21
views
0
recommends
+1 Recommend
0 collections
    0
    shares
      • Record: found
      • Abstract: found
      • Article: found
      Is Open Access

      Talking Convergence: Growing Evidence Links FOXP2 and Retinoic Acid in Shaping Speech-Related Motor Circuitry

      article-commentary

      Read this article at

      Bookmark
          There is no author summary for this article yet. Authors can add summaries to their articles on ScienceOpen to make them more accessible to a non-specialist audience.

          Abstract

          FOXP2 was the first identified monogenic cause of a speech disorder (for review see Graham et al., 2015). However, it remains to be answered how it affects the development of speech related neuronal circuits and what the relevant molecular pathways may be. Devanna et al. (2014) reported a direct interaction between FOXP2 and Retinoic Acid (RA), an important signaling molecule in brain development and neuronal differentiation, in a cell model. What could this finding mean for our understanding of a language-ready brain? Mounting evidence from diverse fields such as linguistics, genetics, brain development, and cellular neurophysiology suggests that the brain's motor circuitry might be the key to the conundrum of language disorders. Human patients with FOXP2 mutations show prominent motor learning deficits at the root of their developmental verbal dyspraxia diagnosis, which is reflected in mouse models of Foxp2 loss (for review see Deriziotis and Fisher, 2013; Graham et al., 2015). Interestingly, in mice with dysregulated RA signaling, similar motor learning impairments have been found (for review see Duester, 2013). Retinoic acid has several prominent roles in brain development and function, for example as a diffusible signaling molecule that changes gene expression via nuclear retinoic acid receptors (RARs, RORs, and RXRs; for review see Maden, 2007). Devanna et al. (2014) demonstrated a direct interaction between FOXP2 and RA signaling in a human neuroblastoma cell line (SH-S5Y5), in which RA exposure initiates differentiation into a dopaminergic-neuron like state (Korecka et al., 2013). When the researchers compared FOXP2 overexpression to RA treatment they found an interesting convergence: In both cases, the cells differentiated from their previously proliferative state into a more neuron-like state and showed elevated expression of RA signaling-related genes, specifically the receptor RARß. A recent review of the same group (van Rhijn and Vernes, 2015) suggests that the motor-learning deficits found in carriers of FoxP2 mutations might be mediated by defects in RA signaling, also in humans. Interestingly, FoxP2 and RA signaling have also been found (separately) to be involved in vocal production learning in a songbird model (Denisenko-Nehrbass et al., 2000; Olson et al., 2011; Roeske et al., 2014; Wohlgemuth et al., 2014). Even though direct evidence of synergies between FoxP2 and RA signaling is still lacking beyond cell models, the hypothesis that they are parts of the same regulatory network (e.g. Benítez-Burraco and Boeckx, 2014; van Rhijn and Vernes, 2015) would fit well with a number of findings, some of which we will address in the following commentary. Striatal development: do FoxP2 and RA converge on specific cell types? The canonical cortico-basal-ganglia-thalamic motor circuit is thought to be crucial for the motor aspects of vocal production. Interestingly, in this canonical circuit FoxP2 (refers to the gene across vertebrate species) and RA receptor expression overlap in several regions (Figures 1A,B). An especially illustrative region in this regard is the striatum, a hub of the motor circuitry that is known to be involved in motor sequencing in mammals (Kravitz and Kreitzer, 2012; Friend and Kravitz, 2014) and also in vocal production learning in songbirds (Bolhuis et al., 2010; Scharff and Petri, 2011). Available data indicates both FoxP2 expression and RA in the development and function of the striatum (van Rhijn and Vernes, 2015). Figure 1 Canonical cortico-basal-ganglia-thalamic circuit of the motor system (A) Sagittal view of the mouse motor system with excitatory (black) and inhibitory projections (red) connections. (B) Functional model (based on Morita 2014) of the motor sequencing circuitry with FoxP2 and RARb expression. IT, Intratelencephalic neurons of the cortical layer V; PT, Pyramidal tract “motor neurons” of layer Vb; Drd1/Drd2, Striatal dopamine-receptor 1/2 expressing MSNs; GPe, Globus Pallidus External; STN, Subthalamic nucleus; SNr, Substantia nigra, pars reticulata. In the developing rodent striatum, two layers of progenitor cells in the ventricular and subventricular zone (SVZ) generate the medium spiny neurons (MSNs), the main cell type of the striatum (Figure 1B). Early during striatal development, a mixed population of progenitor cells in the SVZ starts producing RA, concomitant with the start of FoxP2 and RARß expression in the same region (Li et al., 2000; Ferland et al., 2003; Molotkova et al., 2007; Liao et al., 2008; Garcia-Calero et al., 2015). In general, RA signaling is involved very early in the spatial patterning of the brain, starting with the formation of the telencephalic vesicle; however the precise extent has been controversially discussed (Siegenthaler et al., 2009; Chatzi et al., 2011, 2013; Duester, 2013). More importantly for motor circuit development, RA signaling is involved in the patterning of the striatum and striatonigral projections (Rataj-Baniowska et al., 2015) as well as migration of interneurons from the Ganglionic Eminences (that also produce MSNs) to the cortex (Crandall et al., 2011). Within striatal MSNs, mouse studies imply an interesting convergence of FoxP2 and RA signaling onto a single cell population, namely D1R-MSNs (“direct pathway” neurons) of the dorsolateral striatum (Figure 1B). In mice lacking RARß, the progenitor population prematurely differentiates, specifically reducing the final cell numbers of D1R-MSNs (Rataj-Baniowska et al., 2015). Interestingly, RARβ was also one of the key RA receptors found to be affected in the human neuroblastoma cell study of Devanna et al. (2014). In mice lacking RARß, levels of the D1R-linked signal transduction protein Darpp-32 are greatly reduced (Liao et al., 2008). This is of interest not only because D1R-MSNs preferentially express FoxP2, but also because FoxP2 mutant mouse embryos likewise show strongly reduced Darpp-32 expression levels (Vernes et al., 2011, see supplementary material). One may therefore hypothesize that FoxP2 and RA signaling pathways converge in the development of the D1R-MSNs cell class in the dorsolateral striatum (Figure 1B), impacting the motor circuit's “direct pathway”. If these hypotheses are applied to a computational model of motor sequencing (Morita et al., 2012; Morita, 2014), several possible consequences emerge. On the circuit-function level, FoxP2 and RA disruptions would impair D1R-MSNs function, and hence the “go” signal in motor control (Sippy et al., 2015, Figure 1B). Furthermore, FoxP2 mutations may affect inhibitory feedback to the striatum via a functionally defined population projecting from the Globus Pallidus Externa back to the striatum (Abdi et al., 2015; Dodson et al., 2015; Mallet et al., 2016), leading to additional defects in the “stop” motor control. Interestingly, a recent study found Foxp2 to regulate the development of the projection patterns of thalamic nuclei, another key integrative part of the motor circuitry (Ebisu et al., 2016). Downstream behavioral effects: do similar phenotypes imply convergent mechanisms? In mice, Foxp2 loss impairs vocalizations after birth (e.g. Castellucci et al., 2016; Chen et al., 2016). However, mouse vocalizations are thought to lack a learning component, a key factor in human speech (see French and Fisher, 2014 for a review). In male zebra finches, a vocalization learning model, FoxP2 is expressed throughout the song system (Haesler et al., 2004, 2007; Scharff and Petri, 2011; Mendoza et al., 2015). Expression is especially high in the song-learning related area X of the striatum at the time when songs are acquired (Haesler et al., 2004; Thompson et al., 2013), and FoxP2 knockdown during this period disrupts social modulation of song variability (Murugan et al., 2013). Similar to the mouse striatal network, FoxP2 knockdown in the juvenile zebra finch leads to specific reductions of D1R and Darpp-32 expression in the dorsolateral striatum, specifically in area X (Denisenko-Nehrbass et al., 2000; Haesler et al., 2004, 2007; Olson et al., 2011; Murugan et al., 2013; Roeske et al., 2014). It should be noted that in the songbird striatum, D1R and D2R expression in MSNs overlap to a greater degree than in the mouse (Kubikova et al., 2010). Furthermore, striatal RA signaling plays an important role in the maintenance of birdsong: RA-producing enzymes are located in projection axons to area X, where the MSN-like neurons in turn express RA receptors. Functionally, RA may be important for balancing the input to area X neurons from vocal motor pathway (stereotyped song) versus anterior forebrain pathway (variable song) vocalization centers (Roeske et al., 2014; Wohlgemuth et al., 2014). FoxP2 and RA: synergy in a larger regulatory network for neuronal connectivity? On the molecular level, the gene networks regulated by FoxP2 and RA signaling might overlap in specific parts of the motor circuitry. It is important to separate developmental effects from on-line functional effects: Developmental defects might lead to an impaired network setup (e.g. a smaller and miswired MSN population), while functional effects might impair the learning and function even of healthy networks (e.g. failure to balance synaptic inputs to MSNs). Furthermore, it should be noted that both FoxP2 and RA are broad regulators of large gene networks, also outside the motor circuitry. In this context, a recent study is of particular interest: Chen et al. (2016) found that Foxp2 loss was associated with defects in dendritic spine formation in striatal neurons via de-repression of the autism risk gene Mef2C, leading to defects in corticostriatal connectivity. Research in a different murine cell model suggests that RA treatment indirectly interacts with Mef2C to specify a neuronal fate, via the transcription factor Sp1 (Elmi et al., 2007). It would therefore be highly interesting to see whether Foxp2, RA, Mef2C, and Sp1 are all part of the same regulatory network and influence the development of MSN connectivity. Thus, with their finding of synergies between FOXP2 and RA in a cell model Devanna et al. (2014) made an important first step. Other studies provide indirect evidence that FoxP2 and RA signaling genes could be part of the same regulatory network and co-influence development and function of the speech-motor control circuits, especially the striatum. However, a direct proof beyond cell models is still lacking. Hence future studies in adequate animal models will have to confirm this synergy on the neuronal circuit level, also considering the more extensive regulatory molecular network affecting the development and functioning of speech-motor control circuits. Author contributions DS and MN both reviewed the current literature, wrote the manuscript and prepared the figures with equal contributions. Conflict of interest statement The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

