Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

An epigenetic mechanism mediates developmental nicotine effects on neuronal structure and behavior

This article has been updated

Abstract

Developmental nicotine exposure causes persistent changes in cortical neuron morphology and in behavior. We used microarray screening to identify master transcriptional or epigenetic regulators mediating these effects of nicotine and discovered increases in Ash2l mRNA, encoding a component of a histone methyltransferase complex. We therefore examined genome-wide changes in trimethylation of histone H3 on Lys4 (H3K4me3), a mark induced by the Ash2l complex associated with increased gene transcription. A large proportion of regulated promoter sites were involved in synapse maintenance. We found that Mef2c interacts with Ash2l and mediates changes in H3K4me3. Knockdown of Ash2l or Mef2c abolished nicotine-mediated alterations of dendritic complexity in vitro and in vivo, and attenuated nicotine-dependent changes in passive avoidance behavior. In contrast, overexpression mimicked nicotine-mediated alterations of neuronal structure and passive avoidance behavior. These studies identify Ash2l as a target induced by nicotinic stimulation that couples developmental nicotine exposure to changes in brain epigenetic marks, neuronal structure and behavior.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Morphological changes in cortical neurons induced by developmental nicotine exposure.
Figure 2: Nicotine exposure during development alters expression of a gene involved in histone methylation.
Figure 3: Differential enrichment of H3K4me3 at promoter sites associated with synapse function following developmental nicotine exposure.
Figure 4: The Ash2l–Mef2c complex is regulated by nAChR activity.
Figure 5: The Ash2l–Mef2c complex mediates dendritic remodeling by nicotine in neural progenitor cells and in vivo.
Figure 6: Developmental nicotine-induced changes in passive avoidance behavior require expression of Ash2l and Mef2c.

Similar content being viewed by others

Accession codes

Primary accessions

Gene Expression Omnibus

Referenced accessions

NCBI Reference Sequence

Change history

  • 20 June 2016

    In the version of this article initially published online, the affiliations of Qiaoping Yuan and David Goldman were switched with those of Angelique Bordey. The error has been corrected for the print, PDF and HTML versions of this article.

References

  1. Lozada, A.F. et al. Induction of dendritic spines by β2-containing nicotinic receptors. J. Neurosci. 32, 8391–8400 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Ballesteros-Yáñez, I., Benavides-Piccione, R., Bourgeois, J.P., Changeux, J.P. & DeFelipe, J. Alterations of cortical pyramidal neurons in mice lacking high-affinity nicotinic receptors. Proc. Natl. Acad. Sci. USA 107, 11567–11572 (2010).

    PubMed  PubMed Central  Google Scholar 

  3. Muhammad, A. et al. Prenatal nicotine exposure alters neuroanatomical organization of the developing brain. Synapse 66, 950–954 (2012).

    CAS  PubMed  Google Scholar 

  4. Mychasiuk, R., Muhammad, A., Gibb, R. & Kolb, B. Long-term alterations to dendritic morphology and spine density associated with prenatal exposure to nicotine. Brain Res. 1499, 53–60 (2013).

    CAS  PubMed  Google Scholar 

  5. Sorenson, C.A., Raskin, L.A. & Suh, Y. The effects of prenatal nicotine on radial-arm maze performance in rats. Pharmacol. Biochem. Behav. 40, 991–993 (1991).

    CAS  PubMed  Google Scholar 

  6. Sobrian, S.K., Marr, L. & Ressman, K. Prenatal cocaine and/or nicotine exposure produces depression and anxiety in aging rats. Prog. Neuropsychopharmacol. Biol. Psychiatry 27, 501–518 (2003).

    CAS  PubMed  Google Scholar 

  7. Heath, C.J. & Picciotto, M.R. Nicotine-induced plasticity during development: modulation of the cholinergic system and long-term consequences for circuits involved in attention and sensory processing. Neuropharmacology 56 (suppl. 1), 254–262 (2009).

