick.
increases in ROS production is sustained through long-term acti- 2012. Distinct metabolic programs in activated T cells: opportunities for selec-
tive immunomodulation. Immunol. Rev. 249: 104–115.
vation, which is likely maintained at least in part by lack of increase 12. Palmer, C. S., M. Ostrowski, B. Balderson, N. Christian, and S. M. Crowe. 2015.
in antioxidant enzyme synthesis. Nevertheless, despite sustained Glucose metabolism regulates T cell activation, differentiation, and functions.
Front. Immunol. 6: 1.
increases in cellular ROS levels, our data demonstrated that the 13. Pearce, E. L., and E. J. Pearce. 2013. Metabolic pathways in immune cell ac-
functional and structural integrity of mitochondria is preserved, tivation and quiescence. Immunity 38: 633–643.
which is in line with the previous observations suggesting a 14. MacIver, N. J., R. D. Michalek, and J. C. Rathmell. 2013. Metabolic regulation
of T lymphocytes. Annu. Rev. Immunol. 31: 259–283.
physiological role for the ROS production. 15. Murphy, M. P., and R. M. Siegel. 2013. Mitochondrial ROS fire up T cell ac-
For many cell types, the function of mitochondria goes far beyond tivation. Immunity 38: 201–202.
its role in energy production. Recent studies showed that there is a 16. Sena, L. A., S. Li, A. Jairaman, M. Prakriya, T. Ezponda, D. A. Hildeman,
C. R. Wang, P. T. Schumacker, J. D. Licht, H. Perlman, et al. 2013. Mitochondria
clear correlation between the intracellular organization of mito- are required for antigen-specific T cell activation through reactive oxygen
chondria and the stress response of cells (42, 43). Mitochondria are species signaling. Immunity 38: 225–236.
17. Akkaya, B., A. H. Holstein, C. Isaac, M. P. Maz, D. D. Glass, E. M. Shevach, and
also known to actively participate in programmed cell death (44). M. Akkaya. 2017. Ex-vivo iTreg differentiation revisited: convenient alternatives
Furthermore, recently, we have identified a novel contribution of to existing strategies. J. Immunol. Methods 441: 67–71.
mitochondria in modulating activation-induced cell death in B 18. Akkaya, B., P. Miozzo, A. H. Holstein, E. M. Shevach, S. K. Pierce, and
M. Akkaya. 2016. A simple, versatile antibody-based barcoding method for flow
lymphocytes. Our assays showed that Ag stimulation of B cells cytometry. J. Immunol. 197: 2027–2038.
requires a spatiotemporally distinct second stimulus in the form of 19. Offerdahl, D. K., D. W. Dorward, B. T. Hansen, and M. E. Bloom. 2012. A three-
either cognate B–T interaction or TLR activation. Lack of this dimensional comparison of tick-borne flavivirus infection in mammalian and
tick cell lines. PLoS One 7: e47912.
second signal triggered a cascade of events, including mitochon- 20. Traba, J., P. Miozzo, B. Akkaya, S. K. Pierce, and M. Akkaya. 2016. An opti-
drial swelling, inefficient mitochondrial respiration, and increased mized protocol to analyze glycolysis and mitochondrial respiration
in lymphocytes. J. Vis. Exp.: e54918. Available at: https://www.jove.com/video/
production of ROS, all of which eventually led to the death of the 54918/an-optimized-protocol-to-analyze-glycolysis-mitochondrial-respiration.
cell (7). All of these examples show the importance of accurately 21. Edinger, A. L., and C. B. Thompson. 2002. Akt maintains cell size and survival
assessing mitochondrial changes. by increasing mTOR-dependent nutrient uptake. Mol. Biol. Cell 13: 2276–2288.
22. Kelly, A. P., D. K. Finlay, H. J. Hinton, R. G. Clarke, E. Fiorini, F. Radtke,
Altogether, our findings provide the most comprehensive and D. A. Cantrell. 2007. Notch-induced T cell development requires
mitochondria-centric approach to decipher changes that accompany phosphoinositide-dependent kinase 1. EMBO J. 26: 3441–3450.
The Journal of Immunology 13