          Related collections

          Most cited references37

          • Record: found
          • Abstract: found
          • Article: not found

          Characterization of Foxp2 and Foxp1 mRNA and protein in the developing and mature brain.

          Foxp2 and Foxp1 are recently identified members of the Fox family of winged-helix/forkhead transcription factor genes. A recent study has found that mutations in human FOXP2 produce a severe language disorder. Since Foxp2 appears to be important in language, we wanted to explore the expression of this gene and a homologous gene, Foxp1, in the developing brain. In the present study, we investigated the time course and localization of Foxp2 and Foxp1 mRNA and protein expression in the developing and adult mouse using in situ hybridization and immunohistochemistry. Foxp2 and Foxp1 are expressed as early as E12.5 and persist into adulthood. Foxp2 and Foxp1 were most highly expressed in the developing and mature basal ganglia. Expression of Foxp2 was also observed in the cerebral cortex (layer 6), cerebellum (Purkinje neurons), and thalamus. Foxp1 expression was observed in the cerebral cortex (layers 3-5), hippocampus (CA1), and thalamus. Very little ventricular zone expression was observed for Foxp2 and Foxp1 and the expression of both of these genes occurred following neuronal migration, suggesting a role for these genes in postmigratory neuronal differentiation. Furthermore, we demonstrated the expression of FOXP2 in human fetal brain by RT-PCR, in the perisylvian area of the left and right cerebral hemispheres, as well as in the frontal and occipital cortices. Overall, the widespread expression of Foxp2 in the developing brain makes it difficult to draw specific conclusions about which areas of Foxp2 expression are critical to human language function. Copyright 2003 Wiley-Liss, Inc.
            Bookmark
            • Record: found
            • Abstract: found
            • Article: not found

            Incomplete and Inaccurate Vocal Imitation after Knockdown of FoxP2 in Songbird Basal Ganglia Nucleus Area X

