MeCP2 mutation results in compartment-specific reductions in dendritic branching and spine density in layer 5 motor cortical neurons of YFP-H mice

David P Stuss, Jamie D Boyd, David B Levin, Kerry R Delaney, David P Stuss, Jamie D Boyd, David B Levin, Kerry R Delaney

Abstract

Rett Syndrome (RTT) is a neurodevelopmental disorder predominantly caused by mutations in the X-linked gene MECP2. A primary feature of the syndrome is the impaired maturation and maintenance of excitatory synapses in the central nervous system (CNS). Different RTT mouse models have shown that particular Mecp2 mutations have highly variable effects on neuronal architecture. Distinguishing MeCP2 mutant cellular phenotypes therefore demands analysis of specific mutations in well-defined neuronal subpopulations. We examined a transgenically labeled subset of cortical neurons in YFP-H mice crossed with the Mecp2(tm1.1Jae) mutant line. YFP(+) Layer 5 pyramidal neurons in the motor cortex of wildtype and hemizygous mutant male mice were examined for differences in dendrite morphology and spine density. Total basal dendritic length was decreased by 18.6% due to both shorter dendrites and reduced branching proximal to the soma. Tangential dendrite lengths in the apical tuft were reduced by up to 26.6%. Spine density was reduced by 47.4% in the apical tuft and 54.5% in secondary apical dendrites, but remained unaffected in primary apical and proximal basal dendrites. We also found that MeCP2 mutation reduced the number of YFP(+) cells in YFP-H mice by up to 72% in various cortical regions without affecting the intensity of YFP expression in individual cells. Our results support the view that the effects of MeCP2 mutation are highly context-dependent and cannot be generalized across mutation types and cell populations.

Conflict of interest statement

Competing Interests: The authors have declared that no competing interests exist.

Figures

Figure 1. 3D Sholl analysis of apical…
Figure 1. 3D Sholl analysis of apical and basal arbors in YFP+ L5 motor cortical neurons.
A) Confocal microscopic image stacks from 200 µm tissue sections showing YFP+ dendrites in apical (upper) and basal (lower) compartments (WT male mouse). Scale bar = 50 µm. B) Representative 3D traces from a WT neuron. 10 µm Sholl radii were used in the basal compartment and 20 µm radii in the apical compartment. The pial surface is shown as a dotted line with the apical Sholl origin defined as a point 300 µm from the pial surface. Depth in the z-axis is color-coded.
Figure 2. Dendritic branching and length is…
Figure 2. Dendritic branching and length is reduced in YFP+ MeCP2J mutant neurons.
A, B) Reconstructions of representative WT and mutant YFP+ L5 pyramidal neurons based on the Sholl analysis branch patterns in C1 and D1. Scale bar (A) = 50 µm. C1) Basal compartment Sholl analysis showing the number of dendrite crossings as a function of distance from the soma. Mutant YFP+ neurons have significantly fewer dendritic branches 40–70 µm from the cell body. D1) Apical compartment Sholl analysis of the summed dendritic length per Sholl radius. Mutant YFP+ neurons have less total dendritic length per Sholl radius in L1–2. Dendrite lengths in radii greater than 300 µm correspond to long, laterally extending dendrites that are near and parallel to the pial surface. C2, D2) Percent dendritic length as a function of branch order in basal and apical compartments. The inset in (D2) shows branch order hierarchy. In basal dendrites only, mutant YFP+ neurons exhibit significantly reduced dendritic complexity with proportionally more dendritic length in lower branch orders. C3) A cumulative total length plot for basal dendrites shows that the main divergence in total dendritic length occurs less than ∼120 µm from the soma. The average maximum dendritic length is also reduced in mutant YFP+ neurons. D3) A cumulative total dendritic length plot of apical tuft branches shows no minimal divergence except close to the pial surface, where mutant YFP+ cells extend fewer lateral branches.
