Download as pdf or txt
Download as pdf or txt
You are on page 1of 51

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/311498771

Mitochondrial pyruvate carrier regulates autophagy, inflammation, and


neurodegeneration in experimental models of Parkinsons disease

Article  in  Science translational medicine · December 2016


DOI: 10.1126/scitranslmed.aag2210

CITATIONS READS
27 589

14 authors, including:

Anamitra Ghosh Trevor Tyson


Baylor College of Medicine Van Andel Research Institute
29 PUBLICATIONS   1,127 CITATIONS    40 PUBLICATIONS   367 CITATIONS   

SEE PROFILE SEE PROFILE

Sonia George Erin N. Hildebrandt


Van Andel Research Institute Van Andel Research Institute
30 PUBLICATIONS   393 CITATIONS    13 PUBLICATIONS   151 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Models of Parkinson's disease View project

New insulin sensitizers View project

All content following this page was uploaded by Patrik Brundin on 13 November 2018.

The user has requested enhancement of the downloaded file.


SCIENCE TRANSLATIONAL MEDICINE | RESEARCH ARTICLE

PARKINSON’S DISEASE 2016 © The Authors,


some rights reserved;

Mitochondrial pyruvate carrier regulates autophagy, exclusive licensee


American Association
inflammation, and neurodegeneration in experimental for the Advancement
of Science.
models of Parkinson’s disease
Anamitra Ghosh,1 Trevor Tyson,1 Sonia George,1 Erin N. Hildebrandt,1 Jennifer A. Steiner,1
Zachary Madaj,2 Emily Schulz,1 Emily Machiela,1 William G. McDonald,3 Martha L. Escobar Galvis,1
Jeffrey H. Kordower,1,4 Jeremy M. Van Raamsdonk,1 Jerry R. Colca,3 Patrik Brundin1*

Mitochondrial and autophagic dysfunction as well as neuroinflammation are involved in the pathophysiology of
Parkinson’s disease (PD). We hypothesized that targeting the mitochondrial pyruvate carrier (MPC), a key controller
of cellular metabolism that influences mTOR (mammalian target of rapamycin) activation, might attenuate neuro-
degeneration of nigral dopaminergic neurons in animal models of PD. To test this, we used MSDC-0160, a compound
that specifically targets MPC, to reduce its activity. MSDC-0160 protected against 1-methyl-4-phenylpyridinium (MPP+)
insult in murine and cultured human midbrain dopamine neurons and in an a-synuclein–based Caenorhabditis elegans

Downloaded from http://stm.sciencemag.org/ on December 7, 2016


model. In 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)–treated mice, MSDC-0160 improved locomotor behav-
ior, increased survival of nigral dopaminergic neurons, boosted striatal dopamine levels, and reduced neuroinflamma-
tion. Long-term targeting of MPC preserved motor function, rescued the nigrostriatal pathway, and reduced
neuroinflammation in the slowly progressive Engrailed1 (En1+/−) genetic mouse model of PD. Targeting MPC in
multiple models resulted in modulation of mitochondrial function and mTOR signaling, with normalization of autoph-
agy and a reduction in glial cell activation. Our work demonstrates that changes in metabolic signaling resulting from
targeting MPC were neuroprotective and anti-inflammatory in several PD models, suggesting that MPC may be a
useful therapeutic target in PD.

INTRODUCTION of TZDs or genetically—results in a compensatory increase in utilization


People with Parkinson’s disease (PD) exhibit a range of nonmotor and of other substrates, that is, amino acids and fatty acids, resulting in
motor symptoms (1, 2), with the latter being strongly linked to the alterations in cellular metabolism (20, 21, 26, 27).
degeneration of nigral dopamine neurons. Whereas the disease mecha- For the past two decades, the general consensus has been that TZDs
nisms are incompletely understood, evidence suggests that mitochon- activate the transcription factor peroxisome proliferator–activated
drial deficits, failure of autophagy, and neuroinflammation each plays receptor-g (PPARg) and have multiple effects on mitochondrial func-
a role (3–5). Disease-modifying therapies are not available for PD, tion, including partial inhibition of complex I [for review, see (28)].
and several trials targeting individual pathways implicated in PD patho- The identification of MPC, however, suggests that some of the anti-
genesis have failed (6–14). Therefore, a potentially more powerful diabetic effects and other actions of TZDs might be mediated through
strategy is the targeting of molecules that are upstream in the signaling this protein complex (29). The use of TZDs in diabetes has declined
cascades that modulate altered cellular functions and that affect both over recent years because of several side effects driven by the activa-
neurons and glia in the brain. tion of nuclear receptor PPARg, including fluid retention, weight gain,
The metabolism of all nutrients flows through various molecular and concern over a potential increased risk for urinary bladder cancer
pathways, but ultimately, all are linked to the metabolism of pyruvate (30) and other cancers (31). Notwithstanding the risk for dose-limiting
(15). Recently, a protein complex for transporting pyruvate into the mito- and sometimes serious side effects, TZDs have recently garnered great
chondria, mitochondrial pyruvate carrier (MPC), was discovered in attention in PD. A large retrospective study demonstrated that diabetic
the internal mitochondrial membrane (16–18). MPC contains the proteins subjects prescribed with TZDs have a 29% reduced risk of developing
MPC-1 and MPC-2, is conserved in yeast, flies, worms, and mammals PD within 14 years (32), suggesting that these compounds might affect
(16–18), and is essential for cellular function (17–21). molecular mechanisms involved in PD. Furthermore, TZDs are pro-
The MPC complex is the target of a group of compounds known tective in PD animal models (33, 34). Recent research has shown that
as the thiazolidinedione (TZD) insulin sensitizers (22–25). TZDs slow the disease processes in diabetes and PD share several features, includ-
the entry of pyruvate across the mitochondrial membrane (24), and ing metabolic perturbations and inflammation (35–38).
the effects of TZDs on gluconeogenesis require the MPC complex MSDC-0160 is a member of a new class of compounds that mod-
(21). Knockout of the MPC complex in Drosophila eliminated the pos- ulate MPC and act as insulin sensitizers without activating PPARg
itive effects of TZD treatment in flies on a high-sucrose diet (22). Slow- (22–24, 39). Therefore, MSDC-0160 lacks the negative side effects of
ing the entry of pyruvate at the MPC—for example, by administration the first-generation insulin sensitizers (40). In a phase 2b trial in diabe-
tes, MSDC-0160 effectively lowered glucose and, importantly, caused
1
Center for Neurodegenerative Science, Van Andel Research Institute, Grand Rapids, minimal fluid retention and weight gain (41). Furthermore, MSDC-
MI 49503, USA. 2Bioinformatics and Biostatistics Core, Van Andel Research Institute, 0160 preserved cerebral 2-deoxyglucose uptake after 3 months of use
Grand Rapids, MI 49503, USA. 3Metabolic Solutions Development Company, Kalama- in Alzheimer’s disease patients, suggesting that it engaged the target
zoo, MI 49007, USA. 4Center for Brain Repair, Department of Pathology, Rush Medical
College, Chicago, IL 60612, USA. MPC in the brain after oral administration (42). Given this background,
*Corresponding author. Email: patrik.brundin@vai.org we hypothesized that modulating MPC function could normalize

Ghosh et al., Sci. Transl. Med. 8, 368ra174 (2016) 7 December 2016 1 of 17


SCIENCE TRANSLATIONAL MEDICINE | RESEARCH ARTICLE

molecular pathways that are perturbed in PD and reduce neurodegen- Some mice were pretreated with MSDC-0160, whereas controls received
eration in cellular and animal models of PD. vehicle. Twenty-four hours later, the mice were subjected to a subacute
MPTP regimen, that is, five daily injections of MPTP (25 mg/kg per day
intraperitoneally) (48, 49). During this time, mice received daily oral
RESULTS gavage of MSDC-0160 or vehicle for a total of 11 days (fig. S4A). Four
Protective role of MPC modulation in cellular and days after the last MPTP injection, we tested the mice for spontaneous
nematode models locomotion in an open-field arena and for motor coordination on the
Previous studies have shown that micromolar concentrations of rotarod test. Consistent with previous reports (48, 50, 51), MPTP injec-
MSDC-0160 acutely modulate the metabolism of pyruvate in multiple tions caused significant impairment on both tests, with decreases in dis-
cells types, including neurons, through a direct interaction with MPC tance traveled, speed, and mobility time in the open field, as well as time
(24). We hypothesized that this modulation of MPC function by spent on the rotating rod. However, MSDC-0160 pretreatment attenu-
MSDC-0160 would protect compromised dopaminergic neurons in ated or prevented these deficits (Fig. 2, A and B, and fig. S4B). The
models of PD both in vitro and in vivo. We used several model brains from treated mice were subjected to immunohistochemistry, ste-
systems to evaluate the potential pathways involved. reological counting of neurons, and high-performance liquid chroma-
We found that modulation of MPC shielded cultured human, tography (HPLC) and Western blot analyses. As expected (47, 48),
murine, and invertebrate dopaminergic neurons from 1-methyl-4- MPTP treatment induced a loss of dopaminergic neurons and terminals
phenylpyridinium (MPP+) toxicity (Fig. 1, B, G, and K, and fig. S1, A in the substantia nigra and striatum, respectively (Fig. 2, C to E), a re-

Downloaded from http://stm.sciencemag.org/ on December 7, 2016


and B). Specifically, MSDC-0160 (10 mM) pretreatment (1 hour) pre- duction of cresyl violet–positive nigral neurons (fig. S4C), depletion of
vented the MPP+ (10 mM)–induced loss of both tyrosine hydroxylase striatal dopamine and its metabolite 3,4-dihydroxyphenylacetic acid
(TH)–immunoreactive differentiated Lund human mesencephalic (DOPAC) (Fig. 2, F and G), and a loss of TH expression in the nigro-
(LUHMES) cells (Fig. 1, A and B, and fig. S1A) (43) and TH-immuno- striatal pathway (Fig. 2, H to J). These toxic effects of MPTP were atte-
reactive mouse primary mesencephalic neurons in culture (Fig. 1, F and nuated or blocked by modulation of MPC through pretreatment with
G) (44). We determined that about 11% of the neurons in differentiated MSDC-0160 (Fig. 2, C to J).
LUHMES cultures and 6% of the neurons in the primary murine mes- To address the possibility that the neuroprotection induced by
encephalic culture expressed TH. Further, we found that MSDC-0160 MSDC-0160 was due to inhibition of the conversion of MPTP to
protected only TH-immunoreactive neurons (no change in the number the neurotoxic metabolite MPP+ by glia, we measured MPP+ concen-
of TH-negative cells was observed), which is consistent with the selected trations in striatum 3 hours after the fourth MPTP injection, with or
concentration of MPP+ primarily being toxic to dopamine neurons. In without MSDC-0160 treatment. We found that MSDC-0160 did not
addition, MSDC-0160 counteracted both MPP+-induced shortening of affect striatal MPP+ concentrations (fig. S4D).
neurite length and reduced branching in both LUHMES cells (Fig. 1, C Next, we tested whether modulation of MPC could halt neurode-
to E) and primary dopaminergic cultures of mouse midbrain tissue (Fig. generation in the MPTP mouse model after the insult has already been
1, H to J). To observe the effects of MPC modulation in a whole orga- triggered. We found that starting MSDC-0160 administration 2 days
nism, we exclusively treated Caenorhabditis elegans nematodes after the initial dose of MPTP (in the subacute MPTP regimen) still
expressing green fluorescent protein (GFP) in dopaminergic neurons mitigated motor behavioral deficits (fig. S4G), rescued the loss of do-
with 0.75 mM MPP+ and quantified the resulting neuron loss according paminergic neurons and terminals (Fig. 2, K and L), increased the
to established protocols (45, 46). Treatment with MSDC-0160 (10 or numbers of Nissl-stained neurons (fig. S4F), and restored neuro-
100 mM) prevented the loss of GFP-fluorescent dopaminergic neurons transmitter concentrations (Fig. 2, M and N). Collectively, these results
induced by MPP+ (0.75 mM) in nematodes (Fig. 1K and fig. S1B) (P = demonstrate that modulation of MPC both protected against and
0.0001), whereas 1 mM MSDC-0160 did not. These results indicate that could rescue MPTP-induced neurodegeneration in the subacute MPTP
dopaminergic neurons were protected by MSDC-0160 treatment. mouse model of PD.
To validate the effects of modulating MPC in a progressive PD
Protective role of MPC modulation in mouse models of PD model, we turned to the Engrailed1 heterozygous (En1+/−) mouse.
To determine whether MPC targeting was neuroprotective in mammals, En1+/− mice represent a chronic degeneration genetic model with a
we examined the effects of MSDC-0160 on motor behavior and neuro- PD-like pattern of dopaminergic neurodegeneration that exhibits de-
degeneration in mouse models of PD. To confirm that MSDC-0160 creased autophagy and increased neuroinflammation. The En1+/−
effectively enters the brain, we pretreated C57BL/6J mice with MSDC- mice have a normal complement of dopaminergic neurons at birth,
0160 (30 mg/kg per day via oral gavage) and euthanized the mice 2 or but at 8 weeks after birth, the mice exhibit the first signs of a progres-
4 hours later. Our measurements revealed MSDC-0160 and its hydro- sive loss of nigral dopaminergic neurons, reaching a plateau at 24 weeks
xymetabolite MSDC-0037 both in plasma and in brain tissue (fig. S2, (52). Furthermore, En1+/− mice display motor impairment at 24 to
A and B), confirming that orally administered MSDC-0160 was able 27 weeks of age in the open-field and rotarod tests. Here, we chose to
to gain access to the brain. When expressed as nanogram per gram in initiate the modulation of MPC function at two different ages to de-
the brain and nanogram per milliliter in the plasma, at 2 hours after termine whether the drug would be effective when pathology is mild
an oral dose, there was parent drug (278 ng/g) and hydroxymetabolite or modest. Treatment starting at 3 weeks after birth served as a “mild
(21,100 ng/ml) in the brain, and at the same time point, the plasma pathology stage,” at which there is axonal pathology but no cell death
concentrations were 164 ng/ml for the parent drug and 51,600 ng/ml (Fig. 3A) (53). Treatment starting at 8 weeks after birth served as a
for the metabolite. This suggested that oral administration was suffi- “modest pathology stage,” at which 20% of TH-immunoreactive neu-
cient to provide direct targeting of MPC in brain cells. rons have been lost in En1+/− mice (Fig. 3B).
We treated a separate cohort of mice with the neurotoxin 1-methyl- At the mild pathology stage, En1+/− mice were fed a diet formu-
4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) as a model of PD (47). lated to deliver MSDC-0160 (30 mg/kg) starting at week 3 after birth

Ghosh et al., Sci. Transl. Med. 8, 368ra174 (2016) 7 December 2016 2 of 17


SCIENCE TRANSLATIONAL MEDICINE | RESEARCH ARTICLE

Downloaded from http://stm.sciencemag.org/ on December 7, 2016

Fig. 1. MSDC-0160 protects dopaminergic neurons against MPP+-induced toxicity in cell cultures and C. elegans. LUHMES cells were pretreated with 10 mM MSDC-
0160 (0160) for 1 hour followed by 10 mM MPP+ treatment for 24 hours. (A) Double-label immunocytochemistry for TH and Tuj1 in LUHMES cells. (B) Number of TH-positive
dopaminergic neurons, (C) mean neurite length (mm), (D) neurite branching points in each neuron, and (E) longest neurite length (mm). Primary mesencephalic mouse neurons
were pretreated with 10 mM MSDC-0160 for 1 hour followed by 10 mM MPP+ treatment for 24 hours. (F) Double-label immunocytochemistry for TH and Tuj1 in primary
mesencephalic culture. (G) Number of TH-positive dopaminergic neurons, (H) mean neurite length (mm), (I) neurite branching points in each neuron, and (J) longest neurite
length (mm). C. elegans were synchronized, and L1 larvae were placed in 96-well plates containing OP50 bacteria in liquid culture (6 mg/ml) and MPP+ (0.75 mM) with MSDC-
0160 (1, 10, and 100 mM) for 48 hours. (K) Quantification of percent dopaminergic neuron loss in C. elegans. Scale bars, 50 mm. Data are means ± SEM of three independent
experiments. ****P < 0.0001, ***P < 0.001, **P < 0.01, *P < 0.05, analyzed by one-way analysis of variance (ANOVA) with Tukey’s multiple comparison test.