            Introduction Genetic aberrations of FOXP2 cause developmental verbal dyspraxia (DVD), which is characterized by impaired production of sequenced mouth movements and both expressive and receptive language deficits [1–4]. Brain imaging studies in adult FOXP2 patients implicate the basal ganglia as key affected regions [5–7], and FOXP2 is prominently expressed in the developing human striatum [8]. These findings raise the question whether the speech and language abnormalities observed in individuals with DVD result from erroneous brain development or impaired function of differentiated neural circuits in the postnatal brain, or a combination of both. Human speech and learned vocalizations in oscine birds bear behavioral and neural parallels [9]. Thus songbirds are a suitable model for studying the neural mechanisms of imitative vocal learning, including speech and its pathologies. The FoxP2 expression patterns in songbird and human brains are very similar, with strong expression in the basal ganglia, thalamus, and cerebellum [8,10,11]. Moreover, FoxP2 expression in the basal ganglia song nucleus, Area X, which is important for normal song development [12,13], transiently increases at the time when young zebra finches learn to sing. In adult canaries, FoxP2 expression in Area X is elevated during the late summer months, coincident with the incorporation of most new syllables to their seasonally changing song [10]. FoxP2 is down-regulated in Area X when adult zebra finches sing slightly variable, undirected song, but not when they sing more stereotyped female-directed song [14]. Together, these correlative findings raise the question whether FoxP2 and vocal plasticity are causally related. Using lentivirus-mediated RNA interference (RNAi) during song development, we now show that zebra finches with reduced FoxP2 expression levels in Area X imitated tutor songs incompletely and inaccurately. This effect was already evident during vocal practice in young birds. Moreover, the acoustic structure and the duration of song syllables in adults were abnormally variable, similar to word production in children with DVD [15]. These findings are consistent with a role of FoxP2 during auditory-guided vocal motor learning in songbird basal ganglia. Results Establishing Lentiviral-Mediated RNAi in the Zebra Finch Vocal learning in zebra finches proceeds through characteristic stages. In the sensory phase that commences around 25 d after hatching (post-hatch day [PHD]), young males memorize the song of an adult male tutor. Concomitantly, they start vocalizing the so called “subsong,” consisting of quietly uttered, poorly articulated, and nonstereotypically sequenced syllables [16]. Following intensive vocal practice and improvement toward matching the tutor song during the period of “plastic song,” they eventually imitate the song of their tutor with remarkable fidelity around PHD90. The structural and temporal characteristics of adult “crystallized” song remain essentially stable throughout adult life. To study the function of FoxP2 during song learning of zebra finches, we reduced the levels of FoxP2 expression bilaterally in Area X in vivo, using lentivirus-mediated RNAi. In this approach, short interfering hairpin RNA (shRNA) containing sense and antisense sequences of the target gene connected by a hairpin loop are expressed from a viral vector. The virus stably integrates into the host genome, enabling expression throughout the life of the animal [17]. We designed two different shRNAs (shFoxP2-f and shFoxP2-h) targeting different sequences in the FoxP2 gene. Both hairpins strongly reduced the levels of overexpressed FoxP2 protein in vitro (Figure 1F), but did not change the levels of overexpressed protein levels of FoxP1, the closest homolog of FoxP2. For further control experiments, we generated a shRNA designed not to target any zebra finch gene (shControl). As expected, this nontargeting shRNA did not affect expression of either FoxP2 or FoxP1 in vitro (Figure 1F). Since shFoxP2-f and shFoxP2-h targeted FoxP2 with similar efficiency, both of them were interchangeably used for subsequent in vivo experiments (shFoxP2-f/-h). Figure 1 Establishing Lentivirus-Mediated Knockdown of FoxP2 in Zebra Finch Area X (A) Phase contrast image of a sagittal 50-μm brain section from a male zebra finch. Area X is outlined by white arrows (scale bar indicates 1 mm). The microinjection into Area X is schematized in the inset. (B) Fluorescent microscopy image of (A). Virus-infected cells expressed GFP (green). (C) FoxP2 immunostaining (red; scale bar indicates 10 μm) (D) The neuron shown in (C) also expressed viral GFP from injection with the nontargeting shControl virus. (E) Overlay picture of (C) and (D). (F) Overexpression of zebra finch FoxP2 (left panel) or FoxP1 (right panel), each tagged with the V5 epitope, and one of different hairpin constructs (shFoxP2-f, shFoxP2-h, or shControl) in HEK293 T cells. Western blot analysis using a V5 antibody revealed that shFoxP2-f and shFoxP2-h, but not shControl, efficiently reduced FoxP2 levels. FoxP1 protein levels were unaffected by overexpression of either shRNA. Immunostaining with an actin antibody shows comparable loading of protein samples. (G) Knockdown of FoxP2 in vivo. Immunofluorescent staining with an antibody against FoxP2 on 50-μm brain sections from birds injected with shFoxP2-f/-h in one hemisphere and shControl in the contralateral hemisphere 30 d prior to analysis revealed lower fluorescence levels and fewer cells in knockdown (upper panel) compared to control sections (lower panel). FoxP2-positive cells appear red; virally infected cells express GFP, visible in green (scale bar indicates 20 μm). (H) Quantification of in vivo knockdown efficiency. The fluorescence intensity of FoxP2 immunostaining was measured in images from brain sections injected with shFoxP2-f/-h in one hemisphere and shControl in the contralateral hemisphere 30 d prior to analysis. All antibody incubations were performed simultaneously, and pictures were taken with identical camera settings. Bars represent average intensity levels normalized to the shControl-injected hemisphere (± standard error of the mean [SEM]; two-tailed Mann-Whitney U test, **p 0.5; shFoxP2-f/-h, n = 6, shControl, n = 7). Quantification of Area X volume targeted by virus injection in an equally treated group of birds, but sacrificed at PHD50, confirmed the results obtained for PHD90 (mean volume 20.4% ± STDV 4.0%; two-tailed Mann-Whitney U test, p > 0.6; shControl n = 3 hemispheres from 3 animals; shFoxP2-f/-h, n = 3 hemispheres from 3 animals). To quantify the neuronal extent of lentivirus expression in Area X, we used immunohistochemical staining with the neuronal marker Hu [18] (Figure S2). Of all virus-infected cells, 78.5% ± 3.5% were neurons (mean ± standard error of the mean [SEM]; no significant difference between shFoxP2 and shControl, two-tailed Mann-Whitney U test, p > 0.7; shControl injections n = 3 hemispheres from 3 animals, shFoxP2 injections n = 4 hemispheres from 4 animals;). This result is consistent with Wada et al. [19], who used the same viral constructs in the zebra finch brain in vivo. Among the infected cells were FoxP2-positive spiny neurons, which are assumed to be the most common cell type in Area X [20] (Figure 1C–1E). To quantify FoxP2 knockdown in vivo, we determined FoxP2 protein levels in Area X on PHD50, the time of peak FoxP2 expression [10] in birds injected on PHD23 with shFoxP2-f/-h in one hemisphere and shControl into the contralateral hemisphere. The signal of the immunofluorescent staining with a FoxP2 antibody was significantly lower in knockdown Area X than in control Area X (Figure 1G and 1H). We also assessed FoxP2 mRNA levels after knockdown in Area X. Birds were injected on PHD23 with shFoxP2-f/-h in one hemisphere and shControl in the contralateral hemisphere. On PHD50, we punched out Area X of injected birds and measured FoxP2 mRNA levels by real-time PCR. FoxP2 levels were normalized to two independent RNAs coding for the housekeeping genes Hmbs and Pfkp. FoxP2 mRNA was reduced on average by approximately 70% in the shFoxP2-infected region of Area X compared to the shControl-infected region of Area X (Figure 1I). Of note, RNAi-mediated knockdown approximates FOXP2 levels in DVD patients, since haploinsufficiency, a 50% reduction of functional FOXP2 protein, is apparently the common feature of all reported human FOXP2 mutations [4,21]. To demonstrate that RNAi-mediated gene knockdown can persist in vivo throughout the entire song-learning phase, we used a virus expressing shRNA against the viral reporter GFP (shGFP) in conjunction with the virus expressing a shRNA lacking a target gene (shControl). We injected young zebra finches on PHD23 with equal amounts of equally infectious shGFP and shControl virus in the left and right hemisphere, respectively. More than 3 mo later, on PHD130, the GFP signal in the shGFP-injected hemisphere was still 70.5% ± 5.8% less intense than in the shControl-injected hemisphere (mean ± SEM; n = 2; Figure 1J). To rule out potential side effects of FoxP2 knockdown on cellular survival in Area X, we investigated apoptosis in Area X 6 d after surgery with terminal deoxyribonucleotide transferase-mediated dUTP nick end labeling (TUNEL). The TUNEL method detects genomic DNA double-strand breaks characteristic of apoptotic cells. Of 1,149 GFP-positive cells counted in six hemispheres from three animals, only five were