Figure 3. Spine density is variably affected…
Figure 3. Spine density is variably affected in different dendritic compartments in YFP+ MeCP2J mutant L5 pyramidal neurons.
A–D) Spines were quantified in 20–30 µm segments in four dendritic compartments for WT and mutant YFP+ neurons. Scale bar (A) = 5 µm (all images). A) Apical tuft dendrites were sampled <60 µm from the pial surface. B) Primary apical dendrites were imaged between 100–200 µm from the soma. C) Secondary apical dendrites originating in the same region as (B) were sampled beginning at the branch point from the primary apical dendrite. D) Basal dendrite spine counts were initiated following the first dendritic bifurcation. E) Spine density is reduced by 47.4% in the apical tuft and by 54.5% in secondary apical dendrites.
Figure 4. MeCP2J mutant mice have reduced…
Figure 4. MeCP2J mutant mice have reduced numbers of YFP-expressing cortical neurons.
A, C) Fewer YFP+ cells are observed across the entire cortex in mutant males (200 µm coronal sections, ∼ Bregma −2.2 mm in WT). Some cortical regions are more dramatically affected than others, such as visual cortex (red arrow) and auditory cortex (blue arrow). Sharp boundaries correspond to functional regions (Au, auditory; S, somatosensory; V, visual; RS, retrosplenial). Scale bar = 0.5 mm. Overall cortical YFP expression patterns are shown as a qualitative heat map in E), where white indicates few to no neurons and dark orange indicates the most dense cortical labeling. The number of YFP+ cells is generally higher in rostral and medial region in both genotypes. B) Fluorescent Nissl staining (red) of L5 motor cortex. A limited subset of pyramidal neurons express YFP. Scale bar = 100 µm. D) Neuronal density is increased in frontal (Fr), motor (M) and retrosplenial (RS) cortex in mutant mice. F) The percentage of YFP+ pyramidal neurons is reduced in three cortical regions (F, frontal; M, motor; RS, retrosplenial).
Figure 5. Developmental progression of cortical YFP…
Figure 5. Developmental progression of cortical YFP expression.
Left column: 200 µm coronal sections from YFP-H MeCP2J littermate pairs (∼ Bregma 0.14 mm in WT). The anterior commissure was used as a reference anatomical marker. Right column: maximum intensity projections of confocal fluorescence imaging z-stacks from cingulate cortex. The red box in P9 mutant (rotated 90° clockwise) corresponds to the region imaged in each brain. Single bright YFP+ L5 pyramidal neurons are observed in both genotypes as early as P9. By P12 YFP expression has been strongly upregulated in the WT, after which the number of YFP+ cells continues to increase more slowly until stabilizing to mature levels sometime between P21 and P40. The number of YFP+ cells in mutant males also continues to increase, but at a slower rate, and final levels remain below those seen in the WT. All confocal fluorescence images shown were obtained from the brain section shown except P9 WT, which was obtained from a tissue section 400 µm anterior due to sparse YFP labeling at this age. Left column scale bar = 0.5 mm. Right column scale bar = 100 µm.

References

    1. Amir RE, Van den Veyver IB, Wan M, Tran CQ, Francke U, et al. Rett syndrome is caused by mutations in X-linked MECP2, encoding methyl-CpG-binding protein 2. Nat Genet. 1999;23:185–188.
    1. Hagberg B. Clinical manifestations and stages of Rett syndrome. Ment Retard Dev Disabil Res Rev. 2002;8:61–65.
    1. Chahrour M, Zoghbi HY. The story of Rett syndrome: from clinic to neurobiology. Neuron. 2007;56:422–437.
    1. Asaka Y, Jugloff DG, Zhang L, Eubanks JH, Fitzsimonds RM. Hippocampal synaptic plasticity is impaired in the Mecp2-null mouse model of Rett syndrome. Neurobiol Dis. 2006;21:217–227.
    1. Chao HT, Zoghbi HY, Rosenmund C. MeCP2 controls excitatory synaptic strength by regulating glutamatergic synapse number. Neuron. 2007;56:58–65.
