ArticlePDF Available

Endosomal lipid signaling reshapes the endoplasmic reticulum to control mitochondrial function

Authors:

Abstract

Cells respond to fluctuating nutrient supply by adaptive changes in organelle dynamics and in metabolism. How such changes are orchestrated on a cell-wide scale is unknown. We show that endosomal signaling lipid turnover by MTM1, a phosphatidylinositol 3-phosphate [PI(3)P] 3-phosphatase mutated in X-linked centronuclear myopathy in humans, controls mitochondrial morphology and function by reshaping the endoplasmic reticulum (ER). Starvation-induced endosomal recruitment of MTM1 impairs PI(3)P-dependent contact formation between tubular ER membranes and early endosomes, resulting in the conversion of ER tubules into sheets, the inhibition of mitochondrial fission, and sustained oxidative metabolism. Our results unravel an important role for early endosomal lipid signaling in controlling ER shape and, thereby, mitochondrial form and function to enable cells to adapt to fluctuating nutrient environments.
RESEARCH ARTICLE SUMMARY
CELL BIOLOGY
Endosomal lipid signaling reshapes the endoplasmic
reticulum to control mitochondrial function
Wonyul Jang, Dmytro Puchkov, Paula Samsó, YongTian Liang, Michal Nadler-Holly, Stephan J. Sigrist,
Ulrich Kintscher, Fan Liu, Kamel Mamchaoui, Vincent Mouly, Volker Haucke*
INTRODUCTION: Cells need to react appropri-
ately to nutritional cues. Defects in the rewiring
of metabolism in response to alterations in
nutrient supply have been linked to human
diseases ranging from diabetes to muscle
atrophy. Starvation represses anabolic path-
ways and facilitates catabolic ones, such as the
degradation of macromolecules by autophagy
and endolysosomes. Starvation also promotes
the b-oxidation of fatty acids in mitochondria
to produce adenosine triphosphate (ATP).
Within cells, organelles including lysosomes
and mitochondria undergo changes in shape
and dynamics. These processes are often regu-
lated by phosphoinositide lipids. Phosphoino-
sitides are also involved in the formation of
membrane contacts between organelles and
in the response of cells and tissues to growth
and nutrient signals. How the adaptive changes
that protect mammalian cells and tissues from
starvation-induced damage are coordinated on
a cell-wide scale is unknown.
RATIONALE: Endolysosomal membrane dynam-
ics and function are controlled by phospho-
inositide signaling lipids, most notably by the
synthesis and turnover of phosphatidylinositol
3-phosphate [PI(3)P]. Patients carrying muta-
tions in the gene encoding the lipid phosphatase
MTM1, an enzyme that mediates endosomal
PI(3)P turnover, suffer from X-linked centro-
nuclear myopathy (XLCNM), a severe neuro-
muscular disease characterized by muscle
atrophy, disorganization of mitochondria, and
defects in the organization of the muscle endo-
plasmic reticulum (ER). Given that PI(3)P is
a hallmark of endosomes, we hypothesized
that the control of early endosomal PI(3)P by
MTM1 might serve to orchestrate adaptive
changes in the dynamics of the ER and mito-
chondria in response to altering nutrient
supply.
RESULTS: Working with XLCNM patientderived
myoblasts and engineered cell lines, we found
that nutrient starvation (for example, lack of
amino acids) induced the hydrolysis of PI(3)P
by endosomal recruitment of MTM1. Concom-
itantly, tubular ER membranes were observed
to be converted into ER sheets by live super-
resolution light microscopy. Mechanistically,
loss of early endosomal PI(3)P upon starvation
was found to reduce membrane contacts be-
tween peripheral ER tubules and early endo-
somes. These contacts function as physical
tethers that may transmit pulling forces from
highly motile peripheral endosomes to the
tubular ER. Using proximity labeling proteomic
and functional cell biological experiments we
demonstrated that the ERendosome contacts
were mediated by binding of the related ER
membrane proteins RRBP1 and kinectin 1 to
PI(3)P on endosomes. To study the role of
starvation-induced reshaping of tubular ER
membranes into sheets on mitochondrial form
and function, we combined live imaging with
three-dimensional focused ion beam milling
scanning electron microscopy (FIB-SEM) and
proteomic analysis. We found that starvation-
induced ER reshaping by MTM1 reduced the
rate of mitochondrial fission and promoted
the formation of a hyperfused mitochondrial
network. Genetic manipulations that resulted
in ER sheet expansion caused the formation
of an enlarged mitochondrial network even
in fed cells. Conversely, impaired ER reshap-
ing and reduced mitochondrial network for-
mation were observed in starved myoblasts
from XLCNM patients. Mitochondrial net-
work formation appeared to be critical for
the delivery of fatty acids from lipid droplets
to mitochondria and for oxidative ATP pro-
duction to sustain energy supply in nutrient-
deprived cells.
CONCLUSION: Our data unravel a crucial role for
early endosomal lipid signaling in controlling
ER morphology and, thereby, mitochondrial
form and function to orchestrate the adaptive
response of cells to alterations in nutrient (e.g.,
amino acid) supply. This mechanism operates
independent of autophagy, a cellular self-eating
process typically induced by prolonged starva-
tion. Rather, it resembles an organellar con-
veyor belt, in which the tubular ER serves as a
membrane conduit that transmits nutrient-
triggered changes in endosomal PI(3)P levels
to metabolic organelles to enable metabolic
rewiring. How early endosomal PI(3)P levels
and MTM1 function are controlled by cellular
nutrient status is currently unknown. Defects
in ER shape, mitochondrial morphogenesis,
and cellular ATP depletion caused by loss of
MTM1 function can explain the observed myo-
fiber hypotrophy and defective ER organi-
zation in animal models of XLCNM and in
human patients who often appear under-
nourished. We therefore hypothesize that
dysregulated organelle remodeling may un-
derlie XLCNM caused by MTM1 mutations
in humans.
RESEARCH
Jang et al., Science 378, 1188 (2022) 16 December 2022 1of1
The list of author affiliations is available in the full article online.
*Corresponding author. Email: haucke@fmp-berlin.de
Cite this article as W. Jang et al., Science 378, eabq5209
(2022). DOI: 10.1126/science.abq5209
READ THE FULL ARTICLE AT
https://doi.org/10.1126/science.abq5209
Role of MTM1-mediated endosomal
PI(3)P signaling in mitochondrial
metabolic rewiring through
reshaping of the ER in response
to starvation. In fed cells, early
endosomes form contacts with ER
tubules. Tubular ER membranes
facilitate mitochondrial fission and
serve as a source for lipid droplet
formation. Nutrient starvation
induced hydrolysis of endosomal
PI(3)P by MTM1 reduces mem-
brane contacts between the
tubular ER and early endosomes.
The resulting loss of peripheral ER
tubules induces mitochondrial
network formation and the delivery
of fatty acids to mitochondria to
sustain cellular energy supply.
EE, early endosome; MT, microtubule;
FA, fatty acid; LD, lipid droplet.
Sheet ER
Fed
PI3P
EE
Tubular
ER
MT
Starved
FA
LD from
tubular ER
β-oxidation
FA
Tubular
ER
Mito
ssion
Sheet
ER
MTM1
Downloaded from https://www.science.org at Leibniz-Institut für Molekulare Pharmakologie on December 16, 2022
RESEARCH ARTICLE
CELL BIOLOGY
Endosomal lipid signaling reshapes the endoplasmic
reticulum to control mitochondrial function
Wonyul Jang
1
, Dmytro Puchkov
1
, Paula Samsó
1
, YongTian Liang
2
, Michal Nadler-Holly
1
,
Stephan J. Sigrist
2
, Ulrich Kintscher
3
, Fan Liu
1,3
, Kamel Mamchaoui
4
,
Vincent Mouly
4
, Volker Haucke
1,2,3
*
Cells respond to fluctuating nutrient supply by adaptive changes in organelle dynamics and in metabolism. How
such changes are orchestrated on a cell-wide scale is unknown. We show that endosomal signaling lipid
turnover by MTM1, a phosphatidylinositol 3-phosphate [PI(3)P] 3-phosphatase mutated in X-linked
centronuclear myopathy in humans, controls mitochondrial morphology and function by reshaping the
endoplasmic reticulum (ER). Starvation-induced endosomal recruitment of MTM1 impairs PI(3)P-dependent
contact formation between tubular ER membranes and early endosomes, resulting in the conversion of
ER tubules into sheets, the inhibition of mitochondrial fission, and sustained oxidative metabolism. Our
results unravel an important role for early endosomal lipid signaling in controlling ER shape and, thereby,
mitochondrial form and function to enable cells to adapt to fluctuating nutrient environments.
The ability of cells to react appropriately
to nutritional cues is of fundamental im-
portance for cell physiology, and defects
in the cellular response to altered nutri-
ent supply have been linked to human
diseases ranging from diabetes to muscle
atrophy (14). Among the early changes elic-
ited by nutrient stress are the suppression of
anabolic programs such as protein translation
(5,6) and the concomitant induction of cat-
abolic processes involving the proteasome,
autophagy [e.g., lipophagy, reticulophagy (7)],
endolysosomal turnover of proteins (6,8), and
increased mitochondrial b-oxidation of fatty
acids (9). How these adaptive responses that
protect mammalian cells and tissues from
starvation-induced damage and the induction
of apoptotic cell death are coordinated is un-
known. Endolysosomal membrane dynamics
and function are controlled by the spatiotem-
porally regulated synthesis and turnover of
phosphatidylinositol 3-phosphate [PI(3)P] and
related signaling lipids by phosphatidylinositol
(PI) 3-kinases and 3-phosphatases (10,11).
MTM1, the founding member of the myotubu-
larin family of PI 3-phosphatases, is crucially
involved in PI(3)P homeostasis on endosomes
(1215). Patients carrying mutations in the
MTM1 gene suffer from X-linked centronu-
clear myopathy (XLCNM), a severe, often fatal
disease characterized by muscle weakness
due to myofiber atrophy, disorganization of
mitochondria, and structural defects in the
organization of the sarcoplasmic reticulum, a
specialized form of the endoplasmic reticulum
(ER) found in muscle tissue that is important
for excitation-contraction coupling (16,17).
How XLCNM-linked mutations in endosomal
MTM1 cause such pleiotropic defects in the
organization of the ER and other organelles
has remained elusive but may relate to a thus
far unexplored function of endosomes in
orchestrating adaptive changes in organelle
dynamics.
Results
Nutrient-regulated reshaping of the ER is
controlled by endosomal MTM1
To address the question how XLCNM-linked
mutations in endosomal MTM1 cause defects
in ER organization, we analyzed ER morphol-
ogy (18) in myoblast cell lines from healthy
controls or XLCNM patients (table S1) (1921)
suffering from pronounced or complete loss of
MTM1 protein (fig. S1A) and a resulting in-
crease in PI(3)P levels (15). The ER membrane
is composed of interconnected uniform flat
cisternal sheets, fenestrated sheets with nano-
holes (22,23), and peripheral dynamic narrow
tubules (30 to 60 nm in diameter). ER sheets
and peripheral ER tubules were initially char-
acterized by confocal light microscopy (24)
and have recently been resolved at the nano-
scalebysuper-resolutionimaging(22,25). We
determined the morphology of the peripheral
ER by semiautomated image analysis of cells
expressing mEmerald- or Halo-tagged ver-
sions of the ER membrane protein Sec61bor
cells stained for the ER marker calreticulin
(fig. S1B). This analysis revealed a prominent
accumulation of tubular versus sheet ER in
myoblasts from XLCNM patients compared
with cells from healthy controls (Fig. 1, A and
B). Accumulation of tubular ER in XLCNM
myoblasts was confirmed by time-gated stimu-
lated emission depletion (gSTED) nanoscopy
imagingat50-nmresolution[consistentwith
(22)] (Fig. 1A). Depletion of MTM1 in healthy
controls phenocopied XLCNM myoblasts with
respect to the accumulation of tubular ER and
the concomitant reduction in sheet ER (Fig. 1C
and fig. S1C). These data suggest that the de-
fects in ER morphology observed in XLCNM
patient myoblasts are a consequence of MTM1
loss of function and are consistent with earlier
in vivo data indicating a possible role for MTM1
in the control of ER morphology (17).
A hallmark of MTM1 loss is the accumula-
tion of its substrate lipid PI(3)P (15). We rea-
soned that the observed defects in ER shape in
XLCNM myoblasts are a consequence of ele-
vated PI(3)P levels and, thus, might be rescued
by pharmacological inhibition of PI(3)P syn-
thesis. Consistently, specific inhibition of the
endosomal PI 3-kinase VPS34 in the presence
of VPS34-IN1 or reexpression of active MTM1
sufficed to restore normal ER morphology in
XLCNM patient myoblasts (Fig. 1D and fig. S1,
D and E). An elevated ER sheet-to-tubule ratio
was also observed in VPS34-IN1treated hu-
man HeLa cells imaged by confocal light mi-
croscopy or gSTED nanoscopy at steady state
(Fig. 1E and fig. S1, F and G). In contrast, de-
pletion of phosphatidylinositol 3,5-bisphosphate
[PI(3,5)P
2
], another potential MTM1 substrate
lipid (26), did not affect ER shape (fig. S1, F
and H). Previous data show that PI(3)P is
involved in the cellular response to altered
nutrient supply (2729). We therefore tested
whether PI(3)P metabolism might be regulated
by nutrients. We found endosomal PI(3)P levels
to decline upon cellular nutrient deprivation
[consistent with (29)] (Fig. 1F and fig. S1I),
and this was accompanied by progressive ER
sheet expansion in live or fixed cells imaged
by super-resolution STED or confocal micros-
copy(Fig.1,GandH;fig.S1,JandK;and
Movies 1 and 2). The levels of major ER-
shaping proteins remained unaltered (fig. S3, J
and K). ER sheet expansion was also observed
in cells treated with the potent catalytic mech-
anistic target of rapamycin (mTOR) inhibitor
Torin 1 (fig. S1L), which is often used to mimic
conditions of starvation. Further analysis re-
vealed that deprivation of amino acids (gluta-
mine in particular), rather than growth factors
and glucose, was responsible for ER reshaping
during starvation (Fig. 1I and fig. S1M). In con-
trast, amino acid replenishment failed to rescue
the starvation-induced reduction in cell area
(fig. S1, N and O), suggesting that ER shape and
cell size are controlled by distinct mechanisms.
To probe whether reduced PI(3)P levels
causally underlie the starvation-induced re-
modeling of ER membranes, we depleted cells
of MTM1, a condition under which PI(3)P
accumulates (15), by RNA interference or
RESEARCH
Jang et al., Science 378, eabq5209 (2022) 16 December 2022 1of16
1
Leibniz-Forschungsinstitut für Molekulare Pharmakologie
(FMP), 13125 Berlin, Germany.
2
Department of Biology,
Chemistry, and Pharmacy, Freie Universität Berlin, 14195
Berlin, Germany.
3
Charité-Universitätsmedizin Berlin,
10117 Berlin, Germany.
4
Centre de Recherche en Myologie,
Institut de Myologie, Inserm, Sorbonne Université, 75013
Paris, France.
*Corresponding author. Email: haucke@fmp-berlin.de
Downloaded from https://www.science.org at Leibniz-Institut für Molekulare Pharmakologie on December 16, 2022
CRISPR-Cas9mediated knockout (KO). Loss
of MTM1 potently antagonized the starvation-
induced conversion of peripheral ER tubules
into sheets (Fig. 1J and fig. S2, A to E), a pheno-
type that was rescued by reexpression of cat-
alytically active mCherry-MTM1 (fig. S2F). Further
ultrastructural analysis by three-dimensional
(3D) focused ion beam scanning electron mi-
croscopy (FIB-SEM) showed that even the peri-
nuclear ER, which appeared as flat uniform
sheets in starved wild-type (WT) cells, was
highly fenestrated in starved cells lacking
MTM1 (Fig. 1K, FIB-SEM; fig. S2G; and movie S1).
The total ER volume fraction was unaltered
(fig. S2H). Accumulation of highly fenestrated
sheet ER and ER tubules in MTM1 KO cells
was further evidenced by the reduced average
length of ER profiles determined by SEM anal-
ysis of 2D cross sections (Fig. 1K, SEM).
We also tested whether other members of
the myotubularin family of MTM1-related phos-
phatases affect ER shape. On the basis of their
mRNA expression levels in HeLa cells, we ana-
lyzed four myotubularin-related proteins (MTMRs)
and found that, of these four, only MTMR1, the
family member most closely related to MTM1,
affected ER shape (fig. S2, I to K). These data
suggest a close functional relationship between
endosomal PI(3)P and ER morphology, in par-
ticular the presence of ER tubules in cells. In-
creased levels of endosomal PI(3)P caused by
MTM1 loss of function prevent the starvation-
induced remodeling of ER membranes, and
Jang et al., Science 378, eabq5209 (2022) 16 December 2022 2of16
Fig. 1. Nutrient-regulated
reshaping of the ER is
controlled by endosomal
MTM1. (A) Confocal and STED
live images of Halo-Sec61bin
fed myoblasts from healthy
and XLCNM patients. (B) Ratio
of sheet/total ER area stained
for calreticulin. AB1190 (n=
206), KM1288 (n= 222), NL15
(n= 102). (C) Ratio of sheet/
total ER area in fed healthy
(AB1190) myoblasts at steady-
state treated with control
(siCo/siCo) or MTM1 siRNA
(siMTM1_1/_SP). siCo/siCo
(n= 181), siMTM1_1/_SP (n=
167). (D) Ratio of sheet/total
ER area in XLCNM patient
myoblasts treated with DMSO
(control) or VPS34-IN1 (5 mM,
2 hours). KM1288, DMSO
(n= 113), VPS34-IN1 (n= 148);
NL15, DMSO (n= 80), VPS34-
IN1 (n= 86). (E) Confocal
and STED images of DMSO or
VPS34-IN1-treated (5 mM,
2 hours) HeLa cells expressing
mEmerald-Sec61b.(F) PI(3)P
levels in fed (n= 820) or
starved (EBSS 2 hours; n=
725) HeLa cells. (G) STED
images of fed or starved live
HeLa cells stably expressing
Halo-Sec61b. See Movies 1
and 2. (H) Ratio of sheet/total
ER area in fed or starved
(EBSS 2 hours) HeLa cells
stained for calreticulin. Fed
(n= 312), Starved (n= 320).
(I) Ratio of sheet/total ER area
of HeLa cells expressing
mEmerald-Sec61bexposed to
different nutrients (2 hours). Full (n= 128), (i.e., EBSS only, n= 148), +Dialyzed
FBS (n= 160), +Glucose (4.5 g/liter) (n= 111), all amino acids (+ All AA, n= 171).
(J) Ratio of sheet/total ER area in fed or starved HeLa WT or MTM1 KO cells.
Fed WT (n= 136), KO (n= 123); Starved WT (n= 156), KO (n= 207).
(K) (Top left) Electron microscopy images of starved WT or MTM1 KO HeLa cells.
Purple, ER cisternal cross sections. (Right) Length of ER cross sections (# of
objects: WT = 1018, KO = 1070) from three cells. (Bottom left) 3D FIB-SEM
analysis. 3D reconstruction of ER (purple) and mitochondria (brown). ER sheets
are fenestrated in starved KO cells. See fig. S2G and movie S1. Scale bars:
10 mm (white), 2 mm (yellow), and 1 mm (black). n, total number of cells analyzed
from two (H) or three [(B) to (D), (F), (I), and (J)] independent experiments.
One-way ANOVA with Dunnetts multiple comparisons test (B); two-tailed
Mann-Whitney test [(C), (D), (H), (J), and (K)]; two-tailed unpaired ttest (F);
Kruskal-Wallis test with two-sided Dunns multiple comparison test (I).
**P0.01, ****P0.0001. Data are median ± interquartile range [(B) to (D),
and (H) to (K)] or mean ± SD (F).