TUNEL-positive (Figure S3). ShControl-injected and uninjected animals had similar low levels of apoptotic cells (unpublished data). Thus, FoxP2 is not a gene essential for short-term survival of postmitotic neurons. Since the TUNEL method does not capture any long-term changes in neuronal viability that might follow after reduction of FoxP2, we used the neuronal marker Hu to determine neuronal densities in Area X 30 d after injecting either shFoxP2-f/-h or shControl virus (Figure S4). Neuronal densities in the infected region in Area X did not differ in knockdown and shControl-injected birds (two-tailed Mann-Whitney U test, p > 0.39; shControl, n = 4 hemispheres; shFoxP2-f/-h, n = 3 hemispheres). Density of neurons were also comparable inside and outside of the virus-infected region of Area X for all viruses (two-tailed Mann-Whitney U test, p > 0.6 for both shFoxP2-f/-h and shControl). In sum, these data demonstrate that virus-mediated RNAi can induce specific, long-lasting knockdown of gene expression in zebra finch Area X without causing cell death. Song Imitation of FoxP2 Knockdown Zebra Finches Adult zebra finch song consists of different sound elements, here called syllables, that are separated by silent intervals. Syllables are rendered in a stereotyped sequential order, constituting a motif. During a song bout, a variable number of motifs are sung in short succession. To obtain a first descriptive account of the song of knockdown and control pupils, we measured mean acoustic features for all syllables recorded from all pupils using the software Sound Analysis Pro (SAP) [22]. The features extracted were mean pitch, mean frequency, mean frequency modulation (FM; change of frequency in time), mean entropy, and mean pitch goodness (PG; periodicity of sound), as well as mean duration. The comparison of the distribution of these features across the repertoire of knockdown and control pupils did not reveal any significant differences, indicating that knockdown pupils, control pupils, and tutors sang syllables with similar acoustic features (Figure S5). Next, we analyzed the behavioral consequences of bilateral FoxP2 knockdown in Area X for the outcome of song learning at PHD90. When a juvenile male finch is tutored individually by one adult male, the pupil learns to produce a song that strongly resembles that of his tutor [23]. We therefore determined learning success by the degree of acoustic similarity between pupil and tutor songs. Analysis of song recorded at PHD90 revealed that pupils with experimentally reduced FoxP2 levels in Area X imitated tutor songs with less fidelity than control animals did (see also Audio S1–S6). The comparison of sonograms from shControl-injected (Figure 2A) and shFoxP2-injected pupils (Figure 2B and 2C) with their respective tutors shows the characteristic effects caused by reduction of FoxP2. Typical features of FoxP2 knockdown pupils included syllable omissions (Figure 2B, syllables C, D, F, and G; Figure 2C, syllable B), imprecise copying of syllable duration (Figure 2B, syllable E longer; Figure 2C, syllable D shortened), and inaccurate imitation of spectral characteristics (Figure 2B, syllable E; Figure 2C, syllable D). In addition, in four out of seven knockdown pupils, the motif contained repetitions of individual syllables or syllable pairs (e.g., see Figure 2B and 2C). In contrast, none of the control or tutor motifs contained repeated syllables. Pupils did not reverse the sequential order of syllables in the tutor motifs, except for one control (unpublished data) and one FoxP2 knockdown pupil (Figure 3A). Figure 2 Incomplete Tutor Song Imitation by FoxP2 Knockdown Pupils Each sonogram depicts a typical motif of one animal (scale bars indicate 100 ms, frequency range 0–8,600 Hz). Tutor syllables are underlined with black bars and identified by letters. The identity of pupil syllables was determined by similarity comparison to tutor syllables using SAP software. Imprecisely copied pupil syllables are designated with red italic letters. (A–C) (A) tutor #38 and shControl-injected pupil, (B) tutor #396 and shFoxP2-injected pupil, and (C) tutor #414 and shFoxP2-injected pupil. ShFoxP2-injected pupils copied fewer syllables and the fidelity of syllable imitation was worse than in shControl pupils, reflected by lower SAP scores (similarity/accuracy indicated vertically at the right edge of the sonograms). (D) The mean similarity scores between tutor and pupil motifs were significantly lower in shFoxP2- injected pupils than in shControl- and shGFP-injected pupils (± SEM; two-tailed Mann-Whitney U test, **p 0.5). Figure 3 Inaccurate Tutor Song Imitation by FoxP2 Knockdown Pupils (A) Representative sonograms of FoxP2 knockdown and control pupils both tutored by male 388 (scale bars indicate 100 ms, frequency range = 0–8,600 Hz). Syllables are underlined with black bars and identified by letters. The identity of pupil syllables was determined by similarity comparison to tutor syllables using SAP software. Red italic letters denote imprecisely copied syllables. Inaccurate imitation is particularly evident in the second element of syllable A and the first element of syllable B. Similarity and accuracy scores are indicated vertically at the right edge of the sonograms. (B) Average motif accuracy was significantly lower in shFoxP2 knockdown pupils compared to control pupils, indicating that they imitated their tutors less precisely (±SEM; two-tailed Mann-Whitney U test, **p 0.4). (C) The frequency distribution of identity scores of all syllables from FoxP2 knockdown pupils (dark grey bars)was shifted towards lower scores, compared to control pupils (light grey bars).. This suggests that all syllable types were affected. (Identity scores were obtained from comparison of pupil/tutor syllable pairs; shFoxP2 n = 24 syllables from 7 animals; shControl n = 26 syllables from 7 animals). (D) Comparison of syllable duration and mean acoustic feature values (FM and PG) between pupil syllables and their respective tutor syllables. The divergence of imitated syllables from the tutor model tended to be larger for all acoustic measures in the FoxP2 knockdown pupils (dark grey bars) than in the controls (light grey bars). For average syllable duration and mean entropy measures, the difference was significant (±SEM; two-tailed Mann-Whitney U test, **p 0.8; tutors n = 6, on average 5 syllables per animal; shControl n = 7 animals, on average 4 syllables per animal; shGFP n = 3, on average 5 syllables per animal). (D) Syllable duration varied more from rendition to rendition in knockdown pupils (shFoxP2) than in controls (shControl and shGFP) and tutors, as indicated by a higher mean coefficient of variation of syllable duration (±SEM, two-tailed Mann-Whitney U test, **p 0.7, same animals as [C]). Next, we quantified the variability of syllable duration between different renditions of the same syllable. The coefficient of variation of syllable duration was significantly higher in knockdown than in control pupils and tutors, suggesting imprecise motor coordination on short temporal scales (Figure 4D). Notably, the timing of syllables in control pupils (shControl and shGFP) was as stable as in tutors (Figure 4D). The variability of syllable duration in tutor and control birds varied in the same range as reported previously [30], emphasizing how tightly adult zebra finches normally control syllable duration. Finally, we analyzed the sequential order of syllables over the course of many motifs. To this end, we first annotated sequences of 300 user-defined syllables with the positions in their respective motifs. We then measured the stereotypy of a motif by calculating for each syllable the entropy of its transition distribution. Based on this entropy measure, we generated a sequence consistency score (1 − entropy), which reflects song stereotypy. An entropy score of 0 indicates random syllable order, whereas a score of 1 reflects a fixed syllable order. The mean sequence consistency was similar in shControl and shFoxP2-f/-h animals (Figure S8). Because stereotypy of motif delivery is a hallmark of “crystallized” adult song, it seems plausible that both knockdown animals and controls had reached the end of the sensory-motor learning period [31]. To investigate this question in more detail, we next analyzed the song of knockdown and control pupils recorded at earlier stages of song development. Song Development in FoxP2 Knockdown Zebra Finches To explore the developmental trajectory of song learning in knockdown and control pupils, we analyzed songs recorded during plastic song at PHD65 and towards the end of the learning phase at PHD80. Since syllables are not yet rendered in a stereotyped motif structure at PHD65, we quantified song imitation success and vocal variability on the level of the syllables only. To avoid the necessity of identifying individual syllables based on their morphology, we made use of an automated procedure provided by SAP to compare all song material from a given day to the tutor's typical motif. The vocalizations of pupils were first segmented into syllables. All segments were subsequently compared to the typical motif of the tutor in a pairwise fashion (between 1,000–3,000 comparisons per pupil per day). The output variable of these measurements is an accuracy score, which describes the extent to which the pupil's sounds match those of the tutor (see Materials and Methods and [22] for further details). We found that knockdown pupils imitated their