    1. Chapleau CA, Calfa GD, Lane MC, Albertson AJ, Larimore JL, et al. Dendritic spine pathologies in hippocampal pyramidal neurons from Rett syndrome brain and after expression of Rett-associated MECP2 mutations. Neurobiol Dis. 2009;35:219–233.
    1. Armstrong DD. Rett syndrome neuropathology review 2000. Brain Dev. 2001;23(Suppl 1):S72–S76.
    1. Carter JC, Lanham DC, Pham D, Bibat G, Naidu S, et al. Selective cerebral volume reduction in Rett syndrome: a multiple-approach MR imaging study. AJNR Am J Neuroradiol. 2008;29:436–441.
    1. Subramaniam B, Naidu S, Reiss AL. Neuroanatomy in Rett syndrome: cerebral cortex and posterior fossa. Neurology. 1997;48:399–407.
    1. Belichenko PV, Oldfors A, Hagberg B, Dahlstrom A. Rett syndrome: 3-D confocal microscopy of cortical pyramidal dendrites and afferents. Neuroreport. 1994;5:1509–1513.
    1. Armstrong D, Dunn JK, Antalffy B, Trivedi R. Selective dendritic alterations in the cortex of Rett syndrome. J Neuropathol Exp Neurol. 1995;54:195–201.
    1. Bauman ML, Kemper TL, Arin DM. Microscopic observations of the brain in Rett syndrome. Neuropediatrics. 1995;26:105–108.
    1. Bauman ML, Kemper TL, Arin DM. Pervasive neuroanatomic abnormalities of the brain in three cases of Rett's syndrome. Neurology. 1995;45:1581–1586.
    1. Belichenko PV, Hagberg B, Dahlstrom A. Morphological study of neocortical areas in Rett syndrome. Acta Neuropathol (Berl) 1997;93:50–61.
    1. Armstrong DD, Dunn K, Antalffy B. Decreased dendritic branching in frontal, motor and limbic cortex in Rett syndrome compared with trisomy 21. J Neuropathol Exp Neurol. 1998;57:1013–1017.
    1. Amir RE, Van den Veyver IB, Schultz R, Malicki DM, Tran CQ, et al. Influence of mutation type and X chromosome inactivation on Rett syndrome phenotypes. Ann Neurol. 2000;47:670–679.
    1. Schanen C, Houwink EJ, Dorrani N, Lane J, Everett R, et al. Phenotypic manifestations of MECP2 mutations in classical and atypical Rett syndrome. Am J Med Genet A. 2004;126:129–140.
    1. Charman T, Neilson TC, Mash V, Archer H, Gardiner MT, et al. Dimensional phenotypic analysis and functional categorisation of mutations reveal novel genotype-phenotype associations in Rett syndrome. Eur J Hum Genet. 2005;13:1121–1130.
    1. Neul JL, Fang P, Barrish J, Lane J, Caeg EB, et al. Specific mutations in methyl-CpG-binding protein 2 confer different severity in Rett syndrome. Neurology. 2008;70:1313–1321.
    1. Chen RZ, Akbarian S, Tudor M, Jaenisch R. Deficiency of methyl-CpG binding protein-2 in CNS neurons results in a Rett-like phenotype in mice. Nat Genet. 2001;27:327–331.
    1. Guy J, Hendrich B, Holmes M, Martin JE, Bird A. A mouse Mecp2-null mutation causes neurological symptoms that mimic Rett syndrome. Nat Genet. 2001;27:322–326.
    1. Shahbazian M, Young J, Yuva-Paylor L, Spencer C, Antalffy B, et al. Mice with truncated MeCP2 recapitulate many Rett syndrome features and display hyperacetylation of histone H3. Neuron. 2002;35:243–254.
    1. Pelka GJ, Watson CM, Radziewic T, Hayward M, Lahooti H, et al. Mecp2 deficiency is associated with learning and cognitive deficits and altered gene activity in the hippocampal region of mice. Brain. 2006;129:887–898.