Sheet ER ratio
AB1190
KM1288
NL15
0.0
0.2
0.4
0.6
0.8
1.0
Healthy XLCNM
(steady fed)
****
****
DMSO
VPS34In1
DMSO
VPS34In1
0.0
0.2
0.4
0.6
0.8
1.0
Sheet ER ratio
KM1288 NL15
**** ****
0.0
0.5
1.0
1.5
Sheet ER ratio
****
AB1190
siCo./siCo.
siMTM1_1/_SP
Fed
Starved
0.6
0.7
0.8
0.9
1.0
1.1
Norm. PI(3)P inte nsity
**
WTKOWT
KO
0.0
0.2
0.4
0.6
0.8
1.0
Sheet ER ratio
Fed Starved
****
Fed
Starved
0.0
0.5
1.0
1.5
Sheet ER ratio
****
A
BD
G
IJK
H
C
EF
AB1190
Healthy myoblast
KM1288NL15
XLCNM myoblast
Halo-Sec61βInset
Confocal STED
Inset
mEmerald-Sec61βInset
Confocal STED
Inset
DMSOVPS34In1
Halo-Sec61βInset
Live-STED
WT
KO
0.0
0.5
1.0
1
4
ER length (μm)
****
Starved
Starved-WT Starved-KO
FIB-SEM SEM
Full
-
+Dialyzed FBS
+Glucose (4.5 g/liter)
+All AA
0.0
0.2
0.4
0.6
0.8
1.0
Sheet ER ratio
ns
ns
****
EBSS
RESEARCH |RESEARCH ARTICLE
Downloaded from https://www.science.org at Leibniz-Institut für Molekulare Pharmakologie on December 16, 2022
this may underlie the structural defects in ER
morphology found in muscle cells and tissue
from XLCNM patients (16,17,30,31).
Starvation-induced PI(3)P hydrolysis by
MTM1 at ERearly endosome contacts mediates
ER reshaping
We next sought to understand how PI(3)P hy-
drolysis by MTM1 mediates the starvation-
induced conversion of peripheral ER tubules
into flat uniform sheets. As the ER is subject
to turnover by means of autophagy (32), we
probed whether blockade of autophagosome
formation interferes with starvation-induced
ER sheet expansion. However, ER sheet expan-
sion in starved cells was unaffected by phar-
macological or genetic inhibition of autophagy
in the presence of VPS34-IN1 (15)orknock-
down of the essential autophagy factor ATG5
(fig. S3, A to D). Moreover, loss of MTM1, that
is, a condition in which starvation-induced ER
sheet expansion is perturbed, did not alter the
ability of starved cells to form microtubule-
associated protein 1 light chain 3 (LC3)positive
autophagosomes (fig. S3, E and F), the over-
all levels of the autophagy marker LC3-II (fig.
S3G), or the subcellular localization and ac-
tivation of transcription factor EB (TFEB), a
master regulator of autophagy gene expres-
sion (33) (fig. S3, L and M). MTM1 KO cells
displayed slightly reduced levels of the sheet-
localized ER-phagy receptor FAM134B under
fed and starved conditions (fig. S3, H and I).
Reduced levels of FAM134B have been shown
to result in sheet ER expansion (32), a pheno-
type opposite to the expansion of the tubular
ER observed in MTM1 KO cells (Fig. 1). The
starvation-induced increase in the sheet ER
ratiothusappearstobeindependentofFAM134B-
mediated ER-phagy. Although mechanistic
target of rapamycin complex 1 (mTORC1) ac-
tivity was elevated in fed MTM1 KO cells
[consistent with observations in MTM1 KO
mice (34)], loss of MTM1 did not affect sup-
pression of mTORC1 activity in starved cells
(fig. S3N). Finally, lumenal ER calcium levels
were not significantly altered (fig. S3O), and
no signs of an ER stress response were de-
tected in MTM1 KO cells (fig. S3P). We con-
clude that PI(3)P hydrolysis by MTM1 in starved
cells controls ER shape independently of au-
tophagy, the ER stress response, and ER cal-
cium homeostasis.
Given that PI(3)P is a hallmark of early endo-
somes (10,35) and that the ER makes extensive
contacts with other organelles including the
plasma membrane (36), endosomes, lysosomes,
and mitochondria (37), we hypothesized that
MTM1 specifically acts on peripheral early
endosomes in starved cells and thereby con-
trols ER morphology (Fig. 2A), for example,
through membrane contacts. To test this, we
generated a cell line stably expressing a chimera
between the early endosomal protein Rab5A
and the biotinylating enzyme ascorbate per-
oxidase 2 (APEX2) (38) under the control of
a doxycycline-inducible promoter (fig. S4A).
Proximity labeling and affinity capture re-
vealed a prominent starvation-induced enrich-
ment of endogenous MTM1 on early endosomes,
while early endosomal antigen 1 (EEA1), a PI(3)P-
binding scaffold protein, was depleted in early
endosomes (Fig. 2B and fig. S4B). Nutrient
starvation thus induces the recruitment of
MTM1 to Rab5-containing early endosomes,
likely resulting in PI(3)P hydrolysis and ER
sheet expansion. We further probed this model
by a chemical genetic approach that capitalizes
on the FRB/FKBP system, which enables the
artificial tethering of organelles via rapalog-
induced heterodimerization of chimeras be-
tweenFK506-bindingprotein(FKBP)andthe
FKBP-rapamycin binding (FRB) domain of
mTOR (39). We found that acute rapalog-
induced recruitment of active FKBP-MTM1
to early endosomes tagged with FRB-Rab5A
resulted in massive ER tubule-to-sheet con-
versioninfedcells(Fig.2Candfig.S4C),phe-
nocopying starvation-induced PI(3)P depletion.
Recruitment of active MTM1 to lysosomes (fig.
S4,DandE)orearlyendosomalrecruitment
of catalytically inactive mutant MTM1 (CS) did
not affect ER shape (Fig. 2C). We conclude that
recruitment of catalytically active MTM1 to
early endosomes drives ER sheet expansion
to mount the cellular response to nutrient
starvation.
The role of endosomal MTM1 in controlling
ER shape might depend on the formation of
hitherto molecularly undefined membrane con-
tacts between the tubular ER and highly motile
early endosomes that could provide a force
that aids in keeping ER tubules under tension.
In support of this hypothesis, we observed
PI(3)P and the early endosome marker Rab5A
to tightly colocalize with the tubular ER net-
work in the cell periphery (Fig. 2D). Moreover,
tracking early endosomes and the ER network
in live cells by high-speed spinning disk con-
focal imaging showed that forming ER tubules
are tightly associated with motile early endo-
somes (Fig. 2E and Movie 3), a conclusion fur-
ther supported by the close association of ER
tubuleswithearlyendosomesmarkedbyin-
ternalized bovine serum albumingold (BSA-
gold) in electron micrographs (Fig. 2F). Hence,
the large population of peripheral early endo-
somes may serve to promote ER tubules. Con-
sistently, we found that acute depletion of early
endosomes from the cell periphery by rapalog-
induced endosomal recruitment of the micro-
tubule minus enddirected dynein adaptor
BICD2 caused a near complete loss of the tu-
bular ER network and a concomitant accumu-
lation of perinuclear sheet ER (Fig. 2, G and H;
fig. S4F; and movie S2).
We reasoned that MTM1-mediated hydrolysis
of PI(3)P at early endosomes during starvation
maycontrolERshapeeitherby(i)controlling
early endosome motility or (ii) altering the
number or stability of hitherto unknown phys-
ical contacts between early endosomes and
the ER (fig. S4G). As early endosome motility
was unaffected by nutrient starvation (fig. S4,
H and I), we followed the alternative hypoth-
esis that nutrient regulation of early endo-
somal PI(3)P controls membrane contact sites
between early endosomes and the ER and,
thereby, ER shape. To monitor such contacts,
we determined the fractional overlap between
the peripheral ER and early endosomes by
multicolor STED microscopy (fig. S4J). Starva-
tion reduced the fractional overlap of the ER
with early endosomes (Fig. 2I). The total con-
tact area between the ER and early endosomes
marked by internalized BSA-gold was also
found to be significantly reduced in starved
cells analyzed by electron microscopy (Fig. 2J).
To confirm these findings by another inde-
pendent approach, we generated a stable cell
line that coexpresses equimolar ratios of an
ER membrane targeting domain (ERM) fused
Jang et al., Science 378, eabq5209 (2022) 16 December 2022 3of16
Movie 1. Live imaging of ER dynamics in fed or
starved cells. Live-cell spinning disk confocal
imaging of fed or starved HeLa cells expressing
mEmerald-Sec61b. Videos were acquired at
1 frame/5 min for 150 min and correspond to
fig. S1K.
Movie 2. Live STED imaging of tubular or sheet
ER from fed or starved HeLa cells expressing
Halo-Sec61 b.Live-cell STED imaging of starved
HeLa cells expressing Halo-Sec61b. Videos were
acquired at 1 frame/3 s for 30 s and are related
to Fig. 1G.
RESEARCH |RESEARCH ARTICLE
Downloaded from https://www.science.org at Leibniz-Institut für Molekulare Pharmakologie on December 16, 2022
Jang et al., Science 378, eabq5209 (2022) 16 December 2022 4of16
PI(3)P MergeHalo-Rab5A
Merge with
mEmerald-Sec61β
MTM1
Binding to Rab5A-EE
Fed
Starved
0
1
2
3
4
*
Fed
Starved
EEA1
0.0
0.5
1.0
1.5
****
Binding to Rab5A-EE
BICD2-FRB +
FKBP-Rab5
-
rapalog
0.0
0.2
0.4
0.6
0.8
1.0
Sheet ER ratio
****
FKBP-MTM1 WT
FKBP-MTM1 CS
FRB-Rab5A
-
rapalog -
rapalog
0.0
0.2
0.4
0.6
0.8
1.0
Sheet ER ratio
**** ns
0.0
0.2
0.4
0.6
0.8
1.0
Sheet ER ratio
**** ns
-
rapalog -
rapalog -
rapalog -
rapalog
Fed Starved Fed Starved
Endosomal
FKBP-Rab5
Cytosolic
Rab5-FKBP
FKBP-Rab5
+rapalog
BICD2-
FRB
EE arrest at
perinuclear
Tubular ER
PI3P
EE
Fed
Sheet ER
MTM1
Starved
rapalog
Starved
S
tarve
d
Rab5
FRB FKBP
rapalog
ER
Fed
Fe
d
MT
MT
Tubular
ER
PI3P
EE
Fed
Fe
d
GFP 11
GFP1-10
MTM1
Starved
S
tarve
d
BD
F
C
G
H
K
I
L
J
M
NO
P
A
Norm. split-GFP do ts
(Starved / Fed)
siCo.
siMTM1_1
0.5
0.6
0.7
0.8
0.9
1.0
1.1
# split- GFP dots / Cell
0
50
100
150
200
250 *
Fed
Starved
0 s 0.4 s 0.8 s 1.2 s
mEmerald-
Sec61β
mCh-Rab5
E
HA
Endosomal
FKBP-Rab5 HA
Fed
-rapalog
Starved
-rapalog
Cytosolic
Rab5-FKBP
Fractional overlap between
peripheral ER and EE
Fed
Starved
0.00
0.01
0.02
0.03 *
(STED)
ER-EE contact length (nm)
0
200
400
600
(EM)
*
Fed
Starved
Fig. 2. Starvation-induced PI(3)P hydrolysis by MTM1 at ERearly
endosome contacts mediates ER reshaping. (A) PI(3)P-positive early endo-
somes (EE) may control the tubular ER. (B) Starvation-induced recruitment
of MTM1 to EE. EE-associated MTM1 (left) and EEA1 (right) in fed versus starved
HeLa cells. n= 3 independent experiments. (C) Ratio of sheet/total ER in
HeLa cells coexpressing FRB-iRFP-Rab5A and mRFP-FKBP-MTM1 WT (wild-type)
or CS (inactive mutant) ± rapalog. FKBP-MTM1 WT (n= 162), rapalog (n=
265), FKBP-MTM1 CS (n= 138), rapalog (n= 187). (D) Confocal images of HeLa
cells stably coexpressing Halo-Rab5A (EE) and mEmerald-Sec61b(ER) and
stained for PI(3)P. (E) Time-lapse confocal images of HeLa cells coexpressing
mEmerald- Sec61b(ER) and mCherry-Rab5 (EE). Yellow arrowheads mark motile
EE. (F) Electron micrographs illustrating tubular ER (blue) contacts with EE
(orange) marked by internalized BSA-gold. (G) Rapalog-induced acute depletion
of EE from the periphery. See movie S2. (H) Ratio of sheet/total ER in mock-
(n= 177) versus rapalog-treated (n= 141) HeLa cells as in (G). (I) Fractional
overlap between the peripheral ER and EE in fed or starved cells determined
by STED microscopy. Fed = 118 ROIs; Starved = 166 ROIs (40 to 50 cells from
four experiments). (J) Morphometric analysis of contact length (nanometers)
between ER and BSA-gold labeled EE in fed (42 endosomes) versus starved
cells (43 endosomes). (K) Reconstitution of split-GFP fluorescence by ER-EE
contacts. (L) Normalized number of ER/EE contacts in fed (n=238)versus
starved (n= 257) HeLa cells. (M)RescueofER/EEcontacts in starved MTM1-
depleted HeLa cells. siCo Fed (n=123),Starved(n=77),siMTM1_1Fed
(n=111),Starved(n=85).(N) ERM-FRBrapalogFKBP-Rab5A ER/EE
synthetic tether. (O) Ratio of sheet/total ER in fed or starved cells from
(N). Endosomal FKBP Rab5 Fed (n= 68), rapalog (n=70),Starved(n=
92), rapalog (n= 86); Cytosolic Rab5-FKBP Fed (n= 97), rapalog (n=90),
Starved (n=104),rapalog(n=120).(P) Confocal images from (O) stained
for HA to mark the ER, gray. Scale bars: 2 mm (yellow), 1 mm(black),and
100 nm (white). n, number of cells analyzed from two [(M) and (O)] or three
[(C), (H), and (L)] independent experiments. Two-tailed unpaired ttest
[(B), (C), (I), (L), and (O);] two-tailed Mann-Whitney test [(H) and (J)];
*P0.05, ****P0.0001; ns, nonsignificant. Data are mean ± SD [(B), (I),
(L), and (M)] or median ± interquartile range [(C), (H), (J), and (O)].
RESEARCH |RESEARCH ARTICLE
Downloaded from https://www.science.org at Leibniz-Institut für Molekulare Pharmakologie on December 16, 2022
to GFP
11
(40)alongwithaGFP
1-10
-Rab5A chi-
mera under the control of a doxycycline-inducible
promoter (fig. S4, K and L). Formation of mem-
brane contacts between early endosomes and
the ER in fed cells reconstitutes GFP fluores-
cence (Fig. 2K). ERearly endosome membrane
contacts were greatly reduced in starved cells
(Fig. 2L and fig. S4M), and this effect was re-
verted completely by depletion of MTM1 (Fig.
2M). These data suggest that MTM1-mediated
hydrolysis of PI(3)P at early endosomes reduces
the contacts between these organelles and the
ER (Fig. 2K). We directly tested this model by
examining the effects of semisynthetic tether-
ing of early endosomes to the ER using chem-
ical genetics. Acute rapalog-induced formation
of ERearly endosome tethers mediated by
ERM-FRBrapalogFKBP-Rab5A potently sup-
pressed starvation-induced ER tubule-to-sheet
conversion (Fig. 2, N to P; fig. S5B; and movie
S3) but had no effect on early endosome
motility (fig. S5A). In contrast, soluble, non-
lipidated Rab5A (Rab5-mRFP-FKBP) or cyto-
solic mRFP-FKBP had no effect on ER shape
(Fig. 2O and fig. S5, B and C). Extending the
length of the semisynthetic tether well beyond
the typical distance of 20 to 30 nm observed
for organelle contacts in vivo (36,37)didnot
affect its ability to prevent starvation-induced
ER reshaping (fig. S5, D and E), suggesting
that the exertion of a physical pulling force is
critical for the regulation of ER shape by mem-
brane contacts with early endosomes. Further-
more, the capability of ERM-FRBrapalog
FKBP-Rab5A tethers to prevent starvation-
induced ER tubule-to-sheet conversion was
unaffected by depletion of PI(3)P, indicating
that PI(3)P acts upstream of ERearly endo-
some membrane contact site formation (fig. S5,
F and G). Other organelleslate endosomes
and lysosomes, in particularhave also been
shown to form numerous contacts with the ER
(41,42).Earlyendosomesdisplayanevendis-
tribution throughout cells and into the periphery
where the tubular ER is located and outnumber
lysosomes by up to an order of magnitude. In
contrast, most lysosomes are concentrated in
the perinuclear area at steady state (fig. S6, A
to D). As a consequence, the total number of
membrane contacts between the ER and early
endosomes greatly exceeds that of the ER with
lysosomes (fig. S6, E to H), in spite of the high
relative fraction of lysosomes in touch with the
ER (42,43). Redistribution of lysosomes to the
perinuclear area either through acute rapalog-
induced dynein adaptor recruitment or deple-
tion of PI(3,5)P
2
did not affect ER shape (fig.
S6, I to L; fig. S1H; and movie S4). Sustained
loss of the lysosomal kinesin adaptor Arl8b,
a protein essential for lysosome dispersion,
only marginally increased the sheet ER frac-
tion [fig. S6, M to O; see also (44)], whereas
depletion of the ER-lysosome contact site pro-
tein protrudin (41) was without effect. Hence,
the tubular ER is largely controlled by its mem-
brane contacts with early endosomes (as
demonstrated in this study) and a smaller con-
tribution from lysosomes, possibly dependent
on cell type or conditions (44). Taken together,
our findings unravel a crucial role for MTM1-
mediated PI(3)P hydrolysis in the reduction of
membrane contacts between the ER and early
endosomes to reshape the ER in response to
changing nutrient levels.
Contacts formed by ER membrane
proteinmediated recognition of early
endosomal PI(3)P controls the tubular ER
To identify the molecular machinery that teth-
ers ER tubules to early endosomes in fed cells,
wecapitalizedontheobservationthatinWT
but not in MTM1 KO cells ERearly endosome
contacts are reduced under conditions of nu-
trient starvation (Fig. 3A). Starved WT or MTM1
KO HeLa cells inducibly expressing APEX2-
Rab5A were subjected to combined proximity
labeling and affinity purificationmass spec-
trometry to probe the molecular environment
of Rab5A-containing early endosomes. Subse-
quent biochemical fractionation to enrich for
ER membrane proteins (fig. S7, A to C, and
table S2) and comparison with previously iden-
tified ER surface proteins (45) identified sev-
eral putative ER transmembrane proteins that
might serve as early endosome tethers (fig.
S7D). Analysis of the phenotypic consequences
of cellular depletion of these factors revealed
that only loss of ER ribosome-binding protein
1 (RRBP1) promoted ER tubule-to-sheet con-
version in fed cells (fig. S7E). RRBP1 is local-
ized exclusively to ER membranes (Fig. 3, B
andC)andhasbeenassociatedwithchanges
in ER morphology, although no consensus re-
garding its precise function exists (4648). ER
tubule-to-sheet conversion was exacerbated by
concomitant loss of RRBP1 and its close para-
log kinectin 1 (KTN1) (Fig. 3D and fig. S8, A to
C). Moreover, loss of RRBP1 and KTN1 potent-
ly reduced the number of ERearly endosome
membrane contacts in fed cells (Fig. 3E), sug-
gesting that RRBP1 and KTN1 regulate ER
shapebyactingastethers for early endosomes.
KTN1 and RRBP1 harbor functionally unchar-
acterized lysine-rich regions (LR1, -2, and -3)
in their cytoplasmic domains (Fig. 3F and fig.
S8,AandB)thatcouldconceivablyattachto
PI(3)P. Notably, KTN1 was found to be in con-
tact with PI(3)P-enriched early endosomes by
spatiotemporally resolved interaction proteo-
mics using 2xFYVE-APEX as a probe (49). We
found recombinant RRBP1-LR1-3 (fig. S8D) to
bind to PI(3)P with preference over both PI(4)
P and PI(3,4)P
2
(Fig.3,GandH).Thesedata
suggest that RRBP1 might recognize PI(3)P on
early endosomes to form tethers with the ER.