    CAS  PubMed  Google Scholar 

  8. Jacobsen, L.K., Picciotto, M.R., Heath, C.J., Mencl, W.E. & Gelernter, J. Allelic variation of calsyntenin 2 (CLSTN2) modulates the impact of developmental tobacco smoke exposure on mnemonic processing in adolescents. Biol. Psychiatry 65, 671–679 (2009).

    CAS  PubMed  Google Scholar 

  9. Pauly, J.R., Sparks, J.A., Hauser, K.F. & Pauly, T.H. In utero nicotine exposure causes persistent, gender-dependant changes in locomotor activity and sensitivity to nicotine in C57Bl/6 mice. Int. J. Dev. Neurosci. 22, 329–337 (2004).

    CAS  PubMed  Google Scholar 

  10. Heath, C.J., Horst, N.K. & Picciotto, M.R. Oral nicotine consumption does not affect maternal care or early development in mice but results in modest hyperactivity in adolescence. Physiol. Behav. 101, 764–769 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Dwyer, J.B., McQuown, S.C. & Leslie, F.M. The dynamic effects of nicotine on the developing brain. Pharmacol. Ther. 122, 125–139 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Rampalli, S. et al. p38 MAPK signaling regulates recruitment of Ash2L-containing methyltransferase complexes to specific genes during differentiation. Nat. Struct. Mol. Biol. 14, 1150–1156 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Wysocka, J., Allis, C.D. & Coonrod, S. Histone arginine methylation and its dynamic regulation. Front. Biosci. 11, 344–355 (2006).

    CAS  PubMed  Google Scholar 

  14. Barbosa, A.C. et al. MEF2C, a transcription factor that facilitates learning and memory by negative regulation of synapse numbers and function. Proc. Natl. Acad. Sci. USA 105, 9391–9396 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Cho, E.G. et al. MEF2C enhances dopaminergic neuron differentiation of human embryonic stem cells in a parkinsonian rat model. PLoS One 6, e24027 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Wan, M. et al. The trithorax group protein Ash2l is essential for pluripotency and maintaining open chromatin in embryonic stem cells. J. Biol. Chem. 288, 5039–5048 (2013).

    CAS  PubMed  Google Scholar 

  17. Flavell, S.W. et al. Genome-wide analysis of MEF2 transcriptional program reveals synaptic target genes and neuronal activity-dependent polyadenylation site selection. Neuron 60, 1022–1038 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Gossett, L.A., Kelvin, D.J., Sternberg, E.A. & Olson, E.N. A new myocyte-specific enhancer-binding factor that recognizes a conserved element associated with multiple muscle-specific genes. Mol. Cell. Biol. 9, 5022–5033 (1989).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Potthoff, M.J. & Olson, E.N. MEF2: a central regulator of diverse developmental programs. Development 134, 4131–4140 (2007).

    CAS  PubMed  Google Scholar 

  20. Zweier, M. et al. Cornelia Kraus. Mutations in MEF2C from the 5q14.3q15 microdeletion syndrome region are a frequent cause of severe mental retardation and diminish MECP2 and CDKL5 expression. Hum. Mutat. 31, 722–733 (2010).

    CAS  PubMed  Google Scholar 

  21. Li, H. et al. Transcription factor MEF2C influences neural stem/progenitor cell differentiation and maturation in vivo. Proc. Natl. Acad. Sci. USA 105, 9397–9402 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Sakai, Y. et al. Neuroendocrine phenotypes in a boy with 5q14 deletion syndrome implicate the regulatory roles of myocyte-specific enhancer factor 2C in the postnatal hypothalamus. Eur. J. Med. Genet. 56, 475–483 (2013).