    1. Jentarra GM, Olfers SL, Rice SG, Srivastava N, Homanics GE, et al. Abnormalities of cell packing density and dendritic complexity in the MeCP2 A140V mouse model of Rett syndrome/X-linked mental retardation. BMC Neurosci. 2010;11:19.
    1. Fukuda T, Itoh M, Ichikawa T, Washiyama K, Goto Y. Delayed maturation of neuronal architecture and synaptogenesis in cerebral cortex of Mecp2-deficient mice. J Neuropathol Exp Neurol. 2005;64:537–544.
    1. Moretti P, Levenson JM, Battaglia F, Atkinson R, Teague R, et al. Learning and memory and synaptic plasticity are impaired in a mouse model of Rett syndrome. J Neurosci. 2006;26:319–327.
    1. Smrt RD, Eaves-Egenes J, Barkho BZ, Santistevan NJ, Zhao C, et al. Mecp2 deficiency leads to delayed maturation and altered gene expression in hippocampal neurons. Neurobiol Dis. 2007;27:77–89.
    1. Kishi N, Macklis JD. MECP2 is progressively expressed in post-migratory neurons and is involved in neuronal maturation rather than cell fate decisions. Mol Cell Neurosci. 2004;27:306–321.
    1. Kishi N, Macklis JD. MeCP2 functions largely cell-autonomously, but also non-cell-autonomously, in neuronal maturation and dendritic arborization of cortical pyramidal neurons. Exp Neurol. 2010;222:51–58.
    1. Belichenko NP, Belichenko PV, Li HH, Mobley WC, Francke U. Comparative study of brain morphology in Mecp2 mutant mouse models of Rett syndrome. J Comp Neurol. 2008;508:184–195.
    1. Belichenko PV, Wright EE, Belichenko NP, Masliah E, Li HH, et al. Widespread changes in dendritic and axonal morphology in Mecp2-mutant mouse models of Rett syndrome: evidence for disruption of neuronal networks. J Comp Neurol. 2009;514:240–258.
    1. Belichenko NP, Belichenko PV, Mobley WC. Evidence for both neuronal cell autonomous and nonautonomous effects of methyl-CpG-binding protein 2 in the cerebral cortex of female mice with Mecp2 mutation. Neurobiol Dis. 2009;34:71–77.
    1. Degano AL, Pasterkamp RJ, Ronnett GV. MeCP2 deficiency disrupts axonal guidance, fasciculation, and targeting by altering Semaphorin 3F function. Mol Cell Neurosci. 2009;42:243–254.
    1. Tropea D, Giacometti E, Wilson NR, Beard C, McCurry C, et al. Partial reversal of Rett Syndrome-like symptoms in MeCP2 mutant mice. Proc Natl Acad Sci U S A. 2009;106:2029–2034.
    1. Percy AK, Zoghbi HY, Lewis KR, Jankovic J. Rett syndrome: qualitative and quantitative differentiation from autism. J Child Neurol. 1988;3(Suppl):S65–S67.
    1. Akemann W, Zhong YM, Ichinohe N, Rockland KS, Knöpfel T. Transgenic mice expressing a fluorescent in vivo label in a distinct subpopulation of neocortical layer 5 pyramidal cells. J Comp Neurol. 2004;480:72–88.
    1. Molyneaux BJ, Arlotta P, Fame RM, MacDonald JL, MacQuarrie KL, et al. Novel subtype-specific genes identify distinct subpopulations of callosal projection neurons. J Neurosci. 2009;29:12343–12354.
    1. Sugino K, Hempel CM, Miller MN, Hattox AM, Shapiro P, et al. Molecular taxonomy of major neuronal classes in the adult mouse forebrain. Nat Neurosci. 2006;9:99–107.
    1. Tsiola A, Hamzei-Sichani F, Peterlin Z, Yuste R. Quantitative morphologic classification of layer 5 neurons from mouse primary visual cortex. J Comp Neurol. 2003;461:415–428.