tutors less accurately than control pupils already at PHD65 (Figure 5A). The frequency distribution of accuracy values also suggests that imprecise syllable imitation was not skewed towards particular syllables or syllable types (Figure S9). This result is in line with the observation made earlier for the syllables at PHD90 (Figure 3C). In contrast to control pupils, knockdown pupils did not improve in accuracy after PHD80, suggesting they had reached the end of the learning phase (Figure 5A). Figure 5 Differences in Song Development of FoxP2 Knockdown and Control Pupils (A) We measured the accuracy of syllable imitation in song recordings of the same pupils made at three different ages (PHD65, PHD80, and PHD90), using the automated batch procedure in Sound Analysis Pro. Data points represent mean values (±SEM) of 1,000–3,000 pairwise comparisons between pupil recordings and the tutor model (shFoxP2-f/-h, n = 7 animals for all ages; shControl, n = 5 animals for PHD65, n = 6 animals for PHD80, and n = 7 animals for PHD90). Syllable imitation was already less accurate in FoxP2 knockdown pupils by 65PHD (two-tailed Mann-Whitney U test, PHD65 *p 0.5). The dashed line connecting the data points illustrates the directionality of changes over time. but does not imply a linear relationship. (B) Variance of syllable accuracy values increased with age in knockdown pupils, but not in controls (two-tailed Wilcoxon signed-rank test, PHD65 to PHD90 *p 0.4). This leads to significantly higher variance at PH90 in knockdown pupils compared to control pupils (two-tailed Mann-Whitney U test, PHD90 *p 0.9; n = 5 for shFoxP2-f/-h and n = 7 for shControl), suggesting that up to this age, syllable imitation followed largely similar dynamics. However, from PHD80 to PHD90, accuracy of syllable imitation continued to improve only in control, but not in knockdown pupils (two-tailed Mann-Whitney U test, p 0.8 for both similarity and accuracy). Quantification of FoxP2 knockdown. Young male zebra finches received an injection of shFoxP2-f/-h virus in one hemisphere and an injection with control virus (shControl) in the contralateral hemisphere on PHD23 as described above. For the quantification of protein levels after FoxP2 knockdown, we performed an immunohistological staining with the FoxP2 antibody on 50-μm sections 30 d after virus injection. Immunohistological staining was performed as described [10], but using an antibody dilution of 1:5,000. All sections were processed at the same time with the same batch of antibody solution. Images of stained brain sections were taken with a digital camera using the Simple PCI software (Compix) at 40× magnification. For each section, we acquired multiple Z-stacked images of the virus-infected area (230.3 μm × 230.3 μm), and reconstructed a maximal projection. All images from the same bird were taken with the same microscope and software settings. Finally, we quantified fluorescence intensity levels in the images. The intensity of the green fluorescence from the viral GFP was not significantly different between shFoxP2-f/-h–injected and shControl-injected hemispheres (two-tailed Mann-Whitney U test, p > 0.3). For the quantification of FoxP2 knockdown mRNA levels, young male zebra finches were injected with shFoxP2-f/-h virus in one hemisphere and control virus (shControl) in the contralateral hemisphere on PHD23, as described above. This permitted analysis of FoxP2 knockdown in the same bird while avoiding confounding differences in gene expression levels between birds. On PHD50, we sacrificed the birds and excised the GFP-expressing brain area with a 1-mm–diameter glass capillary (Brand) under a fluorescence dissecting microscope. RNA was extracted with TRIZOL (Invitrogen); yield was determined by UV spectroscopy at 260/280 nm with a Nanodrop device. FoxP2 expression was quantified by real-time PCR using SybrGreen (Applied Biosystems). We determined relative FoxP2 expression levels through normalization to the expression levels of two internal control genes, which were identified in a BLAST homology search for the mouse housekeeping genes Hmbs and Pfkp in the database from the Songbird Neurogenomics Initiative (http://titan.biotec.uiuc.edu/songbird/) and the Songbird Brain Transcriptome Database (http://songbirdtranscriptome.net/). The expression of Hmbs and Pfkp in the left and right hemisphere in both injected and untreated animals was equivalent (numbers indicate fold change between left and right hemispheres; untreated: Hmbs = 1.4 ± 0.5 and Pfkp = 1.3 ± 0.6; injected: Hmbs = 1.0 ± 0.4 and Pfkp = 1.1 ± 0.4, n=5 birds). Relative expression levels were determined with the comparative cycle time (Ct) method. All primers used in this study amplified the cDNA with similar efficiency (E = 1 ± 5%) in a validation experiment. Normalized Ct values from the same animal were calibrated to the shControl-injected hemisphere. FoxP2 expression levels are thus presented as the ratio of expression in shControl- to shFoxP2 -injected hemispheres. Song recording and analysis. Vocalizations were recorded between 9 am and 4 pm on PHDs 65, 80, and between 90 to 93 in absence of the tutor. Quantitative song analysis was performed using the SAP software, version 1.04 [22,51]. We analyzed song at the level of the syllables, the motif, and syntax. We define “syllable” as a continuous sound element, surrounded by silent intervals. The “typical song motif” was defined as the succession of syllables that includes all syllable types (except introductory notes), and occurs in a repeated manner during a song bout. Syntax refers to the sequence of syllables in many successive motifs. Motif analysis. We quantified how well pupils had copied the motif of their tutor using a similarity score and an accuracy score obtained in SAP from ten asymmetric pairwise comparisons of the pupil's typical motif with the tutor motif. In asymmetric comparisons, the most similar sound elements of two motifs are compared, independent of their position within a motif. The smallest unit of comparison are 9.26-ms–long sound intervals (FFT windows). Each interval is characterized by measures for five acoustic features: pitch, FM, amplitude modulation (AM), Wiener entropy, and PG. SAP calculates the Euclidean distance between all interval pairs from two songs, over the course of the motif, and determines a p-value for each interval pair. This p-value is based on p-value estimates derived from the cumulative distribution of Euclidean distances across 250,000 sound-interval pairs, obtained from 25 random pairs of zebra finch songs. Neighboring intervals that pass the p-threshold value (p = 0.1 in this study) form larger similarity segments (70 ms). The amount of sound from the tutor's motif that was included into the similarity segments represents the similarity score; it thus reflects how much of the tutor's song material was found in the pupil's motif. To measure how accurately pupils copied the sound elements of the tutor motif, we used the accuracy score from SAP. The accuracy score is computed locally, across short (9 ms) FFT windows and indicates how well the sound matched to the sound in the tutor song. SAP calculates an average accuracy value of the motif by averaging all accuracy values across the similarity segments. Syllable analysis—manual counting of imitated syllables. For manual counting of imitated syllable types, two individuals who were blind to treatment counted all syllables that matched a tutor syllable by visual inspection of sonograms. Their interobserver reliability was 80%. Syllable analysis—syllable acoustic features. We extracted the mean pitch, mean FM, mean entropy, and mean PG, as well as mean duration from 25 renditions of each syllable. To compare the similarity of individual spectral features between pupil and tutor syllables, we subtracted each mean feature value of each tutor syllable from the mean feature value of the corresponding pupil syllable. Next, we normalized the absolute differences between the values of tutor and pupil syllables to the values of the tutor syllable to obtain the difference of a pupil syllable in a given feature from the tutor syllable in percent. To describe the variability of syllable duration between different renditions, we calculated the coefficient of variation of duration values among 25 renditions of each syllable. Syllable analysis—syllable identity score. We quantified the acoustic similarity between different syllables using symmetric comparisons to obtain syllable identity scores. In contrast to asymmetric comparison, no similarity segments are identified during symmetric comparisons. Instead, the FFT windows are compared sequentially from beginning to the end of the two sounds. Thus, similarity reflects how many sound intervals were above p-value, and accuracy indicates the average (1 − p-value). To comprehensively capture the acoustic similarity between syllables in a single measure we used the product of similarity and accuracy to obtain the syllable identity score. As for the motif analysis the p-threshold value was set to p = 0.1. To quantify how accurately pupils learned individual syllables, we performed ten symmetric comparisons of each pupil syllable with its corresponding tutor syllable. To assess how variable the same pupil performed a particular syllable in multiple renditions of his motif, we compared 20 renditions of each syllable, two at a time. Because minute temporal shifting of FFT windows is allowed in symmetric comparisons (10 