Consistent with this model, we observed that
expression of a mini-RRBP1 truncation pro-
tein variant harboring only the lysine-rich
cytoplasmic domain fused to its ER transmem-
brane anchor sufficed to restore a normal
tubular ER network in HeLa cells depleted of
endogenous RRBP1 and KTN1 (Fig. 3I, WT).
Deletion of lysine-rich regions 1 or 3 or of
the transmembrane anchor rendered trun-
cated mini-RRBP1 inactive (Fig. 3I and fig. S8,
E and F). These data suggest that RRBP1 and
KTN1 mediate recognition of early endosomal
PI 3-phosphates and, possibly, additional fac-
tors, to facilitate contact site formation be-
tween the ER and early endosomes, which
control the tubular ER network in human cells.
MTM1-dependent ER reshaping is required
for mitochondrial network formation
during starvation
Several mechanisms have been shown to con-
tribute to mitochondrial morphogenesis, includ-
ing membrane contacts between mitochondria
and lysosomes (50), late Golgi-derived vesicles
(51), and the tubular ER (37,52). The tubular ER
also directly promotes mitochondrial fission
(53) and acts as a donor organelle for the for-
mation of lipid droplets (LDs) (54,55) that
serve as an energy reservoir in fed cells and
tissues. Under conditions of starvation (e.g.,
deprivation of glutamine and other amino
acids), mitochondria undergo hyperfusion into
tubular networks to protect themselves from
mitophagy (56) and to enable efficient utiliza-
tion of fatty acids (57).
On the basis of these prior works, we hy-
pothesized that the MTM1-mediated starvation-
induced reshaping of the tubular ER into
sheets (Fig. 1) might serve to enable cells to
metabolically adapt to altering nutrient envi-
ronments, for example, by altering mitochon-
drial organization. To test this hypothesis, we
examined the effect of MTM1 loss on mito-
chondrial morphology and respiratory func-
tion. We found mitochondria to become
hyperfused in starved WT but not in MTM1 KO
cells or in WT cells depleted of endogenous
Jang et al., Science 378, eabq5209 (2022) 16 December 2022 5of16
Movie 3. Live imaging of early endosome motility and the tubular ER. Live-cell high-speed spinning
disk confocal imaging of fed HeLa cells expressing mEmerald-Sec61b(labeled gray) and mCherry-Rab5
(labeled yellow). Videos were acquired at 1 frame/0.4 s for 1 min and correspond to Fig. 2E.
RESEARCH |RESEARCH ARTICLE
Downloaded from https://www.science.org at Leibniz-Institut für Molekulare Pharmakologie on December 16, 2022
MTM1 (Fig. 4, A and B, and fig. S9A). Further
ultrastructural analysis by 3D FIB-SEM re-
vealed the formation of an extensive cell-wide
network of hyperfused mitochondria in starved
WT cells. In contrast, starved MTM1 KO cells
displayed an accumulation of small spherical
mitochondria suggestive of exacerbated mito-
chondrial fission and defects in cristae mor-
phology (Fig. 4, C to E, and fig. S9, B to D),
whereas the total mitochondrial volume frac-
tion was unchanged (Fig. 4F). Reduced mito-
chondrial network formation and defective
ER reshaping were also observed in starved
XLCNM patientderived myoblasts (fig. S9,
G and H). Defective mitochondrial morpho-
genesis was not a consequence of the altered
expression of mitochondrial fusion- or fission-
related proteins (58) such as mitofusin 1/2,
OPA1 (59), and DRP1 or its hyperactive form
(pS616-DRP1) in MTM1 KO cells (fig. S3, J
and K). Moreover, acute rapalog-induced for-
mation of ERearly endosome tethers to inhibit
loss of ER tubules prevented the starvation-
induced formation of a hyperfused mitochon-
drial network in WT cells (Fig. 4G and fig. S9,
E and F), thereby phenocopying MTM1 loss.
Conversely, reducing membrane contacts be-
tween the ER and early endosomes by deple-
tion of RRBP1 and KTN1 led to the formation
of hyperfused mitochondrial networks and
expansion of the sheet ER in MTM1 KO cells
(fig. S9, I to K). RRBP1 and KTN1 therefore
act downstream of MTM1-mediated PI(3)P
hydrolysis.
If reshaping of the tubular ER into sheets
causally underlies the formation of a hyper-
fused mitochondrial network, one would expect
experimental manipulations that reshape the
ER to affect mitochondrial morphology. Con-
sistently, we found that cellular depletion of
either Rab10, a factor required for maintenance
of the tubular ER (60); RRBP1 and KTN1; or the
tubular ER-shaping protein reticulon 4 (61)
conditions that result in ER sheet expansion
(fig.S10,A,C,D,andF)cause the formation
of an enlarged mitochondrial network in fed
cells(fig.S10,B,C,E,andF).Mitochondrial
hyperfusion was further induced by overex-
pression of the ER sheetinducing membrane
protein Climp63 (46) (fig. S10, G and H). Final-
ly, we observed increased rates of mitochon-
drial fission in starved MTM1 KO cells (fig.
S10, I to K, and movie S5), that is, conditions
under which ER tubules accumulate. Col-
lectively, these findings establish that the
starvation-induced conversion of ER tubules
to sheets by MTM1-mediated hydrolysis of
PI(3)P at ERearlyendosomecontactsfacili-
tates the formation of a functional mitochon-
drial network.
Defective ER morphogenesis in absence of
endosomal MTM1 impairs mitochondrial
metabolic rewiring during starvation
To probe the physiological consequences of
defective mitochondrial morphogenesis, we
monitored mitochondrial oxygen consumption
and mitochondria-driven adenosine triphos-
phate (ATP) production in WT and MTM1 KO
Jang et al., Science 378, eabq5209 (2022) 16 December 2022 6of16
Fig. 3. Membrane contacts formed by ER ribosome-binding
protein 1 and kinectin 1mediated recognition of early
endosomal PI(3)P control the tubular ER. (A) Starvation-
induced dissociation of ERearly endosome (EE) contacts in
WT but not in MTM1 KO cells. (B) Volcano plot of proximity
biotinylated interactome of APEX2-Rab5A isolated from ER
membranes of starved MTM1 KO versus WT HeLa cells. Dark
purple: RRBP1 and KTN1. (C) Confocal images of HeLa cells
coexpressing mEmerald-Sec61b(ER) and RRBP1 full length
(FL)mCherry. (D) Ratio of sheet/total ER in fed HeLa cells
treated with control (siCo/siCo = 174) or KTN1+RRBP1 siRNAs
(siKTN1/siRRBP1 = 221). (E) Normalized number of ER/EE
contacts in control or KTN1/RRBP1-depleted fed HeLa cells.
n= 3 independent experiments. (F) Schematic illustrating
truncated mini-RRBP1 (amino acids 1 to 150). TM, trans-
membrane anchor; LR1-3, cytoplasmic lysine-rich regions 1 to 3.
(G) GST-RRBP1-LR1-3 binds to PI(3)P liposomes. Supernatant
(S) and liposomal pellet (P) fractions were analyzed by
immunoblotting for GST. No PIP, liposomes lacking phospho-
inositides. (H) Quantified data as in (G) from n= 3 independent
experiments. (I) Ratio of sheet/total ER in control (siCo/siCo)
or KTN1/RRBP1-depleted stable doxycycline-inducible HeLa
cells expressing the indicated truncated RRBP1 protein
(mini-RRBP1) variants [see (F)]. WT, wild-type V5-tagged
mini-RRBP1; delLR1, mutant RRBP1 lacking lysine-rich region 1;
delLR3, mutant RRBP1 lacking lysine-rich region 3; delTM,
mutant RRBP1 lacking its transmembrane domain. siCo/siCo
(n= 111), WT (n= 129), delLR1 (n= 102), delLR3 (n= 120),
delTM (n= 130); siKTN1/siRRBP1 (n= 274), WT (n= 271),
delLR1 (n= 326), delLR3 (n= 263), delTM (n= 288). Scale bars:
2mm (yellow). nindicates the total number of cells analyzed
from three independent experiments. Two-tailed Mann-Whitney
test (D); one-way ANOVA with Tukeys multiple comparisons
(E); one-way ANOVA with Dunnetts multiple comparisons
test (H); Kruskal-Wallis test with two-sided Dunns multiple
comparison test (I). **P0.01, ****P0.0001. Data
are median ± interquartile range [(D) and (I)]; data are mean ±
SD [(E) and (H)].
RRBP1 (FL)-mCh mEmerald-Sec61βMerged
log2 difference (KO / WT)
-log10 p-value
-6 -4 -2 0 2 4 6
0
1
2
3
4
5
KTN1
RRBP1
PI3P
ER
tether
WT MTM1 KO
MTM1
MTM1
Starved
siCo./siCo.
siKTN1/siRRBP1
0.0
0.2
0.4
0.6
0.8
1.0
Sheet ER ratio
****
0.0
0.5
1.0
1.5
2.0
2.5
Norm. split-GFP dots
** **
siKTN1/siRRBP1
siCo./siCo.
-
Starved
Fed
Fed
0.0
0.2
0.4
0.6
0.8
1.0
Sheet ER ratio
********
****
****
****
-
WT
delLR1
delLR3
delTM-
WT
delLR1
delLR3
delTM
siCo./siCo. siKTN1/siRRBP1
reconstituted
mini V5-RRBP1:
TM Lysine-rich
ER lumen
58 145
LR1 LR2 LR3
150aa
B
F
DE
C
G
H
I
A
Norm. liposome bound
0.0
0.5
1.0
1.5
****
****
****
no PIP
PI(3)P
PI(4)P
PI(3,4)P2
PI(3)P
no PIP
PI(4)P
PI(3,4)P2
SPSPSPSP
-55
GST-RRBP1
LR1-3 (1-150aa)
RESEARCH |RESEARCH ARTICLE
Downloaded from https://www.science.org at Leibniz-Institut für Molekulare Pharmakologie on December 16, 2022
cells using Seahorse technology. Starved
MTM1 KO cells displayed severely reduced
basal as well as maximal oxygen consumption
rates resulting in significantly impaired mito-
chondrial ATP synthesis (Fig. 5, A and B, and
fig. S11, A and B). As a consequence, starved
MTM1 KO cells suffered from a pronounced
reduction of total cellular ATP (Fig. 5C). Sim-
ilar results were observed in cells, in which the
ER was artificially tethered to early endo-
somes (Fig. 5D). No significant differences in
mitochondrial basal oxygen consumption or
mitochondrial ATP synthesis were detected
in fed MTM1 KO cells, whereas the maximal
oxygen consumption rate was marginally de-
creased (fig. S11, C to E). The mitochondrial
membrane potential was unaffected by MTM1
loss, irrespective of the nutritional status of
the cells (fig. S11F). As defective mitochondrial
morphogenesis and cellular ATP depletion
are associated with reduced cell viability, we
probed whether MTM1 KO cells might be
poised to undergo apoptosis under conditions
of limited nutrient availability. Starved MTM1
KO cells indeed suffered from increased levels
of cleaved caspase 3 and poly(ADP ribose)
polymerase (PARP), common indicators of
apoptotic cell death (fig. S11, G to I). How-
ever, defective mitochondrial morphogenesis
in MTM1 KO cells was not a secondary con-
sequence of increased apoptosis, as treatment
of KO cells with the pan-caspase inhibitor
Z-VAD-FMK effectively blocked apoptosis (fig.
S11J) but failed to rescue defects in mitochon-
drial morphology (fig. S11K). We note that sim-
ilar apoptotic phenotypes have been reported
in starved OPA1 and mitofusin 1/2 KO cells
defective in mitochondrial fusion (62).
These results indicate that MTM1-mediated
reshaping of the tubular ER into sheets is
required for the formation of a functional mito-
chondrial network and mitochondria-driven
ATP production in starved cells. Previous work
has shown that the effective transfer and uti-
lization of fatty acids (FAs), a major substrate
for mitochondrial ATP synthesis via b-oxidation,
in starved cells requires mitochondria to be
organized into a highly tubulated hyperfused
network (56,57), although the exact underly-
ing molecular mechanism is unclear. We there-
fore hypothesized that impaired mitochondrial
ATP production in MTM1 KO cells might be a
consequence of defective FA trafficking. We
directly tested this by monitoring the fate of
FAs in WT and MTM1 KO cells under different
nutrient conditions. Depending on metabolic
state, cytosolic FAs can be metabolized in mito-
chondria (e.g., during starvation) or stored in
LDs (54,57,63), which exclusively form from
the tubular ER (55,64). Consistently, we found
that in WT cells, the number of LDs inter-
mittently declined at the onset of starvation
(2 hours), likely as a result of increased mito-
chondrial b-oxidation of FAs and blocked LD
formation upon loss of the tubular ER, before
eventually rising (Fig. 5E and fig. S11, L to N)
owing to autophagy-promoted lipid buildup
during sustained long-term (6hours)starva-
tion (57). In contrast, the number and total
Jang et al., Science 378, eabq5209 (2022) 16 December 2022 7of16
0.003 7.280
Volume (μm³)
C
AB
Fed Starved
siCo.
InsetSegmented TOM20
siMTM1_1
Fed Starved
WT KO
Starved
Mean area / mitochondrion
(A.U.)
0
100
200
300
400 **** ***
****
Fed
Starved
Fed
Starved
Fed
Starved
siMTM1_1 KO
siCo.
WT
KO
Starved
0.0
0.5
1.0
4
6
8
Individual mitochondrion
volume (μm³)
Top 3%
Sum of mitochondrial
volume (μm³)
69
mito
0
5
10
15
310
mito
62
%
18
%
Top 3%
E
EtOH
Rapalog
0
200
400
600
Mean area / mitochondrion
(A.U.)
*
Starved
(ER-EE tether)
G
Mitochondrial volume fraction
per cytoplasm
0.00
0.05
0.10
0.15
ns
Starved
WT
KO
(EM)
FD
# of mitochondria
Cumulative
mitochondria volume
(μm³)
0 100 200 300 400
0
5
10
15
Starved- WT
Starved- KO
Fig. 4. MTM1-dependent ER reshaping is required for mitochondrial
network formation during starvation. (A) Confocal images of mitochondria
(TOM20) in fed or starved control (siCo) and MTM1-depleted (siMTM1_1) HeLa
cells. (Bottom) Segmented mitochondria in ROI (yellow, 15 mmby15mm).
(B) Mean area of individual mitochondrion per ROI from fed or starved WT or
MTM1 KD or KO HeLa cells. siCo Fed = 345 ROIs, Starved = 418 ROIs; siMTM1_1
Fed = 354 ROIs, Starved = 440 ROIs; KO Fed = 343 ROIs, Starved = 322 ROIs
from four independent experiments. Each ROI represents a single cell. (C)3D
rendering of mitochondria by FIB-SEM in starved WT and MTM1 KO HeLa
cells. Heat bar reflects individual mitochondrial volumes. Images are represen-
tative of three cells. (D) Cumulative plot of mitochondrial volume distribution
in starved WT versus MTM1 KO cells derived from FIB-SEM. (E) 3D volumes
of individual mitochondria as in (C) were plotted. (Inset) The largest 3% of
mitochondria occupied 62% of the total mitochondrial volume in WT (gray)
but only 18% in KO (red) cells. The total volume of mitochondria in WT cells
(69 mito = 12.95 mm
3
) versus KO cells (310 mito = 13.25 mm
3
) was unchanged.
(F) Mitochondrial volume fraction in starved WT and MTM1 KO HeLa cells (n=
10) analyzed by stereological analysis of thin-sectioned electron micrographs.
(G) Artificial ERearly endosome tethering prevents mitochondrial hyperfusion.
Mean area of individual mitochondria per ROI from starved HeLa cells ±
rapalog (Fig. 2N). EtOH = 251 ROIs, rapalog = 250 ROIs from three independent
experiments. Each ROI represents a single cell. Scale bars: 10 mm (white),
2mm (yellow), 1 mm (black). One-way ANOVA with Tukeys multiple comparisons
(B); two-tailed unpaired ttest [(F) and (G)]; *P0.05, ***P0.001,
****P0.0001; ns, nonsignificant. Data are mean ± SD [(B) and (G)] or
median ± interquartile range (F).
RESEARCH |RESEARCH ARTICLE
Downloaded from https://www.science.org at Leibniz-Institut für Molekulare Pharmakologie on December 16, 2022
volume of LDs increased in MTM1 KO cells at
the onset of starvation (Fig. 5E and fig. S11O).
LD accumulation persisted upon treatment of
MTM1 KO cells with the pan-caspase inhibitor
Z-VAD-FMK (fig. S11P). These results suggest
that starvation-induced ER tubule-to-sheet
conversion mediated by MTM1 orchestrates the
reflux of FAs to mitochondria for b-oxidation
while counteracting their storage in LDs. We
directly probed this model by monitoring FA
trafficking during early stages of starvation
using the fluorescent FA analog BODIPY 558/
568 C
12
(Red C12) (Fig. 5F). Pulse-labeled Red
C12 was efficiently transported to mitochondria
in starved WT cells (Fig. 5G and fig. S11Q), con-
sistent with prior data (57). In contrast, in starved
MTM1 KO cells, Red C12 failed to accumulate in
mitochondria and instead was targeted to LDs
(Fig. 5, G to J). Impaired mitochondrial lipid
and fatty acid catabolism was also clearly shown
by the unbiased quantitative proteomic anal-
ysis of fed or starved WT and MTM1 KO cells
using tandem mass tag (TMT) labeling. These
analyses revealed a down-regulation of pro-
teins involved in mitochondrial respiration and
transport (Fig. 5K, fig. S11R, and tables S3 and
S4), for example, mitochondrial very long-chain
specific acylcoenzyme A dehydrogenase, car-
nitine palmitoyltransferase 2, NADH:ubiquinone
oxidoreductase, and the mitochondrial protein
import factor TIMM17B (Fig. 5L), possibly as
an indirect consequence of the observed struc-
tural mitochondrial defects in starved MTM1
KO cells (Fig. 4). None of these proteins were
altered in KO cells under fed conditions (fig.
S11S), indicative of a specific defect of MTM1 KO
cells to appropriately respond to altered nu-
trient supply. These findings indicate that
MTM1 mediates reshaping of the ER at the
onset of starvation to drive the formation of a
functional mitochondrial network and facili-
tate mitochondrial b-oxidation, which sustains
ATP production. Conversely, defective ER mor-
phogenesis in the absence of endosomal MTM1
impairs mitochondrial metabolic rewiring dur-
ing starvation.
Discussion
How cells and tissues orchestrate adaptive
changes in organelle dynamics and metabo-
lism on a cell-wide scale has remained unclear.
Jang et al., Science 378, eabq5209 (2022) 16 December 2022 8of16
Fig. 5. Defective ER morphogenesis in absence of
endosomal MTM1 impairs mitochondrial metabolic
rewiring during starvation. (A) Basal mitochondrial
oxygen consumption rate (mito-OCR) of starved WT or
MTM1 KO HeLa cells. (B) Mitochondria-dependent
ATP production of starved WT or MTM1 KO HeLa cells.
(C) Normalized total cellular ATP levels of starved
WT (set to 1) or MTM1 KO HeLa cells. n= 3 independent
experiments. (D) Normalized total cellular ATP
levels of fed versus starved HeLa cells ± rapalog
(Fig. 2N). Data for fed cells (rapalog) were set
to 1. n= 5 independent experiments. (E) Number
of BODIPY 493/503labeled lipid droplets (LDs) in
fed or starved (2 hours) WT or MTM1 KO HeLa
cells. WT Fed (n= 157), Starved (n= 181); KO Fed
(n= 153), Starved (n= 148). (F) Schematic depicting
the pulse-chase assay to monitor FA mobilization.