    1. Yu J, Anderson CT, Kiritani T, Sheets PL, Wokosin DL, et al. Local-Circuit Phenotypes of Layer 5 Neurons in Motor-Frontal Cortex of YFP-H Mice. Front Neural Circuits. 2008;2:6.
    1. Hattox AM, Nelson SB. Layer V neurons in mouse cortex projecting to different targets have distinct physiological properties. J Neurophysiol. 2007;98:3330–3340.
    1. Feng G, Mellor RH, Bernstein M, Keller-Peck C, Nguyen QT, et al. Imaging neuronal subsets in transgenic mice expressing multiple spectral variants of GFP. Neuron. 2000;28:41–51.
    1. Zhang S, Boyd J, Delaney K, Murphy TH. Rapid reversible changes in dendritic spine structure in vivo gated by the degree of ischemia. J Neurosci. 2005;25:5333–5338.
    1. Dombeck DA, Khabbaz AN, Collman F, Adelman TL, Tank DW. Imaging large-scale neural activity with cellular resolution in awake, mobile mice. Neuron. 2007;56:43–57.
    1. Enright LE, Zhang S, Murphy TH. Fine mapping of the spatial relationship between acute ischemia and dendritic structure indicates selective vulnerability of layer V neuron dendritic tufts within single neurons in vivo. J Cereb Blood Flow Metab. 2007;27:1185–1200.
    1. Grutzendler J, Kasthuri N, Gan WB. Long-term dendritic spine stability in the adult cortex. Nature. 2002;420:812–816.
    1. Holtmaat AJ, Trachtenberg JT, Wilbrecht L, Shepherd GM, Zhang X, et al. Transient and persistent dendritic spines in the neocortex in vivo. Neuron. 2005;45:279–291.
    1. Li P, Murphy TH. Two-photon imaging during prolonged middle cerebral artery occlusion in mice reveals recovery of dendritic structure after reperfusion. J Neurosci. 2008;28:11970–11979.
    1. Miller MN, Okaty BW, Nelson SB. Region-specific spike-frequency acceleration in layer 5 pyramidal neurons mediated by Kv1 subunits. J Neurosci. 2008;28:13716–13726.
    1. Jordan C, Li HH, Kwan HC, Francke U. Cerebellar gene expression profiles of mouse models for Rett syndrome reveal novel MeCP2 targets. BMC Med Genet. 2007;8:36.
    1. Luikenhuis S, Giacometti E, Beard CF, Jaenisch R. Expression of MeCP2 in postmitotic neurons rescues Rett syndrome in mice. Proc Natl Acad Sci U S A. 2004;101:6033–6038.
    1. Braunschweig D, Simcox T, Samaco RC, LaSalle JM. X-Chromosome inactivation ratios affect wild-type MeCP2 expression within mosaic Rett syndrome and Mecp2−/+ mouse brain. Hum Mol Genet. 2004;13:1275–1286.
    1. Paxinos G, Franklin KBJ. The mouse brain in stereotaxic coordinates. San Diego: Academic Press; 2001.
    1. Caroni P. Overexpression of growth-associated proteins in the neurons of adult transgenic mice. J Neurosci Methods. 1997;71:3–9.
    1. Adams VH, McBryant SJ, Wade PA, Woodcock CL, Hansen JC. Intrinsic disorder and autonomous domain function in the multifunctional nuclear protein, MeCP2. J Biol Chem. 2007;282:15057–15064.
    1. Beirowski B, Berek L, Adalbert R, Wagner D, Grumme DS, et al. Quantitative and qualitative analysis of Wallerian degeneration using restricted axonal labelling in YFP-H mice. J Neurosci Methods. 2004;134:23–35.
    1. Bittner T, Fuhrmann M, Burgold S, Ochs SM, Hoffmann N, et al. Multiple events lead to dendritic spine loss in triple transgenic Alzheimer's disease mice. PLoS One. 2010;5:e15477.