ms in this study), the more variable duration of syllables in FoxP2 knockdown animals did not bias the identity score. The syllable identity score rather reflects spectral differences between syllables. Syntax analysis. For each pupil, we manually annotated sequences of 300 user-defined syllables with the positions in their respective motifs. That is, each syllable of a motif was given a unique integer. Based on these data, we computed the Markov chain for each pupil, i.e., all transition probabilities between syllables. To measure the stereotypy of a motif, we calculated for each syllable the entropy of its transition distribution [52]. Because motif duration differed between birds, these entropy values were rescaled by the maximal possible entropy for each given motif duration. The entropy score for a pupil was then represented by the average of these fractions of maximal entropy over all syllables. Based on this entropy measure, we generated a sequence consistency score (1 − entropy measure), which reflects song stereotypy. An entropy score of 0 indicates random syllable order, whereas a score of 1 reflects a fixed syllable order. Analysis of song development. To determine tutor similarity and vocal variability during plastic song and towards the end of the learning phase, we analyzed songs recorded on PHD65, PHD80, and PHD90–93 (PHD ± 1 d; in one control pupil, recordings were only available from PHD75 instead of PHD80). First, all sound files from one day were segmented into sounds in the feature batch mode of SAP. Here, the pupils' vocalization is separated from nonvocalization background using two thresholds (Wiener entropy and amplitude). The thresholds were adjusted for each pupil individually to obtain an optimal segmentation. We validated the segmentation for each pupil by visual inspection of the segments and confirmed that segments correspond to syllables. Next, all segments from a given day (between 1,000 and 3,000 segments) were automatically compared to the tutor motif. That is, in each comparison, SAP identifies the best possible match to the tutor motif for each segment. Of all segments analyzed from PHD65, PHD80, and PHD90, 11.0% ± 0.9% were less similar to the tutor model than two random zebra finch sounds are to each other, and thus did not receive any accuracy value in SAP. These sounds were found to represent cage noise, mostly. There were no differences between the amount of sounds excluded between knockdown and control pupils for any of the ages (two-tailed Mann-Whitney U test, p > 0.9 for PHD65; p > 0.8 for PHD80; p > 0.7 for PHD90). Supporting Information Audio S1 Example of Song Motif from Tutor #414 (199 KB WMA) Click here for additional data file. Audio S2 Example of Song Motif from Pupil of Tutor #414 (123 KB WMA) Click here for additional data file. Audio S3 Example of Song Motif from Tutor #38 (166 KB WMA) Click here for additional data file. Audio S4 Example of Song Motif from Pupil of Tutor #38 (223 KB WMA) Click here for additional data file. Audio S5 Example of Song Motif from Tutor #396 (245 KB WMA) Click here for additional data file. Audio S6 Example of Song Motif from Pupil of Tutor #396 (370 KB WMA) Click here for additional data file. Figure S1 Timeline of Experiments By PHD20, fathers and older male siblings were removed from family cages to prevent experimental zebra finches from instructive auditory experience prior to the onset of tutoring. At the beginning of the sensory learning period at PHD23, virus was injected bilaterally into Area X. From PHD30 on, injected birds were housed individually in sound-recording chambers together with an adult male zebra finch as tutor. We recorded the song of pupils on PHD65, PHD80, and between PHD90 and 93 using an automated recording system, in absence of the tutor. (31 KB PDF) Click here for additional data file. Figure S2 Immunohistochemical Staining with the Neuronal Marker Hu Identified Virus-Infected Neurons Expressing GFP (A) shows neuronal marker Hu, and (B) shows virus-infected neurons expressing GFP. These neurons appear yellow in the merged image (C) (scale bar indicates 20 μm). (288 KB PDF) Click here for additional data file. Figure S3 Infection with shFoxP2-Virus Did Not Induce Apoptosis (A) We labeled apoptotic cells in 50-μm sagittal sections from PHD29 male zebra finch brains injected with shFoxP2 or shControl virus on PHD23. DNA double-strand breaks characteristic of apoptotic cells were detected using the TUNEL method, visualized with an Alexa568 secondary antibody (red). The filled white arrow points to a TUNEL-labeled cell not infected by shFoxP2-f. (B) The open white arrow points to a shFoxP2-infected cell expressing the viral reporter GFP, but showing no TUNEL labeling (A). (C) DAPI staining identifies cellular nuclei. The apoptotic cell (white arrow) contains fragmented DNA typical of apoptosis. (D) Overlay picture of (A–C). (E) As positive control for the TUNEL method, we treated a section adjacent to that shown in (A–D) for 10 min with DNAse to artificially induce DNA double-strand breaks. (E–H) Numerous cells were now detected, among them a virally infected cell expressing GFP (white arrow in [E–H]). Colors as in (A–D). Scale bar in (A) indicates 10 μm. (658 KB PDF) Click here for additional data file. Figure S4 Neuronal Densities Were Similar in Area X Injected with Either shFoxP2 or shControl Neuronal densities were measured using the neuronal marker Hu in Area X 30 d after injecting either shFoxP2-f/-h or shControl virus. Bar graphs represent the number of neurons/mm2. Neuronal densities in the virus-infected region in Area X were similar in knockdown and shControl-injected birds (two-tailed Mann-Whitney U test, p > 0.39; shControl, n = 4 hemispheres; shFoxP2-f/-h, n = 3 hemispheres). Moreover, there were no differences between inside and outside of the injection site for any of the viruses (two-tailed Mann-Whitney U test, p > 0.6 for both shFoxP2-f/-h and shControl). (54 KB PDF) Click here for additional data file. Figure S5 Syllables from Knockdowns and Control Zebra Finches Were Similar in the Distribution of their Acoustic Features and Their Duration Box plots represent the distribution of mean pitch (A), mean frequency (B), mean frequency modulation (FM) (C), mean entropy (D), mean goodness of pitch (PG) (E), and mean duration (F) across all syllables from tutors and each experimental group (shControl-, shGFP-, and shFoxP2-injected zebra finches). Boxes indicate the interquartile range (IQR) of the distribution; circles and asterisks specify individual values lying beyond the inner (1.5 × IQR) and outer fences (3 × IQR), respectively (n = 40 syllables for tutors; n = 31 syllables for shControl; n = 15 syllables for shGFP; and n = 31 syllables for shFoxP2). Mean syllable acoustic features and syllable duration (each averaged per animal) were not significantly different between groups (ANOVA; n = 6 tutors; n = 7 birds for shControl; n = 3 birds for shGFP; and n = 7 animals for shFoxP2-f/-h). (60 KB PDF) Click here for additional data file. Figure S6 Manual Counting of Syllables Copied by Knockdown and Control Animals All syllables that matched a tutor syllable by visual inspection on a sonogram were counted for shFoxP2- and shControl-injected animals. Bars represent the mean percentage of tutor syllables copied by the pupils (± STDEV, two-tailed Mann-Whitney U test, **p = 0.004; n = 7 animals for both shControl and shFoxP2-f/-h). (54 KB PDF) Click here for additional data file. Figure S7 Both Hairpin Constructs Targeting FoxP2 Affected Song Imitation to the Same Degree Bars indicate the similarity and accuracy scores, respectively, of zebra finches injected with either shFoxP2-f or shFoxP2-h (± SEM; two-tailed Mann-Whitney U test, p > 0.6 for similarity and p > 0.4 for accuracy). (54 KB PDF) Click here for additional data file. Figure S8 The Syllable Sequence within Motifs Was Highly Stereotyped across Many Different Renditions, Both in shFoxP2-Injected and shControl-Injected Birds This is reflected by high sequence consistency scores (two-tailed Mann-Whitney U test, no significant difference between shFoxP2-f/-h and shControl, p > 0.6; n = 7 animals for both shFoxP2-f/-h and shControl). The sequence consistency score (1 − entropy) was calculated based on the entropy of sequences of 300 successive syllables. (54 KB PDF) Click here for additional data file. Figure S9 Frequency Distribution of Syllable Accuracy Scores Was Shifted towards Lower Values in FoxP2 Knockdown Pupils Zebra finch vocalizations recorded at PHD65 were first segmented into sounds corresponding to syllables. The segments from each bird were subsequently compared to their respective tutor motif in a pairwise fashion yielding one accuracy score for each sound segment. To obtain a balanced dataset, we randomly extracted 800 accuracy scores from each bird. Bars represent the relative frequency of accuracy scores. (56 KB PDF) Click here for additional data file. Accession Numbers The GenBank (http://www.ncbi.nlm.nih.gov/Genbank) accession numbers for the genes and gene products discussed in this paper are FoxP1 (AY549152), FoxP2 isoform I (AY549148), FoxP2 isoform IV (AY549151), Hmbs (NM_013551), and Pfkp (NM_019703). The Online Mendelian Inheritance in Man (OMIM; http://www.ncbi.nlm.nih.gov/sites/entrez?db=OMIM) accession number for FOXP2 is 605317.
              Bookmark
              • Record: found
              • Abstract: found
              • Article: found
              Is Open Access