(G) Confocal images of WT and MTM1 KO HeLa cells
pulse-labeled with RedC12 and chased for 2 hours
in EBSS and stained for TOM20. (H) Number of
RedC12-labeled LDs in WT or MTM1 KO HeLa cells
chased for 0 or 2 hours in EBSS. 0h: WT (n= 126),
KO (n= 112); 2h: WT (n= 139), KO (n= 118). (I) Pearson
correlation coefficient of RedC12-labeled FAs and
mitochondria (TOM20) from randomly selected
100 pixel by 100 pixel ROIs in WT or MTM1 KO cells
as in (H). WT = 169 ROIs, KO = 203 ROIs from three
independent experiments. (J) Schematic of trafficking
of pulse-labeled FA RedC12 in WT and MTM1 KO cells
during starvation. (K) Gene Ontology (GO) analysis
of proteins depleted in starved MTM1 KO compared
with WT HeLa cells. Terms related to fatty acid
catabolism and mitochondrial function are highlighted
in red. All of these proteins were unaltered in fed
MTM1 KO cells. (L) Volcano plot of proteins depleted
in starved MTM1 KO compared with WT HeLa cells.
Red dots, proteins enriched from GO analysis shown
in (K). Scale bars: 10 mm. nindicates the total number
of cells analyzed from three independent experiments.
Two-tailed Mann-Whitney test [(B), (E), (H), and (I)];
two-tailed unpaired ttest [(A), (C), and (D)]; ns,
nonsignificant, *P0.05, **P0.01, ***P0.001,
****P0.0001. Data are median ± interquartile range
[(A), (B), (E), (H), and (I)] or mean ± SD [(C) and (D)].
0
50
100
150
200 ****
# RedC12 LD / Cell
WT KO
WT KO
0h 2h
Chase
in EBSS:
0.0
0.2
0.4
0.6
0.8
1.0
Pearson coefficient
(TOM20, RedC12)
WT
KO
Chase 2h
in EBSS
**** Fatty acid
RedC12
β-oxidation
LD
Tubular ER
WT
KO
Starvation
(EBSS 2h)
Pulse RedC12
in EBSS (20min)
Fed
condition
Wash out
Chase
in EBSS (0, 2h)
Fed
Starved 2h
Fed
Starved 2h
0
20
40
60
80
100
WT KO
****
# Bodipy 493/503 LD / Cell
****
0
2
4
6
8
log2 difference (KO / WT)
-4 -2 0 2 4
-log10 p-value
Starved
Proteins from GO list labeled red
COX6B1
ACADVL
NDUFB11
ETFA
TIMM17B
Chase 2h
WTKO
RedC12 TOM20 Merge
WTKO
0.6
0.7
0.8
0.9
1.0
1.1
Norm. total cellular
ATP levels
Starved
**
-0.2
0.0
0.2
0.4
0.6
0.8
1.0
mito-ATP production
(pmol / min / μg)
**
WT KO
Starved
-0.5
0.0
0.5
1.0
1.5
2.0
Basal mito-OCR
(pmol / min / μg)
****
WT KO
Starved
(ER-EE tether)
ns
*
***
0.6
0.7
0.8
0.9
1.0
1.1
1.2
Norm. total cellular
ATP levels
-
rapalog
Fed
-
rapalog
Starved
AB CD
EFG
HI
J
0 4 8 12 16
-log10(P)
GO:0006721: terpenoid metabolic process
R-HSA-6809371: Formation of the cornified envelope
GO:0033539: FA β-oxidation using acyl-CoA DH
GO:0016139: glycoside catabolic process
GO:0006839: mitochondrial transport
R-HSA-6806667: Metabolism of fat-soluble vitamins
WP465: Tryptophan metabolism
GO:0051599: response to hydrostatic pressure
GO:0007029: endoplasmic reticulum organization
ko04142: Lysosome
GO:0061684: chaperone-mediated autophagy
R-HSA-1428517: TCA cycle and ETC
R-HSA-5668914: Diseases of metabolism
GO:0051262: protein tetramerization
R-HSA-9609507: Protein localization
CORUM:387: MCM complex
M5887: NABA BASEMENT MEMBRANES
ko00280: Valine, leucine and isoleucine degradation
GO:0007005: mitochondrion organization
GO:0044242: cellular lipid catabolic process
GO : Down-regulated proteome by KO upon starvation
KL
RESEARCH |RESEARCH ARTICLE
Downloaded from https://www.science.org at Leibniz-Institut für Molekulare Pharmakologie on December 16, 2022
Here we reveal a key role for early endosomal
lipid signaling mediated by MTM1, a lipid
phosphatase mutated in XLCNM in humans
(19,30,31), in controlling the tubular ER and,
thereby, mitochondrial morphology and meta-
bolic function in the acute response to fluctuating
nutrient conditions. This mechanism operates
independent of ER-phagy, a process typically
induced by prolonged starvation (65). We dem-
onstrate that starvation-induced PI(3)P hydro-
lysis by endosomal MTM1 reduces previously
undescribed membrane contacts between pe-
ripheral ER tubules and early endosomes. These
contacts act as physical tethers that may trans-
mit pulling forces from highly motile periph-
eral endosomes to the ER and are mediated by
endosomal PI(3)P binding to RRBP1, a large ER
membrane protein overexpressed in colorectal
cancer (66), and its close paralog kinectin 1.
Whether ERearly endosome membrane con-
tacts also enable material exchange in vivo as
shown for contact sites between the ER and
the plasma membrane, the trans-Golgi network,
or lysosomes (37) remains to be determined.
Loss of ER tubules, possibly in conjunction
with ER-independent mechanisms, drives mito-
chondrial network formation and, directly or
indirectly, facilitates FA transfer to mitochon-
dria to fuel b-oxidation and, thereby, mitochon-
drial ATP production to sustain cellular energy
homeostasis. The precise relationship between
mitochondrial morphogenesis, FA mobilization
to mitochondria, and mitochondrial ATP pro-
duction remains to be defined. Interestingly,
ER sheets are favored over tubules from an
energetic perspective (67,68)and,hence,
should prevail under conditions of nutrient
starvation when cellular energy levels are low.
Consistent with this, it has recently been shown
that the hepatic ER in obese mice is character-
ized by disorganized ER sheets and a predomi-
nance of ER tubules and accompanying defects
in lipid metabolism (69). Our findings thus
identify an organellar conveyor belt, in which
the tubular ERserves as a membrane conduit
that transmits nutrient-triggered changes (i.e.,
in glutamine and other amino acids) in early
endosomal PI(3)P levels to metabolic organ-
elles such as LDs and mitochondria [in agree-
ment with (56)] to enable metabolic rewiring
under conditions of limited nutrient supply
and, possibly, in cancer [e.g., when RRBP1 is
overexpressed (66)]. Defects in ER shape, mito-
chondrial morphogenesis, and cellular ATP de-
pletion caused by loss of MTM1 function can
explain the observed myofiber hypotrophy and
defective sarcoplasmic reticulum organization
in animal models of XLCNM (16,17) and in
human patients who often appear undernour-
ished (19,30,31). Furthermore, it is conceivable
that reduced contact formation between early
endosomesandERtubulesduetoMTM1-
mediated PI(3)P hydrolysis, in addition to its
effects on ER shape and mitochondrial func-
tion, may facilitate endosomal exocytosis of
b-integrins, a mechanism shown to be defec-
tive in XLCNM (15). How precisely early endo-
somal PI(3)P levels and MTM1 function are
controlled by cellular nutrient status remains
poorly understood. VPS34, the main PI(3)P-
synthesizing enzyme on endosomes has been
reported to be stimulated by fed signals (70),
that is, conditions in which MTM1 activity
is repressed.
The endosomal signaling lipidbased path-
way to control oxidative cell metabolism un-
covered in this work may synergize with other
cellular mechanisms that impinge on the dy-
namics of metabolically active organelles. For
example, it has been shown that the function
and localization of lysosomes depend on motor
proteins (8) as well as on their association with
the ER (41) and are regulated by cellular nu-
trient status (71), which in turn affects nutrient
signaling (28,72). Late endosomes (i.e., organ-
elles distinct from the Rab5-positive early en-
dosomes described here) have been shown to
undergo fission at sites of contact with the ER
that are molecularly and functionally distinct
(73)fromtheERearlyendosomecontacts
identified in this study. Finally, mitochondria
lysosome contacts have been shown to regu-
late mitochondrial fission (50). Whether any
of these contacts are subject to nutrient regu-
lation and impact on cell metabolism will need
to be addressed in future studies. Conceivably,
lipid phosphatases including other members of
the myotubularin family (26), many of which
are linked to human disease, may play crucial
physiological roles in the regulation of these
and other membrane contacts.
Materials and methods
Materials
Plasmids
Plasmids used: mEmerald-Sec61b-C1 (Addgene
#90992), mCh-Rab5 (Addgene #49201), ER-
GCaMP6-1-150 (Addgene #86918), tdTomato-
BICD2-FKBP (Addgene #64205), and mCh-Climp63
(mouse) (Addgene #136293). HA-BICD2-FRB was
kindly provided by G. G. Farías. Plasmids for
transient transfection (e.g., mCherry MTM1
WT, mRFP-FKBP-empty, mRFP-FKBP-MTM1
WT, mRFP-FKBP-MTM1 C375S, FRB-iRFP-
Rab5A, mRFP-FKBP-Rab5A, TMEM192-3xHA-
FRB, ERM-2xHA-FRB) were generated with
the pcDNA3.1(+) vector and polymerase chain
reaction(PCR) or ligation-based cloning. Note
that ERM is the N-terminal ER membrane
targeting sequence of residues 1 to 27 of ER-
resident P450 oxidase 2C1. Full-length RRBP1
was amplified from pcDNA4 HisMax-V5-GFP-
RRBP1(Addgene#92150)andinsertedinto
pcDNA3.1(+)-based mCherry expression vec-
tor with tags at its C terminus. PCR-amplified
RRBP (amino acids 1150) was inserted in
pGEX4T-1 vector by ligation-based cloning.
Plasmids for lentivirus transduction [e.g.,
mEmerald-Sec61b, mScarlet-Sec61b,3xHA-
APEX2-Rab5A, Halo-Rab5A, ERM-GFP11-p2a-
GFP1-10-Rab5A, ERM-2xHA-FRB, ERM-2xHA-
(EAAAK)x9-FRB, mRFP-FKBP-Rab5A, Rab5A-
mRFP-FKBP, V5-RRBP1 1-150aa WT, V5-RRBP1
del PB1, V5-RRBP1 del PB3, V5-RRBP1 del TM]
were generated with the pLVX-TetONE puro
vector (Takara, Catalog 631849) by PCR, re-
petitive oligo annealing, or inverse PCR with
Gibson assembly. All constructs were verified
by double-stranded DNA sequencing.
Primary antibodies for immunoblots
Anti-GAPDH (glyceraldehyde-3-phosphate de-
hydrogenase; mouse, Sigma-Aldrich, G8795,
1:1000), anti-MTM1 (goat, Invitrogen, PA5-
17972, 1:250), anti-Calnexin (rabbit, Abcam,
ab75801, 1:2000), anti-calreticulin (rabbit,
Thermo Fisher, PA 3-900, 1:1000), anti-Reticulon4
(Nogo) (mouse, Santa Cruz, sc-271878, 1:1000),
anti-RPL26 (rabbit, Proteintech, 17619-1-AP,
1:1000),anti-RPL7 (rabbit, Proteintech, 14583-
1-AP, 1:1000), anti-EEA1 (mouse, BD biosciences,
610456, 1:500), anti-Histone H1 (rabbit, Abcam,
ab17729, 1:1000), anti-GST (mouse, Thermo
Fisher, MA4-004, 1:1000), anti-KTN1 (rabbit,
Proteintech, 19841-1-AP, 1:500), anti-RRBP1
(rabbit, Proteintech, 22015-1-AP, 1:500), anti-
Clim63 (mouse, Enzo Life Sciences, ENZ-
ABS669-0100, 1:2000), anti-TOM20 (mouse,
Santa Cruz, sc-17764, 1:200), anti-p-DRP1 (S616)
(rabbit, Cell signaling, 3455S, 1:500), anti-
DRP1 (rabbit, Abcam, ab184247, 1:500), anti-
MFN1 (rabbit, Proteintech, 13798-1-AP, 1:500),
anti-MFN2 (rabbit, Proteintech, 12186-1-AP,
1:500), anti-OPA1 (rabbit, Proteintech, 27733-
1-AP, 1:500), anti- GRP78/BIP (rabbit, Proteintech,
11587-1-AP, 1:3000), anti-CHOP (rabbit, Pro-
teintech, 15204-1-AP, 1:500), anti-LC3-II (mouse,
MBL, M152-3, 1:200), anti-Cleaved PARP(Asp214)
(rabbit, Cell Signaling, 9541S, 1:1000), anti-
Cleaved Caspase3(Asp175) (rabbit, Cell Signal-
ing, 9661T, 1:250), anti-V5 (mouse, Invitrogen,
P/N 46-0705, 1:1000), anti-Phospho-p70 S6
Kinase (Thr389) (rabbit, Cell signaling, 9205L,
1:500), anti-p70 S6 KinaseAntibody(rabbit,Cell
Signaling, 9202L, 1:1000), anti-ATG5 (rabbit,
Proteintech, 10181-2-AP, 1:1000).
Primary antibodies for
immunofluorescence
Anti-calreticulin (rabbit, Thermo Fisher, PA
3-900, 1:200), anti-calreticulin (rabbit, Abcam,
ab92516, 1:200), anti-Rab5A (mouse, BD
biosciences, 610724, 1:100), anti-GFP (mouse,
Invitrogen, A-11120, 1:500), anti-LC3-II (mouse,
MBL, M152-3, 1:100), anti-HA (mouse, Santa
Cruz, sc-7392, 1:500), anti-RFP (rabbit, MBL,
PM005, 1:400), anti-LAMP1 (mouse, BD bio-
sciences, 555798, 1:1000), anti-LAMP1 (rabbit,
Cell signaling, 9091P, 1:300), anti-V5 (mouse,
Invitrogen, P/N 46-0705, 1:200), anti-TOM20
(mouse,SantaCruz,sc-17764,1:500),anti-
TOM20 (rabbit, Abcam, ab186734, 1:500),
Jang et al., Science 378, eabq5209 (2022) 16 December 2022 9of16
RESEARCH |RESEARCH ARTICLE
Downloaded from https://www.science.org at Leibniz-Institut für Molekulare Pharmakologie on December 16, 2022
anti-TFEB(rabbit, Biomol, A303-673A, 1:200),
Streptavidin, Alexa Fluor647conjugate(Thermo
Fisher, S21374).
siRNAs
The small interfering RNAs (siRNAs) used
were: Scrambled siControl 5-CGUACGCG-
GAAUACUUCGA-3, or Sigma MISSION Uni-
versal Negative Control (SIC001), siMTM1-1
5-GATGCAAGACCCAGCGTAA-3, siMTM1-2
5-TATGAGTGGGAAACGAAATAA-3,siMTM1-
SP (ON-TARGETplus SMARTpool, Dharmacon,
L-008036-00-0005), siMTMR1-1 5-GAGA-
TAGTGTGCAAGGATA-3, siMTMR1-2 5-CG-
CTGATACCAACAAGACAAA-3,siMTMR2-1
5-GGACATCGATTTCAACTAA-3,siMTMR2-2
5-CGGCCAAGTGTTAATGCTGTT-3, siMTMR2-
35-GTAGAAAGTCTTCGGAATTTA-3,
siMTMR6-1 5-GGACTACAAGATTTGTGAA-3,
siMTMR6-2 5-CGGGACTACAAGATTTGTGAA-
3,siMTMR12-15-CCAGGTGAACAGCTGCTTT-3,
siMTMR12-2 5-GCAAGAGAATTAGCAAACTTA-
3, siMTMR12-3 5- CGCTTCAAACATCAACGA-
CAA-3, siKTN1 (ON-TARGETplus SMARTpool,
Dharmacon, L-010605-00-0005), siERLIN1
(ON-TARGETplus SMARTpool, Dharmacon,
L-015639-01-0005), siERLIN2 (ON-TARGETplus
SMARTpool, Dharmacon, L-017943-01-0005),
siRRBP1 (ON-TARGETplus SMARTpool,
Dharmacon, L-011891-02-0005), siOSBPL8
(ON-TARGETplus SMARTpool, Dharmacon,
L-009508-00-0005), siITPR2 (ON-TARGETplus
SMARTpool, Dharmacon, L-006208-02-0005),
siARL8B (ON-TARGETplus SMARTpool,
Dharmacon, L-020294-01-0005), siRTN4
(ON-TARGETplus SMARTpool, Dharmacon,
L-010721-00-0005), siRab10 (ON-TARGETplus
SMARTpool, Dharmacon, L-010823-00-0005),
siProtrudin 5-CTTCTTGATCCAGCTGGAGG-
3, siATG5 (ON-TARGETplus SMARTpool,
Dharmacon, L-004374-00-0005).
Cell culture
HeLa, human embryonic kidney 293 T
(HEK293T), and Cos7 cells were obtained
from ATCC. Cells were cultured in Dulbeccos
modified Eaglesmedium(DMEM)high
(4.5 g/liter glucose, Thermo Fisher, 41965062)
containing 10% fetal bovine serum (FBS;
Thermo Fisher, 10270106) and 50 units/ml
penicillin, 50 mg/ml streptomycin (Thermo
Fisher, 15070063). Cells were routinely tested
for mycoplasm contamination. The human
myoblast cell line KM1288 was derived from
the deltoid muscle of a patient with XLCNM
and carries a missense genomic mutation
(c.205<T) in exon 4 of the MTM1 gene (20).
The human myoblast cell strain NL15 was
derived from the quadriceps of a patient with
XLCNM and carries the missense genomic
mutation R241C in the MTM1 gene (21). The
control myoblast cell line AB1190 was derived
from the paravertebral muscle of a healthy
individual. AB1190 and KM1288 were immor-
talized as described before (74). Myoblasts were
cultured in the homemade medium: 1 volume
medium 199 (Thermo Fisher, 41150020) + 4
volumes DMEM high, supplemented with
20% FBS, Fétuin (25 mg/ml, Thermo Fisher,
10344026), human epidermal growth factor
(5 ng/ml, Thermo Fisher, PHG0311), basic
fibroblast growth factor (0.5 ng/ml, Thermo
Fisher, PHG0026), insulin (5 mg/ml, Sigma,
91077C-1G), and dexamethasone (0.2 mg/ml,
Sigma, D4902-100mg). To induce starvation, cells
were washed with prewarmed Earles balanced
salt solution (EBSS; Thermo Fisher, 24010043)
five times for 10 s each and then incubated
in EBSS in 5% CO
2
at 37°C for 2 hours, unless
indicated otherwise. This washing step is im-
portant to fully remove remaining nutrients
and growth factors. For steady-state (fed) con-
ditions, cells were incubated overnight in fresh
complete DMEM medium supplemented with
10% FBS. For the experiment in fig. S1M, EBSS
was supplemented, as indicated, with 5% di-
alyzed FBS (One Shot format, Gibco, A3382001)
or insulin (final at 5 mg/ml) or glucose (final at
4.5 g/liter) or sodium pyruvate (final at 1 mM)
(Gibco, 11360070) or MEM essential amino
acids (50x) solution (Sigma, M5550) (final 1x
dilution) or MEM solution or nonessential
amino acids (100x) (Gibco, 11140050) (final 1x
dilution) or L-glutamine (final at 4 mM) (Gibco,
25030081).
CRISPR-Cas9mediated genome engineering
Guide RNAs targeting genomic human MTM1
exon 2 (sgMTM1: 5AGTTGATGCAGAAGC-
CATCC 3) was cloned into Lenti-CRISPRv2
(Addgene plasmid # 52961). HeLa cells were
transfected using FuGene-6 as a transfection
reagent. Cells were t selected with puromycin
(2 mg/ml) for 72 hours. Surviving cells were
diluted and plated into 96-well plates with a
density of 1 cell per well. Expanded colonies
were screened by immunoblotting using anti-
MTM1 antibodies.