Ghosh et al., Sci. Transl. Med. 8, 368ra174 (2016) 7 December 2016 3 of 17


SCIENCE TRANSLATIONAL MEDICINE | RESEARCH ARTICLE

Downloaded from http://stm.sciencemag.org/ on December 7, 2016

Fig. 2. MSDC-0160 improves motor behavior, protects nigrostriatal neurons, and suppresses disease progression in the MPTP mouse model of PD. (A) Open-field
parameters: distance traveled (cm), mean speed (cm/s), and time spent mobile (s). (B) Time spent on rotarod (s). (C) Diaminobenzidine immunohistochemistry for TH in
substantia nigra (SN; upper panel) and striatum (ST; lower panel) of MPTP mice. (D) Stereological counting of TH-positive neurons from the SN. (E) Relative density of TH-positive
neuronal fibers in the ST. Quantification of (F) dopamine (ng/mg) and (G) the dopamine metabolite DOPAC (ng/mg) by HPLC in the ST. (H) Representative Western blots
illustrating the expression of TH in SN and ST. Bar graph showing mean Western blot TH/b-actin in SN (I) and in ST (J). For the disease progression studies, mice were administered
MSDC-0160 (30 mg/kg per day) starting 3 days after MPTP treatment, with continuing treatment for another 7 days. (K) Stereological counting of TH-positive neurons from SN. (L)
Relative density of TH-positive neuronal fibers in ST. Quantification of striatal (M) dopamine and (N) DOPAC by HPLC. Scale bars, 50 mm (SN) and 200 mm (ST). Data are means ±
SEM of 7 to 10 mice per group. ****P < 0.0001, ***P < 0.001, **P < 0.01, *P < 0.05. Data were analyzed by one-way ANOVA with Tukey’s multiple comparison test.

Ghosh et al., Sci. Transl. Med. 8, 368ra174 (2016) 7 December 2016 4 of 17


SCIENCE TRANSLATIONAL MEDICINE | RESEARCH ARTICLE

Downloaded from http://stm.sciencemag.org/ on December 7, 2016

Fig. 3. MSDC-0160 improves motor behavior in the open-field and rotarod tests in the En1+/− genetic mouse model of PD. (A) Schedule for MSDC-0160 treatment of
En1+/− mice in the mild pathology stage. (B) Schedule for MSDC-0160 treatment of En1+/− mice in the modest pathology stage. (C) Distance traveled (cm), (D) mean speed (cm/s),
(E) maximum speed (cm/s), and (F) time on rotarod (s) in the mild pathology stage. WT, wild type. (G) Distance traveled (cm), (H) mean speed (cm/s), (I) maximum speed (cm/s),
and (J) time on rotarod (s) in the modest pathology stage. Data are means ± SEM of 8 to 12 mice per group. ***P < 0.001, **P < 0.01, *P < 0.05, oP < 0.05 to 0.1. Data were analyzed
using linear mixed-effects regression analysis with false discovery rate (FDR)–corrected P values.

Ghosh et al., Sci. Transl. Med. 8, 368ra174 (2016) 7 December 2016 5 of 17


SCIENCE TRANSLATIONAL MEDICINE | RESEARCH ARTICLE

until euthanasia at week 28 or week 48. At the modest pathology stage, On the basis of the literature, we hypothesized that the mammalian
En1+/− mice were fed the same diet starting at week 8 until euthanasia target of rapamycin (mTOR) pathway is involved downstream of
at week 16 or week 28. At each time point, we assayed motor behavior, MPC (29, 56–59). In nematodes, we knocked down genes of interest
and after euthanasia, we performed immunohistochemistry, Western related to MPC and to the mTOR pathway by RNA interference and
blotting, and HPLC analysis to examine dopaminergic markers and neu- evaluated whether the knockdown of these genes affected the protec-
rotransmitter concentrations. For motor behaviors, including sponta- tion of A53T a-synuclein–induced dopaminergic neurons from degen-
neous activity in the open-field and coordination testing on the rotarod, eration by MSDC-0160. MSDC-0160 treatment prevented neuron loss
the MSDC-0160–treated group exhibited improvements on all mea- due to the A53T mutation in a-synuclein in control worms, which were
sures (Fig. 3, C to G, fig. S5, and table S1). fed bacteria expressing the empty vector L4440 (Fig. 5C). Knockdown
For all outcomes, the En1+/− non–MSDC-0160–treated control of BRP44L (the ortholog of MPC-1) prevented neuroprotection by
group of mice performed significantly worse than did the wild-type MSDC-0160 (Fig. 5D). Furthermore, knockdown of AKT-1 (a serine/
control mice (see the En1 genotype variable in table S1 for adjusted threonine kinase that functions upstream of RHEB-1; Fig. 5E), RHEB-
and unadjusted P values). Treatment with MSDC-0160 did not signif- 1 (a guanosine triphosphatase that is a stimulator of the mTOR
icantly affect the behavior of wild-type mice. However, the En1+/− pathway; Fig. 5F), or LET-363 (the C. elegans ortholog of the human
group treated with MSDC-0160 showed evidence of improvement (af- mTOR protein; Fig. 5G) also prevented the neuroprotection exerted by
ter false discovery multiple testing adjustments) in mean maze speed MSDC-0160 in this dopaminergic neuronal loss assay. By contrast,
for the groups that started on MSDC-0160 at 3 and 8 weeks of age knockdown of AAK-1, which is a homolog of adenosine monopho-

Downloaded from http://stm.sciencemag.org/ on December 7, 2016


(3 weeks, P = 0.032; 8 weeks, P = 0.027), mean distance traveled (group sphate (AMP)–activated protein kinase, mimicked the control condition
that started on treatment at 8 weeks of age; P = 0.021), and time spent (Fig. 5H), suggesting that BRP44L does not control this pathway. RNA
mobile in the maze (group that started on treatment at 8 weeks of age; interference knockdown of the genes tested above was confirmed by
P = 0.003). There was little evidence that treatment of En1+/− mice quantitative polymerase chain reaction (fig. S1C). The conclusion of
with MSDC-0160 reduced the amount of time spent frozen in either these experiments is that engaging MPC and the mTOR pathway is ne-
of the groups that started MSDC-0160 treatment at 3 or 8 weeks of age. cessary for the neuroprotective effect of MSDC-0160 in C. elegans.
All other behavioral measures for the En1+/− groups treated with MSDC- In a second set of experiments, we used the mitochondria isolated
0160 exhibited significant improvement (P < 0.05) before the multiple from rat brain and intact cultured LUHMES cells to define the time
testing adjustments, but these apparent improvements did not remain course of the effects of MSDC-0160 on the different components of
significant after adjustment (see the En1 genotype * MSDC treatment the mTOR pathway. Initially, we demonstrated that MSDC-0160 de-
variable in table S1 for adjusted and unadjusted P values). creased pyruvate oxidation in isolated brain mitochondria in a direct
Consistent with the results of the behavioral testing, stereological and immediate fashion (fig. S2, C and D). We then measured the ef-
analyses of immunohistochemically stained sections of the brains of fect of MSDC-0160 on cellular respiration in LUHMES neuronal cells
the En1+/− mice revealed that treatment with this MPC modulator pro- treated with the mitochondrial toxin MPP+ and found that when
tected the TH-immunoreactive (Fig. 4, A, B, and G to I) and Nissl- MSDC-0160 was added 1 hour before or after MPP+ addition,
stained (fig. S6, A to C) dopaminergic neurons of the substantia nigra. MSDC-0160 normalized oxygen consumption (fig. S2E). These data
In addition, MPC modulation with MSDC-0160 increased the expres- show that modulation of MPC has immediate effects on mitochondri-
sion of TH and restored striatal dopamine and DOPAC content in all al function. On the other hand, when LUHMES cells were treated for
groups at each studied time point in the En1+/− mice [Fig. 4, C and D 2, 6, or 24 hours with MSDC-0160, no changes in mTOR activity
(P = 0.0018), E (P = 0.0018), F (P = 0.0168), J to L (P = 0.0001), M (P = [monitored as the ratio of phosphorylated mTOR (p-mTOR)/mTOR]
0.0001), N (P = 0.0014), O (P = 0.0001), and P (P = 0.0009), and fig. S6, could be detected (fig. S3, A and B). These data indicated that the
E (P = 0.0002) and F (P = 0.0001)]. Collectively, these data demonstrate changes in nutrient-sensing pathways were downstream of the meta-
that modulation of MPC was neuroprotective in the slowly progressing bolic action of MPC and upstream of the demonstrated neuroprotec-
En1+/− mouse model of PD. tive effect (fig. S2C). Together, the data obtained in nematodes and
cultured neurons showed that when modulating MPC, functional
The effects of MPC modulation on degeneration of changes in the mTOR pathway (described in detail below) appear with
dopaminergic neurons a delay after the immediate effects seen on mitochondrial function.
MPC, the target of MSDC-0160, is present in nigral dopaminergic Next, we queried the role of the mTOR pathway on the neuropro-
neurons (fig. S7, A and B) (17). Thus, we set out to dissect the molecular tection observed in vivo after MPC modulation in the MPTP and En1+/−
mechanisms underpinning the neuroprotective effect of MPC modula- mouse models of PD. In the MPTP model, using Western blotting of
tion using C. elegans and cultured cells, as well as the MPTP and En1+/− substantia nigra samples, we measured mTOR, p-mTOR (Ser2448), S6,
mouse models of PD. C. elegans expresses brain protein 44–like and phosphorylated S6 (pS6) (Ser235/Ser236), normalized to b-actin
(BRP44L), which is an ortholog of MPC-1. In the C. elegans model, (Fig. 5I). As expected, MPTP injections increased p-mTOR measured
we used worms expressing both GFP and A53T mutant a-synuclein as the p-mTOR/mTOR ratio relative to control conditions (P = 0.04)
in dopaminergic neurons (a point mutation that causes autosomal dom- (Fig. 5J). MPTP treatment also increased the ratio of pS6/S6 (P = 0.03)
inant PD in humans). Aggregated a-synuclein is a major component of (Fig. 5K). Although there was little to no effect under control con-
Lewy bodies in PD (54), and a-synuclein is genetically linked to both ditions, MPC modulation using MSDC-0160 pretreatment reduced
inherited and sporadic forms of PD (55). We visualized dopaminergic the activation of both mTOR and phosphorylation of its downstream
neurons using GFP in worms expressing A53T a-synuclein. These worms substrate, indicating that this regulation was a component of neuro-
showed marked degeneration of dopaminergic neurons after 8 days (P = protection (P = 0.032) (Fig. 5, I to K).
0.05), which was rescued when BRP44L function was modulated by treat- We also returned to the En1+/− mouse model to validate further
ment with 100 mM MSDC-0160 (P = 0.0001) (Fig. 5, A and B). the involvement of the mTOR pathway downstream of MPC. Using

Ghosh et al., Sci. Transl. Med. 8, 368ra174 (2016) 7 December 2016 6 of 17


SCIENCE TRANSLATIONAL MEDICINE | RESEARCH ARTICLE

Downloaded from http://stm.sciencemag.org/ on December 7, 2016

Fig. 4. MSDC-0160 prevents dopaminergic neurodegeneration in the En1+/− genetic mouse model of PD. (A) TH immunostaining in the SN. (B) Stereological counting
of TH-immunoreactive (TH+) neurons in the SN. (C) Representative Western blot illustrating the expression of TH in the SN. (D) Bar graph showing mean Western blot TH/
b-actin ratios in SN, quantification of striatal (E) dopamine (ng/mg), and (F) DOPAC (ng/mg) by HPLC in the mild pathology stage. (G) TH immunostaining in SN (upper panel
pictures are from 16 weeks of age, and lower panel pictures are from 28 weeks of age). Scale bars, 50 mm. (H) Stereological counting of TH-positive neurons in SN from 16 weeks of
age. (I) Stereological counting of TH-positive neurons in SN from 28 weeks of age. (J) Representative Western blot illustrating the expression of TH in SN from 16 weeks of age (top)
and 28 weeks of age (bottom), and bar graphs showing mean Western blot TH/b-actin ratios in SN at 16 weeks (K) and at 28 weeks of age. (L) Quantification of striatal (M) dopamine
and (N) DOPAC by HPLC at 16 weeks of age. Quantification of striatal (O) dopamine and (P) DOPAC by HPLC at 28 weeks of age in the modest pathology stage. Data are means ± SEM
of six to eight mice per group. ****P < 0.0001, ***P < 0.001, **P < 0.01, *P < 0.05. Data were analyzed by two-way ANOVA with Bonferroni’s multiple comparison test.

Ghosh et al., Sci. Transl. Med. 8, 368ra174 (2016) 7 December 2016 7 of 17


SCIENCE TRANSLATIONAL MEDICINE | RESEARCH ARTICLE

Downloaded from http://stm.sciencemag.org/ on December 7, 2016


Fig. 5. MSDC-0160 modulates mTOR signaling in C. elegans and the MPTP mouse model of PD. (A) Images show dopaminergic neuron loss in worm strains BY250 and
JVR203 at 12 days of age. Scale bar, 10 mm. (B) Quantification of dopaminergic neuron loss (percentage of total number of dopamine neurons). (C) Quantification of dopa-
minergic neuron loss when C. elegans were fed bacteria expressing the empty vector L4440 as control. Quantification of dopaminergic neuron loss in C. elegans after knocking
down BRP44L/MPC-1 (D), AKT-1 (E), RHEB-1 (F), LET-363/mTOR (G) and AAK-1/AMPK (H). Mice were coadministered MPTP and MSDC-0160. After 1 day of pretreatment with
MSDC-0160 (30 mg/kg, by oral gavage) and 5 days of cotreatment with MSDC-0160 (30 mg/kg, by oral gavage) and MPTP (25 mg/kg per day intraperitoneally), mice were
euthanized, and SN tissue was dissected and analyzed by Western blot. (I) Representative Western blot illustrating the expression of p-mTOR (Ser2448), total mTOR, pS6 (Ser235/
Ser236), and total S6 in the SN. Bar graphs showing mean Western blot p-mTOR/mTOR ratios relative to b-actin (J) and mean Western blot p-S6/S6 ratios relative to b-actin
(K). (A to H) Data are means ± SEM (n = 3 experiments) and were analyzed by unpaired Student’s t test except (B), which was analyzed by two-way ANOVA with Bonferroni’s
multiple comparison test. (I and J) Data are means ± SEM of six to eight mice per group and were analyzed by two-way ANOVA with Bonferroni’s multiple comparison test.
***P < 0.001, **P < 0.01, *P < 0.05.

Western blotting of substantia nigra samples, we quantitated mTOR, experiments described above, MSDC-0160 treatment of the wild-
p-mTOR (Ser2448 and Ser2481), p70S6kinase, phosphorylated p70S6kinase type mice did not affect mTOR signaling proteins (fig. S7). Notably,
(Thr389), pS6 (Ser235/Ser236), S6, AKT, phosphorylated AKT (Thr308), En1+/− mice also had reduced LC3b/b-actin (Fig. 6, E, H, and K) and
regulated in development and DNA damage responses 1 (REDD1), p62/b-actin ratios (fig. S7G) relative to the wild-type mice, suggest-
LC3b, and p62 (Fig. 6, A to E, and fig. S7). These data show that all ing that the genetic model has a disturbance in autophagy. Treat-
of these pathways were perturbed in the En1+/− mice versus wild-type ment of the En1+/− mice with MSDC-0160 returned the amounts
animals (P value between 0.01 and 0.0001). Treatment with MSDC- of LC3b and p62 proteins to those seen in wild-type mice. Again,
0160 moved all of these changes toward the values observed in wild- there was no significant effect of MSDC-0160 treatment on the expres-
type mice. The changes included those in mTOR signaling—increases sion of these autophagy markers in wild-type mice (Fig. 6, E and K,
in p-mTOR Ser2448/mTOR/b-actin, phosphorylated p70S6kinase/ and fig. S7, C to G).
p70S6kinase/b-actin, pS6/S6/b-actin (Fig. 6, B to D, F, G, I, and J)— Together, the results indicate that modulation of MPC leads to an
and also REDD1/b-actin ratios (fig. S7, C and F). As in the MPTP immediate effect on mitochondrial metabolism associated with a later

Ghosh et al., Sci. Transl. Med. 8, 368ra174 (2016) 7 December 2016 8 of 17


SCIENCE TRANSLATIONAL MEDICINE | RESEARCH ARTICLE

Downloaded from http://stm.sciencemag.org/ on December 7, 2016

Fig. 6. MSDC-0160 down-regulates mTOR signaling and restores autophagy in the En1+/− genetic mouse model of PD. (A) Representative Western blot illustrating the
expression of p-mTOR (Ser2448), mTOR, phosphorylated p70S6kinase (p-p70S6K) (Thr389), p70S6kinase, pS6 (Ser235/Ser236), S6, and LC3b in the SN. Bar graphs showing mean
Western blot p-mTOR/mTOR ratios relative to b-actin (B), mean Western blot p-p70S6K/p70S6K ratios relative to b-actin (C), mean Western blot pS6/S6 ratios relative to b-actin
(D), and mean Western blot LC3b/b-actin ratio (E). (F) Immunostaining for TH and p-mTOR (Ser2448) in SN; insets demonstrate overlap of TH and p-mTOR. (G) Immunostaining
for TH and pS6 (Ser235/Ser236) in the SN. (H) Immunostaining for TH and LC3b. DAPI, 4′,6-diamidino-2-phenylindole. Intensity of p-mTOR (Ser2448) (I), pS6 (Ser235/Ser236)
(J), and LC3b (K) in the TH-positive neurons [a.u. (arbitrary units)]. Scale bars, 10 mm. Data are means ± SEM of six to eight mice per group. ****P < 0.0001, ***P < 0.001,
**P < 0.01, *P < 0.05. Data were analyzed by two-way ANOVA with Bonferroni’s multiple comparison test.

