Accepted Manuscript Withdrawal from cocaine self-administration alters the regulation of protein translation in the nucleus accumbens Michael T. Stefanik, Mike Milovanovic, Craig T. Werner, John C.G. Spainhour, Marina E. Wolf PII:
S0006-3223(18)30113-6
DOI:
10.1016/j.biopsych.2018.02.012
Reference:
BPS 13472
To appear in:
Biological Psychiatry
Received Date: 9 October 2017 Revised Date:
26 January 2018
Accepted Date: 12 February 2018
Please cite this article as: Stefanik M.T., Milovanovic M., Werner C.T., Spainhour J.C.G. & Wolf M.E., Withdrawal from cocaine self-administration alters the regulation of protein translation in the nucleus accumbens, Biological Psychiatry (2018), doi: 10.1016/j.biopsych.2018.02.012. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT
Withdrawal from cocaine self-administration alters the regulation of protein translation in the nucleus accumbens
RI PT
Michael T. Stefanik1, Mike Milovanovic1, Craig T. Werner1, John C.G. Spainhour2, Marina E. Wolf1 1
Department of Neuroscience, The Chicago Medical School at Rosalind Franklin School of Medicine and Science, North Chicago, IL, 60064
2
SC
Wallace H. Coulter Department of Biomedical Engineering, Georgia Tech College of Engineering and Emory School of Medicine, Atlanta, Georgia, 30332
Short title: Cocaine withdrawal alters regulation of translation in the NAc
M AN U
Key Words: AMPA receptor; incubation of cocaine craving; metabolic labeling; nucleus accumbens; protein translation; puromycin
TE D
Corresponding Author: Marina E. Wolf Department of Neuroscience Rosalind Franklin University of Medicine and Science 3333 Green Bay Road North Chicago, IL 60064-3095 USA
[email protected]
EP
Number of words in Abstract: 250
Number of words in body (Introduction, Methods, Results, Discussion): 3999
AC C
Number of Figures: 6 Number of Tables: 0
Supplemental Information: Supplemental Methods, Supplementary Figure 1-3, Supplemental Table 1 and References for Supplemental Methods
1
ACCEPTED MANUSCRIPT
ABSTRACT Background: Cue-induced cocaine craving “incubates” during abstinence from cocaine selfadministration. Expression of incubation ultimately depends upon elevation of homomeric GluA1
RI PT
AMPA receptors (AMPARs) in the nucleus accumbens (NAc). This adaptation requires ongoing protein translation for its maintenance. Aberrant translation is implicated in CNS diseases, but nothing is known about glutamatergic regulation of translation in the drug-naïve NAc or after
SC
incubation.
M AN U
Methods: NAc tissue was obtained from drug-naïve rats and after 1 or >40 days of abstinence from extended-access cocaine or saline self-administration. Newly translated proteins were labeled using
35
S-Met/Cys or puromycin. We compared basal overall translation and its
regulation by mGlu1, mGlu5 and NMDA receptors (NMDARs) in naïve, saline and cocaine rats,
TE D
and compared GluA1 and GluA2 translation by immunoprecipitating puromycin-labeled proteins.
Results: In all groups, overall translation was unaltered by mGlu1 blockade (LY367385) but increased by mGlu5 blockade (MTEP). NMDAR blockade (APV) increased overall translation in
EP
naïve and saline, but not cocaine, rats. Cocaine/late withdrawal rats exhibited greater translation
AC C
of GluA1 (but not GluA2), which was not further affected by NMDAR blockade.
Conclusions: Our results suggest that increased GluA1 translation contributes to the elevated homomeric GluA1 AMPAR levels in NAc that mediate incubation. Additional contributions to incubation-related plasticity may result from loss of the braking influence on translation normally exerted by NMDARs. Apart from elucidating incubation-related adaptations, we found a suppressive effect of mGlu5 on NAc translation regardless of drug exposure, which is opposite to results obtained in hippocampus and points to heterogeneity of translational regulation between brain regions. 2
ACCEPTED MANUSCRIPT
INTRODUCTION A major challenge in treating addiction is the persistence of vulnerability to relapse. Even after prolonged abstinence, cocaine addicts remain susceptible to relapse, often elicited by cues
RI PT
previously associated with drug-taking (1-3). Aspects of this can be studied in rats using the “incubation of craving” model, in which cue-induced craving progressively intensifies (“incubates”) during withdrawal from cocaine self-administration and then remains high for
SC
months (4, 5). This models the situation of humans users when heavy cocaine use is interrupted by incarceration or prolonged hospitalization (6). During this interruption, craving may build and,
M AN U
when the addict returns to the “real world”, drug-related cues may present an elevated relapse risk. Incubation of craving in human drug users has now been observed during abstinence from cocaine and other drugs of abuse (3, 7-9).
In rats, incubation of cocaine craving involves time-dependent changes in multiple brain regions
TE D
(5, 10). However, after >1 month of withdrawal, “incubated” craving depends upon strengthening of AMPA receptor (AMPAR) transmission onto medium spiny neurons (MSNs) in the NAc through accumulation of high-conductance GluA2-lacking, Ca2+-permeable AMPA
EP
receptors (CP-AMPARs; primarily homomeric GluA1) (5). This is accompanied by profound alterations in group I mGluR-mediated synaptic depression (11, 12). Recently, we identified a
AC C
novel linkage between these adaptations and protein translation. NAc slices from “incubated” rats were exposed for ~1 h to inhibitors of translation (anisomycin, cycloheximide or rapamycin) prior to patch-clamp recordings. This normalized the state of synaptic transmission, i.e., the contribution of CP-AMPARs to synaptic transmission returned to control levels and alterations in group I mGluR-mediated synaptic depression were reversed (13). This suggests that ongoing protein translation maintains synaptic adaptations in MSNs directly linked to incubation of craving. Consistent with these findings, our recent results demonstrate that expression of
3
ACCEPTED MANUSCRIPT
incubated cocaine craving during a cue-induced seeking test requires protein translation in the NAc (14).
RI PT
In many brain regions, including hippocampus, cerebellum and ventral tegmental area (VTA), group I mGluRs are well-established regulators of protein translation, with stimulation of translation by these receptors playing a critical role in long-term depression (LTD) (15). In
SC
addition, NMDA receptor (NMDAR) signaling negatively regulates translation in hippocampal and cortical neurons (16-20). Through such studies, aberrant protein translation has been
M AN U
implicated in CNS diseases, including fragile X syndrome and depression, and identified as a potential target for therapeutic intervention (19-25).
Previous studies have demonstrated a role for protein synthesis in the effects of drugs of abuse (e.g. (26-34)) and shown that drugs affect signaling molecules regulating translation (e.g.,
TE D
mTOR; (35, 36)), but nothing is known about upstream regulation of this signaling by glutamate transmission in the NAc. Based on work in other brain regions summarized above, we focused on determining how NMDAR and group I mGluRs regulate translation in the NAc under control
EP
conditions and whether this is altered after incubation of cocaine craving. As protein synthesismediated changes in synaptic connections are believed to underlie long-term alterations in
AC C
neural circuits involved in a variety of behaviors (37), expanding our understanding of the regulation of translation in the NAc should shed light on normal and aberrant learning involving the striatum. Furthermore, if dysregulation of protein translation underlies addiction-related plasticity, normalizing translation may represent a novel therapeutic approach.
4
ACCEPTED MANUSCRIPT
METHODS AND MATERIALS Subjects and drug self-administration All procedures were approved by the Rosalind Franklin University Institutional Animal Care and
RI PT
Use Committee in accordance with the USPHS Guide for Care and Use of Laboratory Animals. Adult male Sprague-Dawley rats underwent extended-access cocaine or saline self-
SC
administration (6 h/d) as described previously (12) and in Supplemental Information.
Metabolic labeling
M AN U
Procedures were adapted from well-established protocols (22, 24) (see Supplementary Fig. 1 for time-line). Rats were decapitated and the NAc (mainly core) was punched from a 2mm brain slice obtained using a brain matrix. Each hemisphere was chopped into 4 pieces to increase surface area, placed into a net-well, and submerged in a manifold containing artificial cerebrospinal fluid (aCSF) saturated with 95% O2/5% CO2. Each hemisphere was counted as
TE D
an individual sample, with hemispheres from the same rat assigned to different experimental groups, enabling paired t-tests (see Statistical Analysis). Tissue recovered in net-wells for 3 h, enabling optimal and stable protein synthesis (Supplementary Figure 2). Actinomycin D (25 µM;
EP
Sigma-Aldrich; St. Louis, MO) was added during the last 30 min of recovery to inhibit new transcription. Then net-wells were transferred to manifolds with aCSF containing actinomycin D 35
S-Met/Cys (EasyTag™ L-[35S]-Methionine, PerkinElmer; Waltham, MA) ±
AC C
and 10 µCi/mL
drug, incubated for 1 h, and snap-frozen. To compare
35
S-Met/Cys incorporation between
experimental groups, labeled tissue was homogenized and processed for autoradiography as described in Supplemental Information.
Measurement of mRNA levels
5
ACCEPTED MANUSCRIPT
RNA was extracted from NAc homogenates or synaptoneurosomes (prepared as described previously (38)) and analyzed by quantitative PCR (qPCR) as described in Supplemental
RI PT
Information.
Puromycin labeling
Procedures were adapted from prior studies (39, 40). Pieces of NAc tissue were prepared
SC
exactly as described for metabolic labeling experiments and placed in net-wells. Tissue was incubated with 25 µM actinomycin D (Sigma-Aldrich) during the last 30 min of the 3-h recovery
M AN U
period, transferred to manifolds with aCSF containing 1 µM puromycin (Sigma-Aldrich) and actinomycin D for 1 h, and then snap-frozen. As described in Supplemental Information, overall puromycin incorporation was determined by immunoblotting; puromycin incorporation into GluA1 or GluA2 was determined by Immunoprecipitation of puromycin-labeled proteins.