    PubMed  Google Scholar 

  23. King, S.L. et al. Conditional expression in corticothalamic efferents reveals a developmental role for nicotinic acetylcholine receptors in modulation of passive avoidance behavior. J. Neurosci. 23, 3837–3843 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Heath, C.J., King, S.L., Gotti, C., Marks, M.J. & Picciotto, M.R. Cortico-thalamic connectivity is vulnerable to nicotine exposure during early postnatal development through α4/β2/α5 nicotinic acetylcholine receptors. Neuropsychopharmacology 35, 2324–2338 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Mazo, A.M., Huang, D.H., Mozer, B.A. & Dawid, I.B. The trithorax gene, a trans-acting regulator of the bithorax complex in Drosophila, encodes a protein with zinc-binding domains. Proc. Natl. Acad. Sci. USA 87, 2112–2116 (1990).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Paro, R. Imprinting a determined state into the chromatin of Drosophila. Trends Genet. 6, 416–421 (1990).

    CAS  PubMed  Google Scholar 

  27. Ikegawa, S., Isomura, M., Koshizuka, Y. & Nakamura, Y. Cloning and characterization of ASH2L and Ash2l, human and mouse homologs of the Drosophila ash2 gene. Cytogenet. Cell Genet. 84, 167–172 (1999).

    CAS  PubMed  Google Scholar 

  28. Dou, Y. et al. Regulation of MLL1 H3K4 methyltransferase activity by its core components. Nat. Struct. Mol. Biol. 13, 713–719 (2006).

    CAS  PubMed  Google Scholar 

  29. Steward, M.M. et al. Molecular regulation of H3K4 trimethylation by ASH2L, a shared subunit of MLL complexes. Nat. Struct. Mol. Biol. 13, 852–854 (2006).

    CAS  PubMed  Google Scholar 

  30. Mackowiak, M., Bator, E., Latusz, J., Mordalska, P. & Wedzony, K. Prenatal MAM administration affects histone H3 methylation in postnatal life in the rat medial prefrontal cortex. Eur. Neuropsychopharmacol. 24, 271–289 (2014).

    CAS  PubMed  Google Scholar 

  31. Cheng, M.K. & Shearn, A. The direct interaction between ASH2, a Drosophila trithorax group protein, and SKTL, a nuclear phosphatidylinositol 4-phosphate 5-kinase, implies a role for phosphatidylinositol 4,5-bisphosphate in maintaining transcriptionally active chromatin. Genetics 167, 1213–1223 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Uhlén, M. et al. Proteomics. Tissue-based map of the human proteome. Science 347, 1260419 (2015).

    PubMed  Google Scholar 

  33. Nadarajah, B. & Parnavelas, J.G. Modes of neuronal migration in the developing cerebral cortex. Nat. Rev. Neurosci. 3, 423–432 (2002).

    CAS  PubMed  Google Scholar 

  34. Goldberg, A.D., Allis, C.D. & Bernstein, E. Epigenetics: a landscape takes shape. Cell 128, 635–638 (2007).

    CAS  PubMed  Google Scholar 

  35. Barski, A. et al. High-resolution profiling of histone methylations in the human genome. Cell 129, 823–837 (2007).

    CAS  PubMed  Google Scholar 

  36. Lauberth, S.M. et al. H3K4me3 interactions with TAF3 regulate preinitiation complex assembly and selective gene activation. Cell 152, 1021–1036 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Tan, C.C. et al. Transcription factor Ap2delta associates with Ash2l and ALR, a trithorax family histone methyltransferase, to activate Hoxc8 transcription. Proc. Natl. Acad. Sci. USA 105, 7472–7477 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Lyons, G.E., Micales, B.K., Schwarz, J., Martin, J.F. & Olson, E.N. Expression of mef2 genes in the mouse central nervous system suggests a role in neuronal maturation. J. Neurosci. 15, 5727–5738 (1995).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Flavell, S.W. et al. Activity-dependent regulation of MEF2 transcription factors suppresses excitatory synapse number. Science 311, 1008–1012 (2006).

    CAS  PubMed  Google Scholar 

  40. Shalizi, A. et al. A calcium-regulated MEF2 sumoylation switch controls postsynaptic differentiation. Science 311, 1012–1017 (2006).

    CAS  PubMed  Google Scholar 

  41. Krmpotic´-Nemanic´, J., Kostovic´ć, I., Kelovic´, Z. & Nemanic´, D. Development of acetylcholinesterase (AChE) staining in human fetal auditory cortex. Acta Otolaryngol. (Stockh.) 89, 388–392 (1980).