Ghosh et al., Sci. Transl. Med. 8, 368ra174 (2016) 7 December 2016 9 of 17


SCIENCE TRANSLATIONAL MEDICINE | RESEARCH ARTICLE

reduction in the overactivation of the mTOR pathway. The data sug- in LPS-activated BV2 cells treated with MSDC-0160. A reduction in
gest that MPC modulation induces autophagy as part of the response mTOR activation caused by LPS was only observed after 24 hours of
that prevents neurodegeneration. incubation with MSDC-0160 (Fig. 7P and fig. S9, B and C). Subse-
quently, LPS-induced activation of S6 was also attenuated by
Role of inflammation in the effects of MPC modulation MSDC-0160 in BV2 cells (Fig. 7Q). However, no changes in the
Given that neuroinflammation is also considered a contributory factor AKT pathway (pAKT/AKT ratio) were detected at any time point
to PD pathogenesis and occurs in several PD animal models (60), we (fig. S9, B and C). This supports the notion that the anti-inflammatory
were curious whether modulation of MPC with MSDC-0160 treatment effects of modulating MPC are downstream of the immediate effects
would directly reduce inflammation. We found that MSDC-0160 re- on metabolism. Modulation of MPC had anti-inflammatory
duced glial fibrillary acidic protein (GFAP; an astrocyte marker), io- consequences in cellular and mouse models of inflammation and
nized calcium-binding adapter molecule 1 (Iba-1; a microglial PD. The earliest measured effects involved direct modulation of pyr-
marker), and inducible nitric oxide synthase (iNOS) expression in uvate metabolism and protection of oxidative metabolism followed by
the substantia nigra of MPTP-treated mice (Fig. 7, A to D, and fig. changes in the mTOR/AKT pathways (increases in mTOR phospho-
S8D). Similarly, in En1+/− mice, treatment with MSDC-0160 at the rylation while AKT activation was reduced).
modest pathology stage attenuated expression of Iba-1, GFAP, and
iNOS in the ventral midbrain, as assayed by immunoblot (Fig. 7, E
to G, and fig. S8F). In addition, immunohistochemical analysis for DISCUSSION

Downloaded from http://stm.sciencemag.org/ on December 7, 2016


Iba-1 and GFAP in sections through the substantia nigra of En1+/− We show that modulating the MPC consistently normalized metabo-
mice showed that MSDC-0160 reduced elevated astrogliosis and mi- lism and downstream molecular pathways that are dysregulated in a
crogliosis (Fig. 7H and fig. S8G). range of cell, nematode, and mouse models of PD (both toxin-induced
To determine in a less complex model whether modulation of and genetic). As a consequence, autophagy and neuroinflammatory
MPC directly affected inflammatory cells, we moved to a microglia changes were normalized, and survival of substantia nigra dopamin-
cell line. As expected (see above), we observed expression of MPC- ergic neurons was promoted.
1 and MPC-2 in mouse BV2 microglia cells (fig. S8, A and B) (61). Modulation of MPC function has been tested as a therapeutic stra-
To induce inflammation in these cells, we used pretreatment (1 hour) tegy in diabetes. The disease processes in diabetes and PD share sev-
with lipopolysaccharide (LPS; 1 mg/ml), which is an endotoxin that in- eral features, with metabolic alterations and inflammation being
duces inflammation in vitro and in vivo. We found that in BV2 cell ly- common to both diseases (35–38, 64) and specific metabolic genes be-
sates, 10 mM MSDC-0160 blocked LPS-induced increases in iNOS ing linked to PD (64). Attempts to modify the course of PD with an
expression (fig. S8C). We also confirmed our findings in mouse primary antidiabetic agent of the TZD class (pioglitazone) have yielded
microglia cells. After harvesting primary microglia (62), we used immu- conflicting results, possibly because of inadequate engagement of the
nofluorescence to determine that MSDC-0160 pretreatment (1 hour) target in the brain (8, 32). A safety and efficacy trial of the antidiabetic
blocked LPS-induced iNOS expression (Fig. 7I). Next, we performed a agent, exendin-4, a glucagon-like peptide-1 agonist, suggested a reduc-
functional Griess assay on supernatant from the treated BV2 cells to de- tion in motor and cognitive decline in PD patients, which was main-
tect the concentration of nitrite and found that MSDC-0160 blocked the tained even 12 months after cessation of treatment (65, 66). These
LPS-induced increase in nitrite (Fig. 7J). Promoter regions of proinflam- findings tentatively support the idea that disease mechanisms in diabe-
matory molecules contain the DNA binding site for nuclear factor kB tes and PD have common features, warranting further exploration of
(NF-kB), a transcription factor involved in inflammatory responses. The molecular therapeutic targets that could be common to the two diseases.
inhibition of NF-kB activation reduces the induction of proinflamma- Here, modulation of MPC by MSDC-0160 normalized mTOR ac-
tory molecules in PD animal models (63). Therefore, we monitored tivity in response to multiple insults (Figs. 5 and 6). mTOR signaling is
whether MPC modulation blocked the entry of p65 into the nucleus. known to be activated in clinical samples from PD patients (67, 68).
Immunofluorescence of BV2 cells stained with antibodies against Iba- Aberrant mTOR signaling is also found in several models of PD, in-
1 and phosphorylated (Ser276) p65 showed that pretreatment with cluding the MPTP toxin rodent model (69–71) and recently in the
MSDC-0160 prevented nuclear accumulation of phosphorylated p65 En1+/− genetic mouse model (53). The mTOR pathway interacts with
(Fig. 7K). Consistent with the immunofluorescence data, we found a stress response protein, RTP801/REDD1, and has been suggested to
increased expression of phosphorylated (Ser276) p65 in the nuclear frac- control neuronal death in PD (72). Rapamycin, which is an inhibitor
tion of BV2 microglial cells (Fig. 7L). Next, we analyzed supernatants of mTOR signaling, promotes longevity, protects dopaminergic and
from BV2 cells on a Meso Scale platform to detect multiple cytokines. other neurons, and stimulates autophagy in the MPTP mouse model
We determined that MSDC-0160 pretreatment lessened the LPS-related of PD (72). In the mTOR signaling complex 1 (mTORC1), mTOR is
induction of expression of IL-1b (Fig. 7M), TNF-a (Fig. 7N), and IL-6 predominantly phosphorylated on Ser2448, whereas in the mTORC2, it
(Fig. 7O). Together, our results from BV2 cell cultures indicated that is predominantly autophosphorylated on Ser2481. In our PD models,
modulating MPC by MSDC-0160 directly in microglia prevented in- we observed increased phosphorylation of mTOR at Ser2448 and
flammation induced by LPS. Ser2481. We found that MSDC-0160 reduced phosphorylation at
Consistent with the direct effect of MSDC-0160 on mitochondrial Ser2448, but not at Ser2481, mimicking the effects of rapamycin treatment,
function in LUHMES cells subjected to MPP+ (fig. S2), MSDC-0160 which inhibits signaling by mTORC1 but not mTORC2. The mTORC1
normalized oxygen consumption in LPS-activated BV2 cells when complex also phosphorylates the ribosomal protein p70S6kinase, an
added 1 hour before or after LPS addition, demonstrating its immediate AGC kinase family member, on its hydrophobic motif site, Thr389 (73).
action on mitochondrial function (fig. S9A). Also consistent with the Our studies showed that the En1+/− mouse exhibits greater phos-
effects in the LUHMES neuronal cultures (fig. S3, A and B), no early phorylation of p70S6kinase and its downstream target molecule ribo-
changes (2 and 6 hours) in the p-mTOR/mTOR ratio could be detected somal S6 and that modulation of MPC reduced the phosphorylation

Ghosh et al., Sci. Transl. Med. 8, 368ra174 (2016) 7 December 2016 10 of 17


SCIENCE TRANSLATIONAL MEDICINE | RESEARCH ARTICLE

Downloaded from http://stm.sciencemag.org/ on December 7, 2016

Fig. 7. MSDC-0160 attenuates inflammation in animal models and in mouse microglial cells. (A) Representative Western blot of Iba-1 and iNOS in the SN. Bar graphs
show mean Iba-1/b-actin (B) and mean iNOS/b-actin in SN (C). (D) Iba-1 immunostaining in SN. Bar graph shows number of Iba-1–positive cells in SN. (E) Representative blot of
Iba-1 and iNOS in SN. Bar graph shows mean Iba-1/b-actin (F) and mean iNOS/b-actin in SN (G). (H) Iba-1 immunostaining in SN. Bar graph shows number of Iba-1–positive
cells in SN. (I) Immunocytochemistry of Iba-1 and iNOS in primary mouse microglial cells. (J) Nitrite measurement and (K) immunocytochemistry of Iba-1 and NF-kB–p65 in BV2
mouse microglial cells. (L) Representative Western blot of NF-kB–p65 from cytosol and nucleus in BV2 cells. Concentration of interleukin-1b (IL-1b) (pg/ml) (M), tumor necrosis
factor–a (TNF-a) (pg/ml) (N), and IL-6 (pg/ml) (O) in BV2 cells. Immunocytochemistry of Iba-1 and p-mTOR (P) and Iba-1 and pS6 (Q) in BV2 cells. Scale bars, 50 mm (D, H, and I),
5 mm (K, P, and Q). (A to H) Data are means ± SEM of six to eight mice per group. (I to Q) Data are means ± SEM (n = 3 experiments). ****P < 0.0001, ***P < 0.001, **P < 0.01, *P <
0.05. Data were analyzed by two-way ANOVA with Bonferroni’s multiple comparison test.

Ghosh et al., Sci. Transl. Med. 8, 368ra174 (2016) 7 December 2016 11 of 17


SCIENCE TRANSLATIONAL MEDICINE | RESEARCH ARTICLE

to wild-type levels. We also observed increased expression of REDD1 (85, 86), and because mTOR inhibitors can inhibit release of inflam-
protein and inhibition of AKT phosphorylation at Thr308 in the En1 matory cytokines from activated macrophages (85, 86), we speculated
+/−
mice. These changes are all relevant to the pathophysiology of PD that MPC modulation would also mitigate mTOR activation in stimu-
because dopaminergic neurons in PD exhibit down-regulation of the lated microglia. In LPS-stimulated BV2 cells, exposure to MSDC-0160
Ser473- and Thr308-phosphorylated forms of AKT (67). REDD1, which caused a down-regulation of p-mTOR (Ser2448) and its target mole-
functions upstream of AKT, is induced during the neurodegenerative cule, pS6 (Fig. 7). As was the case for neurons, these effects were tem-
process (72). Modulation of MPC by MSDC-0160 reduced REDD1 porally downstream from the direct effects of mitochondrial
(fig. S7, C and F). Although this could be due to attenuation of metabolism in the microglia. These results are key because they
p70S6kinase activation, REDD1 was increased in response to multiple showed that MPC was engaged in both neurons and glial cells and
stresses, and it is likely that the upstream modulation of stress path- that down-regulating the mTOR signaling pathway was a consequence
ways is a component of the metabolic modulation. in both cell types. On the basis of our observations, we propose that
Part of the metabolic modulation by MSDC-0160 in all of the strategies that modulate MPC can have dual beneficial effects in PD,
models we examined included increased phosphorylation of AKT at by both improving autophagy in neurons and reducing microglia
Thr308. Although the precise mechanism underlying this change is not activation (fig. S10). We suggest that modulation of MPC by MSDC-
clear, reduction in endoplasmic reticulum stress might play a role (74). 0160 reduced the effects of PD-causing insults, whether a toxin, a mis-
Some of the effects we observed after modulation of MPC function folded protein (a-synuclein), or an En1+/− mutation. The direct interaction
were consistent with previous work using mTOR inhibitors (72). No- of MSDC-0160 with the MPC complex has recently been modeled,

Downloaded from http://stm.sciencemag.org/ on December 7, 2016


tably, whereas those inhibitors directly inhibit mTOR, the action of showing a potentially unique pattern of interaction for both compo-
MSDC-0160 is on upstream metabolic changes deriving from the nents of the complex (87). Downstream events may involve specific
change in the mitochondrial handling of pyruvate (20, 21, 29). Thus, substrates, posttranslational modifications, and redox signals.
the effects of MSDC-0160 are mainly to prevent the activation of There are limitations to our study. It is not yet clear exactly how
mTOR produced by the metabolic changes rather than to directly in- modulation of MPC results in the downstream changes that we have
hibit mTOR kinase activity. This is consistent with our observations of documented. In the rodent PD models, we cannot determine whether
an immediate effect of MPC modulation on pyruvate utilization and ox- the neuroprotection was primarily the consequence of modulation of
ygen consumption, followed by a delayed inhibition of the mTOR MPC in the neurons or the microglia, or both, because our cell culture
pathway. studies clearly demonstrated that the MPC modulator MSDC-0160
It is well established that mTOR signaling is intricately tied to au- could positively affect either cell population when they are grown in
tophagy, which is a conserved homeostatic process by which un- isolation. Similar to what has been published in flies (22), we found
wanted cellular components are degraded by the lysosome (75). that knockdown of MPC-1 in the worm model blocked the response
Several studies have implicated changes in the lysosome autophagy to MSDC-0160, although the knockdown itself did not protect against
pathway in both idiopathic and certain rare genetic forms of PD neurodegeneration. This illustrates a difference between modulating
(76, 77). By boosting lysosomal biogenesis and inducing autophagy, MPC by reversible inhibition and completely diminishing its activity
rapamycin restores normal lysosomal activity and mitigates dopamin- by genetic manipulation. Notably, although the protective effects
ergic neurodegeneration in the MPTP mouse model of PD (78). We produced by MSDC-0160 treatment were consistent in all of the
show here that MSDC-0160 attenuated the activation of mTOR and PD models we tested, some dopaminergic neurons still died in the
its downstream signaling pathway in the subacute MPTP mouse different PD models, suggesting that it may not be possible to fully
model (Fig. 5, I to K). Similarly, we also found down-regulation of prevent neurodegeneration by MPC modulation alone.
mTOR signaling along with reduced LC3b and p62 expression in Antidiabetic agents have recently been proposed as possible disease-
En1+/− mice (Fig. 6 and fig. S6) and that modulation of MPC in this modifying therapies for PD (36, 88). Diabetes, insulin resistance, inflam-
slowly progressing mouse model of PD reversed these changes. Together, mation, and PD are all linked (89), and PD patients exhibit an increased
we found that modulation of MPC normalized autophagy in multiple prevalence of diabetes (90). A recent phase 2, multicenter study using
models of PD, which was likely to be key for preserving the nigral pioglitazone, a member of the first generation of TZDs, failed to show a
dopaminergic neurons. disease-modifying effect in PD (8). A retrospective epidemiological
Autophagic signals modulate inflammatory pathways (79–81), and study, however, determined that persons with diabetes with a prescrip-
neuroinflammation is an integral part of the pathogenesis of PD (82–84). tion for a TZD insulin sensitizer drug were less likely to develop PD
Neuroinflammation was not only evident after LPS or MPTP treat- (32), suggesting that intervention with TZD insulin sensitizers early
ment but was also found in the substantia nigra of En1+/− mice. in the PD disease process, or its prodromal phase, would be more ef-
We hypothesized that modulation of MPC influences inflammatory fective. First-generation TZDs, such as pioglitazone, not only modu-
pathways. We found that modulation of MPC mitigated inflammation late the MPC complex. In contrast to MSDC-0160, they are also direct
in all of the models we examined. In both MPTP-treated mice and activators of the nuclear transcription factor PPARg, which drives
En1-deficient mice, modulation of MPC attenuated the activation of dose-limiting side effects (29, 91). Consequently, the safety profile of
GFAP-positive astrocytes and Iba-1–positive microglia. The reduced MSDC-0160 is more favorable, which allows dosing to higher expo-
inflammatory response was likely not just the result of reduced neu- sures than achieved with pioglitazone and thus makes it more likely to
rodegeneration but also the result of direct effects of MSDC-0160 on significantly engage the target MPC in the brain. The safety of MSDC-
the glial cells. We deduced this from our observations in cultured 0160 has been demonstrated in people with diabetes and Alzheimer’s dis-
mouse BV2 cells, in which we found that modulation of MPC by ease during 3 months of exposure (41, 42), with tentative evidence of
MSDC-0160 blocked nuclear transport of p65 and reduced cytokine target engagement in the brain from [18F]deoxyglucose imaging scans.
release in response to LPS administration. Given that mTOR also In these trials, exposure to circulating MSDC-0160 and its metabolite
plays a crucial role in the regulation of the innate immune system was 50% higher than could be achieved with the highest approved dose