TE D
Statistical analyses
Data were analyzed in Prism (GraphPad Software). Student’s t-tests [paired when comparing differently treated hemispheres from a single brain (Figures 1-3), or unpaired when each rat
EP
generated a single sample (Figures 4,5)] or ANOVA followed by Tukey’s test (Figure 6) were
AC C
conducted with significance set at p≤0.05. Data are presented as mean±SEM.
6
ACCEPTED MANUSCRIPT
RESULTS Measuring protein translation in the NAc To confirm that
35
S-Met/Cys incorporates into newly translated proteins in the NAc, we pre-
RI PT
exposed tissue from naïve animals to the translation inhibitor cycloheximide (60 µM) for 30 min, before moving the tissue into a chamber containing cycloheximide and
35
S-Met/Cys for an
additional 60 min. We then processed tissue for SDS-PAGE/autoradiography. Figures 1A and 1B show incorporation of 35S-Met/Cys into newly synthesized proteins across multiple molecular 35
SC
weights and demonstrate that cycloheximide significantly reduced
S-Met/Cys incorporation
M AN U
(Figure 1C, t3=5.46, p<0.05), validating the assay in NAc tissue.
Regulation of translation in the NAc under control conditions and after incubation We then tested the hypothesis that the overall rate of protein translation is altered in the NAc of rats after incubation of cocaine craving. Tissue was obtained after extended-access self-
TE D
administration of saline or cocaine (Figure 1D) and >40 days of abstinence, a period sufficient for maximal incubation of craving and stable CP-AMPAR elevation in NAc core (10, 42-44). Rats subjected to this cocaine regimen and withdrawal period will hereafter be referred to as 35
S-Met/Cys (60 min) and
EP
“incubated” rats. After 3 h of recovery, tissue was incubated with
processed for SDS-PAGE/autoradiography. Measurements of total protein translation, made by
AC C
scanning the entire lane, were not different between saline controls and “incubated” rats (Figure 1E, t5=0.30, p=0.77). This result indicates that altered translation of a subset of proteins, rather than gross changes in protein translation, must contribute to maintenance of incubation-related neuroadaptations in the NAc (13). Furthermore, it leaves open the possibility that mechanisms regulating protein translation are altered in the NAc after incubation of craving.
To test the latter possibility, we examined two glutamate receptor subtypes – NMDARs and group I mGluRs – that are well-established as regulators of protein translation in several brain 7
ACCEPTED MANUSCRIPT
regions (see Introduction). We utilized the strategy of incubating tissue with receptor antagonists to detect tonic regulation of translation by endogenous glutamate. This strategy has been used to study the regulation of protein synthesis in other brain regions (17, 22) and was preferred to
RI PT
incubating with agonists because NMDAR activation can be associated with toxicity, while effects of the group I mGluR agonist DHPG in our hands were quite variable depending on concentration and time of incubation (data not shown), as has been found previously (Emily
SC
Osterweil, personal communication).
M AN U
First, we examined mGlu5 by co-exposing NAc tissue (from drug-naïve, saline, or “incubated” rats) to vehicle or MTEP, a selective mGlu5 negative allosteric modulator (NAM) (45), and
35
S-
Met/Cys for 60 min. In contrast to previous studies in hippocampus, where mGlu5 activation increases protein translation (46), inhibition of mGlu5 with MTEP increased
35
S-Met/Cys
incorporation in both naïve and saline NAc tissue (Figure 2A,B, naïve: t7=2.20, saline: t7=2.12,
TE D
p<0.05). This is an important result because it shows that protein translation is differentially regulated in these two regions, and thus that regulatory mechanisms operating in the NAc cannot necessarily be inferred from prior studies in hippocampus. Furthermore, the magnitude
EP
of the MTEP-induced increase in translation was similar between saline and cocaine animals (Figure 2B,C, cocaine: t7=2.69, p<0.05). These results suggest that mGlu5 activity suppresses
AC C
protein translation in the NAc and this is not changed after incubation of cocaine craving.
To examine mGlu1, NAc tissue from naïve, saline, or “incubated” rats was labeled with
35
S-
Met/Cys in the presence or absence of LY367385 (50 µM), a selective mGlu1 competitive antagonist (47). LY367385 did not influence overall translation in any group (Figure 2D,E,F, naïve: t7=0.73, saline: t5=0.59, cocaine: t6=0.22). Thus, group I mGluR-mediated translational regulation in the NAc is primarily driven by mGlu5.
8
ACCEPTED MANUSCRIPT
We next asked whether NMDAR transmission regulates protein synthesis in the NAc and if this is altered following incubation of craving. In other brain regions, NMDAR transmission can exert a suppressive influence on translation by stimulating phosphorylation of eukaryotic elongation
with
RI PT
factor 2 (eEF2; see Discussion). NAc tissue from naïve, saline, or “incubated” rats was labeled 35
S-Met/Cys in the presence or absence of the competitive NMDAR antagonist APV (50
µM). APV increased protein synthesis in naïve and saline rats (Figure 3A,B, naïve: t7=2.26,
SC
saline: t6=1.90, p<0.05). This suggests that tonic NMDAR signaling suppresses translation in the NAc, consistent with findings in other regions. Interestingly, this NMDAR-mediated inhibitory
M AN U
tone was lost in the “incubated” rats (t6=1.70, p=0.14; Figure 3C).
GluA1 and GluA2 translation under control conditions and after incubation As described in the Introduction, homomeric GluA1 CP-AMPARs accumulate in NAc core after ~30 days of withdrawal from our cocaine regimen and thereafter are required for expression of
TE D
“incubated” craving (5, 12, 42). Their accumulation is associated with increased GluA1 protein levels in NAc homogenates, LP1 fractions and on the cell surface (42, 48), suggesting a role for
EP
increased GluA1 translation.
Testing this hypothesis requires measuring the translation rate of individual proteins. Using 35
S Met/Cys, performed a highly efficient
AC C
GluA1 as a test case, we labeled tissue using
immunoprecipitation of GluA1 (~95% recovery), subjected the immunoprecipitated material to SDS-PAGE/immunoblotting, and exposed the membrane to a phosphorimager screen for 3-4 weeks (Supplementary Figure 3). Even under these conditions, we could detect only faint isotope incorporation into the GluA1 band that was insufficient for reliable quantification. We therefore explored an alternative method based on the ability of the antibiotic puromycin, which resembles tyrosyl-transfer RNA, to incorporate into newly translated proteins (39, 40, 49, 50). When puromycin incorporates into the growing peptide chain, it interrupts elongation; however, 9
ACCEPTED MANUSCRIPT
if used at a low concentration (1 µM), it labels proteins without substantially altering overall protein synthesis (40).
RI PT
We began by verifying that puromycin and metabolic labeling yielded similar results, using NMDAR regulation of translation as a test case. As observed with metabolic labeling (Figure 3A-C), we found that APV increased translation, measured with puromycin labeling, in NAc
SC
tissue from naïve but not cocaine rats (Figures 3D and 3E, naïve: t12=2.25, p=0.04; cocaine: t7=0.08, p=0.94). To examine whether this method could be used to study the rate of translation
M AN U
of individual proteins, we prepared puromycin-labeled NAc tissue and immunoprecipitated with puromycin antibody to isolate all newly translated (puromycin-containing) proteins. Then, we used immunoblotting to measure GluA1 and GluA2 protein levels in starting material and the puromycin-immunoprecipitated fraction. By comparison to samples immunoprecipitated with IgG (control condition), we identified puromycin-labeled GluA1 and GluA2 bands (Figure 4A,D).
TE D
They migrated slightly above their predicted molecular weight, which is not surprising as structural changes due to puromycin incorporation likely affect protein unfolding/SDS binding and thereby alter migration during SDS-PAGE. As expected, the newly translated pool of each
EP
protein was small compared to the amount present in starting material (Figure 4A,E). Perhaps for this reason, the alternative strategy – immunoprecipitating with GluA1 antibody and
AC C
immunoblotting for puromycin – did not yield a quantifiable signal, paralleling results described above for 35S-labeling (see legend to Supplemental Fig. 3).
Puromycin labeling/pull-down followed by GluA immunoblotting was then used to compare GluA1 and GluA2 translation in the NAc of rats that self-administered saline or cocaine. Rats were killed on WD1 (prior to CP-AMPAR elevation) or on >WD40 (maximal incubation and CPAMPAR accumulation). No group differences were observed on WD1 (Fig. 4B and E, GluA1: t14=0.04, p=0.97, GluA2 t14=0.69, p=0.50). However, “incubated” rats (>WD40) exhibited a 10
ACCEPTED MANUSCRIPT
significant increase in GluA1 but not GluA2 translation (Figure 4C and F, GluA1: t14=1.72, p=0.05, GluA2: t13=0.34, p=0.74). To evaluate a possible role for altered transcription, we used additional “incubated” rats and saline controls. We measured levels of Gria1 and Gria2 mRNA
RI PT
extracted from NAc core or shell homogenates, and from synaptoneurosomes prepared from the entire NAc. We found no group differences in mRNA from homogenates (Figure 5A,C, Gria1: Core t18=0.39, p=0.70, Shell, t18=0.21, p=0.83; Gria2: Core t18=0.01, p=0.99, Shell
SC
t18=0.93, p=0.36). In synaptoneurosomes, we found no group difference in Gria1 mRNA levels (Figure 5B, t20=0.68, p=0.50) but observed a trend towards decreased Gria2 mRNA in cocaine
M AN U
rats (Figure 5D, t19=1.77, p=0.09). Results for Gria3 mRNA were similar to those for Gria2 (Supplementary Table 1). The most important finding is absence of a change in Gria1 mRNA in synaptoneurosomes because this suggests that the enhanced GluA1 translation observed after incubation of cocaine craving (Figure 4C) is not explained by increased Gria1 mRNA levels in dendritic compartments. Together, these findings suggest that elevated translation of GluA1
TE D
contributes to CP-AMPAR accumulation in the NAc of “incubated” rats.