    1. Hansen KF, Sakamoto K, Wayman GA, Impey S, Obrietan K. Transgenic miR132 alters neuronal spine density and impairs novel object recognition memory. PLoS One. 2010;5:e15497.
    1. Landi S, Putignano E, Boggio EA, Pizzorusso T, Ratto GM. The short-time structural plasticity of dendritic spines is altered in a model of Rett syndrome. Sci Rep. 2011;1:45:1–7.
    1. Matter C, Pribadi M, Liu X, Trachtenberg JT. Delta-catenin is required for the maintenance of neural structure and function in mature cortex in vivo. Neuron. 2009;64:320–327.
    1. Migliore M, Shepherd GM. Opinion: an integrated approach to classifying neuronal phenotypes. Nat Rev Neurosci. 2005;6:810–818.
    1. Molnár Z, Cheung AF. Towards the classification of subpopulations of layer V pyramidal projection neurons. Neurosci Res. 2006;55:105–115.
    1. Molyneaux BJ, Arlotta P, Menezes JR, Macklis JD. Neuronal subtype specification in the cerebral cortex. Nat Rev Neurosci. 2007;8:427–437.
    1. Kriaucionis S, Bird A. The major form of MeCP2 has a novel N-terminus generated by alternative splicing. Nucleic Acids Res. 2004;32:1818–1823.
    1. LaSalle JM, Goldstine J, Balmer D, Greco CM. Quantitative localization of heterogeneous methyl-CpG-binding protein 2 (MeCP2) expression phenotypes in normal and Rett syndrome brain by laser scanning cytometry. Hum Mol Genet. 2001;10:1729–1740.
    1. Mullaney BC, Johnston MV, Blue ME. Developmental expression of methyl-CpG binding protein 2 is dynamically regulated in the rodent brain. Neuroscience. 2004;123:939–949.
    1. Samaco RC, Nagarajan RP, Braunschweig D, LaSalle JM. Multiple pathways regulate MeCP2 expression in normal brain development and exhibit defects in autism-spectrum disorders. Hum Mol Genet. 2004;13:629–639.
    1. Reichwald K, Thiesen J, Wiehe T, Weitzel J, Poustka WA, et al. Comparative sequence analysis of the MECP2-locus in human and mouse reveals new transcribed regions. Mamm Genome. 2000;11:182–190.
    1. Coy JF, Sedlacek Z, Bachner D, Delius H, Poustka A. A complex pattern of evolutionary conservation and alternative polyadenylation within the long 3″-untranslated region of the methyl-CpG-binding protein 2 gene (MeCP2) suggests a regulatory role in gene expression. Hum Mol Genet. 1999;8:1253–1262.
    1. Balmer D, Goldstine J, Rao YM, LaSalle JM. Elevated methyl-CpG-binding protein 2 expression is acquired during postnatal human brain development and is correlated with alternative polyadenylation. J Mol Med. 2003;81:61–68.
    1. Zhou Z, Hong EJ, Cohen S, Zhao WN, Ho HY, et al. Brain-specific phosphorylation of MeCP2 regulates activity-dependent Bdnf transcription, dendritic growth, and spine maturation. Neuron. 2006;52:255–269.
    1. Tao J, Hu K, Chang Q, Wu H, Sherman NE, et al. Phosphorylation of MeCP2 at Serine 80 regulates its chromatin association and neurological function. Proc Natl Acad Sci U S A. 2009;106:4882–4887.
    1. Nan X, Campoy FJ, Bird A. MeCP2 is a transcriptional repressor with abundant binding sites in genomic chromatin. Cell. 1997;88:471–481.
    1. Colantuoni C, Jeon OH, Hyder K, Chenchik A, Khimani AH, et al. Gene expression profiling in postmortem Rett Syndrome brain: differential gene expression and patient classification. Neurobiol Dis. 2001;8:847–865.