Ghosh et al., Sci. Transl. Med. 8, 368ra174 (2016) 7 December 2016 12 of 17


SCIENCE TRANSLATIONAL MEDICINE | RESEARCH ARTICLE

of pioglitazone (45 mg, once daily). In response to MSDC-0160, hemo- of varying concentrations, the exact dose ingested by the worm cannot
globin A1c and fasting blood glucose in diabetic patients fell to levels be accurately calculated. Therefore, the maximum concentration that
similar to those of people taking pioglitazone, but fluid retention and had a neuroprotective effect was used in subsequent experiments to
other side effects were less prominent with MSDC-0160 treatment (41). ensure efficacy. About 50 worms were added per well, and the plate
Given the effects of MPC modulation by MSDC-0160 in several was incubated at 20°C for 48 hours. After 48 hours, worms were moved
different PD models in this study and its favorable safety profile in and allowed to recover on unseeded nematode growth medium plates
humans, we believe that targeting MPC in PD is warranted. MPC as for several minutes before being mounted on an agarose pad with 3 mM
a target is attractive because it affects multiple processes that are im- levamisole and analyzed microscopically. Each worm was scored for
plicated in PD pathogenesis, including autophagy and neuroinflam- the presence of GFP-fluorescent dopaminergic neurons in the anterior
mation, through actions on both neurons and glia. These findings of the worm (CEP and ADE neurons). A minimum of 25 worms were
should stimulate the development of other MPC modulators as analyzed for each condition in three independent trials.
potential disease-modifying therapeutic agents in PD and related
neurodegenerative disorders. Measurement of brain and plasma drug concentrations
C57BL/6J mice were orally dosed with MSDC-0160 (30 mg/kg)
suspended in 1% low-viscosity methylcellulose with 0.01% Tween
MATERIALS AND METHODS 80 in distilled water (10 ml/kg) and then euthanized 2 and 4 hours
Study design later. Plasma and whole-brain homogenates were extracted and

Downloaded from http://stm.sciencemag.org/ on December 7, 2016


The goal of this study was to establish evidence, using multiple PD subjected to LC/mass spectrometry measurement of both active
models, that modulation of the mitochondrial pyruvate carrier drug and active alcohol metabolite (MSDC-0037), as previously de-
complex via MSDC-0160 is a viable approach to modify disease pro- scribed (41).
gression in PD. All experiments were blinded. Typically, one individ-
ual would randomize the animals, plates, and slides, and another MPTP mouse model
would analyze them. The minimum sample size for all experiments Ten- to 12-week-old male C57BL/6J mice weighing 24 to 28 g were
was held at six mice per group based on the design of previous studies housed under standard conditions: constant temperature (22°C), hu-
(51, 92). To improve our power, and thus our ability to statistically midity (relative, 30%), and a 12-hour light/dark cycle, with free access
detect smaller effects, many of our analyses included more rodents to food and water. Procedures were performed during daylight hours
per group and/or repeated measures. Further experimental details and were approved and supervised by the Institutional Animal Care
and protocols of each model, including animal care/handling and and Use Committee at the Van Andel Research Institute. Mice were
the number of biological/technical replicates, are in this section, in administered MSDC-0160 (30 mg/kg) by oral gavage beginning 24 hours
the Supplementary Materials, and in the figure legends. before MPTP (Sigma-Aldrich) treatment. Next, mice received five
consecutive doses of MPTP via intraperitoneal injection at 25 mg/kg
C. elegans strains and protocols per day (subacute regimen) along with coadministration of MSDC-
C. elegans strains were cultured and maintained on nematode growth 0160, followed by 6 days of MSDC-0160 treatment. MSDC-0160
medium containing a lawn of OP50 bacteria at 20° or 16°C using was dissolved in 1% methylcellulose with 0.01% Tween 80. Control
established techniques (93). The transgenic strain BY250 (vtIs7 mice received vehicle treatment (1% methylcellulose with 0.01%
[pDAT::GFP(pRB490)]) was provided by the laboratory of R. Blakely Tween 80). In the modest pathology stage, mice were administered
(Florida Atlantic University, Jupiter, FL). The strain TU3401 (sid-1(pk3321) MSDC-0160 (30 mg/kg per day) by oral gavage for 7 days starting
V; uIs69[pCFJ90(myo-2p::mCherry) + unc-119p::sid-1] V) was obtained 3 days after MPTP (25 mg/kg per day) treatment. Seven days after
from the Caenorhabditis Genetics Center, which is funded by the Na- MPTP treatment, mice were euthanized, and tissues were pro-
tional Institutes of Health Office of Research Infrastructure Programs cessed for further evaluation. Mice were randomized to the exper-
(P40 OD010440). The a-synuclein–expressing strain JVR107 was imental groups.
produced by transforming a Pdat-1::a-syn (A53T) construct into wild-
type C. elegans (N2 Bristol strain), generated by T. Iwatsubo (University En1+/− mouse model
of Tokyo, Japan). Additional strains were generated by crossing these The En1+/− heterozygous mice were maintained on an OF1 genetic
strains to generate JVR203 (Pdat-1::a-syn(A53T), Pges-1::RFP; vtIs7 background (52) and under the same conditions as the C57BL/6J
[Pdat-1::GFP(pRB490)]), JVR325 (Pdat-1::a-syn(A53T), Pges-1::RFP; mice. In the mild pathology stage, En1+/− mice were fed a diet of chow
vtIs7[Pdat-1::GFP(pRB490)]; sid-1(pk3321) V; uIs69[pCFJ90(myo-2p:: formulated to deliver MSDC-0160 (30 mg/kg) starting at 3 weeks of
mCherry) + unc-119p::sid-1] V) and JVR326 (vtIs7[Pdat-1::GFP age (at this point, no nigral cell death is detectable) and were eutha-
(pRB490)]; sid-1(pk3321) V; uIs69[pCFJ90(myo-2p::mCherry) + unc- nized at two different time points: 28 and 48 weeks. In the modest
119p::sid-1] V). pathology stage, En1+/− mice were fed a diet of chow formulated to
deliver MSDC-0160 (30 mg/kg) starting at 8 weeks of age (at this
C. elegans MPP+ neurodegeneration assay point, 15 to 20% nigral cell death is detectable) and were euthanized
Worms were synchronized, and L1 larvae (BY250 strain, GFP expres- at either 16 or 28 weeks. Mice were randomized to the groups, and
sion in dopaminergic neurons) were placed in 96-well plates control mice were fed regular chow.
containing OP50 bacteria in liquid culture (6 mg/ml) and MPP+
(0.75 mM) with MSDC-0160 (1, 10, or 100 mM). Wells containing Behavioral experiments
no MPP+ [substituted with water or vehicle (methylcellulose)] and Spontaneous activity data were collected and analyzed by an ANY-maze
with MSDC-0160 alone were also included. Because in these analyzer (Stoelting). The dimensions of the activity chamber (San Diego
experiments worms were soaked in solution containing MSDC-0160 Instruments) were 109 cm by 109 cm by 38 cm. Before any treatment,

Ghosh et al., Sci. Transl. Med. 8, 368ra174 (2016) 7 December 2016 13 of 17


SCIENCE TRANSLATIONAL MEDICINE | RESEARCH ARTICLE

mice were placed daily inside the chamber for 10 min for two consec- plates and supplemented with Advanced Dulbecco’s modified Eagle’s
utive days. Open-field activities were recorded for 10-min test sessions. medium/F12 medium. At 24 hours before the assay, 10 mM MPP+ or
We evaluated a total of seven parameters: total distance traveled, mean 10 mM LPS was added to appropriate wells, and Seahorse probes were
speed, maximum speed, time frozen, time mobile, time immobile, and hydrated with Seahorse calibrant solution (pH 7.4). The day of the
head distance traveled. For the rotarod experiment, a speed of 18 rpm assay, medium was removed and replaced with Seahorse media (pH
for C57BL/6J mice and 10 rpm for En1+/− mice was used. Before testing, 7.4). One hour before the assay, cells were treated with 10 mM MSDC-
each mouse was trained on the rotarod for 5 min on three consecutive 0160 (posttreatment conditions). Cells were then washed twice in Sea-
days. Mice were given a 7- to 10-min rest between rotarod recording horse medium and calibrated in a 37°C non-CO2 incubator for 1 hour.
sessions. All behavior experiments were conducted in a blinded fashion. Oxygen consumption was measured twice basally. Injections of vehi-
cle, 10 mM MSDC, or 10 mM MPP+/LPS (pretreatment conditions)
Measurement of fluorescence intensity and densitometry were then added, and six oxygen consumption rate measurements
A total of three to four sections per brain containing the substantia of 3 min each were taken in a blinded manner. Measurements were
nigra and three mice per group were stained with antibodies directed normalized to 50,000 cells per well. A total of 15 wells per condition
against TH, p-mTOR, pS6, and LC3b. After immunofluorescence were measured, and three independent experiments for each cell line
staining, we took five 40× images from each section, blind-coded were performed.
them, and used the NIS-Elements AR 4.00.08 software (Nikon) to
quantify mean intensity of fluorescence of p-mTOR, pS6, LC3b, and Statistical analysis

Downloaded from http://stm.sciencemag.org/ on December 7, 2016


TH in TH-immunoreactive neurons, as well as the number of LC3b- Except for the behavioral data from En1+/− mice, data were analyzed
immunoreactive punctae in TH-positive nigral neurons (53). with Prism 3.0 software (GraphPad Software). Raw data were first
analyzed using either one-way or two-way ANOVA, and then either
HPLC analysis for striatal monoamine detection Tukey’s or Bonferroni’s multiple comparison test was performed to
Striatal levels of dopamine and DOPAC were quantified as described compare all treatment groups. Differences with P < 0.05 were
previously (48). Briefly, striata tissues were collected, immediately fro- considered significant. Linear mixed-effects models (LMMs) (94) were
zen on dry ice, and stored at −80°C until analysis. On the day of anal- used to determine whether the interaction between the En1+/− geno-
ysis, tissues were sonicated in 0.2 M perchloric acid containing type and the MSDC drug treatment was significant.
isoproterenol (internal standard), and the homogenates were centri- There are two major advantages to using LMMs: first, they appro-
fuged at 20,000g for 15 min at 4°C. Dopamine and DOPAC were priately adjust for repeated measures but allow for the inclusion of
separated isocratically in a C18 reversed-phase column using an HPLC mice without repeated measures, unlike repeated-measures ANOVA;
system with an automatic sampler equipped with a refrigerated tem- second, LMMs adjust for a portion of the within-subject variation via
perature control and electrochemical detector (Thermo Fisher Scientif- random effects. Both of these features should result in superior model
ic). Data acquisition and analysis were performed using the Chromeleon performance and accuracy (95). Each LMM fit included independent
HPLC Software. intercepts to adjust for some of the within-subject variation observed
between mice at the initial time points. Random slopes were also
Pyruvate oxidation assay considered, but there were not enough data to include this effect.
The effect of MSDC-0160 on oxygen consumption rates in brain Analyses for each outcome were stratified by the week in which
mitochondria was analyzed using MitoXpress Xtra (Luxcel). This flu- chow treatment was initiated (that is, 3 or 8 weeks). For each outcome,
orescence reagent, which is quenched by oxygen, allows for the direct, the LMMs were first fit with a three-way interaction between genotype
real-time analysis of oxygen consumption. Mitochondria were (wild-type/En1+/−), time point (28 and 44 weeks or 16 and 28 weeks
prepared by differential centrifugation of freshly prepared Sprague- for 3- and 8-week treatment start times, respectively), and treatment
Dawley rat brain homogenates and resuspended in respiration buffer (control/MSDC). If this interaction was not found to be significant,
containing 250 mM sucrose, 15 mM KCl, 1 mM EGTA, 5 mM MgCl2, then it was removed from the model, and the model was then refit,
and 30 mM K2HPO4 (pH 7.4). MSDC-0160 concentration curves (0, including all two-way interactions (that is, genotype × treatment,
0.1, 0.3, 1.0, 3.0, and 10 mM) were prepared in dimethyl sulfoxide genotype × time, and treatment × time). If the genotype × time inter-
(0.2% final) and assayed in duplicate with 50 mg of mitochondrial pro- action was not significant, then it was removed; likewise, for the treat-
tein per well in a 96-well format (Corning). Malate (5 mM)/pyruvate ment × time interaction. The genotype × treatment interaction was
(5 mM)/adenosine 5′-diphosphate (1 mM), determined empirically retained regardless of significance because this interaction pertained
to be optimal substrate concentrations, were included in the reac- to one of the primary questions of interest, namely, does MSDC-
tion. Time-resolved fluorescence (excitation/emission, 380/645; time- 0160 improve the En1+/− measures. Thus, the minimal final model
resolved 30-ms delay; 100-ms read) was measured over a 2-hour time included at least the following variables: genotype, treatment, time,
course on a BioTek Synergy 2 plate reader. Respiration curves and and the genotype × treatment interaction.
kinetic analysis were conducted using R version 3.2.2 and GraphPad An FDR correction was used to adjust all of the P values calculated
Prism software, respectively. by the 18 LMMs. FDR corrections are less reliant upon the need for all
tests to be independent (96), which is particularly important because
Seahorse assay for cellular oxygen consumption many of our outcome measurements are highly correlated and, thus,
Basal cellular oxygen consumption was measured using the Seahorse not independent.
XFe96 Analyzer. Two days before experiments, BV2 cells were plated Observations were determined to be outliers if they were more
at 1.5 × 105 cells/ml (seeding density) in Seahorse 96-well utility plates than 3 SDs away from the mean of the residuals. All observations that
and supplemented with RMPI 1640 medium. LUHMES cells were met this criterion were investigated to ensure that the data were recorded
plated at 5 × 105 cells/ml (seeding density) in Seahorse 96-well utility accurately (for example, not miskeyed) and that the measurements were

Ghosh et al., Sci. Transl. Med. 8, 368ra174 (2016) 7 December 2016 14 of 17


SCIENCE TRANSLATIONAL MEDICINE | RESEARCH ARTICLE

within the realm of plausibility. Measurements found to have record- S. Rolandelli, U. J. Kang, J. Young, J. Rao, M. M. Cook, L. Severt, K. Boyar, A randomized
clinical trial of high-dosage coenzyme Q10 in early Parkinson disease: No evidence of
ing errors were corrected. No observations were removed. All LMM
benefit. JAMA Neurol. 71, 543–552 (2014).
analyses were performed using R version 3.2.0 (97). 10. B. J. Snow, F. L. Rolfe, M. M. Lockhart, C. M. Frampton, J. D. O’Sullivan, V. Fung,
R. A. J. Smith, M. P. Murphy, K. M. Taylor; Protect Study Group, A double-blind,
placebo-controlled study to assess the mitochondria-targeted antioxidant MitoQ as a
SUPPLEMENTARY MATERIALS disease-modifying therapy in Parkinson’s disease. Mov. Disord. 25, 1670–1674 (2010).
www.sciencetranslationalmedicine.org/cgi/content/full/8/368/368ra174/DC1 11. C. W. Olanow, O. Rascol, R. Hauser, P. D. Feigin, J. Jankovic, A. Lang, W. Langston,
Materials and Methods E. Melamed, W. Poewe, F. Stocchi, E. Tolosa; ADAGIO Study Investigators, A double-blind,
Fig. S1. MSDC-0160 protects dopaminergic neurons against MPP+-induced toxicity in LUHMES delayed-start trial of rasagiline in Parkinson’s disease. N. Engl. J. Med. 361, 1268–1278
cells and C. elegans. (2009).
Fig. S2. Modulation of MPC improves mitochondrial function. 12. Parkinson Study Group PRECEPT Investigators, Mixed lineage kinase inhibitor CEP-1347
Fig. S3. Time-course treatment of MSDC-0160 in LUHMES cells. fails to delay disability in early Parkinson disease. Neurology 69, 1480–1490 (2007).
Fig. S4. MSDC-0160 improves behavioral parameters in the MPTP mouse model. 13. C. W. Olanow, A. H. V. Schapira, P. A. LeWitt, K. Kieburtz, D. Sauer, G. Olivieri, H. Pohlmann,
Fig. S5. MSDC-0160 improves behavioral activities in the En1+/− mice. J. Hubble, TCH346 as a neuroprotective drug in Parkinson’s disease: A double-blind,
Fig. S6. MSDC-0160 prevents dopaminergic neurodegeneration in the En1+/− mice. randomised, controlled trial. Lancet Neurol. 5, 1013–1020 (2006).
Fig. S7. MSDC-0160 down-regulates mTOR signaling pathway and restores autophagy in the 14. NINDS NET-PD Investigators, A randomized, double-blind, futility clinical trial of creatine
En1+/− mice. and minocycline in early Parkinson disease. Neurology 66, 664–671 (2006).
Fig. S8. MSDC-0160 attenuates inflammation in BV2 cells and in vivo animal models. 15. L. R. Gray, S. C. Tompkins, E. B. Taylor, Regulation of pyruvate metabolism and human
Fig. S9. MSDC-0160 improves mitochondrial function and blocks mTOR activation in BV2 cells. disease. Cell. Mol. Life Sci. 71, 2577–2604 (2014).
Fig. S10. Schematic diagram of proposed mechanism of action of MSDC-0160 in glial cells and 16. A. P. Halestrap, The mitochondrial pyruvate carrier: Has it been unearthed at last?