Finally, given its robust effect on overall translation in the NAc, we asked whether NMDAR
EP
blockade, shown to increase dendritic GluA1 translation in hippocampal neurons (17), influenced the translation of GluA1 or GluA2 in NAc, and whether this effect was changed
AC C
following incubation. Using the puromycin immunoprecipitation approach described above, we compared GluA1 and GluA2 translation in NAc tissue from “incubated” and naïve rats (naïve rats were chosen as the control because the effect of APV was most robust in this group) labeled with puromycin in the presence of either aCSF (vehicle control) or APV. Results in Figure 6 confirm increased GluA1 translation in ”incubated” rats (naïve aCSF versus Coc aCSF, Tukey’s post hoc, p=0.04). However, in contrast to its ability to enhance overall translation (Figure 3), APV had no effect on GluA1 translation in naïve rats (naïve aCSF versus naïve APV; p=0.39) and did not further augment GluA1 translation in cocaine rats (cocaine aCSF versus 11
ACCEPTED MANUSCRIPT
cocaine APV; p=0.75). For GluA2 translation, two-way ANOVA revealed a significant main effect of drug history, i.e., cocaine exposure tended to reduce GluA2 translation relative to the drug-naïve condition (F(1,28)=9.04, p<0.01). However, Tukey’s post hoc analysis revealed that
RI PT
GluA2 translation did not differ between naïve and cocaine aCSF groups (p=0.15), similar to results in Figure 4F showing no significant difference between saline and cocaine groups. Likewise, there was no difference between cocaine aCSF and cocaine APV groups (p=0.95),
SC
although there was a marginally significant decrease in newly translated GluA2 in cocaine APV tissue compared to naïve aCSF tissue (Figure 5B, p=0.05). Together, these results suggest that
AC C
EP
TE D
M AN U
cocaine exposure and NMDAR blockade differentially influence protein translation in the NAc.
12
ACCEPTED MANUSCRIPT
DISCUSSION This study is the first to characterize regulation of protein translation in the NAc by glutamate transmission. In addition, we tested the hypothesis that this regulation was altered after
RI PT
incubation of cocaine craving. Our results reveal mGlu5 and NMDAR regulation of translation under control conditions, loss of NMDAR regulation of translation after incubation, and
SC
increased GluA1 translation after incubation.
We suggest that our results primarily reflect translation in MSNs, as these comprise ~90% of
M AN U
neurons in the NAc (48). However, contributions of other cell types cannot be excluded. Furthermore, we suggest that dendritic translation may contribute to our measures of protein translation based on the fact that glutamate inputs synapse on dendrites/spines of MSNs rather than the soma (48) and on similarities between the present mGlu1/5 results and those we have
TE D
recently obtained by monitoring translation in processes of cultured MSNs (see next section).
mGlu5 and translational regulation
In hippocampus, it is well established that group I mGluRs stimulate translation (23, 46). In
EP
contrast, we found an MTEP-induced increase in protein translation in NAc tissue from both drug-naïve and “incubated” rats (Figure 3), consistent with results obtained in cultured NAc
AC C
neurons (51). Since MTEP is an mGlu5 NAM, we cannot determine whether it acts by opposing mGlu5 transmission stimulated by tissue glutamate or by opposing constitutive mGlu5 activity (52), although our results in cultured NAc neurons support the latter possibility (51). Regardless, an important conclusion is that mGlu5 transmission exerts an inhibitory influence on protein translation in NAc tissue, opposite from the direction observed in hippocampus. Different “wiring” may reflect the fact that the NAc is comprised mainly of GABAergic MSNs, whereas hippocampal principal neurons are glutamatergic. While this underscores the importance of characterizing mGluR regulation of translation in the NAc, such differences are not entirely 13
ACCEPTED MANUSCRIPT
unprecedented. For example, mGlu5 stimulation in hippocampus leads to LTD that depends upon dendritic translation in the postsynaptic cell and AMPAR endocytosis (15, 46, 53), while the predominant form of mGlu5-induced LTD in striatum is presynaptically expressed via CB1R
RI PT
stimulation (54, 55). The only study implicating protein translation in striatal mGluR LTD indicated a locus for this translation outside the MSN itself (56), whereas another found no
SC
effect of translation inhibition (57).
We previously showed that mGlu5-mediated synaptic depression is disabled in the NAc of
M AN U
“incubated” rats (11). This is accompanied by decreased coupling of long Homer proteins to mGlu5 (12). Homers are critical for regulating group I mGluR association with downstream signaling pathways (58), and changes in association with long Homers can shift mGlu5 signaling (59-61), including signaling regulating protein translation (46, 62). Despite loss of mGlu5 synaptic depression and reduced association with long Homers after incubation (11), the
TE D
MTEP-induced increase in translation in all groups demonstrates that mGlu5 remains capable of regulating protein translation (Figure 3). This dissociation is consistent with evidence that mGlu5-synaptic depression in NAc MSNs involves the canonical Gq-linked pathway (11, 55)
EP
and is independent of protein translation within the MSN (56, 57), whereas mGlu5 regulation of translation in hippocampus is mediated through a distinct beta-arrestin pathway (63, 64).
AC C
Interestingly, after cocaine exposure and extinction training, mGlu5 remains capable of modulating reinstatement (65-69). It will be important to determine the signaling pathways mediating mGlu5’s disparate roles in regulating MSN function (e.g. (70)).
mGlu1 and translational regulation Both mGlu1 activation and exposure to protein synthesis inhibitors remove CP-AMPARs from NAc synapses of “incubated rats” (11, 13), raising the possibility of overlapping mechanisms. We showed previously that mGlu1 removes CP-AMPARs via a PKC-dependent mechanism 14
ACCEPTED MANUSCRIPT
(11) and here we show no effect of the mGlu1 antagonist LY367385 on overall translation in any group (Figure 4). Nevertheless, we cannot rule out a role for mGlu1 in regulating translation of key proteins. We note that protein translation is implicated in the mGlu1-mediated removal of
RI PT
CP-AMPARs from the VTA of cocaine-exposed animals (71). In the VTA, as in the NAc of “incubated” rats, this mGlu1 LTD involves a swap in which high-conductance CP-AMPARs are removed and replaced with lower conductance GluA2-containing AMPARs (11, 28, 71-73). In
SC
VTA, the swap depends upon mGlu1-stimulated translation of GluA2 (28). Future studies should examine potential mGlu1 regulation of GluA1 or GluA2 translation in the NAc, as we have done
NMDAR and translational regulation
M AN U
here for NMDAR regulation.
NMDAR transmission, including spontaneous transmission (miniature EPSCs), can exert a tonic suppressive effect on translation (16-20). In hippocampus and superior colliculus, this was
TE D
shown to depend on Ca2+ influx through the NMDAR channel and Ca2+-induced Ca2+ release, ultimately leading to phosphorylation of eEF2 (74, 75). Phosphorylation of eEF2 inhibits its activity and reduces general translation, a regulated process that contributes to synaptic
EP
plasticity (76, 77). In cultured NAc neurons, the NMDAR antagonist APV did not alter translation, probably due to low glutamate tone at NMDARs in this system (51). However, in
AC C
NAc tissue from drug-naïve animals, our results with APV suggest a suppressive influence of NMDARs on overall protein translation similar to that observed in other regions (Figure 3A,B,D). Interestingly, the ability of APV to increase overall translation is absent in NAc tissue from “incubated” rats (Figure 3C,E). This alteration in NMDAR function may be related to emergence of GluN3-containing NMDARs in NAc MSN during incubation (78). The presence of this atypical subunit confers low Ca2+ permeability on NMDARs (79), and, as noted above, Ca2+ entry is part of the pathway by which NMDARs reduce translation (74, 75).
15
ACCEPTED MANUSCRIPT
What is the functional consequence of loss of NMDAR regulation of translation? In hippocampal neurons, one of the proteins negatively regulated by NMDAR transmission is GluA1, such that NMDAR blockade increases dendritic GluA1 translation and synaptic insertion of homomeric
RI PT
GluA1 CP-AMPARs (17). Therefore, loss of NMDAR negative regulation could in theory account for increased homomeric GluA1 CP-AMPARs in the NAc after incubation of craving. However, in experiments using puromycin labeling to selectively measure GluA1 translation, we found no
SC
effect of APV in NAc tissue from drug-naïve or “incubated” rats (Figure 6). These results suggest that the functional significance of the loss of NMDAR regulation is not related to CP-
M AN U
AMPAR upregulation during cocaine withdrawal. Rather, our recent results (14) suggest that its significance may be manifest during the expression of incubated craving, i.e., during a cueinduced seeking test in which rats perform an operant response to deliver cues previously paired with cocaine. Specifically, we conducted immunoblotting studies of signaling molecules regulating translation in the NAc after long withdrawal (>47 days) from saline or cocaine self-
TE D
administration (i.e., saline and cocaine groups identical to those in the present study) or identically treated rats subjected to a cue-induced seeking test (saline/test and cocaine/test groups). Several robust changes were observed in the cocaine/test group, including
EP
dephosphorylation of eEF2 relative to the cocaine group and the saline/test group. This is indicative of increased translation. Indeed, we also found that expression of incubated craving
AC C
by the cocaine/test group depends on translation in the NAc (14). It is possible that NMDAR plasticity during incubation (78; see previous paragraph) lessens the ability of NMDAR transmission to maintain eEF2 in a phosphorylated/inactivated state, setting the stage for its dephosphorylation/activation during a cue-induced seeking test (14). Other evidence, however, indicates that altered signaling downstream of mTOR may underlie eEF2 dephosphorylation in this group (14). The picture is further complicated by other changes in translation pathways in the cocaine/test group. Most notably, dephosphorylation of eukaryotic initiation factor 2α (eIF2α), which promotes translation, is observed in the cocaine/test group and this 16
ACCEPTED MANUSCRIPT
dephosphorylation is required for expression of incubation (14). Future studies will probe the relationship between incubation-related alterations in NMDAR regulation of translation,
RI PT
downstream signaling cascades, and cue-induced cocaine seeking.