    Google Scholar 

  42. Krmpotic´-Nemanic´, J., Kostovic´, I., Kelovic´, Z., Nemanic´, D. & Mrzljak, L. Development of the human fetal auditory cortex: growth of afferent fibres. Acta Anat. 116, 69–73 (1983).

    PubMed  Google Scholar 

  43. Franklin, K.B.J. & Paxinos, G. The Mouse Brain in Stereotactic Coordinates (Academic, 1997).

  44. Duque, A. et al. Neuroanatomical changes in a mouse model of early life neglect. Brain Struct. Funct. 217, 459–472 (2012).

    PubMed  Google Scholar 

  45. Zhang, Y. et al. Model-based analysis of ChIP-Seq (MACS). Genome Biol. 9, R137 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  46. Robinson, M.D., McCarthy, D.J. & Smyth, G.K. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics 26, 139–140 (2010).

    CAS  PubMed  Google Scholar 

  47. Hutton, S.R. & Pevny, L.H. Isolation, culture, and differentiation of progenitor cells from the central nervous system. CSH Protoc. http://dx.doi.org/10.1101/pdb.prot5077 (2008).

  48. Hommel, J.D., Sears, R.M., Georgescu, D., Simmons, D.L. & DiLeone, R.J. Local gene knockdown in the brain using viral-mediated RNA interference. Nat. Med. 9, 1539–1544 (2003).

    CAS  PubMed  Google Scholar 

  49. Mineur, Y.S. et al. Nicotine decreases food intake through activation of POMC neurons. Science 332, 1330–1332 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Mineur, Y.S. et al. Cholinergic signaling in the hippocampus regulates social stress resilience and anxiety- and depression-like behavior. Proc. Natl. Acad. Sci. USA 110, 3573–3578 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This work was supported by grants DA14241 (M.R.P.), DA10455 (M.R.P.), NS052519 (F.H.) and NS086329 (F.H.) from the National Institutes of Health, the State of Connecticut, Department of Mental Health and Addiction Services and the Kavli Institute for Neuroscience at Yale.

Author information

Authors and Affiliations

Authors

Contributions

Y.J. designed the study, performed the experiments and data analyses, and wrote and edited the manuscript. Z.Z. and Q.Y. prepared samples and analyzed data for the ChIP-seq experiments. D.G. provided critical resources, experimental design and advice on the ChIP-seq experiments. D.C. performed the imaging studies and data analyses for the diffusion tensor imaging experiments. F.H. provided valuable resources and critical advice on interpretation of the diffusion tensor imaging experiments. Y.S.M. designed shRNA constructs, helped troubleshoot knockdown studies and provided critical advice on behavioral design. A.B. provided critical resources, experimental design and advice on the in utero electroporation studies. L.S.H. performed in utero electroporation studies. A.M.L. performed in utero electroporation surgeries and contributed to data collection. C.J.H. designed and prepared samples for the diffusion tensor imaging experiment and contributed to the microarray experiment. M.R.P. designed the project, assisted in interpretation of all studies, and wrote and edited the manuscript. All authors contributed to editing of the manuscript.

Corresponding author

Correspondence to Marina R Picciotto.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Effects of nicotine exposure from birth to 3 weeks of age on dendritic complexity across cortical regions assessed at 3 months of age.

(a-c) Sholl analysis of the apical dendritic tree in frontal (F), parietal (P) and occipital (O) regions of cortex. Exposure to nicotine in the postnatal period significantly increased dendritic complexity across cortical regions. (a) F(1,80)=6.901, p=0.010321. (b) F(1,30)=16.525; p=0.0000001. (c) F(1,65)=32.586, p=0.000319. (d-f) Layer-specific effects of postnatal-only nicotine exposure on dendritic complexity in superficial (1/2), intermediate (3/4) and deep (5/6) layers of cortex. (d) F(1,20) = 5.625, p=0.027854. (e) F(1,46)=13.719, p=0.000565. (f) F(1, 23)=10.746, p=0.003299. *, p < 0.05. F: Sac, n=28; Nic, n=54; P: Sac, n=36; Nic, n=31; O: Sac, n=20; Nic, n=12; 1/2: Sac, n=9; Nic, n=13; 3/4: Sac, n=18, Nic, n=30; 5/6: Sac, n=10. Nic, n=15.