Downloaded from http://stm.sciencemag.org/ on December 7, 2016


dopaminergic neurons. Cell Metab. 16, 141–143 (2012).
Table S1. LMM results. 17. D. K. Bricker, E. B. Taylor, J. C. Schell, T. Orsak, A. Boutron, Y.-C. Chen, J. E. Cox,
Reference (98) C. M. Cardon, J. G. Van Vranken, N. Dephoure, C. Redin, S. Boudina, S. P. Gygi, M. Brivet,
C. S. Thummel, J. Rutter, A mitochondrial pyruvate carrier required for pyruvate uptake in
yeast, Drosophila, and humans. Science 337, 96–100 (2012).
18. S. Herzig, E. Raemy, S. Montessuit, J.-L. Veuthey, N. Zamboni, B. Westermann, E. R. S. Kunji,
REFERENCES AND NOTES J.-C. Martinou, Identification and functional expression of the mitochondrial pyruvate
1. R. B. Postuma, D. Berg, M. Stern, W. Poewe, C. W. Olanow, W. Oertel, J. Obeso, K. Marek, carrier. Science 337, 93–96 (2012).
I. Litvan, A. E. Lang, G. Halliday, C. G. Goetz, T. Gasser, B. Dubois, P. Chan, B. R. Bloem, 19. P. A. Vigueira, K. S. McCommis, G. G. Schweitzer, M. S. Remedi, K. T. Chambers, X. Fu,
C. H. Adler, G. Deuschl, MDS clinical diagnostic criteria for Parkinson’s disease. Mov. W. G. McDonald, S. L. Cole, J. R. Colca, R. F. Kletzien, S. C. Burgess, B. N. Finck,
Disord. 30, 1591–1601 (2015). Mitochondrial pyruvate carrier 2 hypomorphism in mice leads to defects in glucose-
2. D. Berg, R. B. Postuma, B. Bloem, P. Chan, B. Dubois, T. Gasser, C. G. Goetz, G. M. Halliday, stimulated insulin secretion. Cell Rep. 7, 2042–2053 (2014).
J. Hardy, A. E. Lang, I. Litvan, K. Marek, J. Obeso, W. Oertel, C. W. Olanow, W. Poewe, 20. L. R. Gray, M. R. Sultana, A. J. Rauckhorst, L. Oonthonpan, S. C. Tompkins, A. Sharma, X. Fu,
M. Stern, G. Deuschl, Time to redefine PD? Introductory statement of the MDS Task Force R. Miao, A. D. Pewa, K. S. Brown, E. E. Lane, A. Dohlman, D. Zepeda-Orozco, J. Xie, J. Rutter,
on the definition of Parkinson’s disease. Mov. Disord. 29, 454–462 (2014). A. W. Norris, J. E. Cox, S. C. Burgess, M. J. Potthoff, E. B. Taylor, Hepatic mitochondrial
3. B. J. Ryan, S. Hoek, E. A. Fon, R. Wade-Martins, Mitochondrial dysfunction and mitophagy pyruvate carrier 1 is required for efficient regulation of gluconeogenesis and whole-body
in Parkinson’s: From familial to sporadic disease. Trends Biochem. Sci. 40, 200–210 (2015). glucose homeostasis. Cell Metab. 22, 669–681 (2015).
4. I. Ahmed, Y. Liang, S. Schools, V. L. Dawson, T. M. Dawson, J. M. Savitt, Development 21. K. S. McCommis, Z. Chen, X. Fu, W. G. McDonald, J. R. Colca, R. F. Kletzien, S. C. Burgess,
and characterization of a new Parkinson’s disease model resulting from impaired B. N. Finck, Loss of mitochondrial pyruvate carrier 2 in the liver leads to defects in
autophagy. J. Neurosci. 32, 16503–16509 (2012). gluconeogenesis and compensation via pyruvate-alanine cycling. Cell Metab. 22,
5. C. Cebrián, J. D. Loike, D. Sulzer, Neuroinflammation in Parkinson’s disease animal models: 682–694 (2015).
A cell stress response or a step in neurodegeneration?. Curr. Top. Behav. Neurosci. 22, 22. J. R. Colca, W. G. McDonald, G. S. Cavey, S. L. Cole, D. D. Holewa, A. S. Brightwell-Conrad,
237–270 (2015). C. L. Wolfe, J. S. Wheeler, K. R. Coulter, P. M. Kilkuskie, E. Gracheva, Y. Korshunova,
6. L. V. Kalia, S. K. Kalia, A. E. Lang, Disease-modifying strategies for Parkinson’s disease. M. Trusgnich, R. Karr, S. E. Wiley, A. S. Divakaruni, A. N. Murphy, P. A. Vigueira, B. N. Finck,
Mov. Disord. 30, 1442–1450 (2015). R. F. Kletzien, Identification of a mitochondrial target of thiazolidinedione insulin
7. D. Athauda, T. Foltynie, The ongoing pursuit of neuroprotective therapies in Parkinson sensitizers (mTOT)—Relationship to newly identified mitochondrial pyruvate carrier
disease. Nat. Rev. Neurol. 11, 25–40 (2015). proteins. PLOS ONE 8, e61551 (2013).
8. NINDS Exploratory Trials in Parkinson Disease (NET-PD) FS-ZONE Investigators, 23. J. R. Colca, W. G. McDonald, R. F. Kletzien, Mitochondrial target of thiazolidinediones.
Pioglitazone in early Parkinson’s disease: A phase 2, multicentre, double-blind Diabetes Obes. Metab. 16, 1048–1054 (2014).
randomised trial. Lancet Neurol. 14, 795–803 (2015). 24. A. S. Divakaruni, S. E. Wiley, G. W. Rogers, A. Y. Andreyev, S. Petrosyan, M. Loviscach,
9. Parkinson Study Group QE3 Investigators, M. F. Beal, D. Oakes, I. Shoulson, C. Henchcliffe, E. A. Wall, N. Yadava, A. P. Heuck, D. A. Ferrick, R. R. Henry, W. G. McDonald, J. R. Colca,
W. R. Galpern, R. Haas, J. L. Juncos, J. G. Nutt, T. S. Voss, B. Ravina, C. M. Shults, K. Helles, M. I. Simon, T. P. Ciaraldi, A. N. Murphy, Thiazolidinediones are acute, specific inhibitors of
V. Snively, M. F. Lew, B. Griebner, A. Watts, S. Gao, E. Pourcher, L. Bond, K. Kompoliti, the mitochondrial pyruvate carrier. Proc. Natl. Acad. Sci. U.S.A. 110, 5422–5427 (2013).
P. Agarwal, C. Sia, M. Jog, L. Cole, M. Sultana, R. Kurlan, I. Richard, C. Deeley, C. H. Waters, 25. C. A. Hofmann, J. R. Colca, New oral thiazolidinedione antidiabetic agents act as insulin
A. Figueroa, A. Arkun, M. Brodsky, W. G. Ondo, C. B. Hunter, J. Jimenez-Shahed, A. Palao, sensitizers. Diabetes Care 15, 1075–1078 (1992).
J. M. Miyasaki, J. So, J. Tetrud, L. Reys, K. Smith, C. Singer, A. Blenke, D. S. Russell, C. Cotto, 26. N. M. Vacanti, A. S. Divakaruni, C. R. Green, S. J. Parker, R. R. Henry, T. P. Ciaraldi,
J. H. Friedman, M. Lannon, L. Zhang, E. Drasby, R. Kumar, T. Subramanian, D. S. Ford, A. N. Murphy, C. M. Metallo, Regulation of substrate utilization by the mitochondrial
D. A. Grimes, D. Cote, J. Conway, A. D. Siderowf, M. L. Evatt, B. Sommerfeld, pyruvate carrier. Mol. Cell 56, 425–435 (2014).
A. N. Lieberman, M. S. Okun, R. L. Rodriguez, S. Merritt, C. L. Swartz, W. R. W. Martin, 27. C. Yang, B. Ko, C. T. Hensley, L. Jiang, A. T. Wasti, J. Kim, J. Sudderth, M. A. Calvaruso,
P. King, N. Stover, S. Guthrie, R. L. Watts, A. Ahmed, H. H. Fernandez, A. Winters, Z. Mari, L. Lumata, M. Mitsche, J. Rutter, M. E. Merritt, R. J. DeBerardinis, Glutamine oxidation
T. M. Dawson, B. Dunlop, A. S. Feigin, B. Shannon, M. J. Nirenberg, M. Ogg, S. A. Ellias, maintains the TCA cycle and cell survival during impaired mitochondrial pyruvate
C.-A. Thomas, K. Frei, I. Bodis-Wollner, S. Glazman, T. Mayer, R. A. Hauser, R. Pahwa, transport. Mol. Cell 56, 414–424 (2014).
A. Langhammer, R. Ranawaya, L. Derwent, K. D. Sethi, B. Farrow, R. Prakash, I. Litvan, 28. R. Scatena, G. E. Martorana, P. Bottoni, B. Giardina, Mitochondrial dysfunction by synthetic
A. Robinson, A. Sahay, M. Gartner, V. K. Hinson, S. Markind, M. Pelikan, J. S. Perlmutter, ligands of peroxisome proliferator activated receptors (PPARs). IUBMB Life 56, 477–482
J. Hartlein, E. Molho, S. Evans, C. H. Adler, A. Duffy, M. Lind, L. Elmer, K. Davis, J. Spears, (2004).
S. Wilson, M. A. Leehey, N. Hermanowicz, S. Niswonger, H. A. Shill, S. Obradov, A. Rajput, 29. J. R. Colca, The TZD insulin sensitizer clue provides a new route into diabetes drug
M. Cowper, S. Lessig, D. Song, D. Fontaine, C. Zadikoff, K. Williams, K. A. Blindauer, discovery. Expert Opin. Drug Discov. 10, 1259–1270 (2015).
J. Bergholte, C. S. Propsom, M. A. Stacy, J. Field, D. Mihaila, M. Chilton, E. Y. Uc, J. Sieren, 30. D. Levin, S. Bell, R. Sund, S. A. Hartikainen, J. Tuomilehto, E. Pukkala, I. Keskimäki,
D. K. Simon, L. Kraics, A. Silver, J. T. Boyd, R. W. Hamill, C. Ingvoldstad, J. Young, K. Thomas, E. Badrick, A. G. Renehan, I. E. Buchan, S. L. Bowker, J. K. Minhas-Sandhu, Z. Zafari,
S. K. Kostyk, J. Wojcieszek, R. F. Pfeiffer, M. Panisset, M. Beland, S. G. Reich, M. Cines, C. Marra, J. A. Johnson, B. H. Stricker, A. G. Uitterlinden, A. Hofman, R. Ruiter,
N. Zappala, J. Rivest, R. Zweig, L. P. Lumina, C. L. Hilliard, S. Grill, M. Kellermann, P. Tuite, C. E. de Keyser, T. M. MacDonald, S. H. Wild, P. M. McKeigue, H. M. Colhoun; Scottish

Ghosh et al., Sci. Transl. Med. 8, 368ra174 (2016) 7 December 2016 15 of 17


SCIENCE TRANSLATIONAL MEDICINE | RESEARCH ARTICLE

Diabetes Research Network Epidemiology Group, Diabetes and Cancer Research 54. J. Q. Trojanowski, V. M.-Y. Lee, Aggregation of neurofilament and a-synuclein proteins in
Consortium, Pioglitazone and bladder cancer risk: A multipopulation pooled, cumulative Lewy bodies: Implications for the pathogenesis of Parkinson disease and Lewy body
exposure analysis. Diabetologia 58, 493–504 (2015). dementia. Arch. Neurol. 55, 151–152 (1998).
31. J. D. Lewis, L. A. Habel, C. P. Quesenberry, B. L. Strom, T. Peng, M. M. Hedderson, 55. W. Xu, L. Tan, J.-T. Yu, The link between the SNCA gene and parkinsonism. Neurobiol.
S. F. Ehrlich, R. Mamtani, W. Bilker, D. J. Vaughn, L. Nessel, S. K. Van Den Eeden, A. Ferrara, Aging 36, 1505–1518 (2015).
Pioglitazone use and risk of bladder cancer and other common cancers in persons with 56. J. Zhao, B. Zhai, S. P. Gygi, A. L. Goldberg, mTOR inhibition activates overall protein
diabetes. JAMA 314, 265–277 (2015). degradation by the ubiquitin proteasome system as well as by autophagy. Proc. Natl.
32. R. Brauer, K. Bhaskaran, N. Chaturvedi, D. T. Dexter, L. Smeeth, I. Douglas, Glitazone Acad. Sci. U.S.A. 112, 15790–15797 (2015).
treatment and incidence of Parkinson’s disease among people with diabetes: A 57. Y. Zhang, J. Nicholatos, J. R. Dreier, S. J. H. Ricoult, S. B. Widenmaier, G. S. Hotamisligil,
retrospective cohort study. PLOS Med. 12, e1001854 (2015). D. J. Kwiatkowski, B. D. Manning, Coordinated regulation of protein synthesis and
33. R. Kapadia, J.-H. Yi, R. Vemuganti, Mechanisms of anti-inflammatory and neuroprotective degradation by mTORC1. Nature 513, 440–443 (2014).
actions of PPAR-gamma agonists. Front. Biosci. 13, 1813–1826 (2008). 58. J. L. Jewell, K.-L. Guan, Nutrient signaling to mTOR and cell growth. Trends Biochem. Sci.
34. A. R. Carta, PPAR-g: Therapeutic prospects in Parkinson’s disease. Curr. Drug Targets 14, 38, 233–242 (2013).
743–751 (2013). 59. S. G. Kim, G. R. Buel, J. Blenis, Nutrient regulation of the mTOR complex 1 signaling
35. J. A. Santiago, J. A. Potashkin, Integrative network analysis unveils convergent molecular pathway. Mol. Cells 35, 463–473 (2013).
pathways in Parkinson’s disease and diabetes. PLOS ONE 8, e83940 (2013). 60. P. L. McGeer, E. G. McGeer, Glial reactions in Parkinson’s disease. Mov. Disord. 23, 474–483
36. I. Aviles-Olmos, P. Limousin, A. Lees, T. Foltynie, Parkinson’s disease, insulin resistance and (2008).
novel agents of neuroprotection. Brain 136, 374–384 (2013). 61. E. Blasi, R. Barluzzi, V. Bocchini, R. Mazzolla, F. Bistoni, Immortalization of murine
37. J. A. Santiago, J. A. Potashkin, System-based approaches to decode the molecular links in microglial cells by a v-raf/v-myc carrying retrovirus. J. Neuroimmunol. 27, 229–237 (1990).
Parkinson’s disease and diabetes. Neurobiol. Dis. 72, 84–91 (2014). 62. A. S. Harms, M. G. Tansey, Isolation of murine postnatal brain microglia for phenotypic
38. C. A. Downs, M. S. Faulkner, Toxic stress, inflammation and symptomatology of chronic characterization using magnetic cell separation technology. Methods Mol. Biol. 1041,