Generation of doxycycline-inducible stable
cell lines
Lentivirus was generated by transient trans-
fection of HEK293T cells seeded in 10-cm cell
culture plates at 80 to 90% confluency with
pCMV delta R8.2 (3.5 mg), VSV-G (0.5 mg), and
pLVX-TetONE puro-based constructs (4 mg)
combined with 16 ml of JetPrime in 400 ml
JetPrime buffer. After 16 hours of transfection,
cells were replenished with 7 ml fresh DMEM.
After 48 hours, the supernatant was collected,
and cellular debris was removed by centri-
fugation (3000 rpm, 20 min). Viral super-
natant (4 to 5 ml) was added to the cells with
polybrene (Merck, Cat.#TR-1003-G) at 10 mg/ml.
Cells were incubated with virus for 16 hours
and then replenished with fresh DMEM con-
taining puromycin (2 mg/ml) followed by selec-
tion for 2 to 3 days. After selection, cells were
stabilized for 3 to 4 days in culture without
puromycin. For the induction of protein ex-
pression, doxycycline (1 mg/ml) was added for
16 hours.
Transient transfection
HEK293T and HeLa cells and myoblast cells
were transfected with plasmids using jetPRIME
(PolyPlus, 101000001) or FuGene-6 (Promega,
E2691) or ViaFect (Promega, E4981) according
to the manufacturers protocol, respectively
[e.g., DNA (micrograms): reagent (microliters)
ratio of 1:2]. After 18 to 24 hours of transfection,
cells were further treated and analyzed. For
siRNA transfection, cells were transfected with
the indicated siRNA (20 nM) using jetPRIME
according to the manufacturers protocol. After
48 hours of transfection, cells were further
treated and analyzed.
Dyes and pharmacological inhibitors
Dyes
TMRE (tetramethylrhodamine ethyl ester per-
chlorate): Cell signaling, Mitochondrial mem-
brane potential assay kit, #13296, 200 nM).
Lysotracker Red DND-99: Thermo Fisher,
L7528, 1 mMfor45min.Notethatforfixedcell
samples, buffer containing detergents should
not be used, in order to preserve lysotracker
staining. BODIPY 493/503: Thermo Fisher,
D3922, 2 mM for 20 min in serum free media.
No detergent-containing buffers should be used.
Before treatment, cells were washed twice with
serum-free media. BODIPY 558/568 C
12
:Thermo
Fisher, D3835, 1 mM for 20 min in serum-free
media. Before labeling, cells were washed twice
with serum-free media. For TOM20 antib ody
co-staining, fixed samples were permeabilized
with 20 mM digitonin for 5 min.
MitoTracker Deep Red FM: Thermo Fisher,
M22426, 100 nM for 30 min in full medium.
CellMask Deep Red Plasma membrane Stain:
Thermo Fisher, C10046, fixed cells were stained
for 20 min at 1:2000 dilution. Halo-tag li-
gands (CA-JF646) were chemically synthesized
as described in (75) and used at 100 nM for 10
to 16 hours.
Pharmacological inhibitors
VPS34-IN1 (Selleckchem, S7980, 1mM), apilimod
(Echelon Biosciences, B-0308, 50nM), rapalog
(Takara, A/C Heterodimerizer 635056, 0.5 mM),
nocodazole (Sigma, M1404, 5mM), dimethyl sulf-
oxide (DMSO; Sigma, D2650), thapsigargin
(Sigma, T9033, 3mM), Torin1 (Tocris, Cat. No.
4247, 1 mM), Z-VAD-FMK (Tocris, Cat. No.
2163, 25 mM).
RNA isolation, RT-PCR, and qRT-PCR
Total RNA was isolated using the RNeasy Plus
Mini Kit (Qiagen) according to the manufac-
turers instructions. Quantification of RNA
yields was done with a multimode microplate
reader (SPECTROstar Nano from BMG Labtech).
Jang et al., Science 378, eabq5209 (2022) 16 December 2022 10 of 16
RESEARCH |RESEARCH ARTICLE
Downloaded from https://www.science.org at Leibniz-Institut für Molekulare Pharmakologie on December 16, 2022
A reverse transcription kit (SuperScript IV;
Invitrogen) was used to reverse transcribe
RNA (800 ng) in a 20 ml reaction using oligo(dT)
and random hexamer Primers. Reverse tran-
scription polymerase chain reaction (RT-PCR)
was performed with specific primers set against
each gene with indicated cycles. For quan-
titative reverse transcription polymerase chain
reaction (qRT-PCR), SYBR Green Master Mix
(BioRad) was used according to the manufac-
turers protocol with BioRad CFX Connect.
qRT-PCR primers
GAPDH Fwd: 5-CTTCGCTCTCTGCTCCTCCT-3;
Rev: 5-GTTAAAAGCAGCCCTGGTGA-3,MTM1
Fwd: 5-GTTTGAGATCCTCACGAGATACG-3;
Rev:5-GTCCATCCATCCACGTTAAACTT-3,
MTMR1 Fwd: 5-CCTTGATGTTCCCCTTGGAGT-
3; Rev:5-GTGCCTGTCCGTTAGAAAGAG-3,
MTMR2 Fwd: 5-GTGGAAAGCGAAGCAAA-
GAAG-3;Rev:5-CTTGGCCGGGCATCAAATATAA-
3,MTMR3Fwd:5-GACTGAACAACGCAATCC-
GAC-3;Rev:5-CCTTGAAGTTACATGCTCCCC-3,
MTMR4 Fwd: 5-CCAAGCCAAGGATCTGTTCCC-
3; Rev:5-GCCGGTAGTTAGAGATGGCAA-3,
MTMR5 Fwd: 5-CGACCACACGGAGGTGTTC-
3;Rev:5-GGTTCCCAATCACGTTCTCCA-3,
MTMR6 Fwd: 5-GTTCCCCGGATAGCAAGCAAA-
3;Rev:5-GTGGCTGACTACATCGACAAAT-3,
MTMR7 Fwd: 5-TCCGCTTGGTAGATCGAGTGT-
3;Rev:5-TTTTCCACGAATATGACATGGGT-3,
MTMR8 Fwd: 5-TGCACTCCATCACATTGCCA-
3;Rev:5-GGCACACAAGGTCAGAATCTAA-3,
MTMR9 Fwd: 5-TGAAGCTCTTCGGAAGG-
TAGC-3;Rev:5-GTGGCTGACCACTTCGCATAA-3,
MTMR10 Fwd: 5-ATCCACTTGCCTTCCAGAA-
TACA-3;Rev:5-CACAAGAGCACTGCCGTTAGA-3,
MTMR11 Fwd: 5-GCTGCTCAGAGTTGGTTTTGA-
3;Rev:5-CCCCGAATACTGTTGGGCTT-3,
MTMR12 Fwd: 5-GGCTCCTAAACTGCTTAAA-
CGA-3; Rev:5-GTTGCCTTTGGTCCGTTCCA-3,
MTMR13 Fwd: 5-TCATCGTGGTAGGCTAT-
GACC-3; Rev:5-CCAGGCTGACAAAACAACTCA-3,
MTMR14 Fwd: 5-GGAGTTCTCCCGGACTCAGTA-
3; Rev:5-AACAGTAGTCTCGGCCAAACA-3.
Immunoblot analysis
Cells were lysed with radioimmunoprecipita-
tion assay (RIPA) buffer (50 mM TrisCl pH 7.5,
150 mM NaCl, 1% NP-40, 0.5% sodium deoxy-
cholate, 0.1% SDS) containing protease and
phosphatase inhibitors. Lysates were incu-
bated for 10 min on ice before centrifugation
at 17,000gfor 10 min at 4°C. Protein concen-
tration was measured by Bradford or bicin-
choninic acid (BCA) assays. Cell lysates in
Laemmli sample buffer were boiled for 10 min.
Between 20 and 40 mg of protein was resolved
by SDSpolyacrylamide gel electrophoresis
(SDS-PAGE). Immunoblotting was done on
nitrocellulose membranes. Membranes were
incubated with the indicated primary anti-
bodies at 4°C overnight. The next day, bound
primary antibodies were detected by incu-
bation with IRDye 680/800CW-conjugated or
horseradish peroxidase (HRP)conjugated sec-
ondary antibodies via the Odyssey Fc Imaging
system (LI-COR Biosciences).
Light microscopy
Immunocytochemistry
Cells were seeded on Matrigel-coated coverslips
(BD/Corning), fixed in 4% paraformaldehyde
(PFA) for 15 to 20 min at room temperature (RT).
Cells were washed three times with phosphate-
buffered saline (PBS) before incubation with
3% bovine serum albumin (w/v) in 0.3% PBST
(Triton X-100) for 20 min. Using the same buffer,
the cells were incubated with primary antibodies
for 2 hours at RT, washed three times with 0.3%
PBST, and then incubated with secondary anti-
bodies for 1 hour at RT. After washing with 0.3%
PBST, cells were mounted withHoechst 33258
(Invitrogen, H3569, 1:2000) to counter stain
nuclei. Images were acquired on Zeiss 710 or
780 Laser Scanning Confocal Microscopes
using ZEN.
Note for the specific staining conditions:
For calreticulin staining, cells were fixed with
37°C 4% PFA for 34 min at 37°C. For Rab5A
staining, cells were fixed with 37°C 4% PFA for
8 min at 37°C. For LC3-II staining, cells were
permeabilized with 20 mM digitonin for 5 min
at RT. After permeabilization, all subsequent
procedures were done in PBS.
PI(3)P staining
Cells were fixed in 2% PFA for 15 min at RT,
washed twice in PBS with 50 mM NH
4
Cl, and
permeabilized with 20 mM digitonin in buffer
A (20 mM PIPES pH 6.8 with NaOH, 137 mM
NaCl, 2.7 mM KCl) for 5 min. Note that per-
meabilization is critical for successful PI(3)P
staining, so depending on the batch of digitonin,
one might need to optimize the concentration
and incubation time.
Cells were washed three times in buffer A
before addition of purified GFP (or mCherry)-
2xFYVE
Hrs
at 0.25 mg/ml in buffer A with 5%
normal goat serum, 50 mM NH
4
Cl for 1 hour.
Samples were washed and decorated with anti-
bodies against GFP or RFP in buffer A with 5%
normal goat serum, 50 mM NH
4
Cl for 2 hours.
Cells were washed three times in buffer A, in-
cubated for 1 hour with secondary antibodies
in buffer A with 5% normal goat serum, 50 mM
NH
4
Cl. Samples were washed three times with
buffer A, cells were postfixed for 5 min in 2%
PFA, washed three times in PBS, and mounted
with Hoechst 33258. Fluorescent sum inten-
sity was measured from individual cells and
normalized to the level of PI(3)P detected in
fed cells set to 1.
Live-cell imaging
Cells (6 × 10
3
to 8 × 10
3
) were seeded on
Matrigel-coated 8-well chamber slides (ibidi,
80827). The next day, live-cell imaging was
carried out using Nikon-CSU Yokogawa Spin-
ning disk (CSU-X1) microscope equipped with
an EMCCD Camera (Andor AU-888) at 37°C in
the presence of 5% CO
2
.Imageswereacquired
using Nikon Elements.
Image analysis and quantification
Acquired images were first segmented to re-
solve individual cells by manually drawing each
cell boundary using the Fiji freehand selec-
tion tools combined with Clear Outside option.
Individual cell image files were saved in the in-
put folder and batch processed in CellProfiler
[v4.1.3 (76)] or in Fiji. CellProfiler pipeline
modules used in this study were posted at
https://cellprofiler.org/published-pipelines.
Sheet-to-tubular ER ratio
Total ER (calreticulin or mEmerald-Sec61b)
segmentation was done using a three-class
Otsu threshold followed by conversion into
binary images. ER area was then measured
using the MeasureObjectSizeShape module
in CellProfiler. From the same image, sheet
ER was segmented in a similar manner except
using the Minimum Cross-Entropy threshold.
Depending on signal-to-noise ratio and im-
age quality, threshold settings (e.g., Thresh-
old smoothing scale, correction factor) were
adjusted using the IdentifyObjects module in
CellProfiler. Sheet ER area was then divided
by total ER area using CalculateMath module
in the same pipeline of CellProfiler. Calculated
sheetERarea/totalERarearatiowasdefined
as Sheet ER ratio.
Mean area and number of mitochondria
per ROI
Regions of interest (ROIs; e.g., 15 mmby15mm)
were randomly generated from individual
cells in Fiji, exported to CellProfiler, and
filtered with a Gaussian filter module with
1 or 2 sigma. ROIs were then processed using
the EnhanceOrSuppressFeatures module with
the following options: Operation = Enhance,
Feature type = Neurites,Enhancement
method = Tubeness,Smoothing scale =
1.After binarization, mitochondria were
segmented using the Robust Background
Threshold in the IdentifyObjects CellProfiler
module. The number and area of individual
mitochondria per ROI were calculated using
MeasureObjectSizeShape module in CellProfiler.
Lysosome position
Images were processed in CellProfiler using
the EnhanceOrSuppressFeatures module with
the following options: Operation = Enhance,
Feature type = Speckles,Feature size = 4-6.
Lysosome were then segmented using Robust
Background threshold in the IdentifyObjects
module. The distance between individually
identified lysosomes from the cell centroid
was calculated using the RelateObjects module
Jang et al., Science 378, eabq5209 (2022) 16 December 2022 11 of 16
RESEARCH |RESEARCH ARTICLE
Downloaded from https://www.science.org at Leibniz-Institut für Molekulare Pharmakologie on December 16, 2022
in CellProfiler. The standard deviation of the
distance of lysosomes from the centroid define
the extent of lysosomal dispersion or clustering.
Early endosome distribution
Rectangular ROIs (80 pixels by width pixels rang-
ing from cell center to cell boundary) were ran-
domly selected from individual cells in Fiji and
processed using the EnhanceOrSuppressFeatures
module with the following options: Operation =
Enhance,Feature type = Speckles,Feature
size = 4-6.A Gaussian filter with sigma = 1 or
2 was applied in CellProfiler. Early endosomes
and lysosomes were segmented using two-
classes Otsu and robust background thresholds
using the IdentifyObjects module. To compare
the partitioning of early endosomes and lyso-
somes between the cell center and the periphery,
segmented individual objects were shrunken to
a dot using the ExpandOrShrinkObjects module
with the Shrink objects to a pointoption in
CellProfiler and converted into binary image
files. The MeasureObjectIntensityDistribution
module was then used to measure the inten-
sity distribution from each objects center to
its boundary within a set of bins with the
following options: Bin = 5,Measurement =
Fraction at distance.The relative intensity
fraction among the five bins was calculated
from the binary images in CellProfiler. This
intensity value indirectly represents the frac-
tion of endosomes per bin.
Fluorescent imaging of ERearly
endosome contacts and dynamics
Tubular ER segmentation of 100 pixel by
100 pixel ROIs from the cell periphery were
generated using the Minimum Cross-Entropy
threshold option in the IdentifyObjects mod-
ule, and the ExpandOrShrinkObjects module
with Operation = Skeletonize each object,
and subsequent ExpandOrShrinkObjects
module with Operation = Expand objects
by a specified number of pixelssetting with
1-pixel expansion in CellProfiler. Early endo-
somes and lysosomes from the same image
were segmented as described above. Segmented
individual objects were then shrunken to a dot
using the ExpandOrShrinkObjects module with
Shrink objects to a pointoption, followed
by Operation = Expand objects by a specified
number of pixelssetting with 1-pixel expansion.
To calculate the number of endosomes that
contact the ER, the MeasureObjectNeighbors
module in CellProfiler was operated with
the following options: Method to determine
neighbors = Within a specified distanceset-
ting distance = 1 pixel. If the processed tubular
ER and endosomes overlapped with each other
at least by 1 pixel, a contact was scored.
Split GFP ERearly endosome contact sensor
To quantitatively determine the number ER
earlyendosomemembranecontactsbymeasur-
ing the number of GFP puncta, images were
blurred with a Gaussian filter (sigma = 1 or 2,
depending on the signal-to-noise ratio), back-
ground was subtracted with a rolling ball ra-
dius of 50 pixels, and segmented using the Find
Maxima tool with prominence = 10 in Fiji.
STED image analysis of ERearly
endosome contact
ROIs (8 mmby8mm) were randomly gen-
erated from individual cells in Fiji, ex-
ported to CellProfiler, processed using the
EnhanceOrSuppressFeatures module with
the following options: Operation = Enhance,
Feature type = Speckles,with subsequent
Gaussian filter module with 1. This allowed the
segmentation of early endosomes via robust
background thresholds using the IdentifyObjects
module. Segmented early endosomes were
then converted to binary images using the
ConvertObjectsToImage module. ER images
were filtered with a Gaussian filter module with
1 and converted to binary images using the
Threshold module. The MeasureImageOverlap
module was used to analyze the fractional
overlap between the ER and early endosomes.
Lipid droplets (BODIPY 493/503, Red C12)
Lipid droplet images were blurred with a
Gaussian filter (sigma = 1), enhanced using
the EnhanceOrSuppressFeatures module
(Operation = Enhance, Feature type =
Speckleswith size = 10), and then segmented
by the three-classes Otsu threshold option in
CellProfiler. Segmented objects were further
filtered on the basis of their measured sizes
and intensities using the FilterObjects module
in CellProfiler to minimize spurious detec-
tions. To measure the volume of lipid droplets,
z-stacked confocal images were analyzed using
the 3D objects counter plug-in in Fiji.
Pearson coefficient
To quantitatively assess colocalization between
two channels, two or three randomly chosen
100 pixel by 100 pixel ROIs from the cell pe-
riphery were blurred with a Gaussian filter
(sigma = 1). Pearsons coefficients were calcu-
lated using the Coloc2 plugin in Fiji or the
MeasureColocalization module with corre-
lation option in CellProfiler.
Super-resolution microscopy
Stimulated emission depletion (STED) images
were taken with either a Leica SP8 TCS STED
microscope (Leica Microsystems) for fixed
samplesoraSTEDYCON(AbberiorInstruments
GmbH, Göttingen) for live cell imaging. Leica
SP8 TCS STED microscope was equipped with
a pulsed white-light excitation laser [WLL;
80-p s p uls e width, 80 MHz repetition rate
(NKT Photonics)] and a STED laser for de-
pletion (775 nm). The system was controlled
by the Leica LAS X software. For images of
mEmerald-Sec61b, fixed cells were further
stained for GFP-Booster nanobody conjugated
to Atto647N (Chromotek, gba647n-100, 1:200).
Fluorophore specific excitation (Ex.) and emis-
sion filter (EmF.) settings: Atto647N (Ex.:
640 nm; EmF.: 650 to 700 nm). Time-gated
detection was set from 0.3 to 6 ns. The fluo-
rescence emission signal was collected by hy-
brid detectors (HyD). Images were acquired
with a HC PL APO CS2 100×/1.40 NA oil ob-
jective (Leica Microsystems) in a pixel size of
18.9 nm by 18.9 nm. STEDYCON was mounted
on the Nikon Eclipse Ti research microscope
equipped with a Plan APO 100x/1.45 NA oil
objective (Nikon). Excitation laser with a wave-
lengthof640nmwasusedfortheHaloligand
JF647 fluorescence. The wavelength of the
STED laser was 775 nm. Live cells expressing
Halo-Sec61bwere imaged at 37°C degree with
a pixel size of 20 nm by 20 nm.
Focused ion beam milling scanning electron
microscopy (FIB-SEM)
Cells were fixed with 2% glutaraldehyde (GA)
in PBS, washed with 0.1 M cacodylate buffer,
and embedded in Durcupan following a mod-
ified rOTO protocol with 1% (w/v) OsO4/
1.5% (w/v) K3Fe(CN)6 in 0.1 M cacodylate buf-
fer (pH 7.4); 0.2% (w/v) thiocarbohydrazide in
water; 1% OsO4 in water, 1% aqueous uranyl
acetate with corresponding washes in be-
tween and subsequent acetone dehydration
and resin infiltration. For polymerization, cov-
erslips were mounted onto pre-polymerized
resin blocks and placed into a heating cupboard
for 48 hours. After polymerization, coverslips
were removed by liquid nitrogen treatment,
and blocks were glued to SEM aluminum stubs
with conductive silver epoxy and carbon sputter
coating (30 nm).