Conclusions
These experiments are the first to characterize how glutamatergic transmission regulates
SC
protein translation in the NAc under control conditions and after incubation of cocaine craving. Under control conditions, translation in the NAc showed similar regulation by NMDARs but
M AN U
different regulation by group I mGluRs compared to hippocampus, revealing cell-type differences in regulatory mechanisms. While metabolic labeling revealed no difference in overall protein translation between saline and cocaine animals, the ability of translation inhibition to normalize the elevation of homomeric GluA1 CP-AMPARs that mediates expression of incubation (13) indicates that some proteins must be differentially translated in the NAc of
TE D
“incubated” rats. Indeed, we found elevated translation of GluA1 after incubation, suggesting a simple explanation for increased CP-AMPAR formation. Incubation was also accompanied by loss of NMDAR-mediated suppression of protein translation in the NAc. The functional
EP
consequences of this adaptation remain to be determined, but our results suggest that it is not
AC C
related to increased GluA1 translation.
17
ACCEPTED MANUSCRIPT
Acknowledgements: This research was supported by R01 DA015835 to MEW, postdoctoral NRSA DA040414 to MTS, and predoctoral NRSA DA036950 to CTW. We thank Dr. Emily Osterweil for generously sharing protocols and otherwise advising us on the metabolic labeling
RI PT
experiments, and Dr. Xuan (Anna) Li for help with design of qPCR experiments.
Financial disclosures: The authors report no biomedical financial interests or potential conflicts
AC C
EP
TE D
M AN U
SC
of interest.
18
ACCEPTED MANUSCRIPT FIGURE LEGENDS: Figure 1. Monitoring protein translation in the NAc using
35
S-Met/Cys incorporation reveals no differences in
overall translation between saline controls and rats that have undergone incubation of cocaine craving. (A) Representative autoradiograph shows
35
S-Met/Cys incorporation into proteins of various molecular weights in
lane. (C) Quantification of
RI PT
freshly dissected NAc tissue. (B) Ponceau S stain of the same gel demonstrates equal protein loading in each 35
S-labeled proteins in A, normalized to total protein in each lane determined with
Ponceau S staining, shows that 30 min of incubation with the protein translation inhibitor cycloheximide (Cyclo) significantly reduces protein translation. (D) Mean ± SEM number of cocaine infusions per day over the ten 6-
SC
hour daily self-administration training sessions for all animals (163 total rats). (E) Quantification of
35
S-labeled
proteins (normalized to total protein in each lane) shows that overall translation did not differ between cocaine
M AN U
and saline self-administering rats killed after withdrawal day 40. Numbers in black bars (panels C and E) represent the number of rats in the experiment; for each rat, the NAc from one hemisphere was used for the control group and the NAc from the other hemisphere was used for the experimental group. *p<0.05 versus aCSF control (paired t-test).
TE D
Figure 2. mGlu5 blockade increases protein translation in the NAc of naïve rats, saline controls, and “incubated” rats while mGlu1 blockade does not influence overall translation in the NAc. Including the mGlu5 antagonist MTEP (25 µM) during the
35
S-Met/Cys incorporation period increases protein translation in NAc
EP
tissue from naïve rats (A), as well as rats that self-administered saline (B) or cocaine (C) and were killed after withdrawal day 40. Including the mGlu1 antagonist LY367385 (LY; 50 µM) during the
35
S incorporation period
AC C
does not significantly alter protein translation in NAc tissue from the same experimental groups (D-F). Numbers in black bars represent the number of rats in the experiment; for each rat, the NAc from one hemisphere was used for the control group and the NAc from the other hemisphere was used for the experimental group. *p<0.05 versus aCSF control (paired t-test).
Figure 3. NMDAR blockade increases translation in control rats, but not following prolonged withdrawal from cocaine self-administration. Including the NMDAR antagonist APV (50 µM) during the 19
35
S-Met/Cys or
ACCEPTED MANUSCRIPT puromycin incorporation period increases protein translation in NAc tissue dissected from naïve rats (A, D) and saline self-administering rats killed after withdrawal day 40 (B), but not NAc tissue from cocaine selfadministering rats killed after withdrawal day 40 (C, E). Numbers in black bars represent the number of rats in the experiment. For rats in these metabolic labeling and puromycin experiments, the NAc from one hemisphere
RI PT
was used for the control group and the NAc from the other hemisphere was used for the experimental group. In both cases, data were normalized to total protein in each lane, determined with Ponceau S staining. **p<0.01 and *p<0.05 versus aCSF control (paired t-test).
SC
Figure 4. GluA1 translation in the NAc is unchanged one day after discontinuing cocaine self-administration but is increased after incubation of craving has occurred. NAc tissue was labeled with puromycin and then
M AN U
immunoprecipitated using puromycin antibody or IgG (control condition). Equal amounts (40 µg) of starting material (S), bound material (B; immunoprecipitated material purified by Protein A/G resin), and unbound material (U) were separated by SDS-PAGE and immunoblotted for GluA1 (A) or GluA2 (D). Comparison of bound fractions after puromycin versus IgG immunoprecipitation reveals newly translated GluA1 and GluA2 bands that migrate above their predicted molecular weight (arrowheads in panels A and D), presumably due to
TE D
structural changes associated with puromycin incorporation (see Results). Using this approach, we demonstrated that, compared to saline controls, GluA1 translation is unchanged after 1 day of withdrawal from cocaine self-administration (B) but increased after prolonged withdrawal (>40 days) (C), whereas GluA2
EP
translation is unchanged at both withdrawal times (E,F). For these puromycin experiments, each sample consisted of NAc pooled from both hemispheres of one rat. Numbers in bars represent number of samples/rats
AC C
in each group. *p<0.05 versus saline controls (unpaired t-test).
Figure 5. GluA1 and GluA2 mRNA levels are unchanged in NAc homogenates or synaptoneurosomes after incubation of cocaine craving. (A and C) Quantification of Gria1 and Gria2 mRNA levels in NAc core and NAc shell homogenates indicates no differences between rats killed after >40 days of withdrawal from saline versus cocaine self-administration. (B and D) Likewise, there was no significant difference in Gria1 or Gria2 mRNA levels in synaptoneurosomes prepared from the whole NAc of identically treated saline and cocaine rats, 20
ACCEPTED MANUSCRIPT although a trend towards a reduction in Gria2 mRNA (p=0.09) was observed in the cocaine group. For these experiments, each sample consisted of NAc pooled from both hemispheres of one rat. Numbers in bars represent number of samples/rats in each group. Data analyzed with unpaired t-tests.
RI PT
Figure 6. APV does not influence GluA1 translation in the NAc of naïve or “incubated” rats. NAc tissue from drug-naïve rats or rats killed after >40 days of withdrawal from cocaine self-administration was labeled with puromycin in the presence of aCSF or APV (50 µM). Tissue was immunoprecipitated with puromycin antibody to isolate newly translated proteins, and the immunoprecipitated fraction was immunoblotted for GluA1 (A) or
SC
GluA2 (B). (A) Comparison of GluA1 translation in naïve aCSF and cocaine aCSF tissue confirmed results in Figure 4 demonstrating elevated GluA1 translation after cocaine withdrawal, but APV did not influence GluA1
M AN U
translation in either naive or cocaine groups. (B) A two-way ANOVA revealed a main effect of drug preexposure on GluA2 translation, as indicated by lines over bars (#p<0.05) (see Results for discussion of post hoc comparisons). For these puromycin experiments, each sample consisted of NAc pooled from both hemispheres of one rat. Numbers in bars represent number of samples/rats in each group. *p<0.05 compared
AC C
EP
TE D
to naïve aCSF group (Tukey test).