Downloaded from http://stm.sciencemag.org/ on December 7, 2016


complications in diabetes. World J. Diabetes 6, 554–565 (2015). 33–39 (2013).
39. C. W. Bolten, P. M. Blanner, W. G. McDonald, N. R. Staten, R. A. Mazzarella, G. B. Arhancet, 63. A. Ghosh, A. Roy, X. Liu, J. H. Kordower, E. J. Mufson, D. M. Hartley, S. Ghosh, R. L. Mosley,
M. F. Meier, D. J. Weiss, P. M. Sullivan, A. E. Hromockyj, R. F. Kletzien, J. R. Colca, Insulin H. E. Gendelman, K. Pahan, Selective inhibition of NF-kB activation prevents
sensitizing pharmacology of thiazolidinediones correlates with mitochondrial gene dopaminergic neuronal loss in a mouse model of Parkinson’s disease. Proc. Natl. Acad. Sci.
expression rather than activation of PPAR gamma. Gene Regul. Syst. Bio. 1, 73–82 (2007). U.S.A. 104, 18754–18759 (2007).
40. J. R. Colca, S. P. Tanis, W. G. McDonald, R. F. Kletzien, Insulin sensitizers in 2013: New 64. J. A. Santiago, J. A. Potashkin, Network-based metaanalysis identifies HNF4A and PTBP1
insights for the development of novel therapeutic agents to treat metabolic diseases. as longitudinally dynamic biomarkers for Parkinson’s disease. Proc. Natl. Acad. Sci. U.S.A.
Expert Opin. Investig. Drugs 23, 1–7 (2014). 112, 2257–2262 (2015).
41. J. R. Colca, J. T. VanderLugt, W. J. Adams, A. Shashlo, W. G. McDonald, J. Liang, R. Zhou, 65. I. Aviles-Olmos, J. Dickson, Z. Kefalopoulou, A. Djamshidian, P. Ell, T. Soderlund,
D. G. Orloff, Clinical proof-of-concept study with MSDC-0160, a prototype mTOT- P. Whitton, R. Wyse, T. Isaacs, A. Lees, P. Limousin, T. Foltynie, Exenatide and the
modulating insulin sensitizer. Clin. Pharmacol. Ther. 93, 352–359 (2013). treatment of patients with Parkinson’s disease. J. Clin. Invest. 123, 2730–2736 (2013).
42. R. C. Shah, D. C. Matthews, R. D. Andrews, A. W. Capuano, D. A. Fleischman, 66. I. Aviles-Olmos, J. Dickson, Z. Kefalopoulou, A. Djamshidian, J. Kahan, P. Ell, P. Whitton,
J. T. VanderLugt, J. R. Colca, An evaluation of MSDC-0160, a prototype mTOT modulating R. Wyse, T. Isaacs, A. Lees, P. Limousin, T. Foltynie, Motor and cognitive advantages persist
insulin sensitizer, in patients with mild Alzheimer's disease. Curr. Alzheimer Res. 11, 12 months after exenatide exposure in Parkinson’s disease. J. Parkinsons Dis. 4, 337–344 (2014).
564–573 (2014). 67. C. Malagelada, Z. H. Jin, L. A. Greene, RTP801 is induced in Parkinson’s disease and
43. J. Lotharius, S. Barg, P. Wiekop, C. Lundberg, H. K. Raymon, P. Brundin, Effect of mutant mediates neuron death by inhibiting Akt phosphorylation/activation. J. Neurosci. 28,
a-synuclein on dopamine homeostasis in a new human mesencephalic cell line. 14363–14371 (2008).
J. Biol. Chem. 277, 38884–38894 (2002). 68. C. Malagelada, E. J. Ryu, S. C. Biswas, V. Jackson-Lewis, L. A. Greene, RTP801 is elevated in
44. A. Ghosh, H. Saminathan, A. Kanthasamy, V. Anantharam, H. Jin, G. Sondarva, Parkinson brain substantia nigral neurons and mediates death in cellular models of
D. S. Harischandra, Z. Qian, A. Rana, A. G. Kanthasamy, The peptidyl-prolyl isomerase Pin1 Parkinson’s disease by a mechanism involving mammalian target of rapamycin
up-regulation and proapoptotic function in dopaminergic neurons: Relevance to the inactivation. J. Neurosci. 26, 9996–10005 (2006).
pathogenesis of Parkinson disease. J. Biol. Chem. 288, 21955–21971 (2013). 69. J. Deguil, F. Chavant, C. Lafay-Chebassier, M.-C. Pérault-Pochat, B. Fauconneau, S. Pain,
45. S. Cao, C. C. Gelwix, K. A. Caldwell, G. A. Caldwell, Torsin-mediated protection from Time course of MPTP toxicity on translational control protein expression in mice brain.
cellular stress in the dopaminergic neurons of Caenorhabditis elegans. J. Neurosci. 25, Toxicol. Lett. 196, 51–55 (2010).
3801–3812 (2005). 70. X.-Q. Bao, X.-C. Kong, C. Qian, D. Zhang, FLZ protects dopaminergic neuron through
46. L. A. Berkowitz, S. Hamamichi, A. L. Knight, A. J. Harrington, G. A. Caldwell, K. A. Caldwell, activating protein kinase B/mammalian target of rapamycin pathway and inhibiting
Application of a C. elegans dopamine neuron degeneration assay for the validation of RTP801 expression in Parkinson’s disease models. Neuroscience 202, 396–404 (2012).
potential Parkinson’s disease genes. J. Vis. Exp. 2008, 835 (2008). 71. M. Cao, X. Tan, W. Jin, H. Zheng, W. Xu, Y. Rui, L. Li, J. Cao, X. Wu, G. Cui, K. Ke, Y. Gao,
47. S. Przedborski, V. Jackson-Lewis, R. Djaldetti, G. Liberatore, M. Vila, S. Vukosavic, G. Almer, Upregulation of Ras homolog enriched in the brain (Rheb) in lipopolysaccharide-induced
The parkinsonian toxin MPTP: Action and mechanism. Restor. Neurol. Neurosci. 16, neuroinflammation. Neurochem. Int. 62, 406–417 (2013).
135–142 (2000). 72. C. Malagelada, Z. H. Jin, V. Jackson-Lewis, S. Przedborski, L. A. Greene, Rapamycin protects
48. A. Ghosh, A. Kanthasamy, J. Joseph, V. Anantharam, P. Srivastava, B. P. Dranka, against neuron death in in vitro and in vivo models of Parkinson’s disease. J. Neurosci. 30,
B. Kalyanaraman, A. G. Kanthasamy, Anti-inflammatory and neuroprotective effects of an 1166–1175 (2010).
orally active apocynin derivative in pre-clinical models of Parkinson's disease. 73. D. C. Fingar, J. Blenis, Target of rapamycin (TOR): An integrator of nutrient and growth
J. Neuroinflammation 9, 241 (2012). factor signals and coordinator of cell growth and cell cycle progression. Oncogene 23,
49. V. Jackson-Lewis, S. Przedborski, Protocol for the MPTP mouse model of Parkinson’s 3151–3171 (2004).
disease. Nat. Protoc. 2, 141–151 (2007). 74. H. W. Yung, D. S. Charnock-Jones, G. J. Burton, Regulation of AKT phosphorylation at
50. J. A. L. Hutter-Saunders, H. E. Gendelman, R. L. Mosley, Murine motor and behavior Ser473 and Thr308 by endoplasmic reticulum stress modulates substrate specificity in a
functional evaluations for acute 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) severity dependent manner. PLOS ONE 6, e17894 (2011).
intoxication. J. Neuroimmune Pharmacol. 7, 279–288 (2012). 75. D. J. Klionsky, S. D. Emr, Autophagy as a regulated pathway of cellular degradation.
51. A. Ghosh, K. Chandran, S. V. Kalivendi, J. Joseph, W. E. Antholine, C. J. Hillard, Science 290, 1717–1721 (2000).
A. Kanthasamy, A. Kanthasamy, B. Kalyanaraman, Neuroprotection by a mitochondria- 76. P. Anglade, S. Vyas, E. C. Hirsch, Y. Agid, Apoptosis in dopaminergic neurons of the
targeted drug in a Parkinson’s disease model. Free Radic. Biol. Med. 49, 1674–1684 (2010). human substantia nigra during normal aging. Histol. Histopathol. 12, 603–610 (1997).
52. L. Sonnier, G. Le Pen, A. Hartmann, J.-C. Bizot, F. Trovero, M.-O. Krebs, A. Prochiantz, 77. A. R. Winslow, C.-W. Chen, S. Corrochano, A. Acevedo-Arozena, D. E. Gordon, A. A. Peden,
Progressive loss of dopaminergic neurons in the ventral midbrain of adult mice M. Lichtenberg, F. M. Menzies, B. Ravikumar, S. Imarisio, S. Brown, C. J. O’Kane,
heterozygote for Engrailed1. J. Neurosci. 27, 1063–1071 (2007). D. C. Rubinsztein, a-Synuclein impairs macroautophagy: Implications for Parkinson’s
53. U. Nordström, G. Beauvais, A. Ghosh, B. C. Pulikkaparambil Sasidharan, M. Lundblad, disease. J. Cell Biol. 190, 1023–1037 (2010).
J. Fuchs, R. L. Joshi, J. W. Lipton, A. Roholt, S. Medicetty, T. N. Feinstein, J. A. Steiner, 78. B. Dehay, J. Bové, N. Rodríguez-Muela, C. Perier, A. Recasens, P. Boya, M. Vila, Pathogenic
M. L. Escobar Galvis, A. Prochiantz, P. Brundin, Progressive nigrostriatal terminal lysosomal depletion in Parkinson's disease. J. Neurosci. 30, 12535–12544 (2010).
dysfunction and degeneration in the engrailed1 heterozygous mouse model of 79. H.-F. Zheng, Y.-P. Yang, L.-F. Hu, M.-X. Wang, F. Wang, L.-D. Cao, D. Li, C.-J. Mao,
Parkinson's disease. Neurobiol. Dis. 73, 70–82 (2015). K.-P. Xiong, J.-D. Wang, C.-F. Liu, Autophagic impairment contributes to systemic

Ghosh et al., Sci. Transl. Med. 8, 368ra174 (2016) 7 December 2016 16 of 17


SCIENCE TRANSLATIONAL MEDICINE | RESEARCH ARTICLE

inflammation-induced dopaminergic neuron loss in the midbrain. PLOS ONE 8, e70472 96. Y. Benjamini, D. Yekutieli, The control of the false discovery rate in multiple testing under
(2013). dependency. Ann. Stat. 29, 1165–1188 (2001).
80. E. Aflaki, N. Moaven, D. K. Borger, G. Lopez, W. Westbroek, J. J. Chae, J. Marugan, 97. R Core Team, R: A Language and Environment for Statistical Computing (R Foundation for
S. Patnaik, E. Maniwang, A. N. Gonzalez, E. Sidransky, Lysosomal storage and impaired Statistical Computing, Vienna, Austria, 2013).
autophagy lead to inflammasome activation in Gaucher macrophages. Aging Cell 15, 98. D. Scholz, D. Pöltl, A. Genewsky, M. Weng, T. Waldmann, S. Schildknecht, M. Leist, Rapid,
77–88 (2016). complete and large-scale generation of post-mitotic neurons from the human
81. P. Prajapati, L. Sripada, K. Singh, K. Bhatelia, R. Singh, R. Singh, TNF-a regulates miRNA LUHMES cell line. J. Neurochem. 119, 957–971 (2011).
targeting mitochondrial complex-I and induces cell death in dopaminergic cells. Biochim.
Biophys. Acta 1852, 451–461 (2015). Acknowledgments: We thank R. Wyse, Cure Parkinson’s Trust, for helpful discussions. We
82. P. L. McGeer, S. Itagaki, B. E. Boyes, E. G. McGeer, Reactive microglia are positive for acknowledge the technical expertise of C. Cole, D. Dues, L. Kefene, and D. Marckini. We also
HLA-DR in the substantia nigra of Parkinson’s and Alzheimer’s disease brains. Neurology acknowledge excellent service from the Vivarium Core and the Biostatistics and Bioinformatics
38, 1285–1291 (1988). Core of the Van Andel Research Institute. We thank D. Nadziejka for manuscript editing
83. M. Mogi, M. Harada, P. Riederer, H. Narabayashi, K. Fujita, T. Nagatsu, Tumor necrosis assistance. We also thank J. Das and J. L. Leasure from the University of Houston for assistance
factor-a (TNF-a) increases both in the brain and in the cerebrospinal fluid from with the StereoInvestigator software and Western blot settings in their laboratories. Funding:
parkinsonian patients. Neurosci. Lett. 165, 208–210 (1994). The work presented here was supported by Cure Parkinson’s Trust (to P.B.), Campbell
84. R. J. Dobbs, A. Charlett, A. G. Purkiss, S. M. Dobbs, C. Weller, D. W. Peterson, Association of Foundation (to P.B.), Spica Foundation (to P.B.), and Van Andel Research Institute (to P.B. and
circulating TNF-a and IL-6 with ageing and parkinsonism. Acta Neurol. Scand. 100, 34–41 J.M.V.R.). Author contributions: A.G., J.R.C., J.M.V.R., and P.B. designed the study. A.G., T.T., S.G.,
(1999). E.N.H., E.M., E.S., and W.G.M. carried out the experiments: A.G. did all of the experiments except
85. T. Weichhart, M. D. Säemann, The multiple facets of mTOR in immunity. Trends Immunol. the following: C. elegans studies (T.T.), microglia cultures (E.N.H.), Seahorse assay (E.M.),
30, 218–226 (2009). drug pharmacodynamics (W.G.M.), and stereological cell counts in some experiments (E.S.).
86. F. Schmitz, A. Heit, S. Dreher, K. Eisenächer, J. Mages, T. Haas, A. Krug, K.-P. Janssen, A.G., T.T., S.G., E.N.H., Z.M., W.G.M., M.L.E.G., J.A.S., J.R.C., and J.H.K. performed analyses and

Downloaded from http://stm.sciencemag.org/ on December 7, 2016


C. J. Kirschning, H. Wagner, Mammalian target of rapamycin (mTOR) orchestrates the figure preparation. A.G., J.A.S., T.T., Z.M., M.L.E.G., J.H.K., J.R.C., and P.B. wrote the manuscript.
defense program of innate immune cells. Eur. J. Immunol. 38, 2981–2992 (2008). Competing interests: JRC is the cofounder and significant owner of Metabolic Solutions
87. C. F. Phelix, G. Perry, S. F. McHardy, Molecular docking analysis of mitochondrial pyruvate Development Company (MSDC), which is currently developing MSDC-0160 as a potential
carrier (MPC) 1 and 2 with pyruvate and inhibitors: Thiazolidinediones currently in clinical treatment for Alzheimer’s disease. MSDC-0160 is protected by several patents, including US
trials to delay mild cognitive impairment and Alzheimer’s disease. Alzheimers Dement. 8389556 B2 and EP 2001468 B1, which are assigned to MSDC. P.B. has received paid support as
11, P760–P761 (2015). a consultant from Renovo Neural Inc., Roche, Teva Pharmaceutical Industries, Lundbeck A/S,
88. M. P. Mattson, Interventions that improve body and brain bioenergetics for Parkinson’s AbbVie Inc., IOS Press Partners, and Versant Ventures. In addition, P.B. has received support for
disease risk reduction and therapy. J. Parkinsons Dis. 4, 1–13 (2014). his research from Renovo, Teva, and Lundbeck. P.B. has ownership interests in Acousort AB
89. P. Brundin, R. Wyse, Parkinson disease: Laying the foundations for disease-modifying and ParkCell AB. J.H.K. has received paid commercial support as a consultant from Cellular
therapies in PD. Nat. Rev. Neurol. 11, 553–555 (2015). Dynamics International Inc., Michael J. Fox Foundation, nLife Therapeutics S.L., NsGene,
90. J. A. Santiago, J. A. Potashkin, Shared dysregulated pathways lead to Parkinson’s disease Clintrex, NeuroDerm, and BrainEver. Data and materials availability: All enquiries regarding
and diabetes. Trends Mol. Med. 19, 176–186 (2013). MSDC-0160 should be directed to J.R.C.; MSDC-0160 will be made available through a material
91. M. Stumvoll, H.-U. Häring, Glitazones: Clinical effects and molecular mechanisms. transfer agreement.
Ann. Med. 34, 217–224 (2002).
92. D. C. Wu, V. Jackson-Lewis, M. Vila, K. Tieu, P. Teismann, C. Vadseth, D.-K. Choi, Submitted 22 July 2016
H. Ischiropoulos, S. Przedborski, Blockade of microglial activation is neuroprotective in Accepted 17 November 2016
the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine mouse model of Parkinson disease. Published 7 December 2016
J. Neurosci. 22, 1763–1771 (2002). 10.1126/scitranslmed.aag2210
93. S. Brenner, The genetics of Caenorhabditis elegans. Genetics 77, 71–94 (1974).
94. D. Bates, M. Mächler, B. M. Bolker, S. C. Walker, Fitting linear mixed-effects models using Citation: A. Ghosh, T. Tyson, S. George, E. N. Hildebrandt, J. A. Steiner, Z. Madaj, E. Schulz,
lme4. ArXiv: 1406.5823v1 (2014). E. Machiela, W. G. McDonald, M. L. Escobar Galvis, J. H. Kordower, J. M. Van Raamsdonk,
95. C. Krueger, L. Tian, A comparison of the general linear mixed model and repeated J. R. Colca, P. Brundin, Mitochondrial pyruvate carrier regulates autophagy, inflammation,
measures ANOVA using a dataset with multiple missing data points. Biol. Res. Nurs. 6, and neurodegeneration in experimental models of Parkinson’s disease. Sci. Transl. Med. 8,
151–157 (2004). 368ra174 (2016).

Ghosh et al., Sci. Transl. Med. 8, 368ra174 (2016) 7 December 2016 17 of 17


Mitochondrial pyruvate carrier regulates autophagy,
inflammation, and neurodegeneration in experimental models of
Parkinson's disease
Anamitra Ghosh, Trevor Tyson, Sonia George, Erin N. Hildebrandt,
Jennifer A. Steiner, Zachary Madaj, Emily Schulz, Emily Machiela,
William G. McDonald, Martha L. Escobar Galvis, Jeffrey H.
Kordower, Jeremy M. Van Raamsdonk, Jerry R. Colca and Patrik
Brundin (December 7, 2016)
Science Translational Medicine 8 (368), 368ra174. [doi:
10.1126/scitranslmed.aag2210]
Editor's Summary

A mitochondrial target for slowing Parkinson's disease


Currently, there are no disease-modifying treatments to stall progression of Parkinson's disease
(PD). A drug in development to treat diabetes might provide a new way to slow the progression of PD
according to new work by Ghosh and colleagues. The drug, MSDC-0160, targets a recently identified

Downloaded from http://stm.sciencemag.org/ on December 7, 2016


carrier of pyruvate (a major substrate for energy production) into the mitochondria. Ghosh et al. now
show that this drug, which attenuates the mitochondrial pyruvate carrier, blocks neurodegeneration in
several different cellular and animal models of PD. Furthermore, cellular autophagy was restored, and
neuroinflammation was reduced in two mouse models of PD. These results support continued
investigations into whether the mitochondrial pyruvate carrier will be a useful therapeutic target in PD.

The following resources related to this article are available online at http://stm.sciencemag.org.
This information is current as of December 7, 2016.