Helios 5CX FIB-SEM autoslice and view
workflow was used to section and image the
embedded cells. FIB was run at 30 kV, 0.23 nA,
and 10-nm milling step. Cross section images
were scanned at 1.5 to 2 kV and 86 pA to 0.17 nA.
Dwell time was 5 ms at 3.37-nm pixel resolution
and ICD detection. The resulting 3D stacks
were binned to isotropic 10-nm voxel resolution.
Alignment, segmentation of mitochondr ia, and
analysis of ER networks were performed using
Microscopy Image Browser, 3D visualization
was done using Imaris. To access how con-
tinuous or fenestrated the ER is, the length
of ER cisternae was measured in several dis-
tantly positioned 2D sections from FIB-SEM
stacks. Sections for measurements from each
cell were randomly chosen and were at least
1mm apart from each other.
For 2D analysis of mitochondrial cristae
length, ultrathin sections of cells were collected
onto carbon-coated coverslips and imaged
with a Helios 5CX SEM. The length of cristae
was measured using Image J and normalized
per mitochondrial cross-section perimeter.
Jang et al., Science 378, eabq5209 (2022) 16 December 2022 12 of 16
RESEARCH |RESEARCH ARTICLE
Downloaded from https://www.science.org at Leibniz-Institut für Molekulare Pharmakologie on December 16, 2022
For 2D analysis of ERearly endosome mem-
brane contacts, 5 nm BSA-gold was allowed to
internalize into HeLa cells for 5 min to reach
early endosomes. Cells were washed before fix-
ation with 2% GA. Cells were processed for EM
analysis as described above. Ultrathin sections
mounted onto EM grids were imaged using a
Zeiss 900 transmission electron microscope
and ERearly endosome contact sites were
measured in Fiji.
Volume fraction analysis was performed
on ultrathin sections mounted on wafers and
scanned with a directional back-scatter de-
tector at 2 kV, 0.34 nA, 5-msdwelltime.A
measurement grid (1 mm
2
per point) was super-
imposed over cellular profiles and the refer-
ence volume (cytoplasm), as well as the relative
volumes of the ER, and mitochondria were es-
timated according to stereological principles.
APEX2 proximity labeling experiments
HeLa cells stably expressing doxycycline-
inducible 3xHA-APEX2-Rab5A were pretreated
with doxycycline for one day. Cells were then
incubated with 500 mM phenol-biotin (Sigma,
SML2135) at 37°C for 30 min. To enable ef-
ficient biotinylation, cells were treated with
1 mM hydrogen peroxide and incubated at
RT for 1 min. The reaction was terminated by
removing the medium and the addition of ice-
cold quenching buffer (10 mM sodium azide,
10 mM sodium ascorbate, 5 mM Trolox in PBS).
Cells were washed with PBS, lysed with RIPA
buffer (see above section on immunoblotting),
and centrifuged at 17,000gfor 10 min at 4°C.
The protein concentration of the resulting
supernatant was determined using the Bradford
assay. Then, 150- to 200-mg samples were saved
as whole cell lysate (WCL) control, while 1000
or 1500 mg of lysates were incubated with 40 ml
of Streptavidin Magnetic Beads (Thermo Fisher,
88817) overnight, nutating at 4°C. Magnetic
beads were washed twice with RIPA buffer,
once with 1 M KCl, once with 0.1 M Na
2
CO
3
,
once with 2 M Urea in 10 mM Tris-HCl, pH
8.0, two times with RIPA buffer, and finally
eluted with Laemmli sample buffer in the pres-
ence of 2 mM biotin. WCL and eluate samples
were analyzed by SDS-PAGE gels and immu-
noblotting. For quantification, the band inten-
sities of eluted endogenous MTM1 or EEA1
were normalized to the band intensity of self-
biotinylated Rab5A in Streptavidin-HRP blots
(Fig. 2B and fig. S4B).
ER isolation and biotinylated protein collection
for liquid chromatography combined with mass
spectrometry (LC-MS)
HeLa WT or MTM1 KO cell line stably ex-
pressing doxycycline-inducible 3xHA-APEX2-
Rab5A were seeded in 15-cm dishes with 17 ×
10
5
cells (WT = 6 dishes, KO = 3 dishes) in the
presence of doxycycline for 16 hours before
biotinylation. The next day, cells at 85% con-
fluencyweretreatedwith500mMphenol-biotin
and hydrogen peroxide to induce biotinylation,
quenched, washed once with isolation buffer
(225 mM mannitol, 75 mM sucrose, 30 mM
Tris-HCl pH 7.4, 0.1 mM EGTA), and, finally,
scraped into 1 ml of the same buffer resulting
in 2 ml total volume. Half a milliliter of this
sample was saved (WCL) as a control. The re-
maining 1.5 mlwas homogenized using a 2 ml
Dounce homogenizer (100 to 120 strokes) on
ice in the presence of protease and phosphatase
inhibitors. To isolate light membranes con-
taining the ER, the material was centrifuged
twice at 600gfor 2 min each. The pellet (F1)
including nuclei and cell debris was resus-
pended in 150 ml RIPA buffer. The remaining
supernatant was centrifuged twice at 7000g
for 20 min each. The resulting pellet (F2) in-
cluding the mitochondrial fraction was resus-
pended in 150 ml RIPA buffer and further
fractionated by centrifugation at 20,000gfor
60 min to pellet light ER membranes (F3). This
ER-enriched membrane fraction was resus-
pended in150 ml RIPA buffer. Protein concen-
tration was determined using the Bradford
assay. To isolate biotinylated proteins, 130 mg
ofWCLorF3fractionwasrotatedinthepres-
ence of Streptavidin Magnetic Beads. Eluted
biotinylated proteins were reduced (5 mM
dithiothreitol, 30 min at 55°C), alkylated (15 mM
iodacetamide, 20 min at RT in the dark), and
submitted to LC-MS analysis (samples from
two independent experiments).
LC-MS analysis of affinity-purified samples
Proteins were loaded on SDS-PAGE and sub-
jected to in-gel digestion. In brief, gel bands
were excised, and protein digestion was carried
out using trypsin at an enzyme-to-protein ratio
of 1:100 (w/w) at 37°C overnight. LC-MS mea-
surement was achieved by reverse phase high-
performance liquid chromatography (RP-HPLC)
on a Thermo Scientific Dionex UltiMate 3000
system connected to a PepMap C-18 trap-
column [0.075 mm by 50 mm, 3-mm particle
size, 100-Å pore size (Thermo Scientific)] and a
200 cm mPAC column was used (PharmaFluidics,
Ghent, Belgium) with 750 or 350 nl/min flow
rate with a 120-min gradient. Samples were
analyzed on an Orbitrap Fusion mass spectrom-
eter. MS1 scan were acquired in the Orbitrap
with a range of 375 to 1500 m/z, mass resolution
of 120,000, automatic gain control (AGC) target
value of 4 × 10
5
, and 50-ms maximum injection
time. MS2 scans were acquired in the ion trap
with an AGC target value of 1 × 10
4
and 35-ms
maximum injection time. Precursor ions with
charge states 2 to 4 were isolated with an iso-
lation window of 1.6 m/z and 40 s dynamic ex-
clusion. Precursor ions were fragmented using
higher-energy collisional dissociation with 30%
normalized collision energy. Analysis of the raw
data was done with MaxQuant (MQ) software
version 1.6.2.6. MaxQuant standard settings
were kept as default. In the search parameters,
two missed cleavage sites were included, the
fixed modification was set to cysteine carba-
midomethyl modification, and variable modifi-
cations to methionine oxidation and N-terminal
protein acetylation. The peptide mass tolerance
was set to 4.5 parts per million (ppm) for MS1
scans and 20 ppm for MS2 scans. Match be-
tween runs option was enabled. The database
search was done using Andromeda against
the Human UniProt/Swiss-Prot database with
common contaminants. The false discovery
rate (FDR) was set to 1% for both peptide and
protein level. Protein quantification was done
on the basis of at least two razor and unique
peptides. Label-free quantification and iBAQ
calculation were enabled. Statistical analysis
was done on the ProteinGroupstable with
Perseus version 1.6.7.0. Proteomics data have
been deposited to the ProteomeXchange Con-
sortium via PRIDE and are available via
ProteomeXchange with identifier PXD033846.
Quantitative whole-cell proteomics with
TMT labeling
Fed or starved (2 hours) WT or MTM1 KO HeLa
cells (one 10-cm dish at 90% confluency) col-
lected from three independent experiments
were lysed in 300 ml 8 M urea-lysis-buffer in
50 mM triethylammonium bicarbonate (TEAB)
containing protease inhibitor cocktail (Roche)
and 0.5 ml Benzonase Nuclease HC (Millipore).
Samples were applied to a Bioruptor Pico
(Diagenode) for 10 cycles (30 s on/30 s off) at
4°C.Sampleswerethenincubatedfor30min
at 25°C. Proteins were reduced with 5 mM Tris
(2-carboxyethyl)phosphin-hydrochloride (TCEP)
and alkylated with 40 mM chloroacetamide
(CAA) for 60 min at 37°C in the dark. Protein
digestion was carried out using Lys C at an
enzyme-to-protein ratio of 1:100 (w/w) at 37°C
for 3 hours. After diluting to 2 M urea with
50 mM TEAB buffer, the digestion was con-
tinued with trypsin at an enzyme-to-protein
ratio of 1:100 (w/w) at 37°C and overnight.
Digestion was stopped by adding formic acid
to a final concentration of 1%. Samples were
desalted with C18 Sep-Pak cartridge (Waters)
and quantified with Pierce colorimetric pep-
tide assay (Thermo Fisher Scientific). Peptides
were dried under speed vacuum and stored
at 20°C.
TMT labeling
Peptides were reconstituted in 50 mM TEAB
buffer to a concentration of 2.1 mg/ml. TMT
10-plex reagent (Thermo Fisher Scientific)
was dissolved in 20 ml 100% acetonitrile to
reach 0.2 mg. For each TMT channel, peptides
(100 mg) were labeled with 0.2 mg TMT 10-plex
reagent (two multiplex10plex for fed and
10plex for starved). For the internal standard
(IS), peptides from all samples were mixed to
reach 100 mg total peptides and labeled with
Jang et al., Science 378, eabq5209 (2022) 16 December 2022 13 of 16
RESEARCH |RESEARCH ARTICLE
Downloaded from https://www.science.org at Leibniz-Institut für Molekulare Pharmakologie on December 16, 2022
the TMT 131 channel to be able to compare
between the two plexes. The labeling reac-
tion was carried out for 60 min at RT and
quenched with 55 mM Tris pH 8.0 for 15 min
at RT. All TMT 10-plex labeled samples were
mixed and desalted with C18 Sep-Pak car-
tridge (Waters). The eluted peptides from the
Sep-Pak were dried under speed vacuum and
stored at 20°C.
High-pH prefractionation
The mixed TMT labeled peptides were re-
constituted in 10 mM NH
4
OH buffer with
1% acetonitrile to reach 200 mg. The Peptides
were fractionated by High-pH chromatogra-
phy using a Gemini column (3 mm, C18, 110 Å,
Phenomenex) on an Agilent 1260 Infinity II
system. An 85-min gradient was applied, and
72 fractions were collected and pooled into
12 fractions. Fractions were dried under speed
vacuum until analysis by LC-MS.
LC-MS analysis
Separation of the labeled TMT samples was
achieved by RP-HPLC on a Thermo Scientific
Dionex UltiMate 3000 system, as described
above. Samples were analyzed on an Orbitrap
Fusion Lumos mass spectrometer with FAIMS
Pro device (Thermo Scientific). MS1 and MS2
scans were acquired in the Orbitrap with a
mass resolution of 120,000 and 50,000, respec-
tively, MS1 scan range was set to between 400
and 1600 m/z, standard AGC target, and maxi-
mum injection time was set to auto. Precursor
ions with charge states 2 to 6 were isolated with
an isolation window of 0.7 m/z and dynamic
exclusion of 60 s. MS2 scans were set to custom
AGC target with normalized AGC target of
250%, and maximum injection time was set
to auto. Precursor ions were fragmented using
higher-energy collisional dissociation with 38%
normalized collision energy. Cycle time was set
to 2 s. An internal stepping of CVs 50, 65,
and 85 was used in all runs. Data acquisition
was done with Xcalibur software 4.4 and Instru-
ment Control Software version 3.4. Data analy-
sis for the TMT labeled samples was done in
Proteome Discoverer version 2.5. The TMT
10-plex was set as the quantification method,
and the 131 mass was set as the control chan-
nel. For the Sequest HT search, the following
parameters were applied: MS1 ion mass toler-
ance of 10 ppm and a MS2 mass tolerance of
0.02 Da. Tryptic digestion allowing two missed
cleavages, minimum peptide length of 6 amino
acids and maximum peptide length of 144
amino acids. The following modifications
were included: cysteine carbamidomethylation
(+57.021 Da) as static modification, methionine
oxidation (+15.995 Da) and N-terminal acet-
ylation (+42.011 Da) were set as dynamic modi-
fications. In addition, TMT 6-plex (229.163 Da)
was set as static modification for peptide
N-terminal and for lysine residue. Strict FDR
was set to 0.01, and relaxed FDR was set to
0.05. The search was performed against the
Human UniProt/Swiss-Prot database. Unique
and razor peptides were used for quantifica-
tion, co-isolation threshold was setto 50, and
average reporter S/N to 10. Data were normal-
ized against total peptide amount, and scaling
was done against the control channel average.
The result Proteinsoutput table was exported,
and the statistical analysis was done in Perseus
version 1.6.15.0. Then, Gene Ontology analysis
was conducted using Metascape web tool (77)
(https://metascape.org/gp/index.html#/main/
step1).Theproteomicsdatahavebeendepos-
ited to the ProteomeXchange Consortium via
PRIDE and are available via ProteomeXchange
with identifier PXD033850.
Liposome co-sedimentation assays
Powdered lipids were individually resuspended
in chloroform then mixed together in a glass
sample vial and slowly evaporated with a dry
N
2
stream. Liposome composition was 60%
phosphatidylcholine (Avanti #850375P), 19.8%
phosphatidylethanolamine (Avanti #850725P),
0.2% rhodamine-phosphatidylethanolamine
(Avanti #810150P), 10% cholesterol (Avanti
#700000), and 10% phosphatidylinositol-3-
sphosphate (Avanti #850150P) or 10% phos-
phatidylinositol 4-phosphate (Avanti #850151P)
or 10% phosphatidylinositol 3,4-biphosphate
(Avanti #850153P). The mixture was resus-
pended in 100 ml HEPES-buffered salt solution
(20 mM Hepes-NaOH pH 7.5, 150 mM NaCl)
for 40 min at 37°C, occasionally vigorously
vortexed, and sonicated in a 37°C water bath
five times, 45 s on, 1 min off. Liposomes (25 ml)
were gently mixed with 25 ml(3mgsuspended
with the same buffer) recombinant Escherichia
coli BL21 expressed GST-RRBP1 1-150aa WT
and incubated for 45 min at 25°C. Liposomes
were reisolated by centrifugation at 70,000g
for 15min at 25°C. Supernatant and pellet frac-
tions were dissolved in Laemmli sample buffer
andanalyzedbySDS-PAGE.
Seahorse XFe96 analyzer
The Seahorse XFe96 sensor cartridge was hy-
drated, and 1 × 10
4
cells were seeded on 96-well
plates one day before the assay. The next day,
while cells were fed or starved for 100 min at
37°C incubator without CO
2
,themitochon-
drial stress kit (Agilent, 103015-100) compound
was prepared in complete DMEM or EBSS and
loaded on the cartridge (OligomycinA1: 2 mM;
FCCP 1.7 mM, Rot/AA 1 mM). Assay media did
not include pyruvate. After instrument cali-
brations, cells were transferred to the XFe96
analyzer to record oxygen consumption rate at
37°C. All values were background-subtracted
(i.e., control well without cells). Finally, cells
were lysed with RIPA buffer and protein con-
tents was measured using the BCA assay for
data normalization.
Luminescent ATP assay
Cells (5 × 10
3
to 8 × 10
3
) were seeded in the
black 96-well plate. On the next day, total
cellular ATP levels of fed or starved cells were
analyzed using ATPlite Luminescence Assay
System (PerkinElmer, 6016943) according to
the manufacturers protocol. Luminescence
was measured by TECAN Luminescence plate
reader, and the value was normalized by pro-
tein concentration measured via Bradford or
BCA assay.
Software
Cartoons and schematics were generated using
BioRender and Adobe illustrator.
Statistics and reproducibility
Statistical analysis and graphing were carried out
using Prism 8. Normality testing (DAgostino-
Pearson) was conducted to determine whether
to use parametric or nonparametric statistical
tests. To compare two datasets, normally dis-
tributed data were analyzed by two-tailed un-
paired Studentsttest, whereas non-normally
distributed data were analyzed by two-tailed
Mann-Whitney test. For the comparison of
more than two normally distributed datasets,
we used ordinary one-way analysis of variance
(ANOVA) with either Dunnettsmultiplecom-
parisons (to compare the mean of each column
with a control column) or Tukeysmultiple
comparisons test (to compare the mean of each
column with every other column). To com-
pare more than two non-normally distributed
datasets, we used the Kruskal-Wallis test with
two-sided Dunns multiple comparison test. All
statistical analyses were performed on samples
drawn from at least two or three independent
experiments.
REFERENCES AND NOTES
1. P. Bonaldo, M. Sandri, Cellular and molecular mechanisms of
muscle atrophy. Dis. Model. Mech. 6,2539 (2013).
doi: 10.1242/dmm.010389; pmid: 23268536
2. B. H. Goodpaster, L. M. Sparks, Metabolic flexibility in health
and disease. Cell Metab. 25, 10271036 (2017). doi: 10.1016/
j.cmet.2017.04.015; pmid: 28467922
3. M. Sandri et al., Foxo transcription factors induce the atrophy-
related ubiquitin ligase atrogin-1 and cause skeletal muscle
atrophy. Cell 117, 399412 (2004). doi: 10.1016/S0092-8674
(04)00400-3; pmid: 15109499
4. S. Schiaffino, K. A. Dyar, S. Ciciliot, B. Blaauw, M. Sandri,
Mechanisms regulating skeletal muscle growth and atrophy.
FEBS J. 280, 42944314 (2013). doi: 10.1111/febs.12253;
pmid: 23517348
5. H. An, A. Ordureau, M. Körner, J. A. Paulo, J. W. Harper,
Systematic quantitative analysis of ribosome inventory during
nutrient stress. Nature 583, 303309 (2020). doi: 10.1038/
s41586-020-2446-y; pmid: 32612236
6. A. Efeyan, W. C. Comb, D. M. Sabatini, Nutrient-sensing
mechanisms and pathways. Nature 517, 302310 (2015).
doi: 10.1038/nature14190; pmid: 25592535
7. H. An et al., TEX264 is an endoplasmic reticulum-resident
ATG8-interacting protein critical for ER remodeling during
nutrient stress. Mol. Cell 74, 891908.e10 (2019). doi: 10.1016/
j.molcel.2019.03.034; pmid: 31006537
8. A. Ballabio, J. S. Bonifacino, Lysosomes as dynamic regulators
of cell and organismal homeostasis. Nat. Rev. Mol. Cell Biol.
