ABSTRACT
The hippocampus is associated with essential brain functions, such as learning and memory. Human hippocampal volume is significantly greater than expected compared with that of non-human apes, suggesting a recent expansion. Intermediate progenitors, which are able to undergo multiple rounds of proliferative division before a final neurogenic division, may have played a role in evolutionary hippocampal expansion. To investigate the evolution of gene regulatory networks underpinning hippocampal neurogenesis in apes, we leveraged the differentiation of human and chimpanzee induced pluripotent stem cells into TBR2 (or EOMES)-positive hippocampal intermediate progenitor cells (hpIPCs). We found that the gene networks active in hpIPCs are significantly different between humans and chimpanzees, with ∼2500 genes being differentially expressed. We demonstrate that species-specific transposon-derived enhancers contribute to these transcriptomic differences. Young transposons, predominantly endogenous retroviruses and SINE-Vntr-Alus (SVAs), were co-opted as enhancers in a species-specific manner. Human-specific SVAs provided substrates for thousands of novel TBR2-binding sites, and CRISPR-mediated repression of these SVAs attenuated the expression of ∼25% of the genes that are upregulated in human intermediate progenitors relative to the same cell population in the chimpanzee.
Keywords: Hippocampal development, Transposable elements, Induced pluripotent stem cells, Intermediate progenitors, TBR2, SINE-Vntr-Alus
Highlighted Article: Evolution of human and chimpanzee hippocampal development was mediated by the co-option of young retrotransposons into species-specific enhancers.
INTRODUCTION
The hippocampus is associated with many traits relevant in the context of human evolution. These include traits such as tool use and language, which require social cognition and learning, as well as spatial memory, navigation and episodic memory (Burgess et al., 2002; Eichenbaum, 2017a,b; Squire, 1992; Tomasello and Herrmann, 2010). This region of the brain is also greatly affected by Alzheimer's Disease (AD), a neurodegenerative disorder characterized by cell death, plaques and tangles of misfolded proteins, and cognitive decline (Duyckaerts et al., 2009). It has been hypothesized that the cognitive AD phenotype is uniquely human and that non-human primates, including chimpanzees, do not exhibit AD-related dementia (Edler et al., 2017; Finch and Austad, 2015; Walker and Jucker, 2017). If humans are uniquely susceptible to AD, it is crucial to understand how the human hippocampus differs from that of our closest biological relatives, the chimpanzees.
Human hippocampal volume is 50% greater than expected compared with the hippocampal volumes of non-human apes, possibly indicating a recent hippocampal expansion specific to the human lineage (Barger et al., 2014). However, the evolution of the human hippocampus and the developmental mechanisms driving the human-specific volume increase have not yet been thoroughly studied.
Recent studies have suggested that evolutionary changes to neuronal progenitors may have had an impact on cortical volume in primates by increasing proliferative potential (Martínez-Cerdeño et al., 2006; Rétaux et al., 2013; Florio and Huttner, 2014). A specific class of neuronal progenitors known as intermediate progenitor cells (IPCs) or ‘transit-amplifying cells’ are able to undergo multiple rounds of proliferative division before a final neurogenic division (Englund, 2005; Arnold et al., 2008; Pontious et al., 2008; Hevner, 2019). These cells express the neurodevelopmental transcription factor TBR2 (EOMES) and are found in the sub-ventricular zone of the developing neocortex and hippocampus (Bulfone et al., 1999; Kimura et al., 1999; Englund, 2005; Cipriani et al., 2016). Genetic ablation of TBR2 in these progenitors results in reduced cortical thickness in mice (Sessa et al., 2008), abnormal cortical cell differentiation (Mihalas et al., 2016) and impaired neurogenesis in the hippocampal formation (Hodge et al., 2012). As these IPCs are hypothesized to play a role in neocortical expansion, they may also be involved in the lineage-specific hippocampal expansion seen in humans.
Many of the differences between humans and chimpanzees are due to diverging gene regulatory sequences (Agoglia et al., 2021; Enard et al., 2002; Gokhman et al., 2021; King and Wilson, 1975; Wray, 2007). A recent study comparing gene expression in adult human, chimpanzee and macaque brain regions identified several genes specifically upregulated in the human hippocampus (Sousa et al., 2017). However, transcriptomic differences between primate species during specific time points of hippocampal development have not been investigated.
As samples of developing human and chimpanzee brain tissue are extremely limited, induced pluripotent stem cells (iPSCs) are an ideal system to conduct comparative studies of human and chimpanzee hippocampal development. Previous studies have employed iPSC-derived cortical organoids and iPSC-derived neuronal progenitor cells (NPCs) from human and chimpanzee for comparative and developmental genomic purposes (Marchetto et al., 2019; Mora-Bermúdez et al., 2016). Here, we leverage human and chimpanzee iPSC-derived hippocampal progenitors as models for comparative developmental and genomic studies with the goal of identifying species-specific differences in gene regulation during hippocampal development.
Several papers have recently demonstrated that transposable elements (TEs) can alter existing regulatory elements or generate entirely novel ones, as well as expand in a species- or lineage-specific manner (Sundaram et al., 2014; Playfoot et al., 2021; Okhovat et al., 2020; Marnetto et al., 2018; reviewed by Sundaram and Wysocka, 2020). Species-specific TE expansion and co-option into gene regulatory networks have been demonstrated as a mechanism for evolutionary change (Chuong et al., 2016; Fuentes et al., 2018; Jacques et al., 2013; Lynch et al., 2011, 2015; Miao et al., 2020; Mika et al., 2021; Pontis et al., 2019; Trizzino et al., 2017). Endogenous retroviruses (ERVs) and SINE-Vntr-Alus (SVAs) are among the TE families more frequently associated with gene regulatory activity in the human genome (Chuong et al., 2016; Fuentes et al., 2018; Pontis et al., 2019; Trizzino et al., 2017, 2018). SVAs encompass six subfamilies, denoted as SVA-A through SVA-F. Of the ∼3000 SVA copies in the human genome, nearly half are human specific, including those belonging to the SVA-E and SVA-F subfamilies (Quinn and Bubb, 2014; Wang et al., 2005). The remaining half are also found in other great apes. SVAs are still replication competent and thus able to transpose in the human genome. ERVs are retrotransposons belonging to the long terminal repeat (LTR) group. They are remnants of past retroviral infection events and make up ∼8% of the human genome (Tokuyama et al., 2018). Both SVAs and ERVs were recently found to be enriched within the sequences of active cis-regulatory elements (enhancers and promoters) in hippocampal tissue compared with other human brain regions in which they are predominantly repressed (Trizzino et al., 2018). Therefore, we hypothesize that species-specific ERV and SVA transposon activity may influence the gene regulatory networks necessary for human and chimpanzee hippocampal development.
Given the key function that IPCs played in the evolution of the primate brain (Florio and Huttner, 2014; Martínez-Cerdeño et al., 2006), we sought to identify molecular differences between iPSC-derived human and chimpanzee hippocampal intermediate progenitor cells (hpIPCs) in terms of gene expression and the regulatory activity of non-coding regions. We specifically examined gene expression [by RNA sequencing (RNA-seq) and single-cell RNA-seq (scRNA-seq)], gene regulation [by assay for transposase-accessible chromatin using sequencing (ATAC-seq)] and functional TE activity via CRISPR interference.
We leveraged scRNA-seq to examine the temporal trajectory of the hpIPCs during differentiation in both species. After confirming that the hpIPC differentiated cells express the appropriate neurodevelopmental markers, we conducted a transcriptomic comparison between human and chimpanzee hpIPCs. This analysis revealed over 2500 differentially expressed genes (DEGs). We then used ATAC-seq to conduct extensive analyses of differential chromatin accessibility between human and chimpanzee hpIPCs. In both species, differentially accessible (DA) chromatin regions were more likely than expected to overlap a TE insertion. Furthermore, these regions were found to be both enriched and depleted for specific TE families. Notably, species-specific enrichment for ERV and SVA sequences within differentially accessible genomic sites correlated with species-specific changes in nearby gene expression. This is likely driven by transcription factors binding to the TE-derived regulatory sequences, as we demonstrate for TBR2 and SVA-derived enhancers. Finally, we used CRISPR interference to repress all the accessible SVAs in progenitor-like cells and demonstrated that such repression results in global changes in gene expression and affects hundreds of important neurodevelopmental genes.
This work demonstrates that two young TE families have contributed significantly to the gene regulatory differences between human and chimpanzee hippocampal development, providing insight into how the human hippocampus evolved both its unique cognitive capacity and its susceptibility to neurodegenerative disease.
RESULTS
An iPSC-derived model for human and chimpanzee hpIPCs
We modeled hpIPCs in humans and chimpanzees using three human and three chimpanzee iPSC lines. All six iPSC lines were validated as pluripotent in previous studies (Gallego Romero et al., 2015; Pagliaroli et al., 2021; Pashos et al., 2017; Ward et al., 2018; Yang et al., 2015; Zhang et al., 2015). For both human and chimpanzee iPSCs, we used two female cell lines and one male line.
We used a previously published method to generate hpIPCs from iPSCs (Yu et al., 2014). In this protocol, the stem cells are treated with a medium containing anticaudalizing factors and sonic hedgehog antagonists (DKK1, Noggin and SB431542) to generate forebrain progenitor cell types (Yu et al., 2014). It should be noted that the hpIPCs are distinct from a more general neuronal progenitor cell (pan-NPC) as the pan-NPC medium is supplemented only with FGF2 and B27. Moreover, the hpIPCs can be further induced to generate mature hippocampal CA3 pyramidal or dentate gyrus granule neurons (Sarkar et al., 2018; Yu et al., 2014).
As we were specifically interested in TBR2-positive hpIPCs, we differentiated one human male and one chimpanzee male iPSC line to ensure that differentiation proceeded similarly in each species-specific cell line and we compared the transcriptomes by conducting scRNA-seq at 24 h intervals from the iPSC stage (day 0) to the hpIPC stage (day 5). We assayed a total of 18,935 cells, with 4540 chimpanzee cells and 14,395 human cells. The scRNA-seq data demonstrate that the differentiation follows a similar trajectory in both species (Fig. 1A). Despite the difference in final cell number between each species-specific cell line (Fig. S1A), all six time points overlap closely between the species (Fig. S1B,C). Chimpanzee cells from days 2 and 3 overlap more than expected with human cells from days 3 and 4, respectively, suggesting that the chimpanzee cells are further along in differentiation to hpIPCs at these time points. However, both species align closely again by day 5, which is the time point chosen for all the genomic analyses conducted in the present study.
We further leveraged the scRNA-seq data to examine the expression of known neurodevelopmental markers and saw a progressive increase in the expression of both PAX6 and OTX2 during the 5-day differentiation (Fig. 1B,C), which remained consistent between species (Fig. S1D,E). To investigate any noticeable changes to the cell cycle or proliferation between species, we quantified the number of cells at the G1, S and G2/M phases of the cell cycle within the scRNA-seq dataset based on a set of known cell cycle genes and did not detect significant differences (Fig. 1E).
To ensure that the differentiation resulted in hpIPCs specifically, we conducted western blotting for samples from both species-specific cell lines and confirmed that TBR2 expression in human cells was greatest at days 4 and 5 (Fig. 1D), similarly to OTX2 and PAX6 expression (Fig. 1B,C). In chimpanzee cells, TBR2 was expressed as early as day 2, corroborating evidence from the scRNA-seq data that chimpanzee differentiation initially proceeds faster. However, we did not see evidence that TBR2 expression in chimpanzee cells was greater than TBR2 expression at day 5 in human cells. As OTX2 labels neuronal progenitors and TBR2 and PAX6 mark the intermediate progenitor cell type (Florio and Huttner, 2014; Cipriani et al., 2016; Hevner, 2019), the data demonstrate that the differentiation was successful and comparable between species. Lastly, we conducted chromatin immunoprecipitation (ChIP) followed by sequencing (ChIP-seq) against TBR2 and the active histone mark H3K27ac to determine whether TBR2 was bound at the same genomic sites in both species (Fig. 1F). We identified 3789 TBR2-bound regions in human hpIPCs and observed that nearly all of them also exhibited H3K27ac. Upon translating the coordinates of these regions to the chimpanzee genome, we observed that the 3781 orthologous regions were bound by TBR2 in chimpanzee hpIPCs. Together, these data indicate that our iPSC-derived model is suitable to study gene regulation within TBR2-positive hpIPCs from both human and chimpanzee.
Important neurodevelopmental genes are differentially expressed between human and chimpanzee hpIPCs
To investigate the differences between human and chimpanzee hpIPCs, we first aimed to characterize differential gene expression between the hpIPCs of the two species. After 5 days of treatment with the hpIPC differentiation medium, we collected cells for RNA extraction (Fig. 2A). To ensure statistical power, we conducted bulk RNA-seq on two replicates each of all six cell lines (i.e. three biological replicates and six technical replicates per species). The differentiation, harvesting and RNA processing were performed in mixed batches with samples from both species to prevent batch effects. The libraries were sequenced using an Illumina NextSeq500, generating 100 bp paired-end reads. Non-orthologous genes were omitted from the analysis and a total of 2588 genes were identified as being differentially expressed [false discovery rate (FDR)<0.05 and log2(fold change)>1.5 or <−1.5]. The genes with log2(fold change)>1.5 were more highly expressed in the human hpIPCs (‘Human UP’), whereas the genes with log2(fold change)<−1.5 were more highly expressed in the chimpanzee hpIPCs (‘Chimp UP’). In total, 1686 (65.1%) of the DEGs were ‘Human UP’ and 901 (34.9%) of the DEGs were ‘Chimp UP’ genes (Fig. 2B; Table S1).
Hippocampal neurodevelopmental markers PAX6, OTX2, NEUROD1 and FOXG1 were found to be highly expressed in both species (Figs 1B,C, 2B).
The ‘Human UP’ genes include FOXP2, MTRNR2L8, DHX40, VPS13B, WDFY2 and PURB, all of which are associated with neurodevelopment or neurodegenerative disease (Abrajano et al., 2009; Hickey et al., 2019; Kamboh et al., 2019; Kolehmainen et al., 2003; MacDermot et al., 2005; Mathys et al., 2019; Sin et al., 2015; Taher et al., 2014). The ‘Chimp UP’ genes include HIST1H3A, BCL2L2 and CLIC1, which have been reported to be expressed in the hippocampus and are associated with sleep, AD and neurite outgrowth (Averaimo et al., 2014; Datson et al., 2009; Wei, 2020).
To understand the transcriptional programs driving these differences in gene expression, we conducted an Ingenuity pathway analysis. Three of the top five upstream regulators predicted by the pathway analysis were the transcription factors CREB1, FOXA2 and TBR2 (Fig. 2C). CREB1 is known to regulate genes involved in the nervous system and neurodevelopment (reviewed by Sakamoto et al., 2011), FOXA2 controls dopaminergic neuronal development and disease (Kittappa et al., 2007), and TBR2 plays a crucial role in cortical and hippocampal neurogenesis and is the signature marker of the intermediate progenitor cell population (Cipriani et al., 2016; Englund, 2005). This pathway analysis was consistent with the RNA-seq data as predicted targets of all three transcription factors were also found to be among the 2588 genes differentially expressed between human and chimpanzee hpIPCs (Fig. 2D; Tables S2-S4). The pathway analysis also determined that several of the DEGs are involved in embryonic development (Fig. 2D; Table S5). Overall, these findings indicate that previously characterized neurodevelopmental gene regulatory networks are utilized differently during human and chimpanzee hippocampal development.
Human-specific chromatin accessibility patterns in hpIPCs
After identifying DEGs and the transcriptional networks that may be involved, we sought to identify cis-regulatory differences between the human and chimpanzee hpIPCs. To this end, we conducted ATAC-seq on the hpIPCs from both species. We used the same batches of differentiated iPSCs for the ATAC-seq as we did for the RNA-seq (i.e. from the same batch of differentiation) and generated 100 bp long paired-end Illumina reads.
We first performed a human-centric analysis. We aligned the ATAC-seq reads from all six cell lines to the respective reference genome assemblies (hg19 for the human cell lines, panTro5 for the chimpanzee cell lines) and only retained uniquely mapped reads with high mapping quality (Samtools q=10 filtering). Next, we identified regions of accessible chromatin (peaks; FDR<0.05) in all three human cell lines. Only peaks replicated in all the three human lines were retained. To carry out a proper comparison, we only retained replicated human ATAC-seq peaks with orthologs in the chimpanzee genome (see Materials and Methods; Fig. 3A). This filtering ultimately resulted in 82,235 human ATAC-seq peaks that were replicated in all the human cell lines and with orthologs in the chimpanzee genome. These 82,235 regions were used for downstream analysis. We found that the chromatin accessibility at these regions was highly reproducible across all three human cell lines (Fig. 3B).
Next, we quantified the ATAC-seq read depth for each of the 82,235 regions and used DESeq2 to identify sites exhibiting differential chromatin accessibility between the two species. In total, we identified 3006 DA regions [FDR<0.05; log2(fold change)>1.5 or <−1.5; Tables S6 and S7]. Of these regions, 92.3% were located at least 1 kb away from the closest transcription start site (TSS), suggesting that they could be putative enhancers, whereas the remaining were putative promoters (Fig. 3C). As expected, given that this analysis was performed with a human-centric approach, 90.1% of the DA peaks were significantly more accessible in the human hpIPCs relative to chimpanzee hpIPCs (‘Human UP’; Fig. 3C).
As TE insertions can be a source of cis-regulatory evolution, we examined whether the DA regions were more likely to overlap with a TE than those that were accessible to the same degree in both species (non-DA; Fig. 3D; Tables S9 and S10). We observed that 1335/3006 (44.4%) DA regions overlapped a TE (Table S8). This is significantly higher than what was observed for the non-DA peaks (33.2% overlapped a TE; two-sided Fisher's exact test, P<0.0001; Fig. 3D). This indicates that chromatin regions with human-specific accessibility are significantly more likely to be TE-derived than the regions with accessibility levels conserved between human and chimpanzee.
Next, we associated the nearest gene to each DA region and found that TE-derived DA regions were significantly more likely to be near a DEG relative to non-TE derived DA regions (two-sided Fisher's exact test, P=0.016; Fig. 3E). Finally, we investigated whether specific TE families were overrepresented among the TE-derived DA regions and found enrichment for LTRs. Although LTRs account for ∼16% of all human annotated TEs, they represented 33.9% of the TEs overlapping DA regions in our human-centric analysis (two-Sided Fisher's exact test, P<0.0001; Fig. 3F; Table S11). Of these enriched LTRs, 97.1% were ERVs. Notably, 31.7% of the LTR-derived DA regions were located near a DEG (two-sided Fisher's exact test P<0.0001).
Taken together, these data indicate that there are TE-derived cis-regulatory elements that have significantly greater accessibility in humans than in chimpanzees during hippocampal neurogenesis. These TE insertions preceded the human-chimpanzee split, but the difference in accessibility is species specific, suggesting that the co-option into gene regulatory networks took place after the human-chimpanzee divergence.
Chimpanzee-specific chromatin accessibility patterns in hpIPCs
We repeated the ATAC-seq analysis as described above, but this time with a chimpanzee-centric approach. We started from a set of 72,211 peaks found to be replicated in all the three chimpanzee lines and with orthologs in both species (Fig. 4A,B). With this approach, we identified 3806 ATAC-seq peaks as being DA between human and chimpanzee [FDR<0.05; log2(fold change)>1.5 or <−1.5; Tables S16 and S17], 82% of which displayed greater accessibility in chimpanzee compared with humans (i.e. ‘Chimp UP’; Fig. 4C). Similar to what we observed with the human-centric analysis, 97.9% of the 3806 DA regions were putative enhancers (distance>1 kb from TSS; Fig. 4C).
As seen in the human-centric analysis, chimpanzee-specific DA peaks were more likely to be TE derived than those that were similarly accessible across species (two-sided Fisher's exact test, P<0.0001; Fig. 4D; Tables S18-S20). Of the chimpanzee DA regions, 37.1% overlapped an annotated chimpanzee TE, compared with only 28.1% of the non-DA regions (two-sided Fisher's exact test, P<0.0001; Fig. 4D). However, the TE-derived enhancers in this chimpanzee-centric analysis were no more likely to be located near DEGs than the non-TE derived ones (Fig. 4E).
We found enrichment for LTRs, which account for approximately 16% of the chimpanzee TEs but represented 36.1% of the TE-derived DA regions (two-sided Fisher's exact test, P-value<0.0001; 98.2% were ERVs; Fig. 4F; Table S21), and SVAs, which account for just 0.25% of annotated chimpanzee TEs but represented 1.1% of the TE-derived DA regions (two-sided Fisher's exact test, P<0.0001, Fig. 4G; Table S23). In particular, the SVA-B and SVA-C subfamilies were the most enriched (Fig. 4H). Taken together, these data indicate that chimpanzee ERV and SVA transposons were co-opted into regulatory elements important for the developing chimpanzee hpIPCs.
Genomic features underlying species-specific LTR enrichment at hpIPC enhancers
We aimed to further investigate genomic features potentially underlying the LTR enrichment among the DA hippocampal progenitor regions. To this end, we conducted a motif analysis using the MEME suite (Bailey et al., 2015). Binding motifs for CTCF and EGR2 were detected as being enriched in the LTR-derived DA regions identified from both the human-centric and chimpanzee-centric analyses (Fig. 5A,C). EGR2 is an early response gene involved in learning and memory, in the brain response to stimuli and in hippocampal synaptic plasticity (Cheval et al., 2012; Mukherjee et al., 2021; Poirier et al., 2007). CTCF, a well-known regulator of chromatin structure, has been implicated in various neurodevelopmental disorders (reviewed by Davis et al., 2018).
We identified 63 DEGs located near the human-enriched LTRs (Fig. 5B; Table S12). These included GLUL, a glutamine synthetase hypothesized to provide neuroprotection in AD patients, (Kohane and Wood, 2021 preprint) and DLK1, a Notch ligand involved in sub-ventricular zone neurogenesis (Ferrón et al., 2011), both of which are upregulated in human hpIPCs compared with chimpanzee hpIPCs.
We identified 34 DEGs located near the chimp-enriched LTRs (Fig. 5D; Table S22). For the chimp-centric analysis, the DEGs located near TE-derived enhancers with species-specific accessibility included HIST1H3A and MTRNR2L8, which were highly upregulated in the chimpanzee and human hpIPCs, respectively (Fig. 5D). Importantly, HIST1H3A has been associated with autism spectrum disorders and sleep deprivation (Crawley et al., 2016; Wei, 2020), whereas MTRN2L8 has been reported as being upregulated in AD patients (Mathys et al., 2019).
Human-specific SVAs play a major role in hippocampal neurogenesis
As mentioned earlier, the chimpanzee-centric analysis also identified SVA transposons as being enriched within chimpanzee-specific enhancers (Fig. 6A; Table S23). These SVAs were enriched for the identified binding motifs of the neurodevelopmental factors ASCL1, ZIC1 and KLF8 (Andersen et al., 2014; Aruga, 2004; Yi et al., 2014), as well as the JUN/FOS-AP-1 dimer, which is a known enhancer activator (Raivich, 2008; Raivich and Behrens, 2006) (Fig. 6A).
It is important to note that all the analyses shown so far were exclusively based on genomic sites with characterized orthologs in both species, to ensure an equivalent comparison. However, nearly 2000 SVA copies, including the entire SVA-E and SVA-F subfamilies, are exclusive to the human genome. Given that previous studies found that SVAs are highly enriched in active enhancers and promoters of the human hippocampus (Trizzino et al., 2018), we sought to investigate this further. We focused on the SVA copies exclusively present in the human genome and not in any other primate genome (hereafter human-specific SVAs). We performed sequence-based motif analysis for all the human-specific SVAs and identified the binding motif for the hpIPC signature factor TBR2 as the most enriched (P=10−2103; Fig. 6B). Motifs for other transcription factors associated with hippocampal neurogenesis and function were also recovered (SMAD3, VDR and PLAG1; Fig. 6B).
Next, we annotated all genes found within 50 kb from each human-specific SVA. A total of 2216 genes were recovered using this approach. We performed pathway analysis on this set of genes and found that ‘melatonin degradation’ and ‘nicotine degradation’ were the two most significantly enriched pathways (P=7.6×10−6 and P=2.5×10−5, respectively; Fig. 6C) and PHF8 was recovered as the top upstream regulator for the gene network (Fig. 6C). Notably, both melatonin and nicotine degradation pathways are strongly active in the human hippocampus. PHF8 is a histone demethylase that contributes to the regulation of mTOR. The mTOR pathway is hyperactive in the human hippocampus where it regulates the protein synthesis-dependent plastic changes underlying learning and memory (Bekinschtein et al., 2007; Fortress et al., 2013; Graber et al., 2013). Mutations in the PHF8 gene cause cognitive impairment and intellectual disability (Chen et al., 2018).
Therefore, we sought to determine the contribution of human-specific SVAs to the TBR2-mediated gene regulatory network in human hpIPCs. Using our ATAC-seq data, we identified 1816 human SVAs as accessible in human hpIPCs (Fig. 6D). Of these, nearly a quarter (434) displayed high accessibility, whereas 1382 were moderately accessible (Fig. 6D). Next, we used ChIP-seq to profile TBR2 binding in two human lines at day 5 of hpIPC differentiation. As with the previous sequencing experiments, we generated 100 bp long paired-end reads and only retained uniquely mapping high-quality reads (Samtools q=10 filtering) in order to maximize the chance to properly map reads on repetitive regions. This experiment revealed that 739 of the accessible SVAs showed TBR2 signals in the two human lines (Fig. 6D; Table S13). Notably, 257 (48.3%) of the TBR2-bound SVAs were human specific. (Fig. 6E; Table S14). TBR2-bound human-specific SVAs were located near 37 genes that our RNA-seq analysis identified as being differentially expressed between human and chimpanzee (Fig. 6F; Table S15). These genes include VPS13B (upregulated in humans), which is responsible for a rare developmental disease known as Cohen syndrome (Kolehmainen et al., 2003); NR4A2 (downregulated in humans), which has been implicated in neurodevelopmental language impairment (Reuter et al., 2017); DHX40 (downregulated in humans), which is implicated in AD (Taher et al., 2014); and E2F1 (upregulated in humans), which is a cell-cycle regulator associated with several neurodegenerative diseases such as AD (Zhang et al., 2010).
In summary, these data support a model in which human-specific SVAs provided a substrate for binding sites of TBR2 and other important hippocampal regulators. Therefore, it is likely that they were co-opted in the gene regulatory networks that are active during hippocampal neurogenesis, which led to human-specific regulation of key genes.
CRISPR-mediated SVA repression has massive repercussions on global gene expression
To further assess the contribution of SVA transposons to the gene regulation of hpIPCs, we leveraged CRISPR interference to simultaneously repress most of the active SVAs. We used NCCIT cells treated with retinoic acid (RA) as the experimental system for this purpose. The NCCIT cell line is derived from embryonal carcinoma and thus exhibits a gene expression signature highly similar to human embryonic stem cells (Fuentes et al., 2018; Barnada et al., 2022). Importantly, NCCITs treated for 7 days with RA differentiate into intermediate neural progenitor-like cells expressing both PAX6 and TBR2 (Mandal et al., 2015; Fig. 7A). RNA-seq data [transcripts per million (TPMs)] suggest very strong correlation between RA-treated NCCITs and human day-5 hpIPCs (Pearson correlation P<2.2×10−16, R=0.99). We cloned a stable NCCIT line with a permanently incorporated doxycycline-inducible, catalytically dead Cas9 fused to a repressive KRAB domain (dCas9-KRAB). The KRAB domain deposits repressive histone methylation (H3K9me3) to the regions targeted by dCas9 via single guide RNAs (sgRNAs). Into this same line, we also permanently knocked in two sgRNAs that are able to target >80% of all SVAs (Pontis et al., 2019). Hereafter, we refer to the RA-treated CRISPR line as RA-NCCITs. Remarkably, exposing the RA-NCCITs to doxycycline for 72 h was sufficient to induce dCas9 activation (Fig. 7B) and the deposition of the repressive H3K9me3 on over 2500 previously unmethylated SVAs (Fig. 7C). We next performed RNA-seq on the RA-NCCITs with or without doxycycline treatment (three replicates per condition). First, we observed that the genome-wide expression levels (TPMs) of the RA-NCCITs were highly correlated with those of the iPSC-derived human hpIPCs (Pearson correlation=0.93; P<2.2×10−16). This indicates that RA-treated NCCITs are appropriate to model hpIPCs.
Then, we compared the expression levels of the RA-NCCITs with or without doxycycline treatment and identified 5795 DEGs (FDR<0.05; Fig. 7D; Table S24). Of these genes, 677 were previously identified as differentially expressed upon comparing human with chimpanzee hpIPCs (Fig. 7E; Table S25). In other words, the expression of over a quarter (26.1%) of the genes that exhibited a human-specific expression signature in hpIPCs seem to be under the control of SVA transposons. Importantly, one of these genes is FOXP2, which has been associated with the evolution of language and is implicated in several speech disorders (Enard, 2011; Liégeois et al., 2016; MacDermot et al., 2005).
As the guide RNAs used for this experiment were originally designed to target a DNA sequence shared by the SVAs with the LTR5H family (Pontis et al., 2019), we restricted the analysis to the genes that are associated with human SVAs (i.e. only considering the genes that represent the closest gene to an annotated SVA; hereafter SVA genes). By doing so, we found 611 SVA genes as differentially expressed in RA-NCCITs upon SVA repression. Of these, 90 were previously identified as differentially expressed upon comparing human with chimpanzee hpIPCs. Thus, the expression of these 90 genes can be bona fide considered to be directly regulated by SVA-derived enhancers in both primary hippocampal progenitors and in the NCCIT cell line. Remarkably, the large majority of these genes (72.5%) had decreased expression upon SVA repression. These include SOX2, FGF2, PRDM1, NTRK2 and TFAP2B.
Finally, we examined the genes previously identified as being near a TBR2-bound human-specific SVA in iPSC-derived hpIPCs and found that nearly a third of them lose expression in RA-NCCITs upon SVA repression (Fig. 7F). In summary, our functional experiments indicate a widespread role for human-specific SVA transposons as cis-regulatory elements during hippocampal neurogenesis.
DISCUSSION
The hippocampus is susceptible to specific neurodegenerative disorders such as AD but may have also played an important role in the evolution of human cognition. Spatial memory, which is attributed to the hippocampus, may have contributed to the geographic expansion of ancient humans. Characterizing the human-specific gene regulatory networks of hippocampal development provides insight into its role in human evolution. Although there is no consensus on whether the cognitive phenotypes seen in AD are uniquely specific to humans, understanding the unique properties of the human hippocampus may lead the way to potential treatments. Thus, the work described here is relevant both in terms of evolutionary developmental biology and evolutionary medicine.
To study the evolution of the human hippocampus from a developmental standpoint, we investigated the extent to which TEs contributed to gene expression profiles of human and chimpanzee hpIPCs. TEs account for nearly 50% of the human genome and many elegant studies have established that at least a fraction of TEs can regulate host genes in humans and other primates (Chuong et al., 2013, 2016; Cosby et al., 2021; del Rosario et al., 2014; Du et al., 2016; Fuentes et al., 2018; Jacques et al., 2013; Judd et al., 2021; Lynch et al., 2011, 2015; Mika et al., 2021; Modzelewski et al., 2021; Okhovat et al., 2020; Rayan et al., 2016; Schmidt et al., 2012; Sundaram et al., 2014; Trizzino et al., 2017, 2018; Ward et al., 2018; Xie et al., 2013).
HpIPCs are a transient progenitor population during a crucial developmental stage in the sub-ventricular zone of both the hippocampus and neocortex (Bulfone et al., 1999; Cipriani et al., 2016; Englund, 2005; Kimura et al., 1999), and there is a consensus that this progenitor population may have played a role in the evolution of brain volume in mammals (Florio and Huttner, 2014; Martínez-Cerdeño et al., 2006). To study the developmental evolution of human hpIPCs, we leveraged a comparative approach centered on differentiating human and chimpanzee iPSCs into a neuronal population that closely recapitulates differentiation into hpIPCs, as demonstrated by the high expression of signature markers such as TBR2 and PAX6.
The transcriptomes of human and chimpanzee hpIPCs have not been previously compared, largely due to the limited availability of primary tissue. Here, we carried out this comparison using our iPSC-derived system and identified profound differences between the two species, with over 2500 genes differentially expressed at this stage. These genes include several that were previously associated with cognitive function, language, neurodevelopment and neurodegeneration, many of which are upregulated in humans relative to the chimpanzee (e.g. MTRNR2L8, DHX40, VPS13B, WDFY2 and PURB).
We demonstrate that species-specific enhancers significantly contributed to the gene expression differences that we identified. This is consistent with recent studies that used primary brain tissues from several species to profile species-specific cis-regulatory activity in mammals (Emera et al., 2016; Reilly et al., 2015). Importantly, we found that these species-specific enhancers are enriched for young transposable elements. Several studies have identified evolutionarily young L1 long interspersed nuclear elements (LINEs) as active in the brain during different developmental stages, suggesting that they could serve as alternative promoters for many genes involved in neuronal functions (Coufal et al., 2009; Jönsson et al., 2019; Sur et al., 2017; Thomas et al., 2012; Zhao et al., 2019). Here, we identified other young transposable elements, ERVs and SVAs, as regulators of IPC gene expression. The identification of ERVs as candidate enhancers in human and chimpanzee IPCs is consistent with previous studies that demonstrated that ERVs heavily impact gene regulatory programs during the immune response (Chuong et al., 2013, 2016), in pluripotency maintenance and development (Coluccio et al., 2018; Fuentes et al., 2018; Miao et al., 2020), in the mammalian placenta (Lynch et al., 2011, 2015; Mika et al., 2021), in the primate liver (Trizzino et al., 2017) and in many cancer types (Ito et al., 2020; Ivancevic and Chuong, 2020; Shah et al., 2021). Similar to the L1s, the ERVs also possess a well-defined cis-regulatory architecture (e.g. they have their own promoter), and this may have played a role in the co-option of these TEs as functioning cis-regulatory elements.
The SVAs are particularly interesting from a human evolution standpoint, given that half the known copies are exclusively present in our species. Moreover, SVAs are among the few transposable elements that still exhibit active transposition in the human genome. We and others have previously revealed that SVAs can be sources of enhancers in primates (Playfoot et al., 2021; Pontis et al., 2019; Trizzino et al., 2017, 2018). The repression of some SVAs by specific zinc-finger proteins at specific stages of neuronal development is also a crucial mechanism for successful neurogenesis (Playfoot et al., 2021; Pontis et al., 2019; Turelli et al., 2020). Here, we demonstrate that SVAs are pervasive regulators of hippocampal neurogenesis and they act as enhancers in the hpIPC population. By using CRISPR interference, we show that repressing hundreds of normally ‘de-repressed’ SVAs alters the expression of thousands of genes. Intriguingly, our CRISPR interference experiments revealed that global SVA repression leads to the attenuation of the expression levels of over a quarter of the ∼2500 genes previously identified as showing human-specific expression in hpIPCs. These include crucial neurodevelopmental regulators such as FOXP2, HAND2, MEF2C, SOX2 and SOX4. In normal conditions, these genes are more highly expressed in humans relative to in chimpanzees, but upon SVA repression, these differences were diminished.
In conclusion, our findings indicate that the development of hippocampal neurons has been profoundly affected by the domestication of young transposable elements. These young TEs have been co-opted as functional enhancers and promoters and ultimately rewired the expression of hundreds of crucially important neuroregulators in the developing human hippocampus. The human-specific gene expression and the associated TE-derived enhancers that we identified here may play important roles in both human evolution and neurodegenerative disease.
MATERIALS AND METHODS
Antibodies
The following antibodies were used in this study: anti-TBR2 (15 µg per ChIP, Abcam, ab216870), anti-H3K9me3 (3 µg per ChIP, Abcam, ab8898), anti-H3K21ac (3 µg per ChIP, Abcam, ab4729), anti-TBR2 (1:500 for western blotting, Santa Cruz Biotechnology, sc-293481), anti-GAPDH (1:1000 for western blotting, Cell Signaling Technology, 5174), HRP-conjugated anti-rabbit (1:10,000 for western blotting, Cell Signaling Technology, 7074S) and HRP-conjugated anti-mouse (1:10,000 for western blotting, Cell Signaling Technology, 7076S).
Human and chimpanzee iPSC cultures
The human male iPSC line denoted as SV20 was obtained from the University of Pennsylvania, where it was generated, and validated by the expression of pluripotency markers and differentiation into various cell types in multiple studies (Pagliaroli et al., 2021; Pashos et al., 2017; Yang et al., 2015; Zhang et al., 2015). The human female iPSC line GM 23716 was obtained from the Coriell Institute for Medical Research (Camden, NJ, USA) and validated by the expression of pluripotency markers and differentiation into cranial neural crest cells in a previous study (Pagliaroli et al., 2021). The human female iPSC line 21792 and all three chimpanzee iPSC lines were obtained from the laboratory of Yoav Gilad at the University of Chicago and validated in previous studies (Gallego Romero et al., 2015; Ward et al., 2018).
The iPSC lines were expanded in feeder-free, serum-free mTeSR1 medium (STEMCELL Technologies). Cells were passaged ∼1:10 at 80% confluency using ReLeSR (STEMCELL Technologies) and small cell clusters (50-200 cells) were subsequently plated on tissue culture dishes coated overnight with Geltrex LDEV-Free hESC-qualified reduced growth factor basement membrane matrix (Thermo Fisher Scientific).
hpIPC differentiation
The iPSC lines were differentiated into hpIPCs as previously described (Yu et al., 2014). Three batches consisting of one human and one chimpanzee iPSC line each were cultured until ∼50-70% confluence was reached, and then treated with the hpIPC medium for 5 days prior to collection for RNA-seq, ATAC-seq, ChIP-seq or immunofluorescence. The hpIPC medium consisted of Dulbecco's modified Eagle medium (DMEM)/F12 (Invitrogen), 0.5× N2 (Invitrogen), 0.5× B27 (Invitrogen), DKK1 (0.5 μg/ml, BioLegend), cyclopamine (1 μM, LC Labroatories), Noggin (0.5 μg/ml, BioLegend), SB431542 (10 μM, Selleck Chemicals) and the antibiotics penicillin/streptomycin (1× from 100× stock, Gibco).
Western blotting
For western blotting of total lysates, cells were harvested and washed three times in 1× PBS and lysed in RIPA buffer (50 mM Tris-HCl pH 7.5, 150 mM NaCl, 1% Igepal, 0.5% sodium deoxycholate, 0.1% SDS, 500 µM dithiothreitol) with protease inhibitors (Sigma-Aldrich). Approximately 20 μg of whole cell lysates were loaded in Novex WedgeWell 4-20% Tris-Glycine Gels (Invitrogen) in a Tris-Glycine-SDS buffer (Invitrogen) and separated by gel electrophoresis (SDS-PAGE). The proteins were then transferred to Immun-Blot PVDF membranes (Thermo Fisher Scientific) for antibody probing. Membranes were incubated with 10% bovine serum albumin in TBS with 3% Tween 20 (TBST) for 30 min at room temperature (RT), then incubated for variable times with suitable antibodies diluted in 5% bovine serum albumin in 1× TBST, washed with TBST and incubated with a dilution of 1:10,000 of the secondary antibody for 1 h at RT. The antibody was visualized using Super Signal West Dura Extended Duration Substrate (Thermo Fisher Scientific) and imaged with Amersham Imager 680.
Real-time quantitative PCR
Cells were lysed in TRI-reagent (Invitrogen) and RNA was extracted using the Direct-zol RNA MiniPrep kit (Zymo Research). Approximately 600 ng of template RNA was retrotranscribed into cDNA using RevertAid first strand cDNA synthesis kit (Thermo Fisher Scientific) according to the manufacturer’s directions. Approximately 15 ng of cDNA was used for each real-time quantitative PCR (RT-qPCR) reaction with 0.1 μM of each primer, 10 μl of PowerUp SYBR Green Master Mix (Applied Biosystems) in a final volume of 20 μl, using QuantStudio 3 Real-Time PCR System (Applied Biosystems). Thermal cycling parameters were set as following: 3 min at 95°C, followed by 40 cycles of 10 s at 95°C, 20 s at 63°C, followed by 30 s at 72°C. Each sample was run in triplicate. 18S rRNA was used as a normalizer.
ChIP-seq
Samples from different conditions were processed together to prevent batch effects. Approximately 15 million cells were cross-linked with 1% formaldehyde for 5 min at RT, quenched with 125 mM glycine, harvested and washed twice with 1× PBS. The pellet was resuspended in ChIP lysis buffer (150 mM NaCl, 1% Triton X-100, 0.7% SDS, 500 μM dithiothreitol, 10 mM Tris-HCl, 5 mM EDTA) and chromatin was sheared to an average length of 200-500 bp, using a Covaris S220 Ultrasonicator. The chromatin lysate was diluted with SDS-free ChIP lysis buffer. For ChIP-seq, 10 µg of antibody (3 µg for H3K27ac) was added to 5 µg of sonicated chromatin along with Dynabeads Protein A magnetic beads (Invitrogen) and incubated at 4°C overnight. On day 2, beads were washed twice with each of the following buffers: mixed micelle buffer (150 mM NaCl, 1% Triton X-100, 0.2% SDS, 20 mM Tris-HCl, 5 mM EDTA, 65% sucrose), buffer 500 (500 mM NaCl, 1% Triton X-100, 0.1% sodium deoxycholate, 25 mM HEPES, 10 mM Tris-HCl, 1 mM EDTA) and LiCl/detergent wash buffer (250 mM LiCl, 0.5% sodium deoxycholate, 0.5% NP-40, 10 mM Tris-HCl, 1 mM EDTA). A final wash was performed with 1× Tris-EDTA (TE) buffer. Finally, beads were resuspended in 1× TE buffer containing 1% SDS and incubated at 65°C for 10 min to elute immunocomplexes. Elution was repeated twice, and the samples were further incubated overnight at 65°C to reverse cross-linking, along with the untreated input (5% of the starting material). On day 3, after treatment with 0.5 mg/ml proteinase K for 1 h at 65°C, DNA was purified with Zymo ChIP DNA Clear Concentrator kit and quantified with QUBIT (Invitrogen).
For all ChIP-seq experiments, barcoded libraries were made with NEB ULTRA II DNA Library Prep Kit for Illumina (New England Biolabs) and sequenced on Illumina NextSeq 500, producing 100 bp paired-end reads.
ChIP-seq analyses
After removing the adapters with TrimGalore! (https://www.bioinformatics.babraham.ac.uk/projects/trim_galore/), the sequences were aligned to the reference hg19, using Burrows–Wheeler Alignment tool (BWA), with the MEM algorithm (http://bio-bwa.sourceforge.net/bwa.shtml). Uniquely mapping aligned reads were filtered based on mapping quality (MAPQ>10) to restrict our analysis to higher quality and likely uniquely mapped reads, and PCR duplicates were removed. We called peaks for each individual using MACS2 (Heinz et al., 2010) (H3K27ac) or Homer (http://homer.ucsd.edu/homer/) at 5% FDR with default parameters.
RNA-seq
Cells were lysed in TRI-reagent and total RNA was extracted using Quick RNA Miniprep kit (Zymo Research) according to the manufacturer's instructions. RNA was further quantified using the DeNovix DS-11 Spectrophotometer and the RNA integrity was checked on Bioanalyzer 2100 (Agilent). Only samples with an RNA integrity number above 8.0 were used for transcriptome analysis. RNA libraries were prepared using 1 μg of total RNA input using NEBNext Poly(A) mRNA Magnetic Isolation Module, NEBNext Ultra II Directional RNA Library Prep Kit for Illumina and NEBNext Ultra II DNA Library Prep Kit for Illumina according to the manufacturer's instructions (New England Biolabs).
scRNA-seq
Cells from both species at each time were first incubated with Accutase (STEMCELL Technologies) at 37°C for 7 min. The cells were collected with DMEM/F12 and centrifuged for 5 min at 150 g. The cells were resuspended in 10% fetal bovine serum (Fisher Scientific) in DMEM and strained with a 40 µm cell strainer to create a single-cell suspension. After confirming >90% viability with a Countess III (Invitrogen), the cells were processed with the 10x Genomics Cell Multiplexing Oligo protocol (https://www.10xgenomics.com/support/single-cell-gene-expression/documentation/steps/sample-prep/cell-multiplexing-oligo-labeling-for-single-cell-rna-sequencing-protocols) and 10x Genomics 3′ Cell Plex Kit.
scRNA-seq analyses
10× Cell Ranger (Zheng et al., 2017) was used to demultiplex and map the scRNA-seq data, with the tools cellranger multi and cellranger mkfastq. Seurat 4 (Hao et al., 2021) was used for individual analysis of scRNA-seq data, as well as integration of human and chimpanzee datasets. The human genes used to classify the cells into the G1/S/G2/M phases of the cell cycle were obtained from https://github.com/hbctraining/scRNA-seq/blob/master/lessons/06_SC_SCT_and_integration.md.
RNA-seq analyses
After removing the adapters with TrimGalore!, Kallisto (Bray et al., 2016) was used to count reads mapping to each gene. We analyzed differential gene expression levels with DESeq2 (Love et al., 2014), with the following model: design=∼condition, where condition indicates either Human or Chimpanzee.
ATAC-seq
For ATAC-Seq experiments, 50,000 cells per condition were processed as described in the original ATAC-seq protocol paper (Buenrostro et al., 2013). ATAC-seq data were processed with the same pipeline described for ChIP-seq, with one modification: all mapped reads were offset by +4 bp for the forward-strand and −5 bp for the reverse-strand. Peaks were called using MACS2 (Zhang et al., 2008).
Generation of the NCCIT-dCas9KRAB-SVAsgRNA stable cell line
This cell line was generated in our recent study (Barnada et al., 2022). Briefly, dCas9-KRAB was cloned into a piggyBac transposon containing ampicillin and puromycin resistance, which was obtained from Raquel Fueyo at Stanford University. The piggyBac dCas9-KRAB doxycycline-inducible plasmid and a piggyBac transposase (Cell Signaling Technology) were transfected into NCCIT cells (American Type Culture Collection) at ∼70% confluency using a 6:1 ratio of Fugene HD (Promega) for 48 h in ATCC-formulated RPMI medium (American Type Culture Collection). Two days post-transfection, the medium was changed and the transfected cells were selected using 1 µg puromycin in 1 ml medium. A piggyBac transposon plasmid containing two sgRNAs (SVAsgRNA1, 5′-CTCCCTAATCTCAAGTACCC-3′, and SVAsgRNA2, 5′-TGTTTCAGAGAGCACGGGGT-3′; Integrated DNA Technologies) targeting ∼80% of all annotated SVAs in humans (Pontis et al., 2019) and a piggyBac transposase were transfected into the NCCIT-dCas9KRAB cells using a 6:1 ratio of Fugene HD for 48 h in ATCC-formulated RPMI medium. Two days post-transfection, the medium was changed and the transfected cells were selected using 400 µg geneticin in 1 ml of medium in addition to 1 µg puromycin in 1 ml medium. The NCCIT-dCas9KRAB-SVAsgRNA cell line was maintained in ATCC-formulated RPMI medium supplemented with 10% tetracycline-free fetal bovine serum (Omega Scientific), 1% L-glutamine, 1 µg/ml puromycin and 400 µg/ml geneticin and incubated at 5% CO2, 20% O2 at 37°C.
RA-induced neuronal differentiation and CRISPR interference of NCCIT-dCas9KRAB-SVAsgRNA cells
The NCCIT-dCas9KRAB-SVAsgRNA cells, at ∼20% confluency, were treated with 10 µM RA (Sigma-Aldrich) in 10 ml medium for 1 week to induce neuronal differentiation. At day 4, the medium was refreshed and the cells were additionally treated with 2 µg doxycycline in 1 ml of medium for 3 days. The cells were collected on day 7 of RA treatment (day 3 of doxycycline treatment) for RT-qPCR and genomic experiments. Expression of the doxycycline-inducible dCas9 was verified via western blotting.
Statistical and genomic analyses
All statistical analyses were performed using R v3.3.1 or Graphpad Prism version 9.2.0 for Mac OS X. BEDTools v2.27.1 (Quinlan and Hall, 2010) was used for genomic analyses. Pathway analysis was performed with Ingenuity Pathway Analysis Suite (QIAGEN, https://www.qiagenbioinformatics.com/products/ingenuity-pathway-analysis). Motif analyses were performed using the MEME suite (Bailey et al., 2015), and specifically with the MEME-ChIP application. FASTA files of the regions of interest were produced using BEDTools v2.27.1. Shuffled input sequences were used as background. E-values<0.001 were used as threshold for significance. Orthologous ATAC-seq regions were identified using the University of California Santa Cruz (UCSC) Genome Browser tool LiftOver.
Supplementary Material
Acknowledgements
We are grateful to Dr Yoav Gilad (University of Chicago) and the Yerkes National Primate Research Center (Emory University, Atlanta, GA, USA) for providing all the chimpanzee lines and one of the human lines, and to Dr Joanna Wysocka's lab and in particular Dr Raquel Fueyo (Stanford University), for providing CRISPR reagents and critical advice. We thank the Thomas Jefferson Stem Cell and Regenerative Neuroscience Center for support in the optimization of the iPSC differentiation. We are grateful to Trizzino lab members Luca Pagliaroli, PhD, Chiara Scopa, PhD and Connor Ott for help with specific analyses and experiments. We thank the Genomic Facility at the Wistar Institute (Philadelphia, PA, USA) for Next Generation Illumina Sequencing. The PAX6 antibody developed by A. Kawakami was obtained from the Developmental Hybridoma Bank, created by the National Institute of Child Health and Human Development of the National Institutes of Health and maintained at the University of Iowa, Department of Biology, Iowa City, IA, USA.
Footnotes
Author contributions
Conceptualization: S.P., M.T.; Validation: S.P., S.M.B.; Formal analysis: S.P., S.M.B., C.L., M.T.; Investigation: S.M.B.; Data curation: S.P., J.I.M., M.T.; Writing - original draft: S.P., M.T.; Writing - review & editing: S.P., J.I.M., M.T.; Supervision: J.I.M., M.T.; Project administration: M.T.; Funding acquisition: M.T.
Funding
M.T. was funded by the National Institutes of Health (NIH-NIGMS R35GM138344) and by the G. Harold and Leila Y. Mathers Foundation. Open Access funding provided by Thomas Jefferson University. Deposited in PMC for immediate release.
Data availability
The original genome-wide data generated in this study have been deposited in the GEO database under the accession GSE189347.
Peer review history
The peer review history is available online at https://journals.biologists.com/dev/lookup/doi/10.1242/dev.200413.reviewer-comments.pdf.
References
- Abrajano, J. J., Qureshi, I. A., Gokhan, S., Zheng, D., Bergman, A. and Mehler, M. F. (2009). REST and CoREST modulate neuronal subtype specification, maturation and maintenance. PLoS One 4, e7936. 10.1371/journal.pone.0007936 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Agoglia, R. M., Sun, D., Birey, F., Yoon, S.-J., Miura, Y., Sabatini, K., Paşca, S. P. and Fraser, H. B. (2021). Primate cell fusion disentangles gene regulatory divergence in neurodevelopment. Nature 592, 421-427. 10.1038/s41586-021-03343-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Andersen, J., Urbán, N., Achimastou, A., Ito, A., Simic, M., Ullom, K., Martynoga, B., Lebel, M., Göritz, C., Frisén, J.et al. (2014). A transcriptional mechanism integrating inputs from extracellular signals to activate hippocampal stem cells. Neuron 83, 1085-1097. 10.1016/j.neuron.2014.08.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arnold, S. J., Huang, G.-J., Cheung, A. F. P., Era, T., Nishikawa, S.-I., Bikoff, E. K., Molnár, Z., Robertson, E. J. and Groszer, M. (2008). The T-box transcription factor Eomes/Tbr2 regulates neurogenesis in the cortical subventricular zone. Genes Dev. 22, 2479-2484. 10.1101/gad.475408 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Aruga, J. (2004). The role of Zic genes in neural development. Mol. Cell. Neurosci. 26, 205-221. 10.1016/j.mcn.2004.01.004 [DOI] [PubMed] [Google Scholar]
- Averaimo, S., Gritti, M., Barini, E., Gasparini, L. and Mazzanti, M. (2014). CLIC1 functional expression is required for cAMP-induced neurite elongation in post-natal mouse retinal ganglion cells. J. Neurochem. 131, 444-456. 10.1111/jnc.12832 [DOI] [PubMed] [Google Scholar]
- Bailey, T. L., Johnson, J., Grant, C. E. and Noble, W. S. (2015). The MEME suite. Nucleic Acids Res. 43, W39-W49. 10.1093/nar/gkv416 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barger, N., Hanson, K. L., Teffer, K., Schenker-Ahmed, N. M. and Semendeferi, K. (2014). Evidence for evolutionary specialization in human limbic structures. Front. Hum. Neurosci. 8, 277. 10.3389/fnhum.2014.00277 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barnada, S. M., Isopi, A., Tejada-Martinez, D., Goubert, C., Patoori, S., Pagliaroli, L., Tracewell, M. and Trizzino, M. (2022). Genomic features underlie the co-option of SVA transposons as cis-regulatory elements in human pluripotent stem cells. PLoS Genet. 18, e1010225. 10.1371/journal.pgen.1010225 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bekinschtein, P., Katche, C., Slipczuk, L. N., Igaz, L. M., Cammarota, M., Izquierdo, I. and Medina, J. H. (2007). mTOR signaling in the hippocampus is necessary for memory formation. Neurobiol. Learn. Mem. 87, 303-307. 10.1016/j.nlm.2006.08.007 [DOI] [PubMed] [Google Scholar]
- Bray, N. L., Pimentel, H., Melsted, P. and Pachter, L. (2016). Near-optimal probabilistic RNA-seq quantification. Nat. Biotechnol. 34, 525-527. 10.1038/nbt.3519 [DOI] [PubMed] [Google Scholar]
- Buenrostro, J. D., Giresi, P. G., Zaba, L. C., Chang, H. Y. and Greenleaf, W. J. (2013). Transposition of native chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins and nucleosome position. Nat. Methods 10, 1231-1218. 10.1038/nmeth.2688 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bulfone, A., Martinez, S., Marigo, V., Campanella, M., Basile, A., Quaderi, N., Gattuso, C., Rubenstein, J. L. R. and Ballabio, A. (1999). Expression pattern of the Tbr2 (Eomesodermin) gene during mouse and chick brain development. Mech. Dev. 84, 133-138. 10.1016/S0925-4773(99)00053-2 [DOI] [PubMed] [Google Scholar]
- Burgess, N., Maguire, E. A. and O'Keefe, J. (2002). The human hippocampus and spatial and episodic memory. Neuron 35, 625-641. 10.1016/S0896-6273(02)00830-9 [DOI] [PubMed] [Google Scholar]
- Chen, X., Wang, S., Zhou, Y., Han, Y., Li, S., Xu, Q., Xu, L., Zhu, Z., Deng, Y., Yu, L.et al. (2018). Phf8 histone demethylase deficiency causes cognitive impairments through the mTOR pathway. Nat. Commun. 9, 114. 10.1038/s41467-017-02531-y [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cheval, H., Chagneau, C., Levasseur, G., Veyrac, A., Faucon-Biguet, N., Laroche, S. and Davis, S. (2012). Distinctive features of Egr transcription factor regulation and DNA binding activity in CA1 of the hippocampus in synaptic plasticity and consolidation and reconsolidation of fear memory. Hippocampus 22, 631-642. 10.1002/hipo.20926 [DOI] [PubMed] [Google Scholar]
- Chuong, E. B., Rumi, M. A. K., Soares, M. J. and Baker, J. C. (2013). Endogenous retroviruses function as species-specific enhancer elements in the placenta. Nat. Genet. 45, 325-329. 10.1038/ng.2553 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chuong, E. B., Elde, N. C. and Feschotte, C. (2016). Regulatory evolution of innate immunity through co-option of endogenous retroviruses. Science 351, 1083-1087. 10.1126/science.aad5497 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cipriani, S., Nardelli, J., Verney, C., Delezoide, A.-L., Guimiot, F., Gressens, P. and Adle-Biassette, H. (2016). Dynamic expression patterns of progenitor and pyramidal neuron layer markers in the developing human hippocampus. Cereb. Cortex 26, 1255-1271. 10.1093/cercor/bhv079 [DOI] [PubMed] [Google Scholar]
- Coluccio, A., Ecco, G., Duc, J., Offner, S., Turelli, P. and Trono, D. (2018). Individual retrotransposon integrants are differentially controlled by KZFP/KAP1-dependent histone methylation, DNA methylation and TET-mediated hydroxymethylation in naïve embryonic stem cells. Epigenetics Chromatin 11, 7. 10.1186/s13072-018-0177-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cosby, R. L., Judd, J., Zhang, R., Zhong, A., Garry, N., Pritham, E. J. and Feschotte, C. (2021). Recurrent evolution of vertebrate transcription factors by transposase capture. Science 371, eabc6405. 10.1126/science.abc6405 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Coufal, N. G., Garcia-Perez, J. L., Peng, G. E., Yeo, G. W., Mu, Y., Lovci, M. T., Morell, M., O'Shea, K. S., Moran, J. V. and Gage, F. H. (2009). L1 retrotransposition in human neural progenitor cells. Nature 460, 1127-1131. 10.1038/nature08248 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Crawley, J. N., Heyer, W.-D. and LaSalle, J. M. (2016). Autism and cancer share risk genes, pathways, and drug targets. Trends Genet. 32, 139-146. 10.1016/j.tig.2016.01.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Datson, N. a., Morsink, M. c., Steenbergen, P. j., Aubert, Y., Schlumbohm, C., Fuchs, E. and de Kloet, E. R. (2009). A molecular blueprint of gene expression in hippocampal subregions CA1, CA3, and DG is conserved in the brain of the common marmoset. Hippocampus 19, 739-752. 10.1002/hipo.20555 [DOI] [PubMed] [Google Scholar]
- Davis, L., Onn, I. and Elliott, E. (2018). The emerging roles for the chromatin structure regulators CTCF and cohesin in neurodevelopment and behavior. Cell. Mol. Life Sci. 75, 1205-1214. 10.1007/s00018-017-2706-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- del Rosario, R. C. H., Rayan, N. A. and Prabhakar, S. (2014). Noncoding origins of anthropoid traits and a new null model of transposon functionalization. Genome Res. 24, 1469-1484. 10.1101/gr.168963.113 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Du, J., Leung, A., Trac, C., Lee, M., Parks, B. W., Lusis, A. J., Natarajan, R. and Schones, D. E. (2016). Chromatin variation associated with liver metabolism is mediated by transposable elements. Epigenetics Chromatin 9, 28. 10.1186/s13072-016-0078-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Duyckaerts, C., Delatour, B. and Potier, M.-C. (2009). Classification and basic pathology of Alzheimer disease. Acta Neuropathol. 118, 5-36. 10.1007/s00401-009-0532-1 [DOI] [PubMed] [Google Scholar]
- Edler, M. K., Sherwood, C. C., Meindl, R. S., Hopkins, W. D., Ely, J. J., Erwin, J. M., Mufson, E. J., Hof, P. R. and Raghanti, M. A. (2017). Aged chimpanzees exhibit pathologic hallmarks of Alzheimer's disease. Neurobiol. Aging 59, 107-120. 10.1016/j.neurobiolaging.2017.07.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Eichenbaum, H. (2017a). The role of the hippocampus in navigation is memory. J. Neurophysiol. 117, 1785-1796. 10.1152/jn.00005.2017 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Eichenbaum, H. (2017b). Prefrontal–hippocampal interactions in episodic memory. Nat. Rev. Neurosci. 18, 547-558. 10.1038/nrn.2017.74 [DOI] [PubMed] [Google Scholar]
- Emera, D., Yin, J., Reilly, S. K., Gockley, J. and Noonan, J. P. (2016). Origin and evolution of developmental enhancers in the mammalian neocortex. Proc. Natl. Acad. Sci. USA 113, E2617-E2626. 10.1073/pnas.1603718113 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Enard, W. (2011). FOXP2 and the role of cortico-basal ganglia circuits in speech and language evolution. Curr. Opin. Neurobiol. 21, 415-424. 10.1016/j.conb.2011.04.008 [DOI] [PubMed] [Google Scholar]
- Enard, W., Khaitovich, P., Klose, J., Zöllner, S., Heissig, F., Giavalisco, P., Nieselt-Struwe, K., Muchmore, E., Varki, A., Ravid, R.et al. (2002). Intra- and interspecific variation in primate gene expression patterns. Science 296, 340-343. 10.1126/science.1068996 [DOI] [PubMed] [Google Scholar]
- Englund, C. (2005). Pax6, Tbr2, and Tbr1 are expressed sequentially by radial glia, intermediate progenitor cells, and postmitotic neurons in developing neocortex. J. Neurosci. 25, 247-251. 10.1523/JNEUROSCI.2899-04.2005 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ferrón, S. R., Charalambous, M., Radford, E., McEwen, K., Wildner, H., Hind, E., Morante-Redolat, J. M., Laborda, J., Guillemot, F., Bauer, S. R.et al. (2011). Postnatal loss of Dlk1 imprinting in stem cells and niche astrocytes regulates neurogenesis. Nature 475, 381-385. 10.1038/nature10229 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Finch, C. E. and Austad, S. N. (2015). Commentary: is Alzheimer's disease uniquely human? Neurobiol. Aging 36, 553-555. 10.1016/j.neurobiolaging.2014.10.025 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Florio, M. and Huttner, W. B. (2014). Neural progenitors, neurogenesis and the evolution of the neocortex. Development 141, 2182-2194. 10.1242/dev.090571 [DOI] [PubMed] [Google Scholar]
- Fortress, A. M., Fan, L., Orr, P. T., Zhao, Z. and Frick, K. M. (2013). Estradiol-induced object recognition memory consolidation is dependent on activation of mTOR signaling in the dorsal hippocampus. Learn. Mem. 20, 147-155. 10.1101/lm.026732.112 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fuentes, D. R., Swigut, T. and Wysocka, J. (2018). Systematic perturbation of retroviral LTRs reveals widespread long-range effects on human gene regulation. Elife 7, e35989. 10.7554/eLife.35989 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gallego Romero, I., Pavlovic, B. J., Hernando-Herraez, I., Zhou, X., Ward, M. C., Banovich, N. E., Kagan, C. L., Burnett, J. E., Huang, C. H., Mitrano, A.et al. (2015). A panel of induced pluripotent stem cells from chimpanzees: a resource for comparative functional genomics. Elife 4, e07103. 10.7554/eLife.07103 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gokhman, D., Agoglia, R. M., Kinnebrew, M., Gordon, W., Sun, D., Bajpai, V. K., Naqvi, S., Chen, C., Chan, A., Chen, C.et al. (2021). Human–chimpanzee fused cells reveal cis -regulatory divergence underlying skeletal evolution. Nat. Genet. 53, 467-476. 10.1038/s41588-021-00804-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Graber, T. E., McCamphill, P. K. and Sossin, W. S. (2013). A recollection of mTOR signaling in learning and memory. Learn. Mem. 20, 518-530. 10.1101/lm.027664.112 [DOI] [PubMed] [Google Scholar]
- Hao, Y., Hao, S., Andersen-Nissen, E., Mauck, W. M., Zheng, S., Butler, A., Lee, M. J., Wilk, A. J., Darby, C., Zager, M.et al. (2021). Integrated analysis of multimodal single-cell data. Cell 184, 3573-3587.e29. 10.1016/j.cell.2021.04.048 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Heinz, S., Benner, C., Spann, N., Bertolino, E., Lin, Y. C., Laslo, P., Cheng, J. X., Murre, C., Singh, H. and Glass., C. K. (2010). Simple combinations of lineage-determining transcription factors prime cis-regulatory elements required for macrophage and B cell identities. Mol. Cell 38, 576-589. 10.1016/j.molcel.2010.05.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hevner, R. F. (2019). Intermediate progenitors and Tbr2 in cortical development. J. Anat. 235, 616-625. 10.1111/joa.12939 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hickey, S. L., Berto, S. and Konopka, G. (2019). Chromatin decondensation by FOXP2 promotes human neuron maturation and expression of neurodevelopmental disease genes. Cell Rep. 27, 1699-1711.e9. 10.1016/j.celrep.2019.04.044 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hodge, R. D., Nelson, B. R., Kahoud, R. J., Yang, R., Mussar, K. E., Reiner, S. L. and Hevner, R. F. (2012). Tbr2 Is essential for hippocampal lineage progression from neural stem cells to intermediate progenitors and neurons. J. Neurosci. 32, 6275-6287. 10.1523/JNEUROSCI.0532-12.2012 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ito, J., Kimura, I., Soper, A., Coudray, A., Koyanagi, Y., Nakaoka, H., Inoue, I., Turelli, P., Trono, D. and Sato, K. (2020). Endogenous retroviruses drive KRAB zinc-finger protein family expression for tumor suppression. Sci. Adv. 6, eabc3020. 10.1126/sciadv.abc3020 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ivancevic, A. and Chuong, E. B. (2020). Transposable elements teach T cells new tricks. Proc. Natl. Acad. Sci. USA 117, 9145-9147. 10.1073/pnas.2004493117 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jacques, P.-É., Jeyakani, J. and Bourque, G. (2013). The majority of primate-specific regulatory sequences are derived from transposable elements. PLoS Genet. 9, e1003504. 10.1371/journal.pgen.1003504 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jönsson, M. E., Ludvik Brattås, P., Gustafsson, C., Petri, R., Yudovich, D., Pircs, K., Verschuere, S., Madsen, S., Hansson, J., Larsson, J.et al. (2019). Activation of neuronal genes via LINE-1 elements upon global DNA demethylation in human neural progenitors. Nat. Commun. 10, 3182. 10.1038/s41467-019-11150-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Judd, J., Sanderson, H. and Feschotte, C. (2021). Evolution of mouse circadian enhancers from transposable elements. Genome Biol. 22, 193. 10.1186/s13059-021-02409-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kamboh, M. I., Fan, K.-H., Yan, Q., Beer, J. C., Snitz, B. E., Wang, X., Chang, C.-C. H., Demirci, F. Y., Feingold, E. and Ganguli, M. (2019). Population-based genome-wide association study of cognitive decline in older adults free of dementia: identification of a novel locus for the attention domain. Neurobiol. Aging 84, 239.e15-239.e24. 10.1016/j.neurobiolaging.2019.02.024 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kimura, N., Nakashima, K., Ueno, M., Kiyama, H. and Taga, T. (1999). A novel mammalian T-box-containing gene, Tbr2, expressed in mouse developing brain. Dev. Brain Res. 115, 183-193. 10.1016/S0165-3806(99)00064-4 [DOI] [PubMed] [Google Scholar]
- King, M. and Wilson, A. (1975). Evolution at two levels in humans and chimpanzees: Their macromolecules are so alike that regulatory mutations may account for their biological differences. Science 188, 107-116. 10.1126/science.1090005 [DOI] [PubMed] [Google Scholar]
- Kittappa, R., Chang, W. W., Awatramani, R. B. and McKay, R. D. G. (2007). The foxa2 gene controls the birth and spontaneous degeneration of dopamine neurons in old age. PLoS Biol. 5, e325. 10.1371/journal.pbio.0050325 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kohane, A. A. and Wood, T. R. (2021). Neurodevelopmental clustering of gene expression identifies lipid metabolism genes associated with neuroprotection and neurodegeneration. bioRxiv. 10.1101/2021.09.02.458277 [DOI] [Google Scholar]
- Kolehmainen, J., Black, G. C. M., Saarinen, A., Chandler, K., Clayton-Smith, J., Träskelin, A.-L., Perveen, R., Kivitie-Kallio, S., Norio, R., Warburg, M.et al. (2003). Cohen syndrome is caused by mutations in a novel gene, COH1, encoding a transmembrane protein with a presumed role in vesicle-mediated sorting and intracellular protein transport. Am. J. Hum. Genet. 72, 1359-1369. 10.1086/375454 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liégeois, F. J., Hildebrand, M. S., Bonthrone, A., Turner, S. J., Scheffer, I. E., Bahlo, M., Connelly, A. and Morgan, A. T. (2016). Early neuroimaging markers of FOXP2 intragenic deletion. Sci. Rep. 6, 35192. 10.1038/srep35192 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Love, M. I., Huber, W. and Anders, S. (2014). Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550. 10.1186/s13059-014-0550-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lynch, V. J., Leclerc, R. D., May, G. and Wagner, G. P. (2011). Transposon-mediated rewiring of gene regulatory networks contributed to the evolution of pregnancy in mammals. Nat. Genet. 43, 1154-1159. 10.1038/ng.917 [DOI] [PubMed] [Google Scholar]
- Lynch, V. J., Nnamani, M. C., Kapusta, A., Brayer, K., Plaza, S. L., Mazur, E. C., Emera, D., Sheikh, S. Z., Grützner, F., Bauersachs, S.et al. (2015). Ancient transposable elements transformed the uterine regulatory landscape and transcriptome during the evolution of mammalian pregnancy. Cell Rep. 10, 551-561. 10.1016/j.celrep.2014.12.052 [DOI] [PMC free article] [PubMed] [Google Scholar]
- MacDermot, K. D., Bonora, E., Sykes, N., Coupe, A.-M., Lai, C. S. L., Vernes, S. C., Vargha-Khadem, F., McKenzie, F., Smith, R. L., Monaco, A. P.et al. (2005). Identification of FOXP2 truncation as a novel cause of developmental speech and language deficits. Am. J. Hum. Genet. 76, 1074-1080. 10.1086/430841 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mandal, C., Jung, K. H., Kang, S. C., Choi, M. R., Park, K. S., Chung, I. Y. and Chai, Y. G. (2015). Knocking down of UTX in NCCIT cells enhance cell attachment and promote early neuronal cell differentiation. BioChip J. 9, 182-193. 10.1007/s13206-015-9302-4 [DOI] [Google Scholar]
- Marchetto, M. C., Hrvoj-Mihic, B., Kerman, B. E., Yu, D. X., Vadodaria, K. C., Linker, S. B., Narvaiza, I., Santos, R., Denli, A. M., Mendes, A. P.et al. (2019). Species-specific maturation profiles of human, chimpanzee and bonobo neural cells. Elife 8, e37527. 10.7554/eLife.37527 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Marnetto, D., Mantica, F., Molineris, I., Grassi, E., Pesando, I. and Provero, P. (2018). Evolutionary rewiring of human regulatory networks by waves of genome expansion. Am. J. Hum. Genet. 102, 207-218. 10.1016/j.ajhg.2017.12.014 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martínez-Cerdeño, V., Noctor, S. C. and Kriegstein, A. R. (2006). The role of intermediate progenitor cells in the evolutionary expansion of the cerebral cortex. Cereb. Cortex 16, i152-i161. 10.1093/cercor/bhk017 [DOI] [PubMed] [Google Scholar]
- Mathys, H., Davila-Velderrain, J., Peng, Z., Gao, F., Mohammadi, S., Young, J. Z., Menon, M., He, L., Abdurrob, F., Jiang, X.et al. (2019). Single-cell transcriptomic analysis of Alzheimer's disease. Nature 570, 332-337. 10.1038/s41586-019-1195-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Miao, B., Fu, S., Lyu, C., Gontarz, P., Wang, T. and Zhang, B. (2020). Tissue-specific usage of transposable element-derived promoters in mouse development. Genome Biol. 21, 255. 10.1186/s13059-020-02164-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mihalas, A. B., Elsen, G. E., Bedogni, F., Daza, R. A. M., Ramos-Laguna, K. A., Arnold, S. J. and Hevner, R. F. (2016). Intermediate progenitor cohorts differentially generate cortical layers and require Tbr2 for timely acquisition of neuronal subtype identity. Cell Rep. 16, 92-105. 10.1016/j.celrep.2016.05.072 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mika, K., Marinić, M., Singh, M., Muter, J., Brosens, J. J. and Lynch, V. J. (2021). Evolutionary transcriptomics implicates new genes and pathways in human pregnancy and adverse pregnancy outcomes. Elife 10, e69584. 10.7554/eLife.69584 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Modzelewski, A. J., Shao, W., Chen, J., Lee, A., Qi, X., Noon, M., Tjokro, K., Sales, G., Biton, A., Anand, A.et al. (2021). A mouse-specific retrotransposon drives a conserved Cdk2ap1 isoform essential for development. Cell 184, 5541-5558.e22. 10.1016/j.cell.2021.09.021 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mora-Bermúdez, F., Badsha, F., Kanton, S., Camp, J. G., Vernot, B., Köhler, K., Voigt, B., Okita, K., Maricic, T., He, Z.et al. (2016). Differences and similarities between human and chimpanzee neural progenitors during cerebral cortex development. Elife 5, e18683. 10.7554/eLife.18683 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mukherjee, D., Gonzales, B. J., Ashwal-Fluss, R., Turm, H., Groysman, M. and Citri, A. (2021). Egr2 induction in spiny projection neurons of the ventrolateral striatum contributes to cocaine place preference in mice. Elife 10, e65228. 10.7554/eLife.65228 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Okhovat, M., Nevonen, K. A., Davis, B. A., Michener, P., Ward, S., Milhaven, M., Harshman, L., Sohota, A., Fernandes, J. D., Salama, S. R.et al. (2020). Co-option of the lineage-specific LAVA retrotransposon in the gibbon genome. Proc. Natl. Acad. Sci. USA 117, 19328-19338. 10.1073/pnas.2006038117 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pagliaroli, L., Porazzi, P., Curtis, A. T., Scopa, C., Mikkers, H. M. M., Freund, C., Daxinger, L., Deliard, S., Welsh, S. A., Offley, S.et al. (2021). Inability to switch from ARID1A-BAF to ARID1B-BAF impairs exit from pluripotency and commitment towards neural crest formation in ARID1B-related neurodevelopmental disorders. Nat. Commun. 12, 6469. 10.1038/s41467-021-26810-x [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pashos, E. E., Park, Y., Wang, X., Raghavan, A., Yang, W., Abbey, D., Peters, D. T., Arbelaez, J., Hernandez, M., Kuperwasser, N.et al. (2017). Large, diverse population cohorts of hiPSCs and derived hepatocyte-like cells reveal functional genetic variation at blood lipid-associated loci. Cell Stem Cell 20, 558-570.e10. 10.1016/j.stem.2017.03.017 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Playfoot, C. J., Duc, J., Sheppard, S., Dind, S., Coudray, A., Planet, E. and Trono, D. (2021). Transposable elements and their KZFP controllers are drivers of transcriptional innovation in the developing human brain. Genome Res. 31, 1531-1545. 10.1101/gr.275133.120 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Poirier, R., Cheval, H., Mailhes, C., Charnay, P., Davis, S. and Laroche, S. (2007). Paradoxical role of an Egr transcription factor family member, Egr2/Krox20, in learning and memory. Front. Behav. Neurosci. 1, 6. 10.3389/neuro.08.006.2007 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pontious, A., Kowalczyk, T., Englund, C. and Hevner, R. F. (2008). Role of intermediate progenitor cells in cerebral cortex development. Dev. Neurosci. 30, 24-32. 10.1159/000109848 [DOI] [PubMed] [Google Scholar]
- Pontis, J., Planet, E., Offner, S., Turelli, P., Duc, J., Coudray, A., Theunissen, T. W., Jaenisch, R. and Trono, D. (2019). Hominoid-specific transposable elements and KZFPs facilitate human embryonic genome activation and control transcription in naive human ESCs. Cell Stem Cell 24, 724-735.e5. 10.1016/j.stem.2019.03.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Quinlan, A. R. and Hall, I. M. (2010). BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841-842. 10.1093/bioinformatics/btq033 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Quinn, J. P. and Bubb, V. J. (2014). SVA retrotransposons as modulators of gene expression. Mob. Genet. Elements 4, e32102. 10.4161/mge.32102 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Raivich, G. (2008). c-Jun expression, activation and function in neural cell death, inflammation and repair. J. Neurochem. 107, 898-906. 10.1111/j.1471-4159.2008.05684.x [DOI] [PubMed] [Google Scholar]
- Raivich, G. and Behrens, A. (2006). Role of the AP-1 transcription factor c-Jun in developing, adult and injured brain. Prog. Neurobiol. 78, 347-363. 10.1016/j.pneurobio.2006.03.006 [DOI] [PubMed] [Google Scholar]
- Rayan, N. A., Del Rosario, R. C. H. and Prabhakar, S. (2016). Massive contribution of transposable elements to mammalian regulatory sequences. Semin. Cell Dev. Biol. 57, 51-56. 10.1016/j.semcdb.2016.05.004 [DOI] [PubMed] [Google Scholar]
- Reilly, S. K., Yin, J., Ayoub, A. E., Emera, D., Leng, J., Cotney, J., Sarro, R., Rakic, P. and Noonan, J. P. (2015). Evolutionary changes in promoter and enhancer activity during human corticogenesis. Science 347, 1155-1159. 10.1126/science.1260943 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rétaux, S., Bourrat, F., Joly, J.-S. and Hinaux, H. (2013). Perspectives in evo-devo of the vertebrate brain. In Advances in Evolutionary Developmental Biology (ed. Streelman J. T.), pp. 151-172. John Wiley & Sons, Ltd. [Google Scholar]
- Reuter, M. S., Krumbiegel, M., Schlüter, G., Ekici, A. B., Reis, A. and Zweier, C. (2017). Haploinsufficiency of NR4A2 is associated with a neurodevelopmental phenotype with prominent language impairment. Am. J. Med. Genet. A 173, 2231-2234. 10.1002/ajmg.a.38288 [DOI] [PubMed] [Google Scholar]
- Sakamoto, K., Karelina, K. and Obrietan, K. (2011). CREB: a multifaceted regulator of neuronal plasticity and protection. J. Neurochem. 116, 1-9. 10.1111/j.1471-4159.2010.07080.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sarkar, A., Mei, A., Paquola, A. C. M., Stern, S., Bardy, C., Klug, J. R., Kim, S., Neshat, N., Kim, H. J., Ku, M.et al. (2018). Efficient generation of CA3 neurons from human pluripotent stem cells enables modeling of hippocampal connectivity in vitro. Cell Stem Cell 22, 684-697.e9. 10.1016/j.stem.2018.04.009 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schmidt, D., Schwalie, P. C., Wilson, M. D., Ballester, B., Gonçalves, A., Kutter, C., Brown, G. D., Marshall, A., Flicek, P. and Odom, D. T. (2012). Waves of retrotransposon expansion remodel genome organization and CTCF binding in multiple mammalian lineages. Cell 148, 335-348. 10.1016/j.cell.2011.11.058 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sessa, A., Mao, C.-A., Hadjantonakis, A.-K., Klein, W. H. and Broccoli, V. (2008). Tbr2 directs conversion of radial glia into basal precursors and guides neuronal amplification by indirect neurogenesis in the developing neocortex. Neuron 60, 56-69. 10.1016/j.neuron.2008.09.028 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shah, A. H., Gilbert, M., Ivan, M. E., Komotar, R. J., Heiss, J. and Nath, A. (2021). The role of human endogenous retroviruses in gliomas: from etiological perspectives and therapeutic implications. Neuro Oncol. 23, 1647-1655. 10.1093/neuonc/noab142 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sin, C., Li, H. and Crawford, D. A. (2015). Transcriptional regulation by FOXP1, FOXP2, and FOXP4 dimerization. J. Mol. Neurosci. 55, 437-448. 10.1007/s12031-014-0359-7 [DOI] [PubMed] [Google Scholar]
- Sousa, A. M. M., Zhu, Y., Raghanti, M. A., Kitchen, R. R., Onorati, M., Tebbenkamp, A. T. N., Stutz, B., Meyer, K. A., Li, M., Kawasawa, Y. I.et al. (2017). Molecular and cellular reorganization of neural circuits in the human lineage. Science 358, 1027-1032. 10.1126/science.aan3456 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Squire, L. R. (1992). Memory and the hippocampus: a synthesis from findings with rats, monkeys, and humans. Psychol. Rev. 99, 195-231. 10.1037/0033-295x.99.2.195 [DOI] [PubMed] [Google Scholar]
- Sundaram, V. and Wysocka, J. (2020). Transposable elements as a potent source of diverse cis-regulatory sequences in mammalian genomes. Philos. Trans. R. Soc. B Biol. Sci. 375, 20190347. 10.1098/rstb.2019.0347 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sundaram, V., Cheng, Y., Ma, Z., Li, D., Xing, X., Edge, P., Snyder, M. P. and Wang, T. (2014). Widespread contribution of transposable elements to the innovation of gene regulatory networks. Genome Res. 24, 1963-1976. 10.1101/gr.168872.113 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sur, D., Kustwar, R. K., Budania, S., Mahadevan, A., Hancks, D. C., Yadav, V., Shankar, S. K. and Mandal, P. K. (2017). Detection of the LINE-1 retrotransposon RNA-binding protein ORF1p in different anatomical regions of the human brain. Mob. DNA 8, 17. 10.1186/s13100-017-0101-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Taher, N., McKenzie, C., Garrett, R., Baker, M., Fox, N. and Isaacs, G. D. (2014). Amyloid-β alters the DNA methylation status of cell-fate genes in an Alzheimer's disease model. J. Alzheimer's Dis. 38, 831-844. 10.3233/JAD-131061 [DOI] [PubMed] [Google Scholar]
- Thomas, C. A., Paquola, A. C. M. and Muotri, A. R. (2012). LINE-1 retrotransposition in the nervous system. Annu. Rev. Cell Dev. Biol. 28, 555-573. 10.1146/annurev-cellbio-101011-155822 [DOI] [PubMed] [Google Scholar]
- Tokuyama, M., Kong, Y., Song, E., Jayewickreme, T., Kang, I. and Iwasaki, A. (2018). ERVmap analysis reveals genome-wide transcription of human endogenous retroviruses. Proc. Natl. Acad. Sci. USA 115, 12565-12572. 10.1073/pnas.1814589115 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tomasello, M. and Herrmann, E. (2010). Ape and human cognition: what's the difference? Curr. Dir. Psychol. Sci. 19, 3-8. 10.1177/0963721409359300 [DOI] [Google Scholar]
- Trizzino, M., Park, Y., Holsbach-Beltrame, M., Aracena, K., Mika, K., Caliskan, M., Perry, G. H., Lynch, V. J. and Brown, C. D. (2017). Transposable elements are the primary source of novelty in primate gene regulation. Genome Res. 27, 1623-1633. 10.1101/gr.218149.116 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Trizzino, M., Kapusta, A. and Brown, C. D. (2018). Transposable elements generate regulatory novelty in a tissue-specific fashion. BMC Genom. 19, 468. 10.1186/s12864-018-4850-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Turelli, P., Playfoot, C., Grun, D., Raclot, C., Pontis, J., Coudray, A., Thorball, C., Duc, J., Pankevich, E. V., Deplancke, B.et al. (2020). Primate-restricted KRAB zinc finger proteins and target retrotransposons control gene expression in human neurons. Sci. Adv. 6, eaba3200. 10.1126/sciadv.aba3200 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Walker, L. C. and Jucker, M. (2017). The exceptional vulnerability of humans to Alzheimer's disease. Trends Mol. Med. 23, 534-545. 10.1016/j.molmed.2017.04.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang, H., Xing, J., Grover, D., Hedges, D. J., Han, K., Walker, J. A. and Batzer, M. A. (2005). SVA elements: a hominid-specific retroposon family. J. Mol. Biol. 354, 994-1007. 10.1016/j.jmb.2005.09.085 [DOI] [PubMed] [Google Scholar]
- Ward, M. C., Zhao, S., Luo, K., Pavlovic, B. J., Karimi, M. M., Stephens, M. and Gilad, Y. (2018). Silencing of transposable elements may not be a major driver of regulatory evolution in primate iPSCs. Elife 7, e33084. 10.7554/eLife.33084 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wei, Y. (2020). Comparative transcriptome analysis of the hippocampus from sleep-deprived and Alzheimer's disease mice. Genet. Mol. Biol. 43, e20190052. 10.1590/1678-4685-gmb-2019-0052 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wray, G. A. (2007). The evolutionary significance of cis-regulatory mutations. Nat. Rev. Genet. 8, 206-216. 10.1038/nrg2063 [DOI] [PubMed] [Google Scholar]
- Xie, M., Hong, C., Zhang, B., Lowdon, R. F., Xing, X., Li, D., Zhou, X., Lee, H. J., Maire, C. L., Ligon, K. L.et al. (2013). DNA hypomethylation within specific transposable element families associates with tissue-specific enhancer landscape. Nat. Genet. 45, 836-841. 10.1038/ng.2649 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang, W., Liu, Y., Slovik, K. J., Wu, J. C., Duncan, S. A., Rader, D. J. and Morrisey, E. E. (2015). Generation of iPSCs as a pooled culture using magnetic activated cell sorting of newly reprogrammed cells. PLoS One 10, e0134995. 10.1371/journal.pone.0134995 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yi, R., Chen, B., Zhao, J., Zhan, X., Zhang, L., Liu, X. and Dong, Q. (2014). Krüppel-like factor 8 ameliorates Alzheimer's disease by activating β-catenin. J. Mol. Neurosci. 52, 231-241. 10.1007/s12031-013-0131-4 [DOI] [PubMed] [Google Scholar]
- Yu, D. X., Di Giorgio, F. P., Yao, J., Marchetto, M. C., Brennand, K., Wright, R., Mei, A., Mchenry, L., Lisuk, D., Grasmick, J. M.et al. (2014). Modeling hippocampal neurogenesis using human pluripotent stem cells. Stem Cell Rep. 2, 295-310. 10.1016/j.stemcr.2014.01.009 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang, Y., Liu, T., Meyer, C. A., Eeckhoute, J., Johnson, D. S., Bernstein, B. E., Nusbaum, C., Myers, R. M., Brown, M., Li, W.et al. (2008). Model-based analysis of ChIP-Seq (MACS). Genome Biol. 9, R137. 10.1186/gb-2008-9-9-r137 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang, J., Li, H., Yabut, O., Fitzpatrick, H., D'Arcangelo, G. and Herrup, K. (2010). Cdk5 suppresses the neuronal cell cycle by disrupting the E2F1–DP1 complex. J. Neurosci. 30, 5219-5228. 10.1523/JNEUROSCI.5628-09.2010 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang, H., Xue, C., Shah, R., Bermingham, K., Hinkle, C. C., Li, W., Rodrigues, A., Tabita-Martinez, J., Millar, J. S., Cuchel, M.et al. (2015). Functional analysis and transcriptomic profiling of iPSC-derived macrophages and their application in modeling mendelian disease. Circ. Res. 117, 17-28. 10.1161/CIRCRESAHA.117.305860 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhao, B., Wu, Q., Ye, A. Y., Guo, J., Zheng, X., Yang, X., Yan, L., Liu, Q.-R., Hyde, T. M., Wei, L.et al. (2019). Somatic LINE-1 retrotransposition in cortical neurons and non-brain tissues of Rett patients and healthy individuals. PLoS Genet. 15, e1008043. 10.1371/journal.pgen.1008043 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zheng, G. X. Y., Terry, J. M., Belgrader, P., Ryvkin, P., Bent, Z. W., Wilson, R., Ziraldo, S. B., Wheeler, T. D., McDermott, G. P., Zhu, J.et al. (2017). Massively parallel digital transcriptional profiling of single cells. Nat. Commun. 8, 14049. 10.1038/ncomms14049 [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.