Article Tools Visit the online version of this article to access the personalization and
article tools:
http://stm.sciencemag.org/content/8/368/368ra174

Supplemental "Supplementary Materials"


Materials http://stm.sciencemag.org/content/suppl/2016/12/05/8.368.368ra174.DC1

Permissions Obtain information about reproducing this article:


http://www.sciencemag.org/about/permissions.dtl

Science Translational Medicine (print ISSN 1946-6234; online ISSN 1946-6242) is published
weekly, except the last week in December, by the American Association for the Advancement of
Science, 1200 New York Avenue, NW, Washington, DC 20005. Copyright 2016 by the American
Association for the Advancement of Science; all rights reserved. The title Science Translational
Medicine is a registered trademark of AAAS.
www.sciencetranslationalmedicine.org/cgi/content/full/8/368/368ra174/DC1

Supplementary Materials for


Mitochondrial pyruvate carrier regulates autophagy, inflammation, and
neurodegeneration in experimental models of Parkinson’s disease
Anamitra Ghosh, Trevor Tyson, Sonia George, Erin N. Hildebrandt, Jennifer A. Steiner,
Zachary Madaj, Emily Schulz, Emily Machiela, William G. McDonald,
Martha L. Escobar Galvis, Jeffrey H. Kordower, Jeremy M. Van Raamsdonk,
Jerry R. Colca, Patrik Brundin*

*Corresponding author. Email: patrik.brundin@vai.org

Published 7 December 2016, Sci. Transl. Med. 8, 368ra174 (2016)


DOI: 10.1126/scitranslmed.aag2210

This PDF file includes:

Materials and Methods


Fig. S1. MSDC-0160 protects dopaminergic neurons against MPP+-induced
toxicity in LUHMES cells and C. elegans.
Fig. S2. Modulation of MPC improves mitochondrial function.
Fig. S3. Time-course treatment of MSDC-0160 in LUHMES cells.
Fig. S4. MSDC-0160 improves behavioral parameters in the MPTP mouse model.
Fig. S5. MSDC-0160 improves behavioral activities in the En1+/− mice.
Fig. S6. MSDC-0160 prevents dopaminergic neurodegeneration in the En1+/−
mice.
Fig. S7. MSDC-0160 down-regulates mTOR signaling pathway and restores
autophagy in the En1+/− mice.
Fig. S8. MSDC-0160 attenuates inflammation in BV2 cells and in vivo animal
models.
Fig. S9. MSDC-0160 improves mitochondrial function and blocks mTOR
activation in BV2 cells.
Fig. S10. Schematic diagram of proposed mechanism of action of MSDC-0160 in
glial cells and dopaminergic neurons.
Table S1. LMM results.
Reference (98)
Supplementary Materials

Materials and Methods

Cell culture

Lund human mesencephalic (LUHMES) cells are a subclone of the tetracycline-

controlled, v-myc-over-expressing human mesencephalic-derived cell line MESC2.10

(ATCC, (43)). The LUHMES cells were plated and grown as described by (98 ). Briefly,

plastic cell culture flasks and multi-well plates were pre-coated with 50 μg/mL poly-L-

ornithine and 1 μg/mL fibronectin (Sigma-Aldrich, St. Louis, MO, USA). Cells were

grown at 37°C in a humidified 95% air, 5% CO2 atmosphere using proliferation medium

consisting of Advanced Dulbecco’s modified Eagle’s medium/F12, 1x N-2 supplement

(Invitrogen, Carlsbad, CA), 2 mM L-glutamine (Gibco, Rockville, MD, USA) and 40

ng/mL recombinant basic fibroblast growth factor (R&D Systems, Minneapolis, MN,

USA). Proliferating cells were enzymatically dissociated with trypsin and differentiated

using medium consisting of Advanced Dulbecco’s Modified Eagle’s Medium

(DMEM)/F12, N-2 supplement, 2 mM L-glutamine (all reagents were purchased from

Invitrogen, USA), 1 mM dibutyryl cAMP (Sigma-Aldrich), 1 μg/mL tetracycline (Sigma-

Aldrich) and 2 ng/mL recombinant human GDNF (R&D Systems). We used LUHMES

cells in all experiments at 5-6 d post-differentiation. First, LUHMES cells were pretreated

with 10 μM MSDC-0160 for 1 h followed by 10 μM 1-methyl-4-phenyl-

tetrahydropteridine (MPP+, Sigma-Aldrich) treatment for 24 h. Cells were collected and

processed for immunocytochemistry. For the time course experiments, LUHMES cells
were treated with 10 μM MSDC-0160 for 2, 6, or 24 h. Cells were collected at these time

points and processed for western blotting.

Primary mesencephalic neuronal cultures were prepared from the ventral mesencephalon

of gestational 12- to 13-day-old mouse embryos as described previously (44). Briefly,

mesencephalic tissues were dissected and maintained in ice-cold calcium-free Hanks

balanced salt solution (HBSS) and then dissociated in HBSS containing 0.25% trypsin-

EDTA for 20 min at 37°C. Cultures were maintained in Neurobasal medium with 50x B-

27 supplement, 500 mM L-glutamine, 100 IU/ml penicillin, and 100 μg/mL streptomycin

(all reagents were purchased from Invitrogen). The cells were maintained in a humidified

CO2 incubator (5% CO2 and 37°C) for 24 h. Five days post-culture, primary

mesencephalic neurons were pre-treated with 10 μM MSDC-0160 for 1 h followed by 10

μM MPP+ treatment for 24 h. Cells were collected and processed for

immunocytochemistry.

BV2 cells were purchased from ICLC (Catalog number ATL03001) (61). The media in

which BV2 cells proliferate contained RPMI 1640 supplemented with 10% heat-

inactivated FBS, 2 mM L-glutamine, 50 IU/mL penicillin, and 50 μg/ml streptomycin.

BV2 cells were plated at a density of 200 cells per mm2. After one day, FBS was

removed from the media and the cells were pretreated with 10 μM MSDC-0160. One

hour later, they were treated with 1 μg/mL lipopolysaccharide (LPS) (Escherichia coli

0111:B4, endotoxin content 6.6000000 EU/mg, Sigma-Aldrich). After 24 h of LPS

treatment, supernatant was collected for cytokine analysis or Griess assay and cells were

processed for immunocytochemistry. For the timecourse experiments, BV2 cells were
treated simultaneously with 10 μM MSDC-0160 and 1 μg/mL LPS for 2, 6, or 24 h. Cells

were collected at these time points and processed for western blotting.

For primary microglia isolation, we followed the procedure as described by (62). Briefly,

cortical tissues were dissected out from postnatal-day-3- to day-5-old mouse pups. Tissue

was dissociated using Miltenyi Biotec Neural Tissue Dissociation Kit (P, Miltenyi

Biotec, Bergisch-Gladbach, Germany). The single-cell suspension was incubated in

CD11b microbeads and separated using the magnetic column OctoMACS

Separator. Cells were then plated in 24-well dishes for treatments. Media for primary

microglia contains DMEM/F12 supplemented with 10% heat-inactivated FBS, 50 IU/ml

penicillin and 50 μg/ml streptomycin. Primary microglial cells were plated at a density of

200 cells per mm2. After one day, FBS was removed from the media and the cells were

pretreated with 10 μM MSDC-0160. One hour later, they were treated with 1 μg/mL

LPS. After 24 h of LPS treatment, supernatant was collected for cytokine analysis or

Griess assay and cells were fixed for immunocytochemistry.

C. elegans α-syn-induced neurodegeneration assay and RNAi functional analysis

Worm strains BY250 (GFP expression in dopaminergic neurons) and JVR203 (GFP + α-

syn (A53T) expression in dopaminergic neurons) were synchronized to L4 stage and

grown in liquid culture (6 mg/mL OP50, 120 μM FUDR) treated or untreated with 100

μM MSDC-0160 for 12 d. After this time, worms were scored for neurodegeneration as

described above. RNAi experiments were carried following an identical paradigm using

strains JVR325 and JVR326, in which RNAi is only effective in neuronal cells. Bacterial
RNAi clones were obtained from the Ahringer RNAi library (Source Bioscience). NGM

plates were supplemented with 100 µg/mL carbenicillin and 1 mM IPTG. Bacterial

clones were grown overnight in LB with 100 µg/mL carbenicillin. The following day,

IPTG was added to a final concentration of 1 mM and the bacteria were allowed to grow

(shaking at 37 °C) for another hour. 50 µL of culture was seeded onto each plate. Worms

were synchronized and 100 L1s were added to each NGM RNAi plate and incubated at

20 °C in the dark for 48 h. L4s/early adults were washed from the plates and 50 worms

per well were incubated at 20 °C in M9 with RNAi bacteria (6 mg/mL), FUDR (120 uM),

IPTG (1 mM) and either MSDC-0160 or vehicle (methoxy cellulose). Freshly grown

RNAi bacteria in LB (with 0.1 mM IPTG and 100 μg/mL carbenicillin) was added to the

wells every 3 d. Worms were removed from liquid culture and scored for

neurodegeneration as above after 10 d.

Knockdown of mRNA levels by RNAi was determined using single-worm RT-PCR.

cDNA was made from individual worms using the CellsDirectTM One-Step qRT-PCR

Kit (Invitrogen). qPCR was performed with Power SYBR green from Applied

Biosystems and cycled on an Applied Biosystems Step One Plus system. All reactions

were repeated multiple times: three technical replicates for each of two biological

replicates were performed and all data were normalized to act-1.

Immunocytochemistry

The LUHMES cells, primary mesencephalic neurons, BV2 cells, or primary microglia

cells were fixed with 4% paraformaldehyde in 1X PBS for 20 min and processed for

immunocytochemistry. First, nonspecific sites were blocked with 0.2% bovine serum
albumin, 0.5% Triton X-100, and 0.05% Tween 20 in PBS for 1 h at room temperature.

Cells were then incubated with different primary antibodies: TH (1:1600, EMD

Millipore, Billerica, MA, USA); Tuj-1 (1:1000, Abcam, Cambridge, MA, USA); iNOS

(1:300, Santacruz, Dallas, TX, USA); Iba-1 (1:800, WAKO, Richmond, VA, USA); p-

mTOR (1:300, Cell Signaling, Danvers, MA, USA); p-S6 (1:1000, Cell Signaling,

Danvers, MA, USA) at 4°C overnight. Appropriate secondary antibodies (Alexa Fluor

488 or 594, Invitrogen, Carlsbad, CA, USA) were used followed by incubation with

DAPI to stain the nucleus. The coverslip-containing stained cells were washed twice with

PBS and mounted on slides. Cells were viewed under a NIKON Eclipse Ni-U

fluorescence microscope (Nikon, Melville, NY, USA); images were captured with a

Retiga Exi digital camera using NIS Elements AR 4.00.08 software (Nikon).

Cytokine analysis

Supernatant from treated BV2 cells was collected and stored at -80°C until

analysis. Analysis was performed using the V-PLEX Proinflammatory Panel 1 (mouse)

kit from Meso Scale Discovery (Rockville, MA, USA). Calibration solutions and buffers

were prepared as described by the kit. Supernatants were thawed and 25 μL of sample

was diluted 1:1 with 25 μL Diluent 41 from the kit. Sample was loaded onto the reading

plate and sealed before incubation at RT for 2 h with shaking. After 3 washes with 150

μL wash buffer, 25 μL of detection antibody solution was added to each well. Again, the

plate was sealed and incubated at RT for 2 h with shaking. An additional 3 washes were

performed followed by the addition of 150 μL 2X Read Buffer T. The plate was then read

on the MesoScale QuickPlex SQ120.


Griess assay

Standards in triplicate were used for each plate. Standard mix was made up of 1 mL

media and 1 μL sodium nitrite and added to wells in volumes increasing by 5 μL from 0-

35 μL. Media was added to each well to bring volume up to a total of 100 μL. Sample

supernatant was added in triplicate to remaining wells in 100 μL amounts. Then, 100 μL

of Griess reagent (Sigma, USA) was added to each well and incubated at RT for 10

minutes on a rotator. Absorbance at 540 nm was detected in a Synergy NEO plate reader

(BioTek, Winooski, VT, USA).

Immunohistochemistry

Mice were anesthetized with sodium pentobarbital (130 mg/kg; Sigma) and perfused

transcardially with 4% paraformaldehyde in 0.1 M phosphate buffer. Brains were

removed, postfixed, and cryoprotected. Immunohistochemistry was performed on free-

floating cryomicrotome-cut sections (40 µm in thickness) encompassing substantia nigra

as described previously (51, 53). For tyrosine hydroxylase (TH)-immunohistochemistry

(IHC), free-floating sections were washed with 0.1 M PBS (pH 7.4), incubated with 3%

H2O2 for 10 min to quench endogenous peroxidase activity, and were blocked for 1 h

with 10% normal goat serum and 0.25% Triton X-100. Next, the sections were incubated

overnight at 4 °C with different antibodies: TH (1:1600, EMD Millipore); Iba-1 (1:800,

WAKO); GFAP (1:1000, EMD Millipore). Then, the sections were incubated with

appropriate biotinylated secondary antibodies (goat anti-rabbit, goat anti-mouse, 1:500,

Vector Laboratories, Burlingame, CA, USA) followed by incubation with Avidin/Biotin


ABC reagent (Vector Laboratories). Immunolabeling was revealed using

diaminobenzidine (DAB) which yielded a brown-colored stain visible in bright-field light

microscopy. For double-labeling immunofluorescence, the sections were blocked with

0.2% bovine serum albumin, 0.5% Triton X-100, and 0.05% Tween 20 in PBS for 1 h at

room temperature. Next, sections were incubated with different antibodies: TH (1:1500,

EMD Millipore); p-mTOR (1:200, Cell Signaling); p-S6 (1:800, Cell Signaling); LC3B

(1:500, Cell Signaling), phosphorylated alpha-synuclein serine 129 (1:10000, Abcam).

Appropriate secondary antibodies (Alexa Fluor 488 or 594, Invitrogen) were used

followed by incubation with DAPI to stain the nucleus. The sections were viewed under a

Nikon Eclipse Ni-U fluorescence microscope (Nikon); images were captured with a

Retiga Exi digital camera using NIS Elements AR 4.00.08 software (Nikon).

Stereology and striatal optical density

For the purpose of cell counting, we stained sections with 0.1% cresyl violet. Stained

cells were counted using unbiased sampling and blinded stereology. Following TH-DAB

immunostaining, the TH-ir neurons were counted using unbiased sampling and blinded

stereology as mentioned previously (51, 53). Briefly, the substantia nigra region was

outlined based on the Allen mouse brain atlas (Allen Institute, Seattle, WA, USA) using a

2x objective. Quantitative estimation of TH-ir neurons was performed in every 6th

section of the substantia nigra using the optical fractionator probe in Stereo Investigator

software (MBF Bioscience, Williston, VT, USA). The parameters used include a 60× oil

objective, a counting frame size of 100 × 100, a sampling site of 140 × 140, a dissector

height of 18 μm, 2 μm guard zones, and the Schaffer second estimated co-efficient error

was less than 0.1. In addition, the Gunderson estimated coefficient error was less than
0.1. A total of 6 to 9 animals per group were used and 5 to 7 sections per animal were

counted. Quantification of striatal TH fibers was performed using NIS Elements AR

4.00.08 software (Nikon) in a blinded manner. Briefly, striatal pictures were first taken at

2x magnification, then thresholded. After that, the optical density of fibers was counted

using the NIS Elements AR 4.00.08 software (Nikon).

Western blotting

Cells or brain tissues were collected and resuspended in modified

radioimmunoprecipitation (RIPA) assay buffer containing protease and phosphatase

inhibitor mixture. Cell suspensions were sonicated after resuspension, whereas mouse

brain tissues were homogenized, sonicated, and then centrifuged at 14,000 x g for 45 min

at 4°C. Protein concentrations were estimated using a BCA kit (Thermo Scientific,

Waltham, MA USA). Lysates were separated on 7.5%–15% SDS-polyacrylamide

electrophoresis gels (Bio-Rad). After the separation, proteins were transferred to a

nitrocellulose membrane, and nonspecific binding sites were blocked by treating with

either Odyssey blocking buffer (LI-COR, Lincoln, NE, USA) or TBS with 5% bovine

serum albumin (BSA) followed by antibody incubation: TH (1:1500, EMD Millipore);

phospho (Ser2448)-mTOR (1:1000, Cell Signaling); phospho (Ser2481)-mTOR (1:1000,

Cell Signaling); mTOR (1:1000, Cell Signaling); phospho (Thr389)-p70S6 kinase

(1:1000, Cell Signaling); p70S6 kinase (1:1000, Cell Signaling); phospho (Thr308) AKT

(1:1000, Cell Signaling); AKT (1:1000, Cell Signaling); phospho (Ser235/236)-S6

(1:1500, Cell signaling); S6 (1:1500, Cell signaling); LC3B (1:500, Cell signaling);

REDD1 (1:2000, Bethyl Laboratories, Montgomery, TX, USA); p62 (1:500, Abcam);
iNOS (1:300, Santa Cruz); GFAP (1:1000, EMD Millipore); Iba-1 (1:500, WAKO);

NFκB p65 (1:300, Santa Cruz); and β-actin (1:10,000, Sigma). Secondary IR-680-

conjugated anti-mouse (1:10,000, donkey anti-mouse; Molecular Probes, USA), IR-680-

conjugated anti-goat (1:10,000, donkey anti-goat, Molecular Probes, USA), IRDye 800

donkey anti-rabbit (1:10,000, Rockland, Pottstown, PA, USA) were used. New blot

stripping buffer (LI-COR, USA) was used to reprobe the membrane. Western blot images

were captured with a LI-COR Odyssey machine (LI-COR, USA). The western blot bands

were quantified using ImageJ software (National Institutes of Health, USA). An antibody

that detects β-actin was used to visualize loading controls.