21
ACCEPTED MANUSCRIPT REFERENCES:
AC C
EP
TE D
M AN U
SC
RI PT
1. Childress AR, Mozley PD, McElgin W, Fitzgerald J, Reivich M, O'Brien CP (1999): Limbic activation during cue-induced cocaine craving. The American journal of psychiatry. 156:11-18. 2. Bonson KR, Grant SJ, Contoreggi CS, Links JM, Metcalfe J, Weyl HL, et al. (2002): Neural systems and cue-induced cocaine craving. Neuropsychopharmacology : official publication of the American College of Neuropsychopharmacology. 26:376-386. 3. Parvaz MA, Moeller SJ, Malaker P, Sinha R, Alia-Klein N, Goldstein RZ (2016): Abstinence reverses EEG-indexed attention bias between drug-related and pleasant stimuli in cocaine-addicted individuals. J Psychiatry Neurosci. 41:150358. 4. Lu L, Grimm JW, Hope BT, Shaham Y (2004): Incubation of cocaine craving after withdrawal: a review of preclinical data. Neuropharmacology. 47 Suppl 1:214-226. 5. Wolf ME (2016): Synaptic mechanisms underlying persistent cocaine craving. Nature reviews Neuroscience. 17:351-365. 6. Reichel CM, Bevins RA (2009): Forced abstinence model of relapse to study pharmacological treatments of substance use disorder. Current drug abuse reviews. 2:184-194. 7. Bedi G, Preston KL, Epstein DH, Heishman SJ, Marrone GF, Shaham Y, et al. (2011): Incubation of cue-induced cigarette craving during abstinence in human smokers. Biol Psychiatry. 69:708-711. 8. Li P, Wu P, Xin X, Fan YL, Wang GB, Wang F, et al. (2015): Incubation of alcohol craving during abstinence in patients with alcohol dependence. Addict Biol. 20:513-522. 9. Wang G, Shi J, Chen N, Xu L, Li J, Li P, et al. (2013): Effects of length of abstinence on decision-making and craving in methamphetamine abusers. PLoS One. 8:e68791. 10. Pickens CL, Airavaara M, Theberge F, Fanous S, Hope BT, Shaham Y (2011): Neurobiology of the incubation of drug craving. Trends in neurosciences. 34:411-420. 11. McCutcheon JE, Loweth JA, Ford KA, Marinelli M, Wolf ME, Tseng KY (2011): Group I mGluR activation reverses cocaine-induced accumulation of calcium-permeable AMPA receptors in nucleus accumbens synapses via a protein kinase C-dependent mechanism. The Journal of neuroscience : the official journal of the Society for Neuroscience. 31:14536-14541. 12. Loweth JA, Scheyer AF, Milovanovic M, LaCrosse AL, Flores-Barrera E, Werner CT, et al. (2014): Synaptic depression via mGluR1 positive allosteric modulation suppresses cue-induced cocaine craving. Nature neuroscience. 17:73-80. 13. Scheyer AF, Wolf ME, Tseng KY (2014): A protein synthesis-dependent mechanism sustains calcium-permeable AMPA receptor transmission in nucleus accumbens synapses during withdrawal from cocaine self-administration. The Journal of neuroscience : the official journal of the Society for Neuroscience. 34:3095-3100. 14. Werner CT, Stefanik MT, Milovanovic M, Wolf ME (2015): Regulation of protein translation following prolonged withdrawal from extended-access cocaine self-administration with or with cocaine memory retrieval. Soc Neurosci Abstr. 41. 15. Luscher C, Huber KM (2010): Group 1 mGluR-dependent synaptic long-term depression: mechanisms and implications for circuitry and disease. Neuron. 65:445-459. 16. Sutton MA, Wall NR, Aakalu GN, Schuman EM (2004): Regulation of dendritic protein synthesis by miniature synaptic events. Science. 304:1979-1983. 17. Sutton MA, Ito HT, Cressy P, Kempf C, Woo JC, Schuman EM (2006): Miniature neurotransmission stabilizes synaptic function via tonic suppression of local dendritic protein synthesis. Cell. 125:785-799. 18. Sutton MA, Taylor AM, Ito HT, Pham A, Schuman EM (2007): Postsynaptic decoding of neural activity: eEF2 as a biochemical sensor coupling miniature synaptic transmission to local protein synthesis. Neuron. 55:648-661. 22
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
19. Autry AE, Adachi M, Nosyreva E, Na ES, Los MF, Cheng PF, et al. (2011): NMDA receptor blockade at rest triggers rapid behavioural antidepressant responses. Nature. 475:91-95. 20. Kavalali ET, Monteggia LM (2012): Synaptic mechanisms underlying rapid antidepressant action of ketamine. The American journal of psychiatry. 169:1150-1156. 21. Li N, Lee B, Liu RJ, Banasr M, Dwyer JM, Iwata M, et al. (2010): mTOR-dependent synapse formation underlies the rapid antidepressant effects of NMDA antagonists. Science. 329:959-964. 22. Osterweil EK, Krueger DD, Reinhold K, Bear MF (2010): Hypersensitivity to mGluR5 and ERK1/2 leads to excessive protein synthesis in the hippocampus of a mouse model of fragile X syndrome. The Journal of neuroscience : the official journal of the Society for Neuroscience. 30:15616-15627. 23. Bhakar AL, Dolen G, Bear MF (2012): The pathophysiology of fragile X (and what it teaches us about synapses). Annual review of neuroscience. 35:417-443. 24. Michalon A, Sidorov M, Ballard TM, Ozmen L, Spooren W, Wettstein JG, et al. (2012): Chronic pharmacological mGlu5 inhibition corrects fragile X in adult mice. Neuron. 74:49-56. 25. Richter JD, Bassell GJ, Klann E (2015): Dysregulation and restoration of translational homeostasis in fragile X syndrome. Nature reviews Neuroscience. 16:595-605. 26. Karler R, Finnegan KT, Calder LD (1993): Blockade of behavioral sensitization to cocaine and amphetamine by inhibitors of protein synthesis. Brain research. 603:19-24. 27. Sorg BA, Ulibarri C (1995): Application of a protein synthesis inhibitor into the ventral tegmental area, but not the nucleus accumbens, prevents behavioral sensitization to cocaine. Synapse. 20:217-224. 28. Mameli M, Balland B, Lujan R, Luscher C (2007): Rapid synthesis and synaptic insertion of GluR2 for mGluR-LTD in the ventral tegmental area. Science. 317:530-533. 29. Bailey J, Ma D, Szumlinski KK (2012): Rapamycin attenuates the expression of cocaineinduced place preference and behavioral sensitization. Addict Biol. 17:248-258. 30. Yu F, Zhong P, Liu X, Sun D, Gao HQ, Liu QS (2013): Metabotropic glutamate receptor I (mGluR1) antagonism impairs cocaine-induced conditioned place preference via inhibition of protein synthesis. Neuropsychopharmacology : official publication of the American College of Neuropsychopharmacology. 38:1308-1321. 31. Smith LN, Jedynak JP, Fontenot MR, Hale CF, Dietz KC, Taniguchi M, et al. (2014): Fragile X mental retardation protein regulates synaptic and behavioral plasticity to repeated cocaine administration. Neuron. 82:645-658. 32. Shi X, Miller JS, Harper LJ, Poole RL, Gould TJ, Unterwald EM (2014): Reactivation of cocaine reward memory engages the Akt/GSK3/mTOR signaling pathway and can be disrupted by GSK3 inhibition. Psychopharmacology. 231:3109-3118. 33. Jian M, Luo YX, Xue YX, Han Y, Shi HS, Liu JF, et al. (2014): eIF2alpha dephosphorylation in basolateral amygdala mediates reconsolidation of drug memory. The Journal of neuroscience : the official journal of the Society for Neuroscience. 34:10010-10021. 34. Huang W, Placzek AN, Viana Di Prisco G, Khatiwada S, Sidrauski C, Krnjevic K, et al. (2016): Translational control by eIF2alpha phosphorylation regulates vulnerability to the synaptic and behavioral effects of cocaine. Elife. 5. 35. Dayas CV, Smith DW, Dunkley PR (2012): An emerging role for the Mammalian target of rapamycin in "pathological" protein translation: relevance to cocaine addiction. Front Pharmacol. 3:13. 36. Neasta J, Barak S, Hamida SB, Ron D (2014): mTOR complex 1: a key player in neuroadaptations induced by drugs of abuse. Journal of neurochemistry. 130:172-184. 37. Costa-Mattioli M, Sossin WS, Klann E, Sonenberg N (2009): Translational control of longlasting synaptic plasticity and memory. Neuron. 61:10-26. 38. Most D, Leiter C, Blednov YA, Harris RA, Mayfield RD (2016): Synaptic microRNAs Coordinately Regulate Synaptic mRNAs: Perturbation by Chronic Alcohol Consumption. 23
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
Neuropsychopharmacology : official publication of the American College of Neuropsychopharmacology. 41:538-548. 39. Schmidt EK, Clavarino G, Ceppi M, Pierre P (2009): SUnSET, a nonradioactive method to monitor protein synthesis. Nature methods. 6:275-277. 40. Goodman CA, Hornberger TA (2013): Measuring protein synthesis with SUnSET: a valid alternative to traditional techniques? Exerc Sport Sci Rev. 41:107-115. 41. Ferrario CR, Loweth JA, Milovanovic M, Wang X, Wolf ME (2011): Distribution of AMPA receptor subunits and TARPs in synaptic and extrasynaptic membranes of the adult rat nucleus accumbens. Neuroscience letters. 490:180-184. 42. Conrad KL, Tseng KY, Uejima JL, Reimers JM, Heng LJ, Shaham Y, et al. (2008): Formation of accumbens GluR2-lacking AMPA receptors mediates incubation of cocaine craving. Nature. 454:118-121. 43. Wolf ME, Tseng KY (2012): Calcium-permeable AMPA receptors in the VTA and nucleus accumbens after cocaine exposure: when, how, and why? Frontiers in molecular neuroscience. 5:72. 44. Loweth JA, Tseng KY, Wolf ME (2014): Adaptations in AMPA receptor transmission in the nucleus accumbens contributing to incubation of cocaine craving. Neuropharmacology. 76 Pt B:287300. 45. Cosford ND, Tehrani L, Roppe J, Schweiger E, Smith ND, Anderson J, et al. (2003): 3-[(2Methyl-1,3-thiazol-4-yl)ethynyl]-pyridine: a potent and highly selective metabotropic glutamate subtype 5 receptor antagonist with anxiolytic activity. J Med Chem. 46:204-206. 46. Waung MW, Huber KM (2009): Protein translation in synaptic plasticity: mGluR-LTD, Fragile X. Current opinion in neurobiology. 19:319-326. 47. Clark BP, Baker RS, Goldsworthy J, Harris JR, Kingston AE (1997): (+)-2-methyl-4caboxyphenylglycine (LY367385) selectively antagonises metabotropic glutamate mGluR1 receptors. Bioorg Med Chem Lett. 7:2777-2780. 48. Ferrario CR, Loweth JA, Milovanovic M, Ford KA, Galinanes GL, Heng LJ, et al. (2011): Alterations in AMPA receptor subunits and TARPs in the rat nucleus accumbens related to the formation of Ca(2)(+)-permeable AMPA receptors during the incubation of cocaine craving. Neuropharmacology. 61:1141-1151. 49. Starck SR, Green HM, Alberola-Ila J, Roberts RW (2004): A general approach to detect protein expression in vivo using fluorescent puromycin conjugates. Chem Biol. 11:999-1008. 50. Ge J, Zhang CW, Ng XW, Peng B, Pan S, Du S, et al. (2016): Puromycin Analogues Capable of Multiplexed Imaging and Profiling of Protein Synthesis and Dynamics in Live Cells and Neurons. Angew Chem Int Ed Engl. 55:4933-4937. 51. Stefanik MT, Sakas C, Lee D, Wolf ME (2017): The regulation of protein translation by ionotropic and metabotropic glutamate receptors in co-cultured nucleus accumbens and prefrontal cortex neurons. Submitted. 52. Ango F, Prezeau L, Muller T, Tu JC, Xiao B, Worley PF, et al. (2001): Agonist-independent activation of metabotropic glutamate receptors by the intracellular protein Homer. Nature. 411:962965. 53. Huber KM, Kayser MS, Bear MF (2000): Role for rapid dendritic protein synthesis in hippocampal mGluR-dependent long-term depression. Science. 288:1254-1257. 54. Choi S, Lovinger DM (1997): Decreased frequency but not amplitude of quantal synaptic responses associated with expression of corticostriatal long-term depression. The Journal of neuroscience : the official journal of the Society for Neuroscience. 17:8613-8620. 55. Robbe D, Kopf M, Remaury A, Bockaert J, Manzoni OJ (2002): Endogenous cannabinoids mediate long-term synaptic depression in the nucleus accumbens. Proceedings of the National Academy of Sciences of the United States of America. 99:8384-8388.