              Foxp2 Regulates Gene Networks Implicated in Neurite Outgrowth in the Developing Brain

              Introduction Forkhead-box protein P2 is a highly conserved vertebrate protein, belonging to an important group of transcription factors [1]. By modulating the expression of downstream target genes, forkhead-box proteins influence a diverse array of processes, including cell-cycle regulation, signal transduction, differentiation, patterning and metabolism [2]. They thereby play crucial roles during embryogenesis, in postnatal development and in the mature organism, and many have been linked to disease states [3]. The P subgroup is a divergent branch of forkhead-box proteins that share a distinctive DNA-binding domain located near the C-terminal end of the protein, as well as zinc-finger/leucine-zipper motifs that mediate dimerization, and a glutamine-rich region towards the N-terminus [4], [5]. Functional evidence from multiple species implicates Forkhead-box protein P2 in particularly intriguing aspects of brain development and function [1]. (Here we adopt the standard accepted nomenclature to refer to the protein in different species: FOXP2 in humans, Foxp2 in mice, FoxP2 in other chordates, with the corresponding gene names in italics [6].) In humans, damage to one copy of the FOXP2 gene causes a rare neurodevelopmental disorder, characterised by difficulties mastering sequences of mouth movements during speech, as well as impaired language processing [4], [7], [8]. Heterozygous disruptions of the mouse orthologue (Foxp2) yield dramatic reductions in synaptic plasticity of cortico-striatal brain circuits, associated with deficits in learning of rapid motor skills [9]. Mouse pups with homozygous Foxp2 mutations show more severe neural effects – gross motor impairments, delayed postnatal maturation of the cerebellum and dramatic reductions in emission of ultrasonic vocalisations – against a background of reduced weight-gain and postnatal lethality [9]-[11]. In addition, the avian orthologue (FoxP2) is required for normal vocal learning in songbirds [12], [13]. Selective knockdown of the gene in a key striatal nucleus in juvenile zebrafinches leads to incomplete and inaccurate imitation of tutor songs [14]. Studies of both human FOXP2 and mouse Foxp2 identified similarly strong CNS (central nervous system) expression during embryogenesis, which is confined to neurons (absent from glial cells) and enriched in various brain structures, including deep layers of the developing cortical plate, and parts of the striatum, thalamus and cerebellum [15],[16]. These embryonic expression patterns appear highly concordant in the different species, and show remarkable overlaps with sites of pathology identified by neuroimaging of human children and adults carrying FOXP2 mutations [16], [17]. Neural expression of the gene continues postnatally and into adulthood [4], [15], and is also observed in certain non-neural tissues, most notably the distal alveolar lung epithelium, and the outflow tract and atrium of the cardiovascular system [18]. The above observations of well-conserved and specific CNS expression patterns [15], [16] suggest that Foxp2 is likely to have important functions in neurodevelopment. Nevertheless, as data continue to accumulate regarding its impacts on the postnatal brain [9], [11], [14], the specific roles of Foxp2 in the developing CNS remain largely elusive. One route for gaining insights into the biological processes controlled by a transcription factor is to define the regulatory networks that are directly downstream of it [1]. An efficient strategy for identifying direct targets exploits chromatin immunoprecipitation (ChIP) methods to screen the tissue of interest [19]. Two previous investigations have coupled ChIP with hybridisation to promoter microarrays (ChIP-chip) in order to uncover binding sites of FOXP2 in human foetal brain tissue [20] and in human neurons grown in culture [21]. Both screens were of limited scope – the microarrays in these studies comprised fragments from the 5′ ends of ∼5,000 loci [20], [21], representing a small percentage of the known gene promoters in the genome. Neither study combined ChIP data with large-scale expression analyses. A more recent report used mRNA expression profiling in human neuronal models transfected with different versions of FOXP2 to explore regulatory differences between the human and chimpanzee orthologues, but did not include any ChIP screening for direct targets [22]. In the present study, we performed a systematic large-scale in vivo ChIP-chip screen of the embryonic mouse brain, coupling Foxp2-ChIP to high-density arrays with oligonucleotides tiled across >17,000 promoters. We robustly established the empirical significance of our ChIP results in wild-type brains by determining the null distribution of signals generated by matched control tissue from littermates that expressed no Foxp2 protein. Under strict empirical thresholds that minimised false positive signals, we isolated a set of 264 high-confidence in vivo targets. Gene ontology (GO) analyses of the ChIP-chip data, as well as genome-wide expression profiling and in situ hybridisations of wild-type and mutant mice, converged on neurite outgrowth as one of the most prominent biological themes associated with Foxp2 function in the embryonic CNS. We went on to directly demonstrate, using neuronal cell models and primary neurons from the embryonic mouse brain, that Foxp2 alters expression of neurite-outgrowth targets and thereby influences neurite process length and branch number. Results Genome-wide identification of in vivo Foxp2 targets in embryonic mouse brain In vivo Foxp2-ChIP screening was carried out using brains harvested from embryonic mice. Experiments were performed with mice that were wild-type for Foxp2, as well as homozygous littermates that do not express any Foxp2 protein (Foxp2-S321X mutants; see Materials and Methods) [9]. The different types of sample were screened in parallel, undergoing identical experimental manipulations and data processing. In this context, the homozygous mutant mouse tissue acts as an ideal negative control [21]. Since such samples completely lack Foxp2 protein (see Figure S1 and [9]), fragments that are pulled down by Foxp2-ChIP in these cases give an unbiased empirical indication of background noise and false positive rates yielded by the procedure. Whole mouse brains from wild-type or mutant mice were harvested at embryonic day 16 (E16), corresponding to a timepoint at which particularly high levels of Foxp2 expression are observed in the developing CNS [16]. Chromatin isolated by Foxp2-ChIP was labelled and hybridised to DNA microarrays covering the promoter regions of ∼17,000 mouse transcripts (Agilent Technologies), using total input DNA as a reference sample. Each promoter on these arrays is represented by an average of twenty-five 60-mer probes spanning ∼5.5 kb upstream and ∼2.5 kb downstream of the transcription start site, allowing peak regions of binding to be precisely defined (Figure 1). Moreover, the presence of multiple probes for each promoter scattered throughout the array gives independent enrichment values within the same promoter, which aids discrimination of real biological targets from false positive events. Specifically, since the shearing process during ChIP produces overlapping fragments of chromatin, true targets should show evidence of enrichment for multiple probes across the promoter region, while promoters with only a single enriched probe are most likely to be false positive results. 10.1371/journal.pgen.1002145.g001 Figure 1 In vivo Foxp2 promoter occupancy in embryonic mouse brain. (A) Foxp2-ChIP window enrichment scores of probes across promoters of a subset of putative targets from the neurite outgrowth and axon guidance pathways. The window score (Y axis) is given versus the distance across the promoter region (X axis) - each cross bar represents 1000 bp and each data point represents a single probe (chromosome and position in bp are given below the X axis of each graph). The enrichment in the wild-type experiments is shown by the blue trace, and the pink trace indicates the corresponding values in null mutant controls that lack Foxp2 protein. The predicted start site of the gene (as annotated on UCSC genome browser: http://genome.ucsc.edu/) is given by the black box and arrows denote the direction of transcription. Grey shading indicates the peak area of enrichment and the most likely region for Foxp2 binding to occur. (B) Analysis of in vivo promoter occupancy. DNA isolated via Foxp2-ChIP was PCR amplified using primers directed towards the promoter regions of Nrp2, Sema3a, Nrn1 or the β-actin control. The position of these amplicons within the target promoter region is given by red bars in part A. Results from E16 wild type mice were compared to those from homozygous null mutant littermates. Lane 1 =  wild-type Foxp2-ChIP, lane 2 =  mutant null Foxp2-ChIP, lane 3 =  wild-type total DNA, lane 4 =  mutant total DNA. Target gene promoters were found to be specifically enriched in Foxp2-ChIP samples isolated from wild-type brains compared to those from null mutant brains, unlike the β-actin control promoter. Gels are representative of results from triplicate experiments. In order to identify enriched promoters, Foxp2-ChIP data were analysed as per Materials and Methods. Briefly, array data from independent biological replicates (three independent ChIP experiments hybridised to one each of three array sets) were normalised for each genotype (wild-type or mutant control) separately. Normalised array data (excluding probes with a negative average enrichment across replicate experiments) were subjected to a sliding window analysis, using a similar method to that employed in genome-wide ChIP-chip studies of other forkhead transcription factors [23]. Each probe was assigned a value (window-adjusted score) based on the median fold enrichment of itself and its neighbouring probe on either side (within 500 bp upstream and 500 bp downstream), and then probes were ranked based on this window score. By analysing the distribution of window scores observed in the mutant null control experiments we were able to derive an empirical threshold for significance, which could then be applied to the wild-type data. We found that window scores greater than or equal to 0.974 (corresponding to ∼2-fold enrichment) excluded 99% of the data-points in the mutant null control experiments. When we applied this threshold to data from wild-type experiments, we identified a set of 1,217 promoter regions that were consistently enriched by Foxp2-ChIP over 3 replicates in wild-type mouse brains (Table S1). On inspection of the locations of the enriched probes throughout the mouse genome, no positional bias was observed (Figure S2). Since some of the enriched regions lay close to the transcriptional start site (TSS) of more than one gene, the 1,217 promoter regions corresponded to 1,253 genes. Of note, using the same analysis parameters, only 147 genes were enriched in the mutant null controls, suggesting a low false discovery rate. Nevertheless, in order to minimize false-positive findings, we excluded any enriched genes from the wild-type dataset that also had window scores exceeding the 99% threshold in the mutant control dataset. This filtering process yielded a slightly smaller set of 1,164 putative targets (Table S2). When we applied stricter thresholds to the wild-type data, selecting only those promoters in which at least one probe gave a window score of ≥ 1.5, we identified a shortlist of 259 promoter regions. Since a small number of peak regions lay directly between the TSSs of two different genes, these 259 promoters corresponded to a slightly higher total of 266 genes. Crucially, the same analyses of the entire mutant null control dataset identified only a single gene in the genome with a window score of ≥ 1.5 (the Pigt gene), indicating an extremely low rate of false positives under these stricter selection criteria. We excluded two genes from the strict wild-type shortlist (Pigt and Zfp496) since they contained probes that exceeded the 99% threshold (i.e. window score of >0.974) in mutant null controls (Figure S3). The outcome of these analyses was a final curated shortlist of 264 high-confidence in vivo targets (Table S3). Given that DNA is sheared randomly during the ChIP process, we would expect a true Foxp2 binding event to be represented by a peak of enrichment at a target promoter. This peak would result from the sheared DNA forming a series of overlapping fragments, with the region closest to the binding site showing the highest degree of enrichment (i.e. highest number of fragments pulled down during immunoprecipitation) and with progressively less enrichment observed as the distance to the binding site increases on either side. Figure 1A gives typical examples of the enrichment peaks observed for putative targets from our Foxp2-ChIP dataset. Examination of corresponding data from mutant control experiments emphasises the relative lack of enrichment in nulls that lack Foxp2 protein, indicating that the enrichment in wild-type samples results from highly specific Foxp2-mediated interactions. Furthermore, we followed up a subset of candidates with qPCR, consistently confirming their enrichment (Figure 1B). Enriched regions represented in the shortlist of high-confidence targets were assessed in silico for any over-represented sequence motifs (see Text S1). This analysis did not enforce a priori conditions of motif sequence, other than a length restriction of 8 bases. This meant that rather than limiting our search to occurrences of known patterns in the promoters, we obtained an unbiased list of motifs that were characteristic of the Foxp2-ChIP target promoters. Eight sequences (motifs A-H) were found to be significantly over-represented (p 0.974) are highlighted on the WebGestalt KEGG pathway in red, with those from expression analysis (p 15 images taken for three independent clones of each cell type (Scale bar, 100 µm). To further assess the in vivo relevance of these findings, we examined whether there were corresponding phenotypic effects mediated by functional Foxp2 in neurons of the developing brain. We isolated primary neurons from the ganglionic eminences of E16 mouse brains, matching the region and timepoint used for our original target screening. Here, we aimed to directly test whether the Foxp2-positive neurons derived from the developing basal ganglia show altered neurite outgrowth when the gene is mutated. The assay was facilitated by availability of a mouse model (Foxp2-R552H) in which the protein is expressed at normal levels, but is nevertheless dysfunctional [9]. R552H mice recapitulate an aetiological mutation that causes speech and language deficits in a large human family. This change yields a substitution in the DNA-binding domain which severely impairs the transcription factor function of the mutant protein [30], such that the overall phenotype of homozygous R552H animals is very similar to that observed for mice which completely lack Foxp2 [9]–[11]. However, unlike the Foxp2-null mice, R552H homozygotes still express detectable levels of the protein, allowing us to clearly identify Foxp2-positive cells in our primary cultures via antibody staining. This represents an important measure, given the heterogeneous nature of the dissected material used to generate the primary culture. We again observed obvious differences in neurite outgrowth associated with presence of functional Foxp2 (Figure 6A). A blinded analysis revealed statistically significant changes in quantitative measures of neurite outgrowth for Foxp2-expressing neurons from wild-type embryos as compared to those from homozygous Foxp2-R552H littermates (Figure 6B). In particular, the latter showed significant reductions in total outgrowth (p = 0.001); mean (p 1.5) in wild-type Foxp2-chip. (TIF) Click here for additional data file. Figure S4 EMSA analysis of a novel Foxp2 DNA-binding motif. (TIF) Click here for additional data file. Figure S5 Differential expression of Foxp2 target genes in E16 developing mouse ganglionic eminences. (TIF) Click here for additional data file. Figure S6 Co-expression of Foxp2 and putative target genes in the E16 mouse brain. (TIF) Click here for additional data file. Figure S7 Expression of Foxp2 in medium spiny neurons (MSNs). (TIF) Click here for additional data file. Figure S8 In vivo binding of mouse Foxp2 to intron 1 of Cntnap2 in embryonic brain tissue. (TIF) Click here for additional data file. Table S1 Foxp2-ChIP (window score ≥0.97) complete dataset. (XLS) Click here for additional data file. Table S2 Foxp2-ChIP putative target promoters (curated long list n = 1164, window score ≥0.97). (XLS) Click here for additional data file. Table S3 Foxp2-ChIP putative target promoters (curated short list n = 264, at least 1 window score ≥1.50). (XLS) Click here for additional data file. Table S4 Foxp2-dependent gene expression changes in the E16 mouse brain (p<0.01). (XLS) Click here for additional data file. Table S5 Overlapping target genes from Foxp2-ChIP and expression analysis studies. (XLS) Click here for additional data file. Table S6 Significant over-represented KEGG pathways identified in the Foxp2-ChIP and gene expression datasets. (XLS) Click here for additional data file. Table S7 Cntnap2 semi-quantitative PCR primers. (XLS) Click here for additional data file. Text S1 Supporting Materials and Methods. (DOC) Click here for additional data file.
                Bookmark