Supplementary Fig. 1

Supplementary Fig. 1. MSDC-0160 protects dopaminergic neurons against MPP+-

induced toxicity in LUHMES cells and C. elegans. Six days post-differentiation,

LUHMES cells were pre-treated with 10 μM MSDC-0160 for 1 h followed by 10 μM


MPP+ treatment for 24 h. (A) TH-diaminobenzidine immunocytochemistry in LUHMES

cells. Blue arrows show degenerating TH-ir dopaminergic neurons in the MPP+ treatment

group. C. elegans were synchronized and L1 larvae were placed in 96-well plates

containing OP50 bacteria in liquid culture (6 mg/mL) and MPP+ (0.75 mM) with MSDC-

0160 (1 μM, 10 μM and 100 μM) for 48 h. (B) GFP expressing dopaminergic neurons in

C. elegans. Numbering on the images (1-6) denotes individual dopaminergic neurons.

White asterisk denotes degenerating dopaminergic neurons. (C) Relative mRNA

expression of different genes in C. elegans. Data are means ± SEM (n=3 experiments).

Scale bars, 25 μm.


Supplementary Fig.
F 2

Supp
plementary Fig. 2. Mod
dulation of MPC
M improoves mitochoondrial funcction. Mice

were orally dosed


d with 30 mg
g/kg MSDC--0160 and thhen euthanizzed at 2 or 4 hours after

dosin
ng. The plasm
ma (black do
otted lines; ng/mL)
n and bbrain homoggenates (blacck continuouus

lines;; ng/g) were extracted an


nd analyzed by LC/MS ffor the conceentrations off both parentt

drug (A) and actiive alcohol metabolite


m (B
B). Data forr are means ± SEM (n=3
experiments). Effect of increasing concentrations of MSDC-0160 (one experiment run in

duplicate) on rat brain mitochondria incubated with MitoXpress Xtra reagent (described

in the Materials and Methods section). The concentrations of MSDC-0160 shown were

added at the start of the reaction which contained 5 mM pyruvate. (C) The loss of oxygen

from the medium was measured over time as the increase in relative fluorescence (RFU).

Mean RFU is plotted at each time point. LOESS curves with 95% confidence ribbons are

plotted over top to show smoothed behavior over time. (D) Calculated changes in the

initial rate of oxygen consumption and SEM at the given concentration of MSDC-0160.

Data for are means ± SEM (n=3 experiments). (E) Measurement of oxygen consumption

in differentiated LUHMES cells. Cells were either pre-treated with 10 μM MPP+ or

MSDC-0160 and oxygen consumption was measured basally and for 30 minutes after

injection of vehicle, MSDC-0160, or MPP+. Mean oxygen consumption rate is plotted

for each experimental group at each time point (n=3 experiments). LOESS curves with

95% confidence ribbons are plotted over top to show smoothed behavior over time.
Supplementary Fig. 3

Supplementary Fig. 3. Time-course treatment of MSDC-0160 in LUHMES cells. Six

days post-differentiation, LUHMES cells were treated with 10 μM MSDC-0160 for 2, 6

or 24 h. (A) Representative western blot of phospho-mTOR (Ser 2448), mTOR and β-

actin in LUHMES cells. (B) Mean western blot phospho-mTOR/mTOR ratios relative to

β-actin. Data are means ± SEM (n=3 experiments). Data were analyzed by one-way

ANOVA with Tukey’s multiple comparison test.


Supplementary Fig. 4

Supplementary Fig. 4. MSDC-0160 improves behavioral parameters in the MPTP

mouse model. (A) Treatment schedule of MPTP-injected mice with MSDC-0160 (Pre-

treatment of MSDC-0160). (B) Traced movement of mice using ANY-maze software.


(C) Stereological counting of cresyl violet (CV)-stained neurons in SN in the MSDC-

0160 pre-treatment schedule (D) Measurement of striatal MPP+ (ppb). (E) Treatment

schedule of MPTP-injected mice in which mice were administered MSDC-0160 (30

mg/kg/day) by oral gavage two days post MPTP treatment (25 mg/kg/day) and continued

the treatment for another 7 days (Post-treatment of MSDC-0160). Four days after the last

MPTP injection, mice were tested for spontaneous open field behavioral activities. (F)

Stereological counting of CV-stained neurons in SN in the MSDC-0160 post-treatment

schedule (G) Distance traveled (cm), time mobile (sec), time immobile (sec) and time

freeze (sec). Data are means ± SEM of seven to ten mice per group. ****p < 0.0001,

***p < 0.001, **p < 0.01, *p < 0.05. Data were analyzed by one-way ANOVA with

Tukey’s multiple comparison test except (D) which was analyzed by unpaired student’s t-

test.
Supplementary Figure
F 5

Supp
plementary Fig. 5. MS
SDC-0160 improves
i b
behavioral activities in
n the En1++/−

mice. At the milld pathology 1+/- mice weere fed a dieet of chow fformulated tto
y stage, En1

deliver MSDC-0160 (30 mg//kg) starting at week 3 an


and were testted for motoor functions aat
28 weeks and 44 weeks. (A) Time freeze (sec), (B) Time mobile (sec), (C) Time

immobile (sec), (D) Head distance traveled (cm). At the modest pathology stage, En1+/-

mice were fed a diet mixed with MSDC-0160 (30 mg/kg) starting at week 8 weeks and

were tested for motor functions at 16 weeks and 28 weeks. (E) Time freeze (sec), (F)

Time mobile (sec), (G) Time immobile (sec), (H) Head distance traveled (cm). Data are

means ± SEM of eight to twelve mice per group. ***p < 0.001, **p < 0.01, *p < 0.05,
o
p < 0.05-0.1. Data were analyzed using linear mixed-effects regression analysis false

discovery rate corrected p-value.


Supplementary Fig. 6

Supplementary Fig. 6. MSDC-0160 prevents dopaminergic neurodegeneration in the

En1+/− mice. (A) Stereological counting of cresyl violet (CV)-stained neurons in SN from
3W-28W group. (B) Stereological counting of CV-stained neurons in SN from 8W-16W

group. (C) Stereological counting of CV-stained neurons in SN from 8W-28W group. (D)

Representative western blot illustrating the expression of TH in striatum from 16W (top)

and 28W (bottom). (E) Bar graph showing mean western blot TH/β-actin ratios in ST at

16W. (F) Bar graph showing mean western blot TH/β-actin ratios in ST at 28W. Data are

means ± SEM of six to eight mice per group. ****p < 0.0001, ***p < 0.001, **p < 0.01,

*p < 0.05. Data were analyzed by two-way ANOVA with Bonferroni’s multiple

comparison test.
Supplementary Fig. 7

Supplementary Fig. 7. MSDC-0160 down-regulates mTOR signaling pathway and

restores autophagy in the En1+/− mice. En1+/- mice were fed a diet formulated to

deliver MSDC-0160 (30 mg/kg) starting at week 8 (8W) and were euthanized at week 16

(16W). Substantia nigra (SN) tissues were processed for western blotting. (A)

Representative western blot for MPC-1 and MPC-2 in SN. (B) Immunohistochemistry of

TH and MPC-1 in SN after treatment with MSDC-0160 or control in WT and En1+/-

mice. (C) Representative western blot illustrating the expression of phospho-mTOR (Ser

2481), mTOR, phospho-AKT (Thr 308), AKT, REDD1 and p62 in SN after treatment

with MSDC-0160 or control in WT and En1+/- mice. (D) Bar graph showing mean

western blot phospho-mTOR (Ser 2481)/mTOR ratios in SN relative to β-actin. (E) Bar

graph showing mean western blot phospho-AKT/AKT ratios in SN relative to β-actin. (F)
Bar graph showing mean western blot REDD1/β-actin ratios in SN at 16W. (G) Bar

graph showing mean western blot p62/β-actin ratios in SN at 16W. Scale bars, 5 μm. Data

are means ± SEM of six mice per group. ****p < 0.0001, ***p < 0.001, **p < 0.01, *p

< 0.05. Data were analyzed by two-way ANOVA with Bonferroni’s multiple comparison

test.
Supplementary Fig. 8
Supplementary Fig. 8. MSDC-0160 attenuates inflammation in BV2 cells and in vivo

animal models. (A) Representative western blot illustrating the expression of MPC-1 and

MPC-2 in BV2 cells. (B) Immunocytochemistry of Iba-1 and MPC-1/MPC-2 in BV2

cells. (C) Western blotting of iNOS in BV2 cells. In the MPTP model, MPTP and

MSDC-0160 administration was accomplished as described in Materials and Methods.

(D) Representative western blot illustrating the expression of GFAP in SN. Bar graph

showing mean western blot GFAP/β-actin ratios in SN. (E) DAB immunostaining of

GFAP in SN. In the En1+/- model, MSDC-0160 was administered as described in

Materials and Methods. (F) Representative western blot illustrating the expression of

GFAP in SN. Bar graph showing mean western blot GFAP/β-actin ratios in SN. (G)

DAB immunostaining of GFAP in SN. Scale bars, 5 μm. (A-C) Data are means ± SEM

(n=3 experiments). (D-G) Data are means ± SEM of six to eight mice per group. ****p <

0.0001, ***p < 0.001, **p < 0.01, *p < 0.05. Data were analyzed by two-way ANOVA

with Bonferroni’s multiple comparison test.


Supplementary Fig.
F 9

Supp
plementary Fig. 9. MSD
DC-0160 improves mitoochondrial ffunction and blocks

mTO
OR activatio
on in BV2 ceells. (A) Measurement oof oxygen consumption
n in BV2

cells.. Cells were either pre-trreated with 10


1 μM LPS oor MSDC-01160 and oxyggen

consu
umption wass measured basally
b and for
f 30 minut es after injecction of vehiicle, MSDC--

0160, or LPS. Mean


M oxygen consumptio
on rate is plootted for eachh experimenntal group at

each time point (n


n=3/injected
d group). LO
OESS curvess with 95% cconfidence rribbons are

plotteed over top to


t show smo
oothed behav me.  (B) Timee course treeatment of
vior over tim

MSD
DC-0160 in BV2
B cells. BV2 M LPS and 10 μM
B cells weere co-treateed with 10 μM
MSDC-0160 for 2, 6 or 24 h. Representative western blot of phospho-mTOR (Ser 2448),

mTOR and β-actin in BV2 cells. (C) Mean western blot phospho-mTOR/mTOR ratios

relative to β-actin. Data are means ± SEM (n=3 experiments). *p < 0.05. Data were

analyzed by one-way ANOVA with Tukey’s multiple comparison test.


Supplementary Fig. 10
Supplementary Fig. 10. Schematic diagram of proposed mechanism of action of

MSDC-0160 in glial cells and dopaminergic neurons. MSDC-0160 has a dual effect on

MPC in microglia and dopaminergic neurons resulting in pro-survival signaling.


Tables:

Table 1. LMM results.


Outcome FDR
Estimated Unadjusted 95% CI 95% CI
Measure: Variable Adjusted
Effect P-Values Lower Upper
Trial Start P-Values
Open field
Distance
Wildtype 459.237 < 0.001 < 0.001 381.290 537.184
Traveled: 3
Weeks
EN1 Genotype -199.067 < 0.001 < 0.001 -288.721 -109.413
MSDC Treatment -48.301 0.334 0.431 -145.066 48.464
Weeks -2.942 0.022 0.044 -5.323 -0.561
EN1 Genotype *
145.926 0.033 0.061 16.401 275.451
MSDC Treatment
Open field
Distance
Wildtype 378.459 < 0.001 < 0.001 338.598 418.320
Traveled: 8
Weeks
EN1 Genotype -167.07 < 0.001 < 0.001 -220.541 -113.599
MSDC Treatment -3.501 0.907 0.937 -61.827 54.825
Weeks -0.927 0.488 0.585 -3.518 1.664
EN1 Genotype *
109.261 0.009 0.021 29.822 188.700
MSDC Treatment
Open field
Mean Speed: Wildtype 0.775 < 0.001 < 0.001 0.652 0.898
3 Weeks
EN1 Genotype -0.359 < 0.001 < 0.001 -0.496 -0.222
MSDC Treatment -0.07 0.362 0.461 -0.219 0.079
Weeks -0.006 0.006 0.015 -0.010 -0.002
EN1 Genotype *
0.257 0.015 0.032 0.059 0.455
MSDC Treatment
Open field
Mean Speed: Wildtype 0.593 < 0.001 < 0.001 0.526 0.660
8 Weeks
EN1 Genotype -0.283 < 0.001 < 0.001 -0.373 -0.193
MSDC Treatment -0.002 0.973 0.983 -0.100 0.096
Weeks -0.001 0.512 0.595 -0.005 0.003
EN1 Genotype *
0.174 0.012 0.027 0.041 0.307
MSDC Treatment
Open field
Time
Wildtype 554.888 < 0.001 < 0.001 519.488 590.288
Mobile: 3
Weeks
EN1 Genotype -69.87 < 0.001 < 0.001 -103.813 -35.927
MSDC Treatment 12.922 0.49 0.585 -23.434 49.278
Weeks -1.954 0.011 0.025 -3.381 -0.527
EN1 Genotype *
43.958 0.086 0.129 -5.036 92.952
MSDC Treatment
Open field
Time
Wildtype 528.576 < 0.001 < 0.001 502.298 554.854
Mobile: 8
Weeks
EN1 Genotype -102.43 < 0.001 < 0.001 -136.961 -67.899
MSDC Treatment -1.137 0.953 0.974 -38.798 36.524
Weeks -1.234 0.19 0.256 -3.059 0.591
EN1 Genotype *
87.686 0.001 0.003 36.457 138.915
MSDC Treatment
Open field
Max Speed: Wildtype 5.35 < 0.001 < 0.001 4.397 6.303
3 Weeks
EN1 Genotype -1.743 0.005 0.013 -2.907 -0.579
MSDC Treatment -0.63 0.332 0.431 -1.888 0.628
Weeks -0.043 0.002 0.005 -0.067 -0.019
EN1 Genotype *
1.92 0.031 0.059 0.238 3.602
MSDC Treatment
Open field
Max Speed: Wildtype 3.82 < 0.001 < 0.001 3.355 4.285
8 Weeks
EN1 Genotype -0.973 0.002 0.005 -1.559 -0.387
MSDC Treatment 0.075 0.819 0.886 -0.562 0.712
Weeks -0.027 0.131 0.19 -0.062 0.008
EN1 Genotype *
0.933 0.037 0.067 0.071 1.795
MSDC Treatment
Open field
Time Frozen: Wildtype 466.815 < 0.001 < 0.001 387.288 546.342
3 Weeks
EN1 Genotype 36.526 0.506 0.595 -70.341 143.393
MSDC Treatment -121.337 0.043 0.075 -235.809 -6.865
Weeks -2.781 0.158 0.219 -6.525 0.963
EN1 Genotype *
86.573 0.277 0.368 -67.779 240.925
MSDC Treatment
EN1 Genotype *
3.92 0.144 0.206 -1.170 9.010
Weeks
MSDC Treatment
5.199 0.067 0.104 -0.111 10.509
* Weeks
EN1 Genotype *
MSDC -8.047 0.041 0.073 -15.383 -0.711
Treatment:Weeks
Open field
Wildtype 401.237 < 0.001 < 0.001 369.667 432.807
Time Frozen:
8 Weeks
EN1 Genotype 90.281 < 0.001 < 0.001 49.770 130.792
MSDC Treatment 4.857 0.83 0.887 -39.294 49.008
Weeks -0.37 0.755 0.836 -2.691 1.951
EN1 Genotype *
-58.434 0.06 0.096 -118.334 1.466
MSDC Treatment
Rotarod
Latency: 3 Wildtype 151.333 < 0.001 < 0.001 112.976 189.690
Weeks
EN1 Genotype -107.618 < 0.001 < 0.001 -144.092 -71.144
MSDC Treatment -4.371 0.817 0.886 -41.148 32.406
Weeks -1.791 0.049 0.083 -3.524 -0.058
EN1 Genotype *
57.653 0.031 0.059 7.034 108.272
MSDC Treatment
Rotarod
Latency: 8 Wildtype 175.049 < 0.001 < 0.001 139.067 211.031
Weeks
EN1 Genotype -139.038 < 0.001 < 0.001 -185.367 -92.709
MSDC Treatment 5.161 0.839 0.887 -44.523 54.845
Weeks -2.127 0.146 0.206 -4.971 0.717
EN1 Genotype *
66.548 0.056 0.091 -0.702 133.798
MSDC Treatment

View publication stats

You might also like