21, 101118 (2020). doi: 10.1038/s41580-019-0185-4;
pmid: 31768005
Jang et al., Science 378, eabq5209 (2022) 16 December 2022 14 of 16
RESEARCH |RESEARCH ARTICLE
Downloaded from https://www.science.org at Leibniz-Institut für Molekulare Pharmakologie on December 16, 2022
9. S. M. Houten, S. Violante, F. V. Ventura, R. J. Wanders, The
biochemistry and physiology of mitochondrial fatty acid
b-oxidation and its genetic disorders. Annu. Rev. Physiol. 78,
2344 (2016). doi: 10.1146/annurev-physiol-021115-105045;
pmid: 26474213
10. T. Balla, Phosphoinositides: Tiny lipids with giant impact on cell
regulation. Physiol. Rev. 93, 10191137 (2013). doi: 10.1152/
physrev.00028.2012; pmid: 23899561
11. Y. Posor, W. Jang, V. Haucke, Phosphoinositides as membrane
organizers. Nat. Rev. Mol. Cell Biol. 23, 797816 (2022).
doi: 10.1038/s41580-022-00490-x; pmid: 35589852
12. C. C. Campa et al., Rab11 activity and PtdIns(3)P turnover
removes recycling cargo from endosomes. Nat. Chem. Biol. 14,
801810 (2018). doi: 10.1038/s41589-018-0086-4;
pmid: 29915378
13. C. Cao, J. M. Backer, J. Laporte, E. J. Bedrick,
A. Wandinger-Ness, Sequential actions of myotubularin lipid
phosphatases regulate endosomal PI(3)P and growth factor
receptor trafficking. Mol. Biol. Cell 19, 33343346 (2008).
doi: 10.1091/mbc.e08-04-0367; pmid: 18524850
14. C. Cao, J. Laporte, J. M. Backer, A. Wandinger-Ness, M. P. Stein,
Myotubularin lipid phosphatase binds the hVPS15/hVPS34
lipid kinase complex on endosomes. Traffic 8, 10521067
(2007). doi: 10.1111/j.1600-0854.2007.00586.x;
pmid: 17651088
15. K. Ketel et al., A phosphoinositide conversion mechanism
for exit from endosomes. Nature 529, 408412 (2016).
doi: 10.1038/nature16516; pmid: 26760201
16. L. Al-Qusairi et al., T-tubule disorganization and defective
excitation-contraction coupling in muscle fibers lacking
myotubularin lipid phosphatase. Proc. Natl. Acad. Sci. U.S.A.
106, 1876318768 (2009). doi: 10.1073/pnas.0900705106;
pmid: 19846786
17. L. Amoasii et al., Myotubularin and PtdIns3Premodel the
sarcoplasmic reticulum in muscle in vivo.J. Cell Sci. 126,
18061819 (2013). doi: 10.1242/jcs.118505; pmid: 23444364
18. Y. Shibata, G. K. Voeltz, T. A. Rapoport, Rough sheets and
smooth tubules. Cell 126, 435439 (2006). doi: 10.1016/
j.cell.2006.07.019; pmid: 16901774
19. J. Laporte et al., Mutations in the MTM1 gene implicated in
X-linked myotubular myopathy. Hum. Mol. Genet. 6, 15051511
(1997). doi: 10.1093/hmg/6.9.1505; pmid: 9305655
20. C. R. Piers on et al., Modeling the human MTM1 p.R69C
mutation in murine Mtm1 results in exon 4 skipping and a
less severe myotubular myopathy phenotype. Hum. Mol.
Genet. 21,811825 (2012). doi: 10.1093/hmg/ddr512;
pmid: 22068590
21. O. M. Do rchies et al., Normal innervation and differentiation
of X-linked myotubular myopathy muscle cells in a
nerve-muscle coculture system. Neuromuscul. Disord. 11,
736746 (2001). doi: 10.1016/S0960-8966(01)00221-8;
pmid: 11595516
22. L. K. Schroeder et al., Dynamic nanoscale morphology of the
ER surveyed by STED microscopy. J. Cell Biol. 218,8396
(2019). doi: 10.1083/jcb.201809107; pmid: 30442642
23. G. E. Palade, Studies on the endoplasmic reticulum. II.
Simple dispositions in cells in situ. J. Biophys. Biochem.
Cytol. 1, 567582 (1955). doi: 10.1083/jcb.1.6.567;
pmid: 13278367
24. G. K. Voeltz, W. A. Prinz, Y. Shibata, J. M. Rist, T. A. Rapoport,
A class of membrane proteins shaping the tubular endoplasmic
reticulum. Cell 124, 573586 (2006). doi: 10.1016/
j.cell.2005.11.047; pmid: 16469703
25. J. Nixon-Abell et al., Increased spatiotemporal resolution
reveals highly dynamic dense tubular matrices in the
peripheral ER. Science 354, aaf3928 (2016). doi: 10.1126/
science.aaf3928; pmid: 27789813
26. M. A. Raess, S. Friant, B. S. Cowling, J. Laporte, WANTED -
Dead or alive: Myotubularins, a large disease-associated
protein family. Adv. Biol. Regul. 63,4958 (2017). doi: 10.1016/
j.jbior.2016.09.001; pmid: 27666502
27. E. L. Axe et al., Autophagosome formation from membrane
compartments enriched in phosphatidylinositol 3-phosphate
and dynamically connected to the endoplasmic reticulum.
J. Cell Biol. 182, 685701 (2008). doi: 10.1083/jcb.200803137;
pmid: 18725538
28. Z. Hong et al., PtdIns3P controls mTORC1 signaling through
lysosomal positioning. J. Cell Biol. 216, 42174233 (2017).
doi: 10.1083/jcb.201611073; pmid: 29030394
29. M. J. Munson et al., mTOR activates the VPS34-UVRAG
complex to regulate autolysosomal tubulation and cell survival.
EMBO J. 34, 22722290 (2015). doi: 10.15252/
embj.201590992; pmid: 26139536
30. J. Laporte et al., A gene mutated in X-linked myotubular
myopathy defines a new putative tyrosine phosphatase family
conserved in yeast. Nat. Genet. 13, 175182 (1996).
doi: 10.1038/ng0696-175; pmid: 8640223
31. M. W. Lawlor, J. J. Dowling, X-linked myotubular myopathy.
Neuromuscul. Disord. 31, 10041012 (2021). doi: 10.1016/
j.nmd.2021.08.003; pmid: 34736623
32. A. Khaminets et al., Regulation of endoplasmic reticulum
turnover by selective autophagy. Nature 522, 354358 (2015).
doi: 10.1038/nature14498; pmid: 26040720
33. C. Settembre et al., TFEB links autophagy to lysosomal
biogenesis. Science 332, 14291433 (2011). doi: 10.1126/
science.1204592; pmid: 21617040
34. K. M. Fetalvero et al., Defective autophagy and mTORC1
signaling in myotubularin null mice. Mol. Cell. Biol. 33,98110
(2013). doi: 10.1128/MCB.01075-12; pmid: 23109424
35. A. L. Marat, V. Haucke, Phosphatidylinositol 3-phosphates-at
the interface between cell signalling and membrane traffic.
EMBO J. 35, 561579 (2016). doi: 10.15252/embj.201593564;
pmid: 26888746
36. Y. Wu et al., Contacts between the endoplasmic reticulum
and other membranes in neurons. Proc. Natl. Acad. Sci. U.S.A.
114, E4859E4867 (2017). doi: 10.1073/pnas.1701078114;
pmid: 28559323
37. H. Wu, P. Carvalho, G. K. Voeltz, Here, there, and everywhere:
The importance of ER membrane contact sites. Science
361, eaan5835 (2018). doi: 10.1126/science.aan5835;
pmid: 30072511
38. S. S. Lam et al., Directed evolution of APEX2 for electron
microscopy and proximity labeling. Nat. Methods 12,5154
(2015). doi: 10.1038/nmeth.3179; pmid: 25419960
39. A. Fegan, B. White, J. C. Carlson, C. R. Wagner, Chemically
controlled protein assembly: Techniques and applications.
Chem. Rev. 110, 33153336 (2010). doi: 10.1021/cr8002888;
pmid: 20353181
40. T. Calì, M. Brini, Quantification of organelle contact sites by
split-GFP-based contact site sensors (SPLICS) in living cells.
Nat. Protoc. 16, 52875308 (2021). doi: 10.1038/
s41596-021-00614-1; pmid: 34686857
41. C. Raib org et al., Repeated ER-endosome cont acts promote
endosome translocation and neurite outgrowth. Nature
520,234238 (2015). doi: 10.1038/nature14359;
pmid: 25855459
42. A. M. Valm et al., Applying systems-level spectral imaging
and analysis to reveal the organelle interactome. Nature
546, 162167 (2017). doi: 10.1038/nature22369;
pmid: 28538724
43. A. A. Rowland, P. J. Chitwood, M. J. Phillips, G. K. Voeltz, ER
contact sites define the position and timing of endosome
fission. Cell 159, 10271041 (2014). doi: 10.1016/
j.cell.2014.10.023; pmid: 25416943
44. M. Lu et al., The structure and global distribution of the
endoplasmic reticulum network are actively regulated
by lysosomes. Sci. Adv. 6, eabc7209 (2020). doi: 10.1126/
sciadv.abc7209; pmid: 33328230
45. V. Hung et al., Proteomic mapping of cytosol-facing outer
mitochondrial and ER membranes in living human cells by
proximity biotinylation. eLife 6, e24463 (2017). doi: 10.7554/
eLife.24463; pmid: 28441135
46. Y. Shibata et al., Mechanisms determining the morphology of
the peripheral ER. Cell 143, 774788 (2010). doi: 10.1016/
j.cell.2010.11.007; pmid: 21111237
47. P. Zheng et al., ER proteins decipher the tubulin code to
regulate organelle distribution. Nature 601, 132138 (2022).
doi: 10.1038/s41586-021-04204-9; pmid: 34912111
48. C. S. Janot a et al., Shielding of actin by the endoplasmic
reticulum impacts nuclear positioning. Nat. Commun. 13,
2763 (2022). doi: 10.1038/ s41467-02 2-3038 8-3;
pmid: 35589708
49. B. T. Lobi ngier et al., An approach to spatiotemporally
resolve protein interaction networks in living cells. Cell 169,
350360.e12 (2017). doi: 10.1016/j.cell.2017.03.022;
pmid: 28388416
50. Y. C. Wong, D. Ysselstein, D. Krainc, Mitochondrialysosome
contacts regulate mitochondrial fission via RAB7 GTP
hydrolysis. Nature 554, 382386 (2018). doi: 10.1038/
nature25486; pmid: 29364868
51. S. Nagashima et al., Golgi-derived PI(4)P-containing vesicles
drive late steps of mitochondrial division. Science 367,
13661371 (2020). doi: 10.1126/science.aax6089;
pmid: 32193326
52. S. C. Lewis, L. F. Uchiyama, J. Nunnari, ER-mitochondria
contacts couple mtDNA synthesis with mitochondrial division
in human cells. Science 353, aaf5549 (2016). doi: 10.1126/
science.aaf5549; pmid: 27418514
53. J. R. Friedman et al., ER tubules mark sites of mitochondrial
division. Science 334, 358362 (2011). doi: 10.1126/
science.1207385; pmid: 21885730
54. T. C. Walther, J. Chung, R. V. Farese Jr., Lipid droplet
biogenesis. Annu. Rev. Cell Dev. Biol. 33, 491510 (2017).
doi: 10.1146/annurev-cellbio-100616-060608; pmid: 28793795
55. A. Santinho et al., Membrane curvature catalyzes lipid droplet
assembly. Curr. Biol. 30, 24812494.e6 (2020). doi: 10.1016/
j.cub.2020.04.066; pmid: 32442467
56. A. S. Rambold, B. Kostelecky, N. Elia, J. Lippincott-Schwartz,
Tubular network formation protects mitochondria from
autophagosomal degradation during nutrient starvation.
Proc. Natl. Acad. Sci. U.S.A. 108, 1019010195 (2011).
doi: 10.1073/pnas.1107402108; pmid: 21646527
57. A. S. Rambold, S. Cohen, J. Lippincott-Schwartz, Fatty acid
trafficking in starved cells: Regulation by lipid droplet lipolysis,
autophagy, and mitochondrial fusion dynamics. Dev. Cell 32,
678692 (2015). doi: 10.1016/j.devcel.2015.01.029;
pmid: 25752962
58. M. Giacomello, A. Pyakurel, C. Glytsou, L. Scorrano, The cell
biology of mitochondrial membrane dynamics. Nat. Rev. Mol.
Cell Biol. 21,204224 (2020). doi: 10.1038/s41580-020-0210-7;
pmid: 32071438
59. T. W ai et al., Imbalanced OPA1 processing and mitochondrial
fragmentation cause heart failure in mice. Science 350,
aad0116 (2015). doi: 10.1126/science.aad0116;
pmid: 26785494
60. A. R. English, G. K. Voeltz, Rab10 GTPase regulates ER
dynamics and morphology. Nat. Cell Biol. 15, 169178 (2013).
doi: 10.1038/ncb2647; pmid: 23263280
61. J. E. Lee, P. I. Cathey, H. Wu, R. Parker, G. K. Voeltz,
Endoplasmic reticulum contact sites regulate the dynamics of
membraneless organelles. Science 367, eaay7108 (2020).
doi: 10.1126/science.aay7108; pmid: 32001628
62. L. C. Gomes, G. Di Benedetto, L. Scorrano, During autophagy
mitochondria elongate, are spared from degradation and
sustain cell viability. Nat. Cell Biol. 13, 589598 (2011).
doi: 10.1038/ncb2220; pmid: 21478857
63. W. M. Henne, M. L. Reese, J. M. Goodman, The assembly of
lipid droplets and their roles in challenged cells. EMBO J. 37,
e98947 (2018). doi: 10.15252/embj.201898947;
pmid: 29789390
64. M. A. Roberts, A. Segura-Roman, J. A. Olzmann, Organelle
biogenesis: ER shape influences lipid droplet nucleation.
Curr. Biol. 30, R770R773 (2020). doi: 10.1016/
j.cub.2020.05.027; pmid: 32634419
65. J. R. Liang et al., A genome-wide ER-phagy screen highlights
key roles of mitochondrial metabolism and ER-resident
UFMylation. Cell 180, 11601177.e20 (2020). doi: 10.1016/
j.cell.2020.02.017; pmid: 32160526
66. Y. Pan et al., Endoplasmic reticulum ribosome-binding protein
1, RRBP1, promotes progression of colorectal cancer and
predicts an unfavourable prognosis. Br. J. Cancer 113, 763772
(2015). doi: 10.1038/bjc.2015.260; pmid: 26196185
67. A. H. Bahrami, G. Hummer, Formation and stability of lipid
membrane nanotubes. ACS Nano 11, 95589565 (2017).
doi: 10.1021/acsnano.7b05542; pmid: 28873296
68. T. Shemesh et al., A model for the generation and interconversion
of ER morphologies. Proc. Natl. Acad. Sci. U.S.A. 111,
E5243E5251 (2014). doi: 10.1073/pnas.1419997111;
pmid: 25404289
69. G. Parlakgül et al., Regulation of liver subcellular architecture
controls metabolic homeostasis. Nature 603, 736742 (2022).
doi: 10.1038/s41586-022-04488-5; pmid: 35264794
70. D. S. Hirsch et al., Insulin activation of vacuolar protein sorting
34 mediates localized phosphatidylinositol 3-phosphate
production at lamellipodia and activation of mTOR/S6K1.
Cell. Signal. 26, 12581268 (2014). doi: 10.1016/
j.cellsig.2014.02.009; pmid: 24582588
71. V. I. Korolchuk et al., Lysosomal positioning coordinates
cellular nutrient responses. Nat. Cell Biol. 13, 453460 (2011).
doi: 10.1038/ncb2204; pmid: 21394080
72. A. L. Marat et al., mTORC1 activity repression by late
endosomal phosphatidylinositol 3,4-bisphosphate. Science
356, 968972 (2017). doi: 10.1126/science.aaf8310;
pmid: 28572395
73. M. J. Hoyer et al., A novel class of ER membrane proteins
regulates ER-associated endosome fission. Cell 175, 254265.
e14 (2018). doi: 10.1016/j.cell.2018.08.030; pmid: 30220460
74. K. Mamchaoui et al., Immortalized pathological human
myoblasts: Towards a universal tool for the study of
Jang et al., Science 378, eabq5209 (2022) 16 December 2022 15 of 16
RESEARCH |RESEARCH ARTICLE
Downloaded from https://www.science.org at Leibniz-Institut für Molekulare Pharmakologie on December 16, 2022
neuromuscular disorders. Skelet. Muscle 1, 34 (2011).
doi: 10.1186/2044-5040-1-34; pmid: 22040608
75. F. Bottanelli et al., Two-colour live-cell nanoscale imaging of
intracellular targets. Nat. Commun. 7, 10778 (2016).
doi: 10.1038/ncomms10778; pmid: 26940217
76. C. McQuin et al., CellProfiler 3.0: Next-generation image
processing for biology. PLOS Biol. 16, e2005970 (2018).
doi: 10.1371/journal.pbio.2005970; pmid: 29969450
77. Y. Zho u et al., Metascape p rovides a biolog ist-oriented
resource for the an alysis of syste ms-level dataset s. Nat.
Commun. 10,1523(2019).doi:10.10 38/s41467-019-092 34-6;
pmid: 30944313
ACKNOWL EDGME NTS
We thank members of the Haucke lab for discussion and Y. Posor
for critical reading of the manuscript. We thank S. Zillmann,
D. Löwe, M. Mühlbauer, and C. Schmidt for expert technical
assistance and M. Lehmann, C. Schmied, and J. Eichhorst for aid
with microscopy and image analysis. We also thank G. G. Farías
and G. Voeltz for plasmids and protocols and the MyoLine platform
of the Institute of Myology in Paris for aid in the generation
of myoblast cell lines. Funding: This work was funded by a
Leibniz-German Academic Exchange Service (DAAD) Research
Fellowship (57423756) (W.J.), the Postdoctoral Fellowship
Program (Nurturing Next-generation Researchers) of the National
Research Foundation of Korea (NRF) (2018R1A6A3A03010583)
(W.J.), and Deutsche Forschungsgemeinschaft (TRR186/ A08)
(V.H.). Author contributions: Conceptualization: W.J. and
V.H. Investigation: all cell and molecular biology: W.J.; electron
microscopy: D.P. with W.J.; quantitative proteomics: W.J. with
M.N.-H. and F.L.; Seahorse analysis: W.J., Y.L., S.J.S., and U.K.;
XLCNM myoblast patient cells: K.M. and V.M.; CRISPR: W.J. and
P.S. Funding acquisition: W.J. and V.H. Project administration: V.H.
Supervision: V.H., F.L., U.K., and S.J.S. Writing original draft:
W.J. and V.H. Writing review & editing: all authors. Competing
interests: The authors declare no competing financial interests.
Data and materials availability: All data are available in the main
text or the supplementary materials. Proteomics data have been
deposited to the ProteomeXchange Consortium via PRIDE and
are available via ProteomeXchange with identifier PXD033846.
Materials and reagents are available from the corresponding author
upon request. License information: Copyright © 2022 the
authors, some rights reserved; exclusive licensee American
Association for the Advancement of Science. No claim to original
US government works. https://www.science.org/about/science-
licenses-journal-article-reuse
SUPPLEMENTARY MATERIALS
science.org/doi/10.1126/science.abq5209
Figs. S1 to S11
Tables S1 to S4
MDAR Reproducibility Checklist
Movies S1 to S5
View/request a protocol for this paper from Bio-protocol.