    1. Yasui DH, Peddada S, Bieda MC, Vallero RO, Hogart A, et al. Integrated epigenomic analyses of neuronal MeCP2 reveal a role for long-range interaction with active genes. Proc Natl Acad Sci U S A. 2007;104:19416–19421.
    1. Skene PJ, Illingworth RS, Webb S, Kerr AR, James KD, et al. Neuronal MeCP2 is expressed at near histone-octamer levels and globally alters the chromatin state. Mol Cell. 2010;37:457–468.
    1. Wu H, Tao J, Chen PJ, Shahab A, Ge W, et al. Genome-wide analysis reveals methyl-CpG-binding protein 2-dependent regulation of microRNAs in a mouse model of Rett syndrome. Proc Natl Acad Sci U S A 2010
    1. Young JI, Hong EP, Castle JC, Crespo-Barreto J, Bowman AB, et al. Regulation of RNA splicing by the methylation-dependent transcriptional repressor methyl-CpG binding protein 2. Proc Natl Acad Sci U S A. 2005;102:17551–17558.
    1. Chahrour M, Jung SY, Shaw C, Zhou X, Wong ST, et al. MeCP2, a key contributor to neurological disease, activates and represses transcription. Science. 2008;320:1224–1229.
    1. Urdinguio RG, Lopez-Serra L, Lopez-Nieva P, Alaminos M, Diaz-Uriarte R, et al. Mecp2-null mice provide new neuronal targets for Rett syndrome. PLoS One. 2008;3:e3669.
    1. Ben-Shachar S, Chahrour M, Thaller C, Shaw CA, Zoghbi HY. Mouse models of MeCP2 disorders share gene expression changes in the cerebellum and hypothalamus. Hum Mol Genet. 2009;18:2431–2442.
    1. Gibson JH, Slobedman B, Harikrishnan KN, Williamson SL, Minchenko D, et al. Downstream targets of methyl CpG binding protein 2 and their abnormal expression in the frontal cortex of the human Rett syndrome brain. BMC Neurosci. 2010;11:53.
    1. Nan X, Ng HH, Johnson CA, Laherty CD, Turner BM, et al. Transcriptional repression by the methyl-CpG-binding protein MeCP2 involves a histone deacetylase complex. Nature. 1998;393:386–389.
    1. Peddada S, Yasui DH, LaSalle JM. Inhibitors of differentiation (ID1, ID2, ID3 and ID4) genes are neuronal targets of MeCP2 that are elevated in Rett syndrome. Hum Mol Genet. 2006;15:2003–2014.
    1. Walters MC, Fiering S, Eidemiller J, Magis W, Groudine M, et al. Enhancers increase the probability but not the level of gene expression. Proc Natl Acad Sci U S A. 1995;92:7125–7129.
    1. Graubert TA, Hug BA, Wesselschmidt R, Hsieh CL, Ryan TM, et al. Stochastic, stage-specific mechanisms account for the variegation of a human globin transgene. Nucleic Acids Res. 1998;26:2849–2858.
    1. O'Leary DD, Stanfield BB. Occipital cortical neurons with transient pyramidal tract axons extend and maintain collaterals to subcortical but not intracortical targets. Brain Res. 1985;336:326–333.
    1. Stanfield BB, O'Leary DD. The transient corticospinal projection from the occipital cortex during the postnatal development of the rat. J Comp Neurol. 1985;238:236–248.
    1. Thong IG, Dreher B. The development of the corticotectal pathway in the albino rat: transient projections from the visual and motor cortices. Neurosci Lett. 1987;80:275–282.
    1. Cassel S, Revel MO, Kelche C, Zwiller J. Expression of the methyl-CpG-binding protein MeCP2 in rat brain. An ontogenetic study. Neurobiol Dis. 2004;15:206–211.
    1. Cohen DR, Matarazzo V, Palmer AM, Tu Y, Jeon OH, et al. Expression of MeCP2 in olfactory receptor neurons is developmentally regulated and occurs before synaptogenesis. Mol Cell Neurosci. 2003;22:417–429.

Source: PubMed

3
Prenumerera