24
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
56. Yin HH, Davis MI, Ronesi JA, Lovinger DM (2006): The role of protein synthesis in striatal long-term depression. The Journal of neuroscience : the official journal of the Society for Neuroscience. 26:11811-11820. 57. Jung KM, Sepers M, Henstridge CM, Lassalle O, Neuhofer D, Martin H, et al. (2012): Uncoupling of the endocannabinoid signalling complex in a mouse model of fragile X syndrome. Nat Commun. 3:1080. 58. Szumlinski KK, Ary AW, Lominac KD (2008): Homers regulate drug-induced neuroplasticity: implications for addiction. Biochem Pharmacol. 75:112-133. 59. Kammermeier PJ, Xiao B, Tu JC, Worley PF, Ikeda SR (2000): Homer proteins regulate coupling of group I metabotropic glutamate receptors to N-type calcium and M-type potassium channels. The Journal of neuroscience : the official journal of the Society for Neuroscience. 20:72387245. 60. Kammermeier PJ, Worley PF (2007): Homer 1a uncouples metabotropic glutamate receptor 5 from postsynaptic effectors. Proceedings of the National Academy of Sciences of the United States of America. 104:6055-6060. 61. Kammermeier PJ (2008): Endogenous homer proteins regulate metabotropic glutamate receptor signaling in neurons. The Journal of neuroscience : the official journal of the Society for Neuroscience. 28:8560-8567. 62. Ronesi JA, Collins KA, Hays SA, Tsai NP, Guo W, Birnbaum SG, et al. (2012): Disrupted Homer scaffolds mediate abnormal mGluR5 function in a mouse model of fragile X syndrome. Nature neuroscience. 15:431-440, S431. 63. Eng AG, Kelver DA, Hedrick TP, Swanson GT (2016): Transduction of group I mGluRmediated synaptic plasticity by beta-arrestin2 signalling. Nat Commun. 7:13571. 64. Stoppel LJ, Auerbach BD, Senter RK, Preza AR, Lefkowitz RJ, Bear MF (2017): beta-Arrestin2 Couples Metabotropic Glutamate Receptor 5 to Neuronal Protein Synthesis and Is a Potential Target to Treat Fragile X. Cell Rep. 18:2807-2814. 65. Kufahl PR, Hood LE, Nemirovsky NE, Barabas P, Halstengard C, Villa A, et al. (2012): Positive Allosteric Modulation of mGluR5 Accelerates Extinction Learning but Not Relearning Following Methamphetamine Self-Administration. Front Pharmacol. 3:194. 66. Wang X, Moussawi K, Knackstedt L, Shen H, Kalivas PW (2013): Role of mGluR5 neurotransmission in reinstated cocaine-seeking. Addict Biol. 18:40-49. 67. Knackstedt LA, Trantham-Davidson HL, Schwendt M (2014): The role of ventral and dorsal striatum mGluR5 in relapse to cocaine-seeking and extinction learning. Addict Biol. 19:87-101. 68. Bird MK, Lohmann P, West B, Brown RM, Kirchhoff J, Raymond CR, et al. (2014): The mGlu5 receptor regulates extinction of cocaine-driven behaviours. Drug Alcohol Depend. 137:83-89. 69. Knackstedt LA, Schwendt M (2016): mGlu5 Receptors and Relapse to Cocaine-Seeking: The Role of Receptor Trafficking in Postrelapse Extinction Learning Deficits. Neural Plast. 2016:9312508. 70. Mao L, Yang L, Tang Q, Samdani S, Zhang G, Wang JQ (2005): The scaffold protein Homer1b/c links metabotropic glutamate receptor 5 to extracellular signal-regulated protein kinase cascades in neurons. The Journal of neuroscience : the official journal of the Society for Neuroscience. 25:2741-2752. 71. Bellone C, Luscher C (2006): Cocaine triggered AMPA receptor redistribution is reversed in vivo by mGluR-dependent long-term depression. Nature neuroscience. 9:636-641. 72. Bellone C, Luscher C (2005): mGluRs induce a long-term depression in the ventral tegmental area that involves a switch of the subunit composition of AMPA receptors. The European journal of neuroscience. 21:1280-1288. 73. Mameli M, Halbout B, Creton C, Engblom D, Parkitna JR, Spanagel R, et al. (2009): Cocaineevoked synaptic plasticity: persistence in the VTA triggers adaptations in the NAc. Nature neuroscience. 12:1036-1041. 25
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
74. Scheetz AJ, Nairn AC, Constantine-Paton M (2000): NMDA receptor-mediated control of protein synthesis at developing synapses. Nature neuroscience. 3:211-216. 75. Reese AL, Kavalali ET (2015): Spontaneous neurotransmission signals through store-driven Ca(2+) transients to maintain synaptic homeostasis. Elife. 4. 76. Heise C, Gardoni F, Culotta L, di Luca M, Verpelli C, Sala C (2014): Elongation factor-2 phosphorylation in dendrites and the regulation of dendritic mRNA translation in neurons. Front Cell Neurosci. 8:35. 77. Taha E, Gildish I, Gal-Ben-Ari S, Rosenblum K (2013): The role of eEF2 pathway in learning and synaptic plasticity. Neurobiology of learning and memory. 105:100-106. 78. Christian DT, Tseng, K.Y., Wolf, M.E. (2017): Extended access cocaine self-administration leads to increased GluN3-containing NMDA receptor function in the rat nucleus accumbens. Society for Neuroscience 79. Paoletti P, Bellone C, Zhou Q (2013): NMDA receptor subunit diversity: impact on receptor properties, synaptic plasticity and disease. Nature reviews Neuroscience. 14:383-400.
26
ACCEPTED MANUSCRIPT
B
Cyclo
aCSF
Cyclo
C
50 0
EP
150 100
35
100
*
50
0
Coc Sal
1
2
3
4
5
6
7
8
Self-Administration Day
9
10
150
100
50
0
6
Sal
4
aCSF Cyclo
E S Met-Cys/total (% saline)
200
35
D
150
Total Protein
TE D
S Met-Cys
AC C
35
M AN U
SC
RI PT
aCSF
Infusions
A
S Met-Cys/total (% aCSF)
Figure 1
Coc
ACCEPTED MANUSCRIPT
150
35
*
100
M AN U
50
*
8
8 aCSF MTEP
aCSF MTEP
TE D
150
Cocaine
SC
100
0
C
Saline
*
E
Naïve
AC C
S Met-Cys/total (% aCSF)
D
B
Naïve
Saline
8 aCSF MTEP
F
Cocaine
EP
35
S Met-Cys/total (% aCSF)
A
RI PT
Figure 2
50
0
8
aCSF
6 LY
aCSF
7 LY
aCSF
LY
ACCEPTED MANUSCRIPT
35
A
Naïve
S Met-Cys/total (% aCSF)
150
S Met-Cys
B
**
C
Saline
*
SC M AN U
50
0
7
8 aCSF
aCSF
APV
7 aCSF
APV
150
Naïve
*
E
Cocaine
EP
Puro/total (% aCSF)
D
TE D
Puromycin
100
AC C
35
100
Cocaine
RI PT
Figure 3
50
0
13 aCSF
8 APV
aCSF
APV
APV
ACCEPTED MANUSCRIPT
GluA1
B
IgG
GluA1 (% saline)
kD S B U B U
220 120 100 80 60 50 40
WD1
100
50
30 20
M AN U
E
WD1
F
150
8
8
>WD40
EP
GluA2 (% saline)
S BU B U
AC C
40
IgG
TE D
Puro
60 50
*
Saline Cocaine
GluA2
120 100 80
8
GluA1
20
kD 220
8
0
30
D
C
150
>WD40
RI PT
A
Puro
SC
Figure 4
100
50
0
GluA2
8
8
7
8
Figure 5
Saline Cocaine
150
mRNA levels (% saline)
A
Gria1
B
0
10
10
M AN U
50
10
10
Core Shell Homogenates
TE D
Gria2
AC C
100
12
10
Synaptoneurosomes
D
EP
150
mRNA levels (% saline)
C
SC
100
RI PT
ACCEPTED MANUSCRIPT
50
0
10
10
10
10
Core Shell Homogenates
11
10
Synaptoneurosomes
ACCEPTED MANUSCRIPT
100 50 0
EP
GluA1
AC C
B
M AN U
150
TE D
GluA1 (% naive aCSF)
Naive aCSF Naive APV Coc aCSF Coc APV
*
200
8
8
8
8
#
150
GluA2 (% naive aCSF)
A
SC
RI PT
Figure 6
100
*
50
0 GluA2
8
8
8
8
ACCEPTED MANUSCRIPT Stefanik et al.
Supplement
Withdrawal From Cocaine Self-administration Alters the Regulation of Protein Translation in the Nucleus Accumbens
RI PT
Supplemental Information
Subjects and drug self-administration
All experimental procedures were approved by the Rosalind Franklin University Institutional Animal Care and Use Committee in accordance with the USPHS Guide for Care and Use of
SC
Laboratory Animals. A total of 93 male Sprague-Dawley rats were used (250-275g on arrival; Envigo, Indianapolis, IN). Rats were group housed upon arrival (3/cage) in temperature- and
M AN U
humidity-controlled conditions with a 12-hour reverse light/dark cycle (lights off at 8:00 A.M.) until surgery and then single-housed for the remainder of the experiment. Food and water were available ad libitum. Rats were acclimated to the vivarium for at least 1 week prior to surgeries to implant intra-jugular catheters. As described previously (1), animals were anesthetized using
TE D
a cocktail of ketamine HCl (80mg/kg) and Xylazine (10mg/kg) (i.p.). A Silastic catheter (PlasticsOne, Roanoke, VA) was inserted and secured in the right jugular vein, and connected to a subcutaneous back mount, exiting the back in the mid-scapular region. Intravenous
EP
catheters were flushed daily with cefazolin (0.2mL of 0.1 g/mL in sterile 0.9% saline) to prevent infection and maintain catheter patency. Rats recovered for one week before self-administration
AC C
training. Rats were trained to nose-poke to self-administer saline or cocaine (FR1 schedule, 0.5 mg/kg/infusion, 6 h/day for 10 days). Infusions were paired with a light cue. Following the final self-administration session, rats were returned to home cages for either 1 or >40 days before conducting experiments. All training occurred during the dark cycle.
Metabolic labeling A schematic of the experimental timeline is depicted in Supplemental Figure S1. Details of tissue processing are provided in the Methods and Materials section of the main text. Previous
1
ACCEPTED MANUSCRIPT Stefanik et al.
Supplement
work in hippocampal tissue has demonstrated that several hours of post-slicing recovery is necessary to achieve stable basal levels of protein translation (2). As shown in Supplemental Figure S2, we found that optimal conditions for the measurement of protein translation in NAc
Following recovery, tissue was incubated with
35
RI PT
tissue required at least 3 hours of post-dissection recovery, in line with this previous work (2). S-Met/Cys in the presence or absence of test
drugs. Based on related studies (2-4), the following drug concentrations were used:
SC
cycloheximide (Sigma-Aldrich, St. Louis, MO), 60 µM; LY367385 (Tocris; Bristol, U.K.), 50 µM; 3-((2-Methyl-1,3-thiazol-4-yl)ethynyl)pyridine hydrochloride (MTEP; Tocris), 25 µM; and APV
M AN U
(Tocris), 50 µM. The vehicle for test drugs was always <1% DMSO.
Autoradiography Tissue labeled with
35
S-Met/Cys was homogenized in ice-cold homogenization buffer containing
1% Triton X-100 and proteins were TCA-precipitated, as previously described (2). A portion of
TE D
the sample was processed for SDS-PAGE, transferred to nitrocellulose, and exposed to a phosphorimager screen for 24 h before reading with a phosphorimager (Typhoon; GE Healthcare, Little Chalfont, UK). The same membrane was then stained with Ponceau S to
software; a ratio of
EP
visualize total protein in each lane. Optical density was analyzed using TotalLab Quant 35
S-incorporation to total protein was calculated for each lane and used as a
AC C
measure of overall translation in each sample.
Synaptoneurosome preparation After 55 days of withdrawal from saline or cocaine self-administration, rats were decapitated and the NAc was dissected from 2 mm coronal slices prepared with a brain matrix. Immediately following dissection, synaptoneurosomes were prepared as previously described (5), with slight modifications. Briefly, NAc punches were homogenized in 500 μL of homogenization buffer [HB; 20 mM HEPES, 0.5 mM EGTA, 1X Preasome Inhibitor Cocktail Set 1 (Calbiochem; Temecula,
2
ACCEPTED MANUSCRIPT Stefanik et al.
Supplement
CA)]. Homogenates were passed through a 100 μm-pore filter and then through a 5 μm filter (Millipore). After homogenates were passed through each filter, the filters were washed with 50 μL of HB and the washes were added to the homogenates to maximize yield. Homogenates
RI PT
were centrifuged at 14,000 x g for 20 minutes at 4oC. The pellet containing the synaptoneurosomes was frozen on dry ice, stored at -80oC, and finally lysed in a lysis buffer [0.605 g Tris-HCl, 0.25 g sodium deoxycholate, 0.876 g NaCl, 1 μg/mL PMSF, 5 mL 20% SDS,
SC
1X Protease Inhibitor Cocktail Set 1 (Calbiochem) in 100 mL dH2O] for immunoblotting.
M AN U
Whole cell RNA extraction, reverse transcription, and quantitative (q)PCR RNA was extracted using TRI Reagent and the corresponding protocol (Invitrogen, Waltham, MA). Briefly, freshly dissected NAc tissue was homogenized in 50 μl TRI Reagent, brought up to 400 μL with TRI Reagent and incubated for 5 min at room temperature (RT). Homogenates were then centrifuged at 12,000 x g for 10 min at 4°C and aqueous supernatants transferred to
TE D
new tubes. Next, 50 μL of bromochloropropane (BCP) was added. Samples were shaken vigorously, incubated for 5 min at RT and centrifuged at 12,000 x g for 15 min at 4oC. About 75% of the clear, aqueous RNA phase was transferred into a clean tube. Then, 150 μL of
EP
nuclease free water and 50 μL BCP were added. Samples were shaken vigorously, incubated for 10 min at RT and centrifuged at 12,000 x g for 15 min at 4°C. About 75% of the aqueous
AC C
RNA layer was transferred to a new tube and 3M sodium acetate (1:10) and 1 μL glycogen were added and mixed. For RNA precipitation, isopropanol (1:1) was then added and samples were incubated overnight at -20°C. The next day, samples were centrifuged at 12,000 x g for 10 min at 4°C. After discarding the supernatant, the RNA pellet was washed with 70% ethanol and centrifuged at 7,500 x g for 5 min at 4°C. Ethanol was removed and pellets were allowed to airdry for 5 min. RNA pellets were then dissolved in 50 μL nuclease free water and stored at -80°C. Reverse transcription of RNA was completed using the commercially available SuperScript III Reverse Transcriptase kit (Invitrogen). Manufacturer instructions were followed.
3
ACCEPTED MANUSCRIPT Stefanik et al.
Supplement
qPCR assays were performed using Fam-labeled probes for target genes and a Viclabeled probe for the housekeeping control gene (GAPDH), along with TaqMan Gene Expression Master Mix (Applied Biosystems). qPCR reactions were run in a ViiA7 instrument
RI PT
(Applied Biosystems). Samples were run in duplicate and relative expression was calculated
sequences.
Synaptoneurosome RNA extraction and qPCR
SC
using the ΔΔCt with GAPDH as the control. See Supplemental Table S1 for mRNA primer
M AN U
Synaptoneurosome RNA was extracted and precipitated as above but smaller volumes were used for processing. Reverse transcription of synaptoneurosome RNA was completed using the commercially available SuperScript III Reverse Transcriptase kit (Invitrogen). Manufacturer instructions were followed.
TaqMan PreAmp Master Mix Kit (Applied Biosystems, Waltham, MA) was used for cDNA
TE D
pre-amplification. The PreAmp Reaction was made up as follows: synaptoneurosome cDNA (6 μL), primer sets (1.5 μL each), and PreAmp Mix (15 μL). The final volume was brought to 30 μL with nuclease free water. cDNA was pre-amplified in a Mastercycler Gradient Thermal Cycler
EP
(Ependorf, Hauppauge, NY) using the following program settings: 95ºC hold for 10 min, 14 denaturation cycles at 90ºC for 15 sec, and annealing and extension at 60ºC for 4 min. The pre-
AC C
amplification products were then diluted 5X with nuclease free water and stored at -20ºC. Uniformity of the pre-amplification was verified by comparing cDNA templates from the preamplified and unamplified NAc samples. ΔΔCt values were within the range of ±1.5 for all target genes between the pre-amplified and unamplified samples. qPCR assays were performed as described in the previous section but using actin as the housekeeping control gene.
4
ACCEPTED MANUSCRIPT Stefanik et al.
Supplement
Immunoprecipitation Using immunoprecipitation methods previously described (1, 6), we quantified levels of GluA1 and GluA2 AMPAR subunits that had incorporated either 35S-Met/Cys or puromycin. Briefly, rats
RI PT
were decapitated, brains rapidly removed, and NAc (mainly core) dissected on ice from a 2 mm coronal section obtained using a brain matrix. Membranes were lysed by sonication and sedimented by centrifugation at 23,000 x g for 2 min at 4°C. The pellet was solubilized with 1%
SC
Triton X-100 in 10 mM HEPES pH 7.4 containing 2 mM EDTA, 2 mM EGTA, protease and phosphatase inhibitors I and II (Calbiochem; Temecula, CA) for 45 minutes at 37°C. Insoluble material was removed by centrifugation at 23,000 x g for 2 min at 4°C. The supernatant was
M AN U
stored at -80°C until use. For immunoprecipitation, 3 μg of antibody (GluA1, Abcam, #ab109450 or Millipore, #MAB2263; GluA2, NeuroMab, #75-002; puromycin, Millipore, #MABE343 or Kerafast, #3RH11) or an equal amount of control IgG was incubated with 20 μL of A/G agarose slurry (Pierce, Rockford, IL) for 4 h at 4°C. The pellet was collected by centrifugation at 3000 x g
TE D
for 1 min and washed 3 times with TBS plus 0.1% Tween 20. 100 μL of membrane prep was incubated with the washed pellet overnight at 4°C. The agarose bound antibody was pelleted by centrifugation at 3000 x g for 1 min. This created two fractions, the bound (pellet), and
EP
unbound (supernatant). The unbound fraction was then subjected to another round of immunoprecipitation. Two rounds of immunoprecipitation pulled down >95% of the target. After
AC C
the final immunoprecipitation, the unbound fraction was mixed with an equal volume of sample treatment buffer (Invitrogen, Carlesbad, CA) and heated to 100°C for 4 min. For Western analysis, samples were run on a 4-12% Bis-Tris gel (Invitrogen) and transferred to PVDF membranes for GluA1 and GluA2 immunoblotting. Band densities were determined using ImageQuant software.
5
ACCEPTED MANUSCRIPT Stefanik et al.
Supplement
Statistical analysis For rats in metabolic labeling and puromycin experiments (Figures 1-3), the NAc from one hemisphere was used for the control group, the NAc from the other hemisphere was used for
RI PT
the experimental group, and data were analyzed using a paired t-test. For puromycin immunoprecipitation experiments and qPCR experiments (Figures 4-6), both hemispheres were pooled in order to yield enough sample to immunoprecipitate or quantify mRNA for GluA1 and
SC
GluA2. These samples were analyzed either with an unpaired t-test or ANOVA with Tukey’s
AC C
EP
TE D
M AN U
post-hoc test, as appropriate.
6
ACCEPTED MANUSCRIPT Supplement
RI PT
Stefanik et al.
Supplementary Figure S1. Schematic of the experimental timeline for
35
S-Met/Cys
labeling. Freshly dissected NAc tissue was allowed 3 hours of post-dissection recovery in net-
then transferred to a different net-well containing
35
SC
wells containing oxygenated aCSF, based on data shown in Supplemental Figure 2. Tissue was S-Met/Cys for 60 minutes. The transcription
inhibitor actinomycin D (25µM) was present during the last 30 min of the aCSF incubation and 35
S-Met/Cys incubation. Tissue was then processed for SDS-PAGE,
M AN U
throughout the
immunoblotting and autoradiography as described in the Methods and Materials section of the
AC C
EP
TE D
main text and in Supplemental Methods (above).
7
ACCEPTED MANUSCRIPT Supplement
M AN U
SC
RI PT
Stefanik et al.
Supplementary Figure S2. Three hours of recovery after NAc tissue preparation yields maximal and stable
35
S-Met/Cys labeling. Freshly dissected NAc tissue was incubated for
varying times (1-4 hours) in oxygenated aCSF, with addition of the transcription inhibitor actinomycin D for the last 30 min of the incubation. Tissue was then processed for SDS-PAGE, immunoblotting and autoradiography as described in the Methods and Materials section of the
TE D
main text and in Supplemental Methods (above). Samples were run in triplicate. Optical density of the entire lane was measured for each sample, and results normalized to total protein in the lane measured with Ponceau S staining. Data are presented as mean ± SEM, with the 4 hour
AC C
EP
value set as 100%.
8
ACCEPTED MANUSCRIPT Supplement
Supplementary Figure S3. Incubation of NAc tissue with
35
RI PT
Stefanik et al.
S-Met/Cys does not result in
sufficient incorporation of isotope to quantify newly translated GluA1 protein. NAc tissue was dissected from rats that self-administered saline or cocaine (6 h/day for 10 days) and then S-Met/Cys (10 μCi/mL) as
35
SC
underwent >40 days of withdrawal. Tissue was incubated with
described in the Methods and Materials section of the main text in order to tag newly translated
M AN U
proteins. Then, GluA1 was immunoprecipitated as described in Supplemental Methods. Near quantitative immunoprecipitation of GluA1 was achieved, i.e., GluA1 levels in the bound fraction were ~95% of those detected in an equal amount of starting material (data not shown). Immunoprecipitated material was processed for SDS-PAGE and immunoblotted for GluA1 (bottom
images,
labeled
“Total
GluA1
chemiluminescence”).
Specificity
of
the
immunoprecipitation was demonstrated by comparing tissue from saline and cocaine rats immunoprecipitated with GluA1 antibody (6 lanes on the left) to tissue immunoprecipitated with
TE D
control IgG (right-hand lane). The same immunoblot was exposed to a phosphorimager screen for >3 weeks, with the goal of detecting newly translated GluA1 (top images, labeled “35S autoradiography”). Despite robust GluA1 levels in all lanes as revealed by immunoblotting, autoradiography failed to detect a quantifiable band at the molecular weight where GluA1
EP
should migrate, even after exposure to a phosphorimager screen for >3 weeks, although by eye we could observe an extremely faint band at ~100kD. These results indicate that only a very 35
S-Met/Cys labeling
AC C
small portion of the GluA1 pool is newly translated, as expected, and that
therefore does not result in sufficient isotope incorporation into GluA1 to compare rates of GluA1 translation between experimental groups. In parallel experiments not shown, NAc tissue was labeled with puromycin, GluA1 was immunoprecipitated, and the immunoprecipitated material was immunoblotted with puromycin antibody to detect newly translated GluA1. Here we failed to detect any signal in the immunoblot. We speculate that a faint band was detectable after 35S-labeling but not after puromycin labeling because puromycin labeling likely reduces the efficiency of the GluA1 pull-down by modifying the protein whereas incorporation of
35
S into the
GluA1 protein does not. Regardless of the labeling method, however, results demonstrate that pull-down with GluA1 antibody followed by autoradiography (35S) or immunoblotting (puromycin)
9
ACCEPTED MANUSCRIPT Stefanik et al.
Supplement
does not yield a quantifiable measure of newly translated GluA1 protein. Therefore, for experiments shown in Fig. 4 of the main text, we took the different strategy of labeling with puromycin, immunoprecipitating all newly translated proteins with puromycin, and then
AC C
EP
TE D
M AN U
SC
RI PT
immunoblotting for GluA1 (or GluA2).
10
ACCEPTED MANUSCRIPT Stefanik et al.
Supplement
Protein
Gene
TaqMan probea
GluA1
gria1
Rn00709588_m1
GluA2
gria2
Rn00568514_m1
gria3
Rn00583547_m1
GAPDH
Gapdh
Rn01775763_g1
Actin
Actb
Rn00667869_m1
GluA3
b
a
RI PT
Supplementary Table S1: List of qPCR probes and primers
TaqMan catalog number for ordering from Applied Biosystems. bSimilar to results described in
SC
the main text for Gria2, Gria3 mRNA levels in homogenates prepared from NAc core or shell did not differ between saline controls and cocaine “incubated” rats (core: t16=0.06, p=0.95; shell: t18=0.14, p=0.89), whereas a trend towards a reduction in the cocaine group was found in
AC C
EP
TE D
M AN U
synaptoneurosomes prepared from the entire NAc (t20=1.60, p=0.13).
11
ACCEPTED MANUSCRIPT Stefanik et al.
Supplement
Supplemental References Loweth JA, Scheyer AF, Milovanovic M, LaCrosse AL, Flores-Barrera E, Werner CT, et al. (2014): Synaptic depression via mGluR1 positive allosteric modulation suppresses cueinduced cocaine craving. Nature neuroscience. 17:73-80.
2.
Osterweil EK, Krueger DD, Reinhold K, Bear MF (2010): Hypersensitivity to mGluR5 and ERK1/2 leads to excessive protein synthesis in the hippocampus of a mouse model of fragile X syndrome. The Journal of neuroscience : the official journal of the Society for Neuroscience. 30:15616-15627.
3.
Sutton MA, Ito HT, Cressy P, Kempf C, Woo JC, Schuman EM (2006): Miniature neurotransmission stabilizes synaptic function via tonic suppression of local dendritic protein synthesis. Cell. 125:785-799.
4.
Scheyer AF, Wolf ME, Tseng KY (2014): A protein synthesis-dependent mechanism sustains calcium-permeable AMPA receptor transmission in nucleus accumbens synapses during withdrawal from cocaine self-administration. The Journal of neuroscience : the official journal of the Society for Neuroscience. 34:3095-3100.
5.
Most D, Leiter C, Blednov YA, Harris RA, Mayfield RD (2016): Synaptic microRNAs Coordinately Regulate Synaptic mRNAs: Perturbation by Chronic Alcohol Consumption. Neuropsychopharmacology : official publication of the American College of Neuropsychopharmacology. 41:538-548.
6.
Ferrario CR, Loweth JA, Milovanovic M, Wang X, Wolf ME (2011): Distribution of AMPA receptor subunits and TARPs in synaptic and extrasynaptic membranes of the adult rat nucleus accumbens. Neuroscience letters. 490:180-184.
AC C
EP
TE D
M AN U
SC
RI PT
1.
12