Submitted 13 April 2022; resubmitted 23 September 2022
Accepted 25 October 2022
10.1126/science.abq5209
Jang et al., Science 378, eabq5209 (2022) 16 December 2022 16 of 16
RESEARCH |RESEARCH ARTICLE
Downloaded from https://www.science.org at Leibniz-Institut für Molekulare Pharmakologie on December 16, 2022
Use of this article is subject to the Terms of service
Science (ISSN ) is published by the American Association for the Advancement of Science. 1200 New York Avenue NW, Washington, DC
20005. The title Science is a registered trademark of AAAS.
Copyright © 2022 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim
to original U.S. Government Works
Endosomal lipid signaling reshapes the endoplasmic reticulum to control
mitochondrial function
Wonyul JangDmytro PuchkovPaula SamsóYongTian LiangMichal Nadler-HollyStephan J. SigristUlrich KintscherFan
LiuKamel MamchaouiVincent MoulyVolker Haucke
Science, 378 (6625), eabq5209. • DOI: 10.1126/science.abq5209
A lipid-triggered signal in starvation
Nutrient starvation triggers changes in metabolism that are coordinated across the cell and its organelles. Jang et
al. studied how endosomal signaling lipid turnover through MTM1, a phosphoinositide 3-phosphatase mutated in X-
linked centronuclear myopathy in humans, reshapes the endoplasmic reticulum to control mitochondrial morphology
and oxidative metabolism (see the Perspective by Zanellati and Cohen). A lipid-controlled organellar relay transmits
nutrient-triggered changes in endosomal signaling lipid levels to mitochondria to enable metabolic rewiring. —SMH
View the article online
https://www.science.org/doi/10.1126/science.abq5209
Permissions
https://www.science.org/help/reprints-and-permissions
Downloaded from https://www.science.org at Leibniz-Institut für Molekulare Pharmakologie on December 16, 2022
... The APEX2 proximity labeling assay was performed as described previously, 127 with minor modifications. Briefly, protein biotinylation reactions in 60 10-cm dishes of MEFs with stable expression of APEX2-PDZD8 were treated with DMEM (10 mL per dish) containing 500 mM biotinyl tyramide at 37°C for 30 min, followed by the addition of 1 mM hydrogen peroxide and incubated at room temperature for 1 min. ...
Article
Full-text available
The shift of carbon utilization from primarily glucose to other nutrients is a fundamental metabolic adaptation to cope with decreased blood glucose levels and the consequent decline in glucose oxidation. AMP-activated protein kinase (AMPK) plays crucial roles in this metabolic adaptation. However, the underlying mechanism is not fully understood. Here, we show that PDZ domain containing 8 (PDZD8), which we identify as a new substrate of AMPK activated in low glucose, is required for the low glucose-promoted glutaminolysis. AMPK phosphorylates PDZD8 at threonine 527 (T527) and promotes the interaction of PDZD8 with and activation of glutaminase 1 (GLS1), a rate-limiting enzyme of glutaminolysis. In vivo, the AMPK-PDZD8-GLS1 axis is required for the enhancement of glutaminolysis as tested in the skeletal muscle tissues, which occurs earlier than the increase in fatty acid utilization during fasting. The enhanced glutaminolysis is also observed in macrophages in low glucose or under acute lipopolysaccharide (LPS) treatment. Consistent with a requirement of heightened glutaminolysis, the PDZD8-T527A mutation dampens the secretion of pro-inflammatory cytokines in macrophages in mice treated with LPS. Together, we have revealed an AMPK-PDZD8-GLS1 axis that promotes glutaminolysis ahead of increased fatty acid utilization under glucose shortage.
... In contrast, the KD of RTN4, a reticulon family member not involved in NCE 5 , did not have an inhibitory effect, but instead appeared to slightly increase the contacts near to the PM. This observation is in line with the reported effects of RTN4 KD on general ER morphology 5,12 , which are however unrelated to NCE 5 . ...
Article
Full-text available
One open question in the biology of growth factor receptors is how a quantitative input (i.e., ligand concentration) is decoded by the cell to produce specific response(s). Here, we show that an EGFR endocytic mechanism, non-clathrin endocytosis (NCE), which is activated only at high ligand concentrations and targets receptor to degradation, requires a tripartite organelle platform involving the plasma membrane (PM), endoplasmic reticulum (ER) and mitochondria. At these contact sites, EGFR-dependent, ER-generated Ca²⁺ oscillations are sensed by mitochondria, leading to increased metabolism and ATP production. Locally released ATP is required for cortical actin remodeling and EGFR-NCE vesicle fission. The same biochemical circuitry is also needed for an effector function of EGFR, i.e., collective motility. The multiorganelle signaling platform herein described mediates direct communication between EGFR signaling and mitochondrial metabolism, and is predicted to have a broad impact on cell physiology as it is activated by another growth factor receptor, HGFR/MET.
... Such EE-ER proximity relies at least in part on a change in endosomal dynamics driven by modifications of their molecular composition and/or morphodynamics. This is notably illustrated by the spatiotemporal myotubularin (MTM1)-dependent dephosphorylation of early endosomal PI3P at EE-ER contact sites [26]. ...
Article
Full-text available
In eukaryotes, the spatiotemporal control of endolysosomal organelles is central to the maintenance of homeostasis. By providing an interface between the cytoplasm and external environment, the endolysosomal system is placed at the forefront of the response to a wide range of stresses faced by cells. Endosomes are equipped with a dedicated set of membrane‐associated proteins that ensure endosomal functions as well as crosstalk with the secretory or the autophagy pathways. Morphodynamical processes operate through local spatialization of subdomains, enabling specific remodeling and membrane contact capabilities. Consequently, the plasticity of endolysosomal organelles can be considered a robust and flexible tool exploited by cells to cope with homeostatic deviations. In this review, we provide insights into how the cellular responses to various stresses (osmotic, UV, nutrient deprivation, or pathogen infections) rely on the adaptation of the endolysosomal system morphodynamics.
... Recent studies have also shown that PI(3)P-positive endosomes have a critical role in regulating synaptic vesicle recycling and neurotransmission 55 . In addition, recent studies have also shown that nutrient-sensitive PI(3)P-mediated lipid signaling on endosomes regulates ER shape and mitochondrial function 56 . However, the source, identity, and mechanism(s) regulating PI(3)P formation on endosomes in these studies remains unclear. ...
Preprint
Full-text available
Rab4 GTPase organizes endosomal sorting essential for maintaining the balance between recycling and degradative pathways. Rab4 localizes to many cargos whose transport in neurons is critical for regulating neurotransmission and neuronal health. Furthermore, elevated Rab4 levels in the CNS are associated with synaptic atrophy and neurodegeneration in Drosophila and humans, respectively. However, how the transport of Rab4-associated vesicles is regulated in neurons remains unknown. Using in vivo time-lapse imaging of Drosophila larvae, we show that activation of insulin signaling via Dilp2 and dInR increases the anterograde velocity, run length, and flux of Rab4 vesicles in the axons. Molecularly, we show that activation of neuronal insulin signaling further activates Vps34, elevates the levels of PI(3)P on Rab4-associated vesicles, recruits Klp98A (a PI(3)P-binding kinesin-3 motor) and activates their anterograde transport. Together, these observations delineate the role of insulin signaling in regulating axonal transport and synaptic homeostasis. Highlights Dilp2-mediated insulin signaling activates anterograde transport of Rab4 vesicles Vps34 regulates anterograde velocity, run length and flux of Rab4 vesicles Local PI(3)P signaling on Rab4 vesicles regulates their motility in the axons PI(3)P production upon acute insulin stimulation recruits Klp98A on Rab4 vesicles
Article
Junctions between the endoplasmic reticulum (ER) and the outer membrane of the nuclear envelope (NE) physically connect both organelles. These ER–NE junctions are essential for supplying the NE with lipids and proteins synthesized in the ER. However, little is known about the structure of these ER–NE junctions. Here, we systematically study the ultrastructure of ER–NE junctions in cryo-fixed mammalian cells staged in anaphase, telophase, and interphase by correlating live cell imaging with three-dimensional electron microscopy. Our results show that ER–NE junctions in interphase cells have a pronounced hourglass shape with a constricted neck of 7–20 nm width. This morphology is significantly distinct from that of junctions within the ER network, and their morphology emerges as early as telophase. The highly constricted ER–NE junctions are seen in several mammalian cell types, but not in budding yeast. We speculate that the unique and highly constricted ER–NE junctions are regulated via novel mechanisms that contribute to ER-to-NE lipid and protein traffic in higher eukaryotes.
Preprint
Full-text available
Peripheral endoplasmic reticulum (ER) tubules move along microtubules to interact with various organelles through membrane contact sites (MCS). Traditionally, ER moves by either sliding along stable microtubules via molecular motors or attaching to the plus ends of dynamic microtubules through tip attachment complexes (TAC). A recently discovered third process, hitchhiking, involves motile vesicles pulling ER tubules along microtubules. Previous research showed that ER hitchhikes on Rab5- and Rab7-marked endosomes, but it is uncertain if other Rab-vesicles can do the same. In U2OS cells, we screened Rabs for their ability to cotransport with ER tubules and found that ER hitchhikes on post-Golgi vesicles marked by Rab6 (isoforms a and b). Rab6-ER hitchhiking occurs independently of ER-endolysosome contacts and TAC-mediated ER movement. Disrupting either Rab6 or the motility of Rab6-vesicles reduces overall ER movement. Conversely, relocating these vesicles to the cell periphery causes peripheral ER accumulation, indicating that Rab6-vesicle motility is crucial for a subset of ER movements. Proximal post-Golgi vesicles marked by TGN46 are involved in Rab6-ER hitchhiking, while other post-Golgi vesicles (Rabs 8/10/11/13/14) are not essential for ER movement. Our further analysis finds that ER to Golgi vesicles marked by Rab1 are also capable of driving a subset of ER movements. Taken together, our findings suggest that ER hitchhiking on Rab-vesicles is a significant mode of ER movement. SIGNIFICANCE STATEMENT Peripheral endoplasmic reticulum tubules move on microtubules by either attaching to motors (cargo adaptor-mediated), dynamic microtubule-plus ends (tip attachment complexes) or motile vesicles (hitchhiking) but the prevalence of each mode is not clear Post-Golgi vesicles marked by Rab6/TGN46 and ER to Golgi vesicles marked by Rab1 drive ER movements ER hitchhiking on multiple classes of vesicles (endolysosomal, post-Golgi and ER to Golgi) marked by Rabs plays a prominent role in ER movement
Article
Full-text available
Nuclear position is central to cell polarization, and its disruption is associated with various pathologies. The nucleus is moved away from the leading edge of migrating cells through its connection to moving dorsal actin cables, and the absence of connections to immobile ventral stress fibers. It is unclear how these asymmetric nucleo-cytoskeleton connections are established. Here, using an in vitro wound assay, we find that remodeling of endoplasmic reticulum (ER) impacts nuclear positioning through the formation of a barrier that shields immobile ventral stress fibers. The remodeling of ER and perinuclear ER accumulation is mediated by the ER shaping protein Climp-63. Furthermore, ectopic recruitment of the ER to stress fibers restores nuclear positioning in the absence of Climp-63. Our findings suggest that the ER mediates asymmetric nucleo-cytoskeleton connections to position the nucleus.
Article
Full-text available
Cells display complex intracellular organization by compartmentalization of metabolic processes into organelles, yet the resolution of these structures in the native tissue context and their functional consequences are not well understood. Here we resolved the three-dimensional structural organization of organelles in large (more than 2.8 × 105 µm3) volumes of intact liver tissue (15 partial or full hepatocytes per condition) at high resolution (8 nm isotropic pixel size) using enhanced focused ion beam scanning electron microscopy1,2 imaging followed by deep-learning-based automated image segmentation and 3D reconstruction. We also performed a comparative analysis of subcellular structures in liver tissue of lean and obese mice and found substantial alterations, particularly in hepatic endoplasmic reticulum (ER), which undergoes massive structural reorganization characterized by marked disorganization of stacks of ER sheets3 and predominance of ER tubules. Finally, we demonstrated the functional importance of these structural changes by monitoring the effects of experimental recovery of the subcellular organization on cellular and systemic metabolism. We conclude that the hepatic subcellular organization of the ER architecture are highly dynamic, integrated with the metabolic state and critical for adaptive homeostasis and tissue health. Detailed reconstruction using enhanced focused ion beam scanning electron microscopy imaging and deep-learning-based automated segmentation demonstrates that hepatocyte subcellular organelle architecture regulates metabolism.
Article
Full-text available
Organelles move along differentially modified microtubules to establish and maintain their proper distributions and functions 1,2 . However, how cells interpret these post-translational microtubule modification codes to selectively regulate organelle positioning remains largely unknown. The endoplasmic reticulum (ER) is an interconnected network of diverse morphologies that extends promiscuously throughout the cytoplasm ³ , forming abundant contacts with other organelles ⁴ . Dysregulation of endoplasmic reticulum morphology is tightly linked to neurologic disorders and cancer 5,6 . Here we demonstrate that three membrane-bound endoplasmic reticulum proteins preferentially interact with different microtubule populations, with CLIMP63 binding centrosome microtubules, kinectin (KTN1) binding perinuclear polyglutamylated microtubules, and p180 binding glutamylated microtubules. Knockout of these proteins or manipulation of microtubule populations and glutamylation status results in marked changes in endoplasmic reticulum positioning, leading to similar redistributions of other organelles. During nutrient starvation, cells modulate CLIMP63 protein levels and p180–microtubule binding to bidirectionally move endoplasmic reticulum and lysosomes for proper autophagic responses.
Article
Full-text available
X-linked myotubular myopathy (XLMTM) is a severe congenital muscle disease caused by mutation in the MTM1 gene. MTM1 encodes myotubularin (MTM1), an endosomal phosphatase that acts to dephosphorylate key second messenger lipids PI3P and PI3,5P2. XLMTM is clinically characterized by profound muscle weakness and associated with multiple disabilities (including ventilator and wheelchair dependence) and early death in most affected individuals. The disease is classically defined by characteristic changes observed on muscle biopsy, including centrally located nuclei, myofiber hypotrophy, and organelle disorganization. In this review, we highlight the clinical and pathologic features of the disease, present concepts related to disease pathomechanisms, and present recent advances in therapy development.
Article
Full-text available
The endoplasmic reticulum (ER) comprises morphologically and functionally distinct domains: sheets and interconnected tubules. These domains undergo dynamic reshaping in response to changes in the cellular environment. However, the mechanisms behind this rapid remodeling are largely unknown. Here, we report that ER remodeling is actively driven by lysosomes, following lysosome repositioning in response to changes in nutritional status: The anchorage of lysosomes to ER growth tips is critical for ER tubule elongation and connection. We validate this causal link via the chemo- and optogenetically driven repositioning of lysosomes, which leads to both a redistribution of the ER tubules and a change of its global morphology. Therefore, lysosomes sense metabolic change in the cell and regulate ER tubule distribution accordingly. Dysfunction in this mechanism during axonal extension may lead to axonal growth defects. Our results demonstrate a critical role of lysosome-regulated ER dynamics and reshaping in nutrient responses and neuronal development.
Article
Full-text available
Mammalian cells reorganize their proteomes in response to nutrient stress through translational suppression and degradative mechanisms using the proteasome and autophagy systems1,2. Ribosomes are central targets of this response, as they are responsible for translation and subject to lysosomal turnover during nutrient stress3,4,5. The abundance of ribosomal (r)-proteins (around 6% of the proteome; 10⁷ copies per cell)6,7 and their high arginine and lysine content has led to the hypothesis that they are selectively used as a source of basic amino acids during nutrient stress through autophagy4,7. However, the relative contributions of translational and degradative mechanisms to the control of r-protein abundance during acute stress responses is poorly understood, as is the extent to which r-proteins are used to generate amino acids when specific building blocks are limited⁷. Here, we integrate quantitative global translatome and degradome proteomics⁸ with genetically encoded Ribo–Keima⁵ and Ribo–Halo reporters to interrogate r-protein homeostasis with and without active autophagy. In conditions of acute nutrient stress, cells strongly suppress the translation of r-proteins, but, notably, r-protein degradation occurs largely through non-autophagic pathways. Simultaneously, the decrease in r-protein abundance is compensated for by a reduced dilution of pre-existing ribosomes and a reduction in cell volume, thereby maintaining the density of ribosomes within single cells. Withdrawal of basic or hydrophobic amino acids induces translational repression without differential induction of ribophagy, indicating that ribophagy is not used to selectively produce basic amino acids during acute nutrient stress. We present a quantitative framework that describes the contributions of biosynthetic and degradative mechanisms to r-protein abundance and proteome remodelling in conditions of nutrient stress.
Article
Phosphoinositides are signalling lipids derived from phosphatidylinositol, a ubiquitous phospholipid in the cytoplasmic leaflet of eukaryotic membranes. Initially discovered for their roles in cell signalling, phosphoinositides are now widely recognized as key integrators of membrane dynamics that broadly impact on all aspects of cell physiology and on disease. The past decade has witnessed a vast expansion of our knowledge of phosphoinositide biology. On the endocytic and exocytic routes, phosphoinositides direct the inward and outward flow of membrane as vesicular traffic is coupled to the conversion of phosphoinositides. Moreover, recent findings on the roles of phosphoinositides in autophagy and the endolysosomal system challenge our view of lysosome biology. The non-vesicular exchange of lipids, ions and metabolites at membrane contact sites in between organelles has also been found to depend on phosphoinositides. Here we review our current understanding of how phosphoinositides shape and direct membrane dynamics to impact on cell physiology, and provide an overview of emerging concepts in phosphoinositide regulation. Phosphoinositides are signalling, membrane lipids derived from phosphatidylinositol, whose intracellular distribution and interconversion via phosphoinositide kinases and phosphatases is tightly coupled to membrane dynamics. Accordingly, phosphoinositides are now recognized as key regulators of endocytic and exocytic traffic, the autolysosomal system, and membrane contact site organization and function.
Article
Membrane contact sites between organelles are essential for maintaining cellular homeostasis, which requires the continuous exchange of signaling molecules, ions, nutrients and lipids. Alterations of different contact sites are associated with a wide spectrum of human diseases. However, visualizing and quantifying these contact sites remains a challenge. This protocol describes the use of split-GFP-based contact site sensors (SPLICS) in microscopy applications for mapping organelle contact sites both in fixed and living cells. SPLICS sensors are engineered to express equimolar amounts of the organelle-targeted nonfluorescent β11 and GFP1-10 portions of the split-GFP protein in a single vector, and are capable of reconstituting fluorescence when two opposing membranes come into proximity. Reconstitution will occur only over the cell volume at defined contact sites resulting in a bright signal that can be detected easily and quantified automatically with specific custom-made plugins. The use of minimal targeting sequences facilitates targeting specificity and membrane coverage, avoiding artifacts due to full-length fusion protein overexpression and, thus, possible perturbations of the cell’s physiology. SPLICS sensors engineered to simultaneously detect multiple contact sites within the same cell have been generated by exploiting the ability of the β11 GFP fragment to reconstitute different color-shifted variants of the GFP1-10 fragment. Here, we describe a detailed protocol to set up SPLICS expression in living cells (2–3 d), detection and acquisition (1 d), and automated quantification with custom plugins (1–2 d). We also advise on construct design and characterization for novel organelle contacts. Split-GFP-based contact site sensors (SPLICS) are capable of reconstituting fluorescence when two opposing membranes come into close proximity. SPLICS can be used to quantify membrane contact sites between organelles in both fixed and living cells.
Article
Lipid droplets (LDs) are neutral lipid storage organelles assembled at the endoplasmic reticulum (ER). A new study reveals that the high membrane curvature of ER tubules catalyzes the nucleation of a neutral lipid lens, an early step in LD biogenesis.
Article
Lipid droplet (LD) biogenesis begins in the endoplasmic reticulum (ER) bilayer, but how the ER topology impacts this process is unclear. An early step in LD formation is nucleation, wherein free neutral lipids, mainly triacylglycerols (TGs) and sterol esters (SEs), condense into a nascent LD. How this transition occurs is poorly known. Here, we found that LDs preferably assemble at ER tubules, with higher curvature than ER sheets, independently of the LD assembly protein seipin. Indeed, the critical TG concentration required for initiating LD assembly is lower at curved versus flat membrane regions. In agreement with this finding, flat ER regions bear higher amounts of free TGs than tubular ones and present less LDs. By using an in vitro approach, we discovered that the presence of free TGs in tubules is energetically unfavorable, leading to outflow of TGs to flat membrane regions or condensation into LDs. Accordingly, in vitro LD nucleation can be achieved by the sole increase of membrane curvature. In contrast to TGs, the presence of free SEs is favored at tubules and increasing SE levels is inhibitory to the curvature-induced nucleation of TG LDs. Finally, we found that seipin is enriched at ER tubules and controls the condensation process, preventing excessive tubule-induced nucleation. The absence of seipin provokes erratic LD nucleation events determined by the abundance of ER tubules. In summary, our data indicate that membrane curvature catalyzes LD assembly.