                Author and article information

                Contributors
                Journal
                Front Neurosci
                Front Neurosci
                Front. Neurosci.
                Frontiers in Neuroscience
                Frontiers Media S.A.
                1662-4548
                1662-453X
                25 January 2017
                2017
                : 11
                : 19
                Affiliations
                [1] 1Max Planck Institute for Psycholinguistics Nijmegen, Netherlands
                [2] 2Department of Cognitive Neuroscience, Radboud University Medical Center, Donders Institute for Brain, Cognition, and Behaviour Nijmegen, Netherlands
                Author notes

                Edited by: Christian Hansel, University of Chicago, USA

                Reviewed by: Constance Scharff, Free University of Berlin, Germany

                *Correspondence: Dirk Schubert d.schubert@ 123456donders.ru.nl
                Article
                10.3389/fnins.2017.00019
                5263127
                afb5b3f0-d5b3-440c-9298-f4ca1ff3b86c
                Copyright © 2017 Negwer and Schubert.

                This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) or licensor are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

                History
                : 27 May 2016
                : 10 January 2017
                Page count
                Figures: 1, Tables: 0, Equations: 0, References: 44, Pages: 4, Words: 3179
                Categories
                Neuroscience
                Frontiers Commentary

                Neurosciences
                foxp2,retinoic acid,motor circuits,speech disorder,striatum,dopamine,rarb,globus pallidus
                Neurosciences
                foxp2, retinoic acid, motor circuits, speech disorder, striatum, dopamine, rarb, globus pallidus

                Comments

                Comment on this article