23. Zeng, H., K. Yang, C. Cloer, G. Neale, P. Vogel, and H. Chi. 2013. mTORC1 34. Nagy, G., M. Barcza, N. Gonchoroff, P. E. Phillips, and A. Perl. 2004. Nitric
couples immune signals and metabolic programming to establish T(reg)-cell oxide-dependent mitochondrial biogenesis generates Ca2+ signaling profile of
function. Nature 499: 485–490. lupus T cells. J. Immunol. 173: 3676–3683.
24. Mookerjee, S. A., D. G. Nicholls, and M. D. Brand. 2016. Determining maximum 35. Nagy, G., A. Koncz, and A. Perl. 2003. T cell activation-induced mitochondrial
glycolytic capacity using extracellular flux measurements. PLoS One 11: e0152016. hyperpolarization is mediated by Ca2+- and redox-dependent production of
25. Renner, K., A. L. Geiselhöringer, M. Fante, C. Bruss, S. Färber, nitric oxide. J. Immunol. 171: 5188–5197.
G. Schönhammer, K. Peter, K. Singer, R. Andreesen, P. Hoffmann, et al. 2015. 36. Fernandez, D. R., T. Telarico, E. Bonilla, Q. Li, S. Banerjee, F. A. Middleton,
Metabolic plasticity of human T cells: preserved cytokine production under P. E. Phillips, M. K. Crow, S. Oess, W. Muller-Esterl, and A. Perl. 2009. Acti-
glucose deprivation or mitochondrial restriction, but 2-deoxy-glucose affects vation of mammalian target of rapamycin controls the loss of TCRzeta in lupus
effector functions. Eur. J. Immunol. 45: 2504–2516. T cells through HRES-1/Rab4-regulated lysosomal degradation. J. Immunol.
26. Gergely, P., Jr., C. Grossman, B. Niland, F. Puskas, H. Neupane, F. Allam, 182: 2063–2073.
K. Banki, P. E. Phillips, and A. Perl. 2002. Mitochondrial hyperpolarization and 37. Caza, T. N., D. R. Fernandez, G. Talaber, Z. Oaks, M. Haas, M. P. Madaio,
ATP depletion in patients with systemic lupus erythematosus. Arthritis Rheum. Z. W. Lai, G. Miklossy, R. R. Singh, D. M. Chudakov, et al. 2014. HRES-1/
46: 175–190. Rab4-mediated depletion of Drp1 impairs mitochondrial homeostasis and rep-
27. Banki, K., E. Hutter, N. J. Gonchoroff, and A. Perl. 1999. Elevation of mito- resents a target for treatment in SLE. Ann. Rheum. Dis. 73: 1888–1897.
chondrial transmembrane potential and reactive oxygen intermediate levels are 38. Talaber, G., G. Miklossy, Z. Oaks, Y. Liu, S. A. Tooze, D. M. Chudakov,
early events and occur independently from activation of caspases in Fas K. Banki, and A. Perl. 2014. HRES-1/Rab4 promotes the formation of LC3(+)
signaling. J. Immunol. 162: 1466–1479. autophagosomes and the accumulation of mitochondria during autophagy. PLoS
28. Holmström, K. M., and T. Finkel. 2014. Cellular mechanisms and physiological One 9: e84392.
consequences of redox-dependent signalling. Nat. Rev. Mol. Cell Biol. 15: 411–421. 39. Lai, Z. W., R. Kelly, T. Winans, I. Marchena, A. Shadakshari, J. Yu, M. Dawood,
29. Sesso, A., J. E. Belizário, M. M. Marques, M. L. Higuchi, R. I. Schumacher, R. Garcia, H. Tily, L. Francis, et al. 2018. Sirolimus in patients with clinically
A. Colquhoun, E. Ito, and J. Kawakami. 2012. Mitochondrial swelling and in- active systemic lupus erythematosus resistant to, or intolerant of, conventional
cipient outer membrane rupture in preapoptotic and apoptotic cells. Anat. Rec. medications: a single-arm, open-label, phase 1/2 trial. Lancet 391: 1186–1196.
(Hoboken) 295: 1647–1659. 40. Poot, M., Y. Z. Zhang, J. A. Krämer, K. S. Wells, L. J. Jones, D. K. Hanzel,
30. Sun, M. G., J. Williams, C. Munoz-Pinedo, G. A. Perkins, J. M. Brown, A. G. Lugade, V. L. Singer, and R. P. Haugland. 1996. Analysis of mitochondrial
M. H. Ellisman, D. R. Green, and T. G. Frey. 2007. Correlated three-dimensional morphology and function with novel fixable fluorescent stains. J. Histochem.

Downloaded from http://www.jimmunol.org/ by guest on October 29, 2018


light and electron microscopy reveals transformation of mitochondria during Cytochem. 44: 1363–1372.
apoptosis. Nat. Cell Biol. 9: 1057–1065. 41. Meisinger, C., M. T. Ryan, K. Hill, K. Model, J. H. Lim, A. Sickmann,
31. Verdin, E., M. D. Hirschey, L. W. Finley, and M. C. Haigis. 2010. Sirtuin reg- H. Müller, H. E. Meyer, R. Wagner, and N. Pfanner. 2001. Protein import
ulation of mitochondria: energy production, apoptosis, and signaling. Trends channel of the outer mitochondrial membrane: a highly stable Tom40-Tom22
Biochem. Sci. 35: 669–675. core structure differentially interacts with preproteins, small tom proteins, and
32. Garrabou, G., A. Soriano, S. López, J. P. Guallar, M. Giralt, F. Villarroya, import receptors. Mol. Cell. Biol. 21: 2337–2348.
J. A. Martı́nez, J. Casademont, F. Cardellach, J. Mensa, and O. Miró. 2007. 42. van der Bliek, A. M., Q. Shen, and S. Kawajiri. 2013. Mechanisms of mito-
Reversible inhibition of mitochondrial protein synthesis during linezolid-related chondrial fission and fusion. Cold Spring Harb. Perspect. Biol. 5: a011072.
hyperlactatemia. Antimicrob. Agents Chemother. 51: 962–967. 43. Youle, R. J., and A. M. van der Bliek. 2012. Mitochondrial fission, fusion, and
33. Burbulla, L. F., J. C. Fitzgerald, K. Stegen, J. Westermeier, A. K. Thost, H. Kato, stress. Science 337: 1062–1065.
D. Mokranjac, J. Sauerwald, L. M. Martins, D. Woitalla, et al. 2014. Mito- 44. Pradelli, L. A., M. Bénéteau, and J. E. Ricci. 2010. Mitochondrial control of
chondrial proteolytic stress induced by loss of mortalin function is rescued by caspase-dependent and -independent cell death. Cell. Mol. Life Sci. 67: 1589–
Parkin and PINK1. Cell Death Dis. 5: e1180. 1597.

Vous aimerez peut-être aussi