Supplementary Figure 2 Volcano plot showing all probe sets evaluated in the microarray study.

(a) Genes whose expression levels were significantly different between developmentally nicotine exposed animals and controls are shown as red dots. (b) 6 of 15 probe sets were significantly altered in independent samples at 3 months of age following nicotine treatment throughout the pre- and postnatal period compared to the control group: F(1,8)=7.77 for Ash2l, p=0.02365035; F(1,8)=6.797 for Chsy3, p=0.03127061; F(1,8)=0.64 for Zfp91, p=0.44681333; F(1,8)=0.41 for Cflar, p=0.53987164; F(1,8)=26.538 for Zcchc11, p=0.00087271; F(1,8)=5.18 for Cep192, p=0.05240060; F(1,8)=1.24 for Alkbh1, p=0.29780695; F(1,8)=0.72 for Gmeb1, p=0.42080639; F(1,8)=0.36 for Unc13b, p=0.56511006; F(1,8)=0.035 for Duox1, p=0.85625319; F(1,8)=8.82 for Scula2, p=0.01787528; F(1,8)=1.84 for Zfp597, p=0.21198878; F(1,8)=1.18 for Ctnnal1, p=0.30899997; F(1,8)=0.65 for Ntrk, p=0.44341660; F(1,8)=7.397 for Tmem107, p=0.02625917. (*p <0.05, ** p <0.01 with Sidak’s test; # p <0.05 with LSD test for multiple comparisons). 2 to 4 animals were pooled for each biological replicate. Five Biological replicates were used for each condition (Sac: n = 5; Nic: n = 5). A total of 26 animals were used (Sac: n = 14;Nic: n = 12).

Supplementary Figure 3 Changes in H3K4me3 associated with the promoter sites of multiple gene loci following developmental nicotine exposure.

(a) Gene ontology (GO) analysis identified significantly regulated gene groups, all of which are related to glutamatergic synaptic function. (b) Gene structure and coordination of Ank1 loci associated with H3K4me3 depicted in Fig. 3c. (c) Whisker plot showing verification of changes in histone H3Me3K4 levels associated with gene loci identified in the ChIP-seq analysis by ChIP-PCR in independent samples from subjects treated with nicotine from birth to 21 days (postnatal-only) and evaluated at 3 months of age: F(1,8)= 11.02for Eif4a, p= 0.01054575; F(1,8)= 6.028 for Izumo1, p= 0.03961426; F(1,8)= 1.406 for Gpr19, p= 0.26974334; F(1,8)= 21.7 for Litaf, p= 0.00162698; F(1,8)= 15.839 for kcnq1, p= 0.00406270; F(1,8)= 19.147 for Lage3, p= 0.00236227; F(1,8)=17.512 for Fbxw4, p= 0.00305973; F(1,8)= 16.594 for Fgf12, p= 0.00356573; F(1,8)= 77.18 for Sepsecs, p= 0.00002212; F(1,8)= 9.979 for Rin2, p= 0.01341607; F(1,8)= 64.634 for Rabgap1l, p= 0.00004215; F(1,8)= 81.61 for Ano2, p= 0.00001803; F(1,8)= 17.505 for Apool, p= 0.00306323; F(1,8)= 99.602for Lipc, p= 0.00000862; F(1,8)= 45.502 for Cdk5rap2, p= 0.00014572; F(1,8)= 565.869 for Ing4, p= 0.00000001; F(1,8)= 118.256 for Ank3, p= 0.00000452; F(1,8)= 43.276 for Ntm, p= 0.00017324; F(1,8)= 487.52 for Zfp658, p= 0.00000002; F(1,8)= 8.77 for Ybx3, p= 0.01810693; F(1,7)= 41.57 for Sorcs1, p= 0.00019884; F(1,8)=153.564 for Lars2, p= 0.00000168; F(1,8)= 50.843 for Ank1, p= 0.00009898; F(1,8)= 242.093 for Acacb, p= 0.00000029; F(1,8)= 223.283 for Mdga2, p= 0.00000040; F(1,8)= 110.084 for Chl1, p= 0.00000592; F(1,8)= 23.981 for Auts2, p= 0.00119824; F(1,8)= 0.483 for Mbnl1, p= 0.50674655; F(1,8)=10.846 for Cpeb1, p= 0.01096747; F(1,8)= 15.121 for Zfp65, p= 0.00461852; F(1,8)= 5.928 for Chd9, p= 0.04089817; F(1,8)= 24.034 for Syt4, p= 0.00119008; F(1,8)=7.836 for Sp110, p= 0.02322340; F(1,8)= 51.276 for Sorbs2, p= 0.00009607; F(1,8)= 24.889 for Slc35a2, p= 0.00106755; F(1,8)= 19.501 for Mef2c, p= 0.00223843 (*p <0.05 with LSD test for multiple comparisons). Each replicate was a pool of 2-4 brain samples and 5 replicates were used for each condition (Sac: n = 5 pools from 14 animals; Nic: n = 5 pools from 12 animals). (d) Ash2l and Mef2c binding sites overlap with sites of H3K4me3 enrichment. Among these, GO analysis of 108 genomic sites identified as differentially enriched following nicotine exposure reveals that overlapped genomic sites are associated with synapse related functions.

Supplementary Figure 4 Regulation of Mef2c locus by nicotine treatment in vivo.

(a) Mef2c mRNA levels were significantly elevated at 21 days of age, immediately after nicotine exposure was completed F(1,8)=19.237, p=0.00232999. 5 Biological replicates per each condition from pooled female animals; Nic = 11 Sac = 15. (b) Histone H3 acetylation associated with the Mef2c locus was significantly increased as a result of nicotine exposure during development F(1,8)=118.802, p=0.00000445. 5 Biological replicates per each condition from pooled female animals; Nic = 12 Sac = 14. (c) Nicotine exposure during development significantly increased the level of H3K4me3 associated with the Mef2c locus (F(5,24)=10.403, p=0.000021) with Tukey's multiple comparison test. Frontal sac vs Frontal nic, p=0.001277; Parietal sac vs Parietal nic, p=0.005846; Occipital sac vs Occipital nic, p=0.011952. 5 Biological replicates per each condition from pooled female animals; Nic = 12 Sac = 12. *p < 0.05, *** p <0.01 with Sidak’s test.

Supplementary Figure 5 Evaluation of shRNA-mediated knock down of Ash2l and Mef2c protein levels in neural progenitor cells.

a) shRNA targeting Ash2l. b) shRNA targeting Mef2c. Original Western blots presented in supplementary Figure 7.

Supplementary Figure 6 Spread following in utero electroporation.

Example of the extent of shRNA spread and of GFP expression in a layer 6 cortical pyramidal neuron following in utero electroporation of shRNAs.

Supplementary Figure 7 Original images of representative western blot images in Figure 4 and Supplementary Figure 5.

(a-d) indicate uncropped LICOR machine scanned gel image with annotation. (e) Scanned film image of Figure 4 immunopreciptiation experiment. (f-g) Nicotine induced Wdr5 and Rbbp5 expression blot: original scanned image from LICOR machine. (i-j) Scanned images from LICOR machine for shRNA knockdown efficiency experiment presented in supplementary Figure 5.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–7 and Supplementary Tables 1–5 (PDF 1165 kb)

Supplementary Methods Checklist (PDF 277 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Jung, Y., Hsieh, L., Lee, A. et al. An epigenetic mechanism mediates developmental nicotine effects on neuronal structure and behavior. Nat Neurosci 19, 905–914 (2016). https://doi.org/10.1038/nn.4315

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.4315

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing