Next Article in Journal
GM-CSF and IL-33 Orchestrate Polynucleation and Polyploidy of Resident Murine Alveolar Macrophages in a Murine Model of Allergic Asthma
Next Article in Special Issue
Molecular Regulation of the Polycomb Repressive-Deubiquitinase
Previous Article in Journal
New Treatment Addressing the Pathogenesis of Psoriasis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Emerging Roles of Single-Cell Multi-Omics in Studying Developmental Temporal Patterning

Department of Basic Neuroscience, University of Lausanne, 1005 Lausanne, Switzerland
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2020, 21(20), 7491; https://doi.org/10.3390/ijms21207491
Submission received: 16 September 2020 / Revised: 6 October 2020 / Accepted: 8 October 2020 / Published: 11 October 2020

Abstract

:
The complexity of brain structure and function is rooted in the precise spatial and temporal regulation of selective developmental events. During neurogenesis, both vertebrates and invertebrates generate a wide variety of specialized cell types through the expansion and specification of a restricted set of neuronal progenitors. Temporal patterning of neural progenitors rests on fine regulation between cell-intrinsic and cell-extrinsic mechanisms. The rapid emergence of high-throughput single-cell technologies combined with elaborate computational analysis has started to provide us with unprecedented biological insights related to temporal patterning in the developing central nervous system (CNS). Here, we present an overview of recent advances in Drosophila and vertebrates, focusing both on cell-intrinsic mechanisms and environmental influences. We then describe the various multi-omics approaches that have strongly contributed to our current understanding and discuss perspectives on the various -omics approaches that hold great potential for the future of temporal patterning research.

1. Background

During CNS development, undifferentiated cells initially divide to generate highly diverse cell types, within and across species. The intricate process implicated in the generation of cellular, structural and functional diversity is regulated in both space and time. Relying on the presence of molecular cues at the interface of brain structures as well as on the presence of gradients, cells assemble to generate complex 3D spatial domains [1,2,3,4]. Temporal progression provides an additional dimension to developmental regulation. Indeed, the production and proper assembly of CNS cells into functional networks rely on the interplay between intrinsic temporally regulated programs and extracellular timing signals. The tight control of these two mechanisms permits the generation of a wide variety of cell types from a restricted and relatively homogeneous set of neural progenitors. Neural progenitor cells undergo temporal fate specification and sequentially produce distinct types of neurons, which progressively populate the developing CNS. One of these temporal transitions occurs upon the proliferative (symmetric) to neurogenic (asymmetric) switch in the division mode of first neuroepithelial cells (NEs) and then apical radial glial cells (aRGs) [5]. While cell-intrinsic temporally regulated programs were long shown to play a main role in the temporal tuning of these events, a growing body of evidence suggests the additional involvement of extrinsic cues.
In this review, we will discuss key principles defining the temporal progression of cell identity and the extracellular timing signals driving CNS development. We will first summarize the general mechanisms underlying temporal progression in identity through a detailed analysis of Drosophila CNS and mouse corticogenesis development (Figure 1). We will then describe multi-omics and bioinformatic models that best address this process and highlight the major pending questions.

2. Overview of CNS Neurogenesis

2.1. Drosophila Ventral Nerve Cord Neurogenesis

The Drosophila CNS comprises the central brain, the ventral nerve cord (VNC) and the optic lobe (OL) (Figure 1a). The VNC is arguably best described and will be used as a reference structure here. The embryonic VNC is segmented into a series of repeated units termed neuromeres; each is divided into two hemisegments. Each hemisegment of the VNC is composed of approximately thirty unique neural stem cells (NSCs) or neuroblasts (NBs, an asymmetrically dividing NSC) organized in rows and columns [6]. Each NB, independent of its location in the nervous system, gives birth to a stereotypical series of neurons via three different types of asymmetric divisions. Thus, the NB population is self-renewed upon each division, additionally yielding a distinct cell type: (i) Type-0 divisions directly give birth to a differentiating neuron, (ii) Type-1 divisions produce ganglion mother cells that further divide once into two neurons each, and (iii) Type-2 divisions generate intermediate neural progenitors (INP), which, in turn, further give rise to ganglion mother cells over a prolonged series of asymmetric divisions [7,8,9]. Neurogenesis occurs in two sequential waves: (i) during the embryonic phase and (ii) throughout the larval and pupal stages, in preparation for the final metamorphosis [10,11]. Embryonic neurogenesis is a one-day long event during which distinct NB types transition through a cascade of temporal transcription factors (tTFs) giving rise to a series of restricted lineages. The short length of this neurogenic wave underlies the precise NB-intrinsic temporal progression. At the end of embryogenesis, the majority of NBs undergo apoptosis, while few enter G1 quiescence [12]. By the beginning of the larval stage, surviving NBs exit quiescence and initiate a second, longer wave of neurogenesis, spanning both the larval and pupal stages [10,11,13].

2.2. Mouse Neocortex Neurogenesis

The developing mouse neocortex emerges within the dorsal pallium (Figure 1b). Here, NEs undergo symmetric divisions to amplify the pool of aRGs in the juxtaventricular germinal zone called the ventricular zone (VZ). Next, aRGs initiate neurogenesis through a series of asymmetric divisions. The first neurogenic events are dominated by direct neurogenesis, during which aRGs directly produce excitatory neurons. This phase is followed by the emergence of indirect neurogenesis, generating intermediate transiently amplifying progenitors (IP) [14,15]. These cells are located dorsally to the VZ in the subventricular zone (SVZ). Gliogenesis follows corticogenesis when aRGs undergo self-consuming symmetric divisions to generate glial cells [16]. Meanwhile, ventral pallium-derived interneurons migrate tangentially and ultimately populate the neocortex. Readers further interested in this topic can refer to a recent detailed review [17].
Our current understanding of the mechanisms of mouse neocortical temporal progression is still significantly restricted in comparison to the Drosophila. However, several studies have thoroughly described the stereotypical aRG transition from deep to superficial layer neuron-producing cells [16,18]. Recent studies have identified a core set of genes and postmitotic transcriptional events crucial to determine the final fate of differentiating neurons [19]. On the other hand, the timing of gliogenesis has not been clearly associated with a precise laminar position preference, as the majority of astrocytes are born from grey matter localized proliferation, with a small degree of 3D dispersion from their clonal origin [20,21].

3. Regulation of CNS Temporal Patterning

Much of our current knowledge on the temporal controls over the generation of neuronal diversity is derived from studies in the fruit fly Drosophila melanogaster. Indeed, temporal patterning was first described in Drosophila VNC NBs [22,23,24], occurring at both the embryonic and larval stages, during which several important temporal transitions take place. The successive expression of tTFs during precise temporal windows is necessary and sufficient to specify distinct neuronal fates. Unlike vertebrates, all these mechanisms have long been demonstrated to rely on progenitor-intrinsic programs via transcriptional cross-regulation between tTFs, along with other independent mechanisms [22,23,24,25,26]. However, recent studies have revealed the additional role of extrinsic cues in the regulation of Drosophila NB patterning, thus pointing in the direction of evolutionarily conserved strategies shared by both invertebrates and vertebrates.

3.1. Epigenetic and Transcriptional Events Regulating Phases of Temporal Progression

Understanding how cells measure and react to the progression of time (temporal progression) is not straightforward. Considering the Drosophila nervous system development, a limited set of feed-forward cross-regulated transcription factors (TFs) appears to be sufficient to generate a reproducible temporal sequence of transcriptional events [27,28,29,30,31]. For the sake of brevity, we will exclusively highlight a few examples of temporal patterning progression in the Drosophila CNS to extrapolate the general mechanisms of cell-autonomous temporal patterning. For further descriptions, we refer the readers to these reviews [13,27,32].
As already described for a series of biological events [33,34,35], the synchronization of patterning events requires the existence of molecular timers in each cell [36]. During Drosophila CNS development, this process has often been described as a series of activating timers. Here, temporal progression is paced by the expression and accumulation of the tTF itself. When a needed threshold is reached, the expression of the next activator in the cascade is induced. Thus, Pdm expression onset is mostly determined by the transcription of Kr, which is itself activated by Hb. Pdm expression would, in turn, determine the activation of the Cas-expressing phase. Isolated early- and late-forming NB lines are capable of recapitulating the sequential gene expression of Hb → Kr → Pdm → Cas, supporting the hypothesis of a coordinated action between gene regulatory interactions with an independent cell-intrinsic molecular temporal clock [24].
Recent evidence suggests that a more robust regulatory mechanism of the stereotypical temporal patterning cascade in VNC NBs occurs by means of the decay of tTF repressor timers [37] (Figure 2a). In this case, temporal progression is more sensitive to the upstream expression of a tTF repressor rather than the expression of activating timers. As the upstream repressor expression decays to a sufficiently low level, NBs transition in the following temporal step, representing a permanent and less costly mechanism for gating temporal progression [38]. As an example, Pdm expression regulation is more sensitive to the upstream repressor Hb than to the deletion of the upstream activator Kr. Similarly, Cas expression timing is advanced in Kr mutants, in contrast to Pdm mutants [37].
The tight control of gene transcription depends upon the binding of TFs onto cis regulatory elements (CREs) (Figure 2b), from which transcription initiation complexes are recruited. As shown in the developing Drosophila, many tTF-CRE pairs are required for temporal regulation [28,29]. For instance, three evolutionarily conserved regulatory regions within the Hb enhancer region (HG4-1, HG4-3 and HG4-7) have been functionally confirmed to specifically recapitulate Hb expression in NB stages 9 and 10, but not in stage 11 [28].
On the other hand, gene silencing in Drosophila is mediated through the recruitment of the Polycomb complex by DNA elements termed Polycomb responsive elements (PRE) [39] (Figure 2c). PREs have been implicated in the repression of hundreds of developmentally relevant loci, many of which are widely conserved across vertebrates and invertebrates, and include several developmental key regulators, such as Hox genes [40,41]. Interestingly, a recent study demonstrated the localization of the Drosophila pleiotropic repressive complex (PhoRC) in both enhancers and PRE elements during development. Here, BiTS-ChIP-seq (batch isolation of tissue-specific chromatin for immunoprecipitation) and high-throughput sequencing in Drosophila embryonic tissue was used to investigate the association of PRE elements with genes in the active state of transcription and further confirmed when isolated in a reporter construct [42]. In light of these results, the Polycomb complex and TFs appear to be in direct competition for gene regulation.
Temporal progression can also be controlled by limiting the number of physically available DNA regions. Thus, epigenetic modifications control TF activity by regulating accessible and inaccessible chromatin regions (reviewed in [43]). The modifications occurring in the chromatin landscape during neural differentiation transitions have been studied in both Drosophila [44] and murine stem cells [45].
A complete and highly detailed map of the 3D chromatin structural changes characterizing neuronal differentiation in mouse neural cell lines has also been generated [45]. These findings highlight chromatin remodelling during neuronal differentiation, including at the small-scale level of topologically associated domains (TADs). Here, chromatin accessibility is favoured across active regions by decreasing the strength of TAD interactions [45].
In vertebrates, chromatin accessibility is further regulated by two distinct chromatin-modifying complexes: Polycomb repressive complex 1 (PRC1) and Polycomb repressive complex 2 (PRC2). PRC2-mediated gene-silencing occurs through its enzymatic subunits EZH1 and EZH2, which methylate (di- and tri-) the Lys residue of histone H3 (H3K27me2/3). The PRC1 complex catalyses the monoubiquitylation of Lys 119 of histone H2A (H2AK119ub) by the RING finger domain and was implicated in gene regulation by compacting chromatin in an enzymatic activity-independent manner [46,47]. Neurogenic competence was shown to be regulated by both complexes as the removal of RING1B and EZH2, in the developing cortex, prolonged the neurogenic phase, hence delaying the onset of astrogenesis [48]. PRC2 repressive function is also implicated in the fine-tuning between self-renewal and differentiation of cortical progenitor cells. By removing EZH2 expression prior to the onset of neurogenesis, cortical development was characterized by an acceleration of the developmental timing, accompanied by a drastic reduction in the neuronal output and up-regulation in gene expression [49]. A recent study provided additional evidence on its involvement in the epigenetic regulation of temporal progression. In this case, H3k27 histone modification by PRC2 was shown to control mouse aRG temporal progression [19]. Single-cell RNA sequencing (scRNA-seq) was performed on isochronic cohorts of mouse aRGs at various developmental time points, identifying a specific early expression of PRC2 components during corticogenesis. To functionally address the role of this complex, a knockout mouse was generated to abrogate the expression of PRC2 specifically in the cortex. Loss of function resulted in an anticipation of the neurogenic program, therefore reducing the total corticogenic period and the overall cortical thickness [19].

3.2. Cell Cycle as a Switching Event in Temporal Patterning

During CNS development, a key readout for the temporal progression of progenitors is the identity of daughter cells. Yet, the contribution of the cell cycle to this process remains elusive. Several observations have excluded the cell cycle as the main mode of determining tTF progression. Even during stereotypical and fast-paced Drosophila VNC development, each tTF is expressed for a variable number of cell divisions, from a single division for Hb to two or three divisions for other factors [13,24,50]. In addition, NBs can still recapitulate the typical tTF expression sequence when isolated in vitro and arrested in G2, after reaching the Kr expression stage. In light of these results, cell cycle progression appears to be exclusively required by the temporal steps preceding Kr expression [25,51,52,53].
Although the progression of temporal identity does not primarily rely on the cell cycle, it is not completely independent of it. Indeed, temporal progression in NBs serves the purpose of regulating the cellular output resulting from cell division. tTF expression may nonetheless reflect cell cycle events. Such a mechanism has already been described with the activation of one tTF regulator, the orphan nuclear receptor Svp (Nr2f1/2 paralog). Svp is transcribed during cytokinesis and represses Hb, which, in turn, promotes temporal progression in cellular identity acquisition [26,54].
The overall uncoupling of the cell cycle and temporal progression is also observed in mammalian cortical aRGs, the main self-renewing progenitors of the developing cortex [55]. These progenitor cells progress through a temporal program, even upon cell cycle arrest in vivo [55]. Indeed, the temporal transcriptional signatures were maintained in aRGs even when cell cycle arrest was forced via electroporation of the Cdk2 inhibitor [55].

3.3. Extracellular Regulation of Temporal Patterning

In contrast to embryonic stages, the regulation of postembryonic neurogenesis in Drosophila involves progenitor-extrinsic signals [56] (Figure 3a). Indeed, recent lines of evidence have shown extrinsic controls over the developmental events at play during Drosophila larval stages. The properly timed exit from quiescence and cell cycle re-entry of larval NBs are the first events that are partly regulated by extrinsic cues via nutrition-dependent stimuli.
During larval feeding, the liver-like fat body starts secreting a growth factor/mitogen (FBDM) upon the detection of dietary amino acids [57]. In response to these signals, neuroblast-neighbouring subperineurial glial cells (SPGs) produce insulin/insulin-like growth factor (IGF) peptides (dILPs), thus driving NB reactivation [58,59].
Suppression of the cell growth inhibitor Hippo pathway is an additional mechanism through which nutritional cues and glia prompt NB enlargement and cell cycle re-entry. Both NBs and glial cells express the cell surface transmembrane proteins Crumbs and Echinoid that are necessary to activate the Hippo pathway. As a result, NB enlargement and proliferation are hampered. The loss of Crumbs, Echinoid or the core kinase hippo is sufficient to trigger premature NB reactivation from quiescence, therefore stimulating proliferation [60,61]. Glia also negatively regulate the proliferation of NBs by secreting the Drosophila Anachronism (Ana) glycoprotein. The mechanism through which NB quiescence is maintained remains elusive [62].
In addition to nutrition and glia-dependent regulation, temporal patterning of embryonic and larval NBs by extrinsic hormonal cues has been demonstrated in recent studies. The steroid hormone ecdysone is the key regulator that mediates an early-to-late gene expression transition in central brain type II NBs. During early larval stages, Chinmo- and Imp-expressing INPs give rise to neurons and glia. In contrast, later larval lineages solely produce neurons and express Broad and Syncrip. By ensuring the down- and upregulation of the early and later expressed genes, respectively, the ecdysone signalling pathway contributes to the neuronal diversity produced by each larval NB [26].
As previously discussed, in the developing mammalian neocortex, cell-intrinsic temporal programs are crucial for the formation of functional radial units. However, growing evidence further suggests the implication of extrinsic cues on the temporal “tuning” of these events (Figure 3b). For example, fibroblast growth factor 2 (FGF2) promotes cortical progenitor proliferation, leading to increased neuronal and glial output and expansion of the cerebral cortex [63]. Similarly, the Wnt/β-catenin signalling pathway stimulates proliferative divisions, expanding the precursor populations [64]. In line with these results, a recent study demonstrated that inhibition of the Wnt/β-catenin pathway driven by progenitor hyperpolarization promotes a forward shift in the temporal progression of aRG division modes [65]. Meninge-derived retinoic acid (RA) has been suggested to participate in the regulation of progenitor division. The loss of RA signalling in Foxc1 mutants, which lack forebrain meninges, results in prolonged proliferative divisions. Thus, cortical development is hampered by the reduced provision of intermediate progenitors (IPs) and neurons [66]. Finally, the Notch signalling pathway perturbs the neurogenic period by inhibiting neurogenesis [67]. The lack of some effectors involved in Notch signalling leads to premature production of neurons at the expense of neural precursors [68].
By the end of the neurogenic period, progenitors undergo a final temporal transition and switch from neurogenesis to gliogenesis [5]. Several environmental cues act in concert to regulate the appropriate timing of gliogenesis onset [69]. In vitro experiments in which cortical precursors were cultured on embryonic or postnatal cortical slices intriguingly led to the production of neurons and astrocytes, respectively. In this regard, environmental temporal cues appear to strongly influence the neurogenic-to-gliogenic transition [70]. Moreover, several members of the IL-6 neurotrophic cytokine subfamily play a central role in gliogenesis initiation. Through interaction with several coreceptors, the JAK-STAT pathway is consequentially activated, ultimately triggering gliogenesis [71,72,73,74]. The functional interaction between bone morphogenic proteins (BMPs) and Notch ligands together with the aforementioned JAK-STAT pathway also favours the gliogenic switch. The activation of downstream TFs by BMPs promotes the transcription of glial genes, such as Gfap, thus inducing gliogenesis [75,76,77]. Notch signalling is a well-documented pathway involved in the temporal transition from neurogenesis to gliogenesis. The neurogenic-to-gliogenic switch may occur either via activation of the Gfap promoter or by repression of neurogenic bHLHs by Hes proteins when the JAK-STAT pathway is conjointly activated [78,79]. Altogether, these studies unambiguously demonstrate a relevant contribution of cell-extrinsic signals in the temporal course of corticogenesis.

3.4. Perspective on Single-Cell Omics Approaches to Investigate CNS Temporal Patterning

Studies in Drosophila have provided most of our understanding of temporal progression and have inspired parallel investigations on vertebrate models. In recent years, the advent of single-cell omics approaches has provided access to the complexity of vertebrate systems thanks to high-throughput sampling. In this last chapter, we will discuss these methods and their potential for better understanding vertebrate temporal progression, most notably in the neocortex.

3.4.1. Single-Cell RNA Sequencing

The wish to decipher the relation between genotypes and phenotypes has long been present in biology and medicine. In light of this interest, next-generation sequencing technologies underwent important progress in the recent years. As our understanding of cellular complexity and variability advanced, conventional bulk sequencing was subsequently adapted towards transcriptome sequencing at the single-cell level. Almost ten years ago, scRNA-seq started to be used to interrogate the transcriptomic profile associated with single isolated cells. For example, this approach was recently applied to identify and functionally characterize neuron-specific primordial transcriptional programs, as they dynamically unfold during neuronal differentiation [80]. In this study, researchers isolated neurons born at specific time points in the developing neocortex using the FlashTag technique [81], followed by scRNA-seq. A similar strategy was posteriorly used to investigate the temporal progression of cortical development by identifying evolutionarily conserved, temporally patterned genes driving aRGs from internally directed to exteroceptive states [19]. In accordance with other studies, the translation of laminar-specific gene mRNA in progenitors was shown to be halted until their further progression in differentiation. Consequently, progenitor daughter cells are transcriptionally primed into a specific differentiation path [82].
A similar approach has also been used in developing Drosophila to characterize the transcriptional profiles assigned to specific NB types [83]. NB-specific permanent labelling and scRNA sequencing of the ALad1, ALl1, ALv1 and ALv2 Antennal lobe NB lines and the R14D11 and R41D10 Mushroom body NB lines identified a common neuronal fate-enriched set of 75 TFs as well as 68 mushroom body- and antennal lobe-specific TFs, whose expression patterns were previously unknown. More recently, a comparable method was applied to track time-specific diversity across NB lines [84]. Intermediate and ventral column-derived cells and NBs, neurons and glia were isolated at four independent, 2-hour split time intervals using the DIV-MARIS protocol. Here, in addition to confirming the spatiotemporal pattern of already known protein-encoding temporally specific genes, the authors identified strong spatiotemporal patterning in long noncoding RNA (lncRNA). These data highlight the important role these molecules could play during critical early neurogenic stages.

3.4.2. Spatial Transcriptomics

Temporal patterning often results from its interaction with spatially regulated factors across the tissue. However, most scRNA-seq techniques lead to the loss of spatial information. The need for combining transcriptomic and spatial information led to the emergence of spatial transcriptomics, in which a high number of mRNA molecules can be measured with various degrees of spatial resolution. For example, in Slide-seq [85], RNA molecules are transferred from the tissue on a slide covered in uniquely barcoded microbeads of 10 µm diameter, defining its spatial resolution limit. The microbeads are then covered with DNA oligonucleotides that can bind to 3′ RNA poly-A ends [85], allowing cDNA amplification and sequencing. Requiring up to 10-μm-thick cryosections, this technique provides a clear enough resolution to distinguish monocellular layers, such as ependymal cells from the hippocampus [85]. Additional methods based on the multiplexed variation of single-molecule FISH, such as MERFISH and seqFISH [86,87,88,89], provide high sequencing depth and subcellular resolution. In both of these methods, a set of probes covering the full transcriptome is designed and subsequently sequenced in multiple sequential rounds of hybridization.

3.4.3. Single-Cell ATAC Sequencing

As previously discussed, both cis elements and epigenetic states have important intrinsic roles in the complex temporal patterning. A precise mapping of cis regulatory elements along with temporally controlled expression or repression of tTFs during cortical development has not yet been performed.
High-throughput chromatin investigation has been implemented in the context of in vitro models for mammalian stem cells undergoing neuronal differentiation [45]. In this case, mouse embryonic stem cells, neuronal progenitors and cortical neurons were used to perform RNA-seq, ATAC-seq and Hi-C. The progressive maturation towards a neurogenic and neuronal program was investigated by characterizing associated regulatory elements and structural signatures [45]. Both examples used bulk methods to analyse the chromatin accessibility and structure associated with transcriptional profiles, finding interesting chromatin structural signatures for specific developmental trajectories. The main drawback of this approach is the requirement of a homogeneous cellular population from which the molecular material is acquired. Despite its success, it reduces the possibly high and significant heterogeneity of developing cell populations, often subjected to dynamic changes. Thus, translating ATAC-seq towards a single-cell specificity could resolve such complexity [90,91].
Therefore, the implementation of multimodal modelling by combining scRNA-seq and scATAC-seq could provide a more comprehensive picture of single-cell diversity [92]. The systematic investigation of chromosomic structure modification during cortical development will eventually reveal the regulatory elements controlling the temporal progression of such complex systems.

3.4.4. Lineage Barcoding

Studies on the temporal patterning of the Drosophila CNS have provided a detailed description of the robust association of lineages and the respective neuronal types generated. On the contrary, this lineage relationship has not yet been fully addressed in vertebrate neuronal development. Indeed, clear evidence of the generation of diverse cortical cell types through specific lineages is lacking. Future endeavours should focus on identifying single-lineage derived cells, in the context of single-cell transcriptomic studies, by assigning inheritable barcodes to originally labelled aRGs.
New technologies now allow such complex cellular lineage reconstructions [93]. Readers can refer to a recent detailed review on this topic [94]. One recent example is the genetically modified lineage tracing MARC1 mouse line [95]. Based on CRISPR/CAS9 principles, this mouse line incorporates in its genome a set of loci that are translated into self-homing guide RNA molecules constitutively expressed. In the presence of the CAS9 protein, the self-homing guide RNA will drive CAS9 to target their own loci, inducing random mutations that will be inherited by all further generated daughter cells. Random mutations will likely disrupt the self-homing guide RNA properties, preventing further mutations of each locus. Due to the multitude of randomly mutating loci and the diversity in their mutation efficiency, some mutate very fast, while others do so only after multiple days. As a result, each cell in the lineage will yield a unique mutation profile able to recapitulate its lineage history. Further improvement in the experimental design of such approaches remains an important challenge, as the full reconstruction of the cell type lineage history will only be achieved once these mutation profiles are compared at the single-cell level.

3.4.5. Toward High-Throughput Single-Cell Proteomics and Metabolomics

A major advantage of proteomic analysis is the more accurate reflection of the biological processes ongoing in the cell. Most proteomic techniques have so far been performed in bulk analysis, shedding a light on the association of proteins and post-translational modifications with specific functions and pathologies [96,97]. However, major efforts employed in improving the sensibility and multiplexing of these techniques are revolutionizing the nascent field of single-cell proteomics.
One branch of proteomic analysis focuses on the identification of proteins thanks to antibodies. Miniaturized enzyme-linked immunosorbent assays (ELISA) can quantitatively identify protein levels at the single-cell resolution, targeting up to 40 proteins over tens of thousands of cells [98,99,100]. Alternatively, cytometry-based methods (e.g., Fluorescence flow cytometry, mass cytometry) can semi-quantitatively detect around 30 proteins within a single-cell suspension [101,102]. Interestingly, by combining a similar approach with scRNAseq, researchers developed a method for analysing high-throughput single-cell transcriptome and protein abundance for a few selected targets [103].
Mass spectrometry-based methods (e.g., Time of Flight, Electro-Spray Ionization, Orbitrap) give more high-throughput proteomic analysis, identifying up to thousands proteins for each sample by their respective weight [104]. A reduction in working volumes and further multiplexing would give access to single-cell resolution, thus representing a challenge currently being tackled by multiple independent laboratories [105,106,107,108,109].
Recently, researchers have directly applied mass cytometry to histological preparations, providing spatial information to proteomics analysis. Thus, histological preparations are labelled with heavy metal tagged antibodies and measured by mass cytometry upon laser ablation, allowing single-cell and subcellular resolutions. On top of that, this method has also been successfully coupled to RNA single-molecule imaging. In this case, the analysis of both RNA and protein in the same sample was shown to be possible by simultaneously combining single-cell detection of proteins and mRNA for a few target genes and proteins [110,111].
Beyond proteomics analysis, mass spectrometry serves as a appropriate strategy for analyzing metabolites abundance, such as amino acids, sugars, and lipids [112]. These molecules give direct information about ongoing biological processes [113]. Metabolite structures are highly heterogeneous, have a fast-turnover and are currently largely uncharacterized, making their analysis and sample preparation particularly arduous. Therefore, several methods have been recently developed to generate single-cell high-throughput and spatial metabolomics datasets [113,114,115]. In the not-so-distant future, technological advances could extend our ability to describe the heterogeneity of cell types beyond genes and proteins, allowing researchers to tailor their analysis for the modality according to their questions and needs.

3.4.6. Integrating Datasets

The size of omics studies and databases has been expanding at a rapid pace, progressively increasing the variety of structures, conditions, time points and species analysed. Developing methods able to unify heterogeneous datasets is therefore an important current topic in the -omics field. Few methods manage to combine different datasets without compromising the integrity and interpretation of the data. For instance, it is crucial to correct for batch effects when comparing different studies, while still maintaining the feature diversity within the same dataset [116,117,118,119]. The LIGER method tries to circumvent this issue by employing integrative non-negative matrix factorization (iNMF) to identify shared and dataset-specific factors, i.e., genes defining a particular cell type. The dataset-specific effects are then adjusted by a tuning parameter, which reflects the differences in the analysed datasets. Another interesting method is based on the mutual nearest neighbour (MNN) matching, which corrects datasets by recognizing cells that cluster the closest between distinct datasets [116]. Implemented together with canonical correlation analysis (CCA), employed in Seurat V3.0, it determines groups of genes with similar differential expression across batches. Vectors of difference are then used to determine and batch correct anchor points to further transform the datasets [117]. However, the remaining key challenge consists in developing integrative computational tools able to integrate multiple datasets with different experimental contexts and measurement modalities (Figure 4, Table 1). The emergence of these new tools will offer many exciting opportunities in single-cell biology.

4. Conclusions

Temporal regulation over the biological processes that underlie the accurate assembly of a functional CNS is becoming increasingly understood. This temporal regulation is subject to fine-tuning by cell-intrinsic and extrinsic mechanisms. As discussed in this review, pioneering research in the Drosophila system has facilitated major advances in the identification of cell-intrinsic dependent means through which NBs specify distinct neuronal types and generate neuronal diversity [23,24,27,29,122].
In contrast to Drosophila, the temporal dynamics through which mammalian progenitors progressively induce distinct neuronal fates remain poorly understood. Further investigations on the epigenetic mechanisms through which temporal progression is regulated will also deepen our understanding of cell-intrinsic aspects.
In addition to cell-intrinsic temporal regulation, deciphering the role of extrinsic modulation on temporal patterning is fundamental to understanding CNS assembly. The major challenges faced are the loss of surrounding environmental and spatial information that occurs when investigating these complex mechanisms using in vitro systems. Improved methods, in which omics analyses are combined with functional assays while maintaining cellular location, will provide significant insights into the extrinsic controls of temporal patterning. The rapid emergence of new single-cell omics approaches and the growing number of available scRNA-seq datasets have revealed the progenitor transcriptomic profiles over the course of development [19,120]. In silico analyses consisting of integrating several sources of single-cell omics datasets with additional functional analysis may help overcome the above limitations. The integration of multifaceted datasets represents a critical step in understanding and predicting cellular trajectories in the developing CNS.

Author Contributions

A.L., E.M., and L.T. wrote the manuscript together. A.L., E.M. equally contributed to the final manuscript. All authors read and approved the final manuscript.

Funding

Work in the Telley laboratory is supported by an ERC starting grant CERDEV_759112 and an FNS grant 31003A_182676 EM is a postdoctoral fellow supported by the FNS, AL is a doctoral student supported by an ERC grant and LT is supported by an ERC starting grant.

Acknowledgments

We thank Denis Jabaudon, Ilaria Vitali and Gulistan Agirman for scientific discussion, advice and constructive input.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

APApical progenitor
aRGApical radial glia
BMPBone morphogenic protein
CasCastor
CCACanonical correlation analysis
CNSCentral nervous system
CPCortical plate
CRECis regulatory element
dILPsInsulin-like growth factors like peptides
EEmbryonic day
FGF2Fibroblast growth factor 2
FISHFluorescence in situ hybridization
GMCGanglion mother cell
HbHunchback
iNMFIntegrative non-negative matrix factorization
INPIntermediate neural progenitors
IPIntermediate progenitors
KrKrüppel
LIGERLinked inference of genomic experimental relationships
LncRNALong noncoding RNA
MNNMutual nearest neighbor
NBNeuroblasts
NEsNeuroephitelial cells
NSCNeural stem cells
OLOptic lobe
PRC1Polycomb repressive complex 1
PRC2Polycomb repressive complex 2
PREPolycomb responsive element
PhoRCPleiotropic repressive complex
RARetinoic acid
scRNA-seqSingle-cells RNA sequencing
scATAC-seqSingle-cellsSingle cell Assay for Transposase-Accessible Chromatin with high throughput sequencing
SPGSubperineurial glial cells
SVZSubventricular zone
TADTopologically associating domains
TFTranscription factor
tTFTemporal transcription factors
VNCVentral nerve cord
VZVentricular zone.

References

  1. Sansom, S.N.; Livesey, F.J. Gradients in the Brain: The Control of the Development of Form and Function in the Cerebral Cortex. Cold Spring Harb. Perspect. Biol. 2009, 1, a002519. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Pierani, A.; Wassef, M. Cerebral cortex development: From progenitors patterning to neocortical size during evolution. Dev. Growth Differ. 2009, 51, 325–342. [Google Scholar] [CrossRef] [PubMed]
  3. Alfano, C.; Studer, M. Neocortical arealization: Evolution, mechanisms, and open questions. Dev. Neurobiol. 2013, 73, 411–447. [Google Scholar] [CrossRef] [PubMed]
  4. Aboitiz, F.; Zamorano, F. Neural progenitors, patterning and ecology in neocortical origins. Front. Neuroanat. 2013, 7, 38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Gao, P.; Postiglione, M.P.; Krieger, T.G.; Hernandez, L.; Wang, C.; Han, Z.; Streicher, C.; Papusheva, E.; Insolera, R.; Chugh, K.; et al. Deterministic progenitor behavior and unitary production of neurons in the neocortex. Cell 2014, 159, 775–788. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Urbach, R.; Technau, G.M. Neuroblast formation and patterning during early brain development in Drosophila. BioEssays 2004, 26, 739–751. [Google Scholar] [CrossRef]
  7. Schmidt, H.; Rickert, C.; Bossing, T.; Vef, O.; Urban, J.; Technau, G.M. The embryonic central nervous system lineages of Drosophila melanogaster: II. Neuroblast lineages derived from the dorsal part of the nueroectoderm. Dev. Biol. 1997, 189, 186–204. [Google Scholar] [CrossRef] [Green Version]
  8. Ito, M.; Masuda, N.; Shinomiya, K.; Endo, K.; Ito, K. Systematic analysis of neural projections reveals clonal composition of the Drosophila brain. Curr. Biol. 2013, 23, 644–655. [Google Scholar] [CrossRef] [Green Version]
  9. Schmid, A.; Chiba, A.; Doe, C.Q. Clonal analysis of Drosophila embryonic neuroblasts: Neural cell types, axon projections and muscle targets. Development 1999, 126, 4653–4689. [Google Scholar]
  10. Truman, J.W.; Bate, M. Spatial and temporal patterns of neurogenesis in the central nervous system of Drosophila melanogaster. Dev. Biol. 1988, 125, 145–157. [Google Scholar] [CrossRef] [Green Version]
  11. Prokop, A.; Technau, G.M. The origin of postembryonic neuroblasts in the ventral nerve cord of Drosophila melanogaster. Development 1991, 111, 79–88. [Google Scholar] [PubMed]
  12. Tsuji, T.; Hasegawa, E.; Isshiki, T. Neuroblast entry into quiescence is regulated intrinsically by the combined action of spatial Hox proteins and temporal identity factors. Development 2008, 135, 3859–3869. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Doe, C.Q. Temporal Patterning in the Drosophila CNS. Annu. Rev. Cell Dev. Biol. 2017, 33, 219–240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Haubensak, W.; Attardo, A.; Denk, W.; Huttner, W.B. From The Cover: Neurons arise in the basal neuroepithelium of the early mammalian telencephalon: A major site of neurogenesis. Proc. Natl. Acad. Sci. USA 2004, 101, 3196–3201. [Google Scholar] [CrossRef] [Green Version]
  15. Florio, M.; Huttner, W.B.; Aaku-Saraste, E.; Hellwig, A.; Huttner, W.B.; Arai, Y.; Pulvers, J.N.; Haffner, C.; Schilling, B.; Nüsslein, I.; et al. Neural progenitors, neurogenesis and the evolution of the neocortex. Development 2014, 141, 664–679. [Google Scholar] [CrossRef] [Green Version]
  16. Toma, K.; Hanashima, C. Switching modes in corticogenesis: Mechanisms of neuronal subtype transitions and integration in the cerebral cortex. Front. Neurosci. 2015, 9, 274. [Google Scholar] [CrossRef] [Green Version]
  17. Lim, L.; Mi, D.; Llorca, A.; Marín, O. Development and Functional Diversification of Cortical Interneurons. Neuron 2018, 100, 294–313. [Google Scholar] [CrossRef] [Green Version]
  18. Angevine, J.B.; Sidman, R.L. Autoradiographic study of cell migration during histogenesis of cerebral cortex in the mouse. Nature 1961, 192, 766–768. [Google Scholar] [CrossRef]
  19. Telley, L.; Agirman, G.; Prados, J.; Amberg, N.; Fièvre, S.; Oberst, P.; Bartolini, G.; Vitali, I.; Cadilhac, C.; Hippenmeyer, S.; et al. Temporal patterning of apical progenitors and their daughter neurons in the developing neocortex. Science 2019, 364, eaav2522. [Google Scholar] [CrossRef] [Green Version]
  20. Ge, W.P.; Miyawaki, A.; Gage, F.H.; Jan, Y.N.; Jan, L.Y. Local generation of glia is a major astrocyte source in postnatal cortex. Nature 2012, 484, 376–380. [Google Scholar] [CrossRef] [Green Version]
  21. Clavreul, S.; Abdeladim, L.; Hernández-Garzón, E.; Niculescu, D.; Durand, J.; Ieng, S.H.; Barry, R.; Bonvento, G.; Beaurepaire, E.; Livet, J.; et al. Cortical astrocytes develop in a plastic manner at both clonal and cellular levels. Nat. Commun. 2019, 10, 4884. [Google Scholar] [CrossRef] [PubMed]
  22. Kambadur, R.; Koizumi, K.; Stivers, C.; Nagle, J.; Poole, S.J.; Odenwald, W.F. Regulation of POU genes by castor and hunchback establishes layered compartments in the Drosophila CNS. Genes Dev. 1998, 12, 246–260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Brody, T.; Odenwald, W.F. Programmed Transformations in Neuroblast Gene Expression during Drosophila CNS Lineage Development. Dev. Biol. 2000, 226, 34–44. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Isshiki, T.; Pearson, B.; Holbrook, S.; Doe, C.Q. Drosophila Neuroblasts Sequentially Express Transcription Factors which Specify the Temporal Identity of Their Neuronal Progeny. Cell 2001, 106, 511–521. [Google Scholar] [CrossRef] [Green Version]
  25. Kohwi, M.; Doe, C.Q. Temporal fate specification and neural progenitor competence during development. Nat. Rev. Neurosci. 2013, 14, 823–838. [Google Scholar] [CrossRef] [PubMed]
  26. Syed, M.H.; Mark, B.; Doe, C.Q. Steroid hormone induction of temporal gene expression in Drosophila brain neuroblasts generates Summary: Hormone induction of temporal gene expression in neural progenitors. Elife 2017, 6, e26287. [Google Scholar] [CrossRef] [PubMed]
  27. Miyares, R.L.; Lee, T. Temporal control of Drosophila central nervous system development. Curr. Opin. Neurobiol. 2019, 56, 24–32. [Google Scholar] [CrossRef]
  28. Hirono, K.; Margolis, J.S.; Posakony, J.W.; Doe, C.Q. Identification of hunchback cis-regulatory DNA conferring temporal expression in neuroblasts and neurons. Gene Expr. Patterns 2012, 12, 11–17. [Google Scholar] [CrossRef] [Green Version]
  29. Ross, J.; Kuzin, A.; Brody, T.; Odenwald, W.F. Mutational analysis of a Drosophila neuroblast enhancer governing nubbin expression during CNS development. Genesis 2018, 56, e23237. [Google Scholar] [CrossRef] [Green Version]
  30. Barr, I.; Guo, F. Home|kerkida.net. Bio Protoc. 2015, 5, 1–8. [Google Scholar] [CrossRef]
  31. Oberst, P.; Agirman, G.; Jabaudon, D. Principles of progenitor temporal patterning in the developing invertebrate and vertebrate nervous system. Curr. Opin. Neurobiol. 2019, 56, 185–193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Li, X.; Chen, Z.; Desplan, C. Temporal Patterning of Neural Progenitors in Drosophila. In Current Topics in Developmental Biology; Elsevier Inc.: Amsterdam, The Netherlands, 2013; Volume 105, pp. 69–96. ISBN 9780123969682. [Google Scholar]
  33. Murray, A.W. Recycling the Cell Cycle: Cyclins Revisited. Cell 2004, 116, 221–234. [Google Scholar] [CrossRef] [Green Version]
  34. Reppert, S.M.; Weaver, D.R. Coordination of circadian timing in mammals. Nature 2002, 418, 935–941. [Google Scholar] [CrossRef] [PubMed]
  35. Simon, I.; Barnett, J.; Hannett, N.; Harbison, C.T.; Rinaldi, N.J.; Volkert, T.L.; Wyrick, J.J.; Zeitlinger, J.; Gifford, D.K.; Jaakkola, T.S.; et al. Serial regulation of transcriptional regulators in the yeast cell cycle. Cell 2001, 106, 697–708. [Google Scholar] [CrossRef] [Green Version]
  36. Pourquié, O. Clocks regulating developmental processes. Curr. Opin. Neurobiol. 1998, 8, 665–670. [Google Scholar] [CrossRef]
  37. Averbukh, I.; Lai, S.L.; Doe, C.Q.; Barkai, N. A repressor-decay timer for robust temporal patterning in embryonic drosophila neuroblast lineages. Elife 2018, 7, e38631. [Google Scholar] [CrossRef]
  38. Rappaport, N.; Winter, S.; Barkai, N. The ups and downs of biological timers. Theor. Biol. Med. Model. 2005, 2, 22. [Google Scholar] [CrossRef] [Green Version]
  39. Bauer, M.; Trupke, J.; Ringrose, L. The quest for mammalian Polycomb response elements: Are we there yet? Chromosoma 2016, 125, 471–496. [Google Scholar] [CrossRef] [Green Version]
  40. Ringrose, L. Polycomb comes of age: Genome-wide profiling of target sites. Curr. Opin. Cell Biol. 2007, 19, 290–297. [Google Scholar] [CrossRef]
  41. Simon, J.A.; Kingston, R.E. Occupying Chromatin: Polycomb Mechanisms for Getting to Genomic Targets, Stopping Transcriptional Traffic, and Staying Put. Mol. Cell 2013, 49, 808–824. [Google Scholar] [CrossRef] [Green Version]
  42. Erceg, J.; Pakozdi, T.; Marco-Ferreres, R.; Ghavi-Helm, Y.; Girardot, C.; Bracken, A.P.; Furlong, E.E.M. Dual functionality of cis-regulatory elements as developmental enhancers and Polycomb response elements. Genes Dev. 2017, 31, 590–602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Albert, M.; Huttner, W.B. Epigenetic and transcriptional pre-patterning-An emerging theme in cortical neurogenesis. Front. Neurosci. 2018, 12, 359. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Ye, Y.; Li, M.; Gu, L.; Chen, X.; Shi, J.; Zhang, X.; Jiang, C. Chromatin remodeling during in vivo neural stem cells differentiating to neurons in early Drosophila embryos. Cell Death Differ. 2017, 24, 409–420. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Bonev, B.; Cohen, N.M.; Szabo, Q.; Hugnot, J.-P.; Tanay, A.; Cavalli, G.; Bonev, B.; Cohen, N.M.; Szabo, Q.; Fritsch, L.; et al. Multiscale 3D Genome Rewiring during Mouse Article Multiscale 3D Genome Rewiring during Mouse Neural Development. Cell 2017, 171, 557.e1–557.e24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Simon, J.A.; Kingston, R.E. Mechanisms of Polycomb gene silencing: Knowns and unknowns. Nat. Rev. Mol. Cell Biol. 2009, 10, 697–708. [Google Scholar] [CrossRef] [PubMed]
  47. Margueron, R.; Reinberg, D. The Polycomb complex PRC2 and its mark in life. Nature 2011, 469, 343–349. [Google Scholar] [CrossRef] [Green Version]
  48. Hirabayashi, Y.; Suzki, N.; Tsuboi, M.; Endo, T.A.; Toyoda, T.; Shinga, J.; Koseki, H.; Vidal, M.; Gotoh, Y. Polycomb Limits the Neurogenic Competence of Neural Precursor Cells to Promote Astrogenic Fate Transition. Neuron 2009, 63, 600–613. [Google Scholar] [CrossRef] [Green Version]
  49. Pereira, J.D.; Sansom, S.N.; Smith, J.; Dobenecker, M.-W.; Tarakhovsky, A.; Livesey, F.J. Ezh2, the histone methyltransferase of PRC2, regulates the balance between self-renewal and differentiation in the cerebral cortex. Proc. Natl. Acad. Sci. USA 2010, 107, 15957–15962. [Google Scholar] [CrossRef] [Green Version]
  50. Baumgardt, M.; Karlsson, D.; Terriente, J.; Díaz-Benjumea, F.J.; Thor, S. Neuronal Subtype Specification within a Lineage by Opposing Temporal Feed-Forward Loops. Cell 2009, 139, 969–982. [Google Scholar] [CrossRef] [Green Version]
  51. Grosskortenhaus, R.; Pearson, B.J.; Marusich, A.; Doe, C.Q. Regulation of Temporal Identity Transitions in Drosophila Neuroblasts. Dev. Cell 2005, 8, 193–202. [Google Scholar] [CrossRef] [Green Version]
  52. Kanai, M.I.; Okabe, M.; Hiromi, Y. Seven-up controls switching of transcription factors that specify temporal identities of drosophila neuroblasts. Dev. Cell 2005, 8, 203–213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Mettler, U. Timing of identity: Spatiotemporal regulation of hunchback in neuroblast lineages of Drosophila by Seven-up and Prospero. Development 2006, 133, 429–437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Maurange, C.; Cheng, L.; Gould, A.P. Temporal Transcription Factors and Their Targets Schedule the End of Neural Proliferation in Drosophila. Cell 2008, 133, 891–902. [Google Scholar] [CrossRef] [Green Version]
  55. Okamoto, M.; Miyata, T.; Konno, D.; Ueda, H.R.; Kasukawa, T.; Hashimoto, M.; Matsuzaki, F.; Kawaguchi, A. Cell-cycle-independent transitions in temporal identity of mammalian neural progenitor cells. Nat. Commun. 2016, 7, 11349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Syed, M.H.; Mark, B.; Doe, C.Q. Playing Well with Others: Extrinsic Cues Regulate Neural Progenitor Temporal Identity to Generate Neuronal Diversity. Trends Genet. 2017, 33, 933–942. [Google Scholar] [CrossRef]
  57. Britton, J.S.; Edgar, B.A. Environmental control of the cell cycle in Drosophila: Nutrition activates mitotic and endoreplicative cells by distinct mechanisms. Development 1998, 125, 2149–2158. [Google Scholar]
  58. Chell, J.M.; Brand, A.H. Nutrition-responsive glia control exit of neural stem cells from quiescence. Cell 2010, 143, 1161–1173. [Google Scholar] [CrossRef] [Green Version]
  59. Sousa-Nunes, R.; Yee, L.L.; Gould, A.P. Fat cells reactivate quiescent neuroblasts via TOR and glial insulin relays in Drosophila. Nature 2011, 471, 508–513. [Google Scholar] [CrossRef] [Green Version]
  60. Ding, R.; Weynans, K.; Bossing, T.; Barros, C.S.; Berger, C. The Hippo signalling pathway maintains quiescence in Drosophila neural stem cells. Nat. Commun. 2016, 7, 10510. [Google Scholar] [CrossRef] [Green Version]
  61. Poon, C.L.C.; Mitchell, K.A.; Kondo, S.; Cheng, L.Y.; Harvey, K.F. The Hippo Pathway Regulates Neuroblasts and Brain Size in Drosophila melanogaster. Curr. Biol. 2016, 26, 1034–1042. [Google Scholar] [CrossRef] [Green Version]
  62. Ebens, A.J.; Garren, H.; Cheyette, B.N.R.; Zipursky, S.L. The Drosophila anachronism locus: A glycoprotein secreted by glia inhibits neuroblast proliferation. Cell 1993, 74, 15–27. [Google Scholar] [CrossRef]
  63. Vaccarino, F.M.; Schwartz, M.L.; Raballo, R.; Nilsen, J.; Rhee, J.; Zhou, M.; Doetschman, T.; Coffin, J.D.; Wyland, J.J.; Hung, Y.U.E. Erratum: Changes in cerebral cortex size are governed by fibroblast growth factor during embryogenesis. Nat. Neurosci. 1999, 2, 848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Chenn, A.; Walsh, C.A. Regulation of cerebral cortical size by control of cell cycle exit in neural precursors. Science 2002, 297, 365–369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Vitali, I.; Fièvre, S.; Telley, L.; Oberst, P.; Bariselli, S.; Frangeul, L.; Baumann, N.; McMahon, J.J.; Klingler, E.; Bocchi, R.; et al. Progenitor Hyperpolarization Regulates the Sequential Generation of Neuronal Subtypes in the Developing Neocortex. Cell 2018, 174, 1264–1276.e15. [Google Scholar] [CrossRef] [PubMed]
  66. Siegenthaler, J.A.; Ashique, A.M.; Zarbalis, K.; Patterson, K.P.; Hecht, J.H.; Kane, M.A.; Folias, A.E.; Choe, Y.; May, S.R.; Kume, T.; et al. Retinoic Acid from the Meninges Regulates Cortical Neuron Generation. Cell 2009, 139, 597–609. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Gaiano, N.; Nye, J.S.; Fishell, G. Radial Glial Identity Is Promoted by Notch1 Signaling in the Murine Forebrain. Neuron 2000, 26, 395–404. [Google Scholar] [CrossRef] [Green Version]
  68. Yoon, K.; Gaiano, N. Notch signaling in the mammalian central nervous system: Insights from mouse mutants. Nat. Neurosci. 2005, 8, 709–715. [Google Scholar] [CrossRef]
  69. Miller, F.D.; Gauthier, A.S. Timing Is Everything: Making Neurons versus Glia in the Developing Cortex. Neuron 2007, 54, 357–369. [Google Scholar] [CrossRef] [Green Version]
  70. Morrow, T.; Song, M.R.; Ghosh, A. Sequential specification of neurons and glia by developmentally regulated extracellular factors. Development 2001, 128, 3585–3594. [Google Scholar]
  71. Bonni, A.; Sun, Y.; Nadal-Vicens, M.; Bhatt, A.; Frank, D.A.; Rozovsky, I.; Stahl, N.; Yancopoulos, G.D.; Greenberg, M.E. Regulation of gliogenesis in the central nervous system by the JAK-STAT signaling pathway. Science 1997, 278, 477–483. [Google Scholar] [CrossRef]
  72. Koblar, S.A.; Turnley, A.M.; Classon, B.J.; Reid, K.L.; Ware, C.B.; Cheema, S.S.; Murphy, M.; Bartlett, P.F. Neural precursor differentiation into astrocytes requires signaling through the leukemia inhibitory factor receptor. Proc. Natl. Acad. Sci. USA 1998, 95, 3178–3181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Nakashima, K.; Wiese, S.; Yanagisawa, M.; Arakawa, H.; Kimura, N.; Hisatsune, T.; Yoshida, K.; Kishimoto, T.; Sendtner, M.; Taga, T. Developmental requirement of gp130 signaling in neuronal survival and astrocyte differentiation. J. Neurosci. 1999, 19, 5429–5434. [Google Scholar] [CrossRef] [PubMed]
  74. Barnabé-Heider, F.; Wasylnka, J.A.; Fernandes, K.J.L.; Porsche, C.; Sendtner, M.; Kaplan, D.R.; Miller, F.D. Evidence that Embryonic Neurons Regulate the Onset of Cortical Gliogenesis via Cardiotrophin-1. Neuron 2005, 48, 253–265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Li, W.; Cogswell, C.A.; LoTurco, J.J. Neuronal differentiation of precursors in the neocortical ventricular zone is triggered by BMP. J. Neurosci. 1998, 18, 8853–8862. [Google Scholar] [CrossRef]
  76. Nakashima, K. Synergistic Signaling in Fetal Brain by STAT3-Smad1 Complex Bridged by p300. Science 1999, 284, 479–482. [Google Scholar] [CrossRef]
  77. Nakashima, K.; Takizawa, T.; Ochiai, W.; Yanagisawa, M.; Hisatsune, T.; Nakafuku, M.; Miyazono, K.; Kishimoto, T.; Kageyama, R.; Taga, T. BMP2-mediated alteration in the developmental pathway of fetal mouse brain cells from neurogenesis to astrocytogenesis. Proc. Natl. Acad. Sci. USA 2001, 98, 5868–5873. [Google Scholar] [CrossRef] [Green Version]
  78. Ge, W.; Martinowich, K.; Wu, X.; He, F.; Miyamoto, A.; Fan, G.; Weinmaster, G.; Sun, Y.E. Notch signaling promotes astrogliogenesis via direct CSL-mediated glial gene activation. J. Neurosci. Res. 2002, 69, 848–860. [Google Scholar] [CrossRef]
  79. Kamakura, S.; Oishi, K.; Yoshimatsu, T.; Nakafuku, M.; Masuyama, N.; Gotoh, Y. Hes binding to STAT3 mediates crosstalk between Notch and JAK–STAT signalling. Nat. Cell Biol. 2004, 6, 547–554. [Google Scholar] [CrossRef]
  80. Telley, L.; Govindan, S.; Prados, J.; Stevant, I.; Nef, S.; Dermitzakis, E.; Dayer, A.; Jabaudon, D. Sequential transcriptional waves direct the differentiation of newborn neurons in the mouse neocortex. Science 2016, 351, 1443–1446. [Google Scholar] [CrossRef]
  81. Govindan, S.; Oberst, P.; Jabaudon, D. In vivo pulse labeling of isochronic cohorts of cells in the central nervous system using FlashTag. Nat. Protoc. 2018, 13, 2297–2311. [Google Scholar] [CrossRef]
  82. Zahr, S.K.; Yang, G.; Kazan, H.; Borrett, M.J.; Yuzwa, S.A.; Voronova, A.; Kaplan, D.R.; Miller, F.D. A Translational Repression Complex in Developing Mammalian Neural Stem Cells that Regulates Neuronal Specification. Neuron 2018, 97, 520–537.e6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Yang, C.; Fu, C.; Sugino, K.; Liu, Z.; Ren, Q.; Liu, L.; Yao, X.; Lee, L.P.; Lee, T. Transcriptomes of lineage-specific Drosophila neuroblasts profiled by genetic targeting and robotic sorting. Development 2016, 143, 411–421. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. McCorkindale, A.L.; Wahle, P.; Werner, S.; Jungreis, I.; Menzel, P.; Shukla, C.J.; Abreu, R.L.P.; Irizarry, R.A.; Meyer, I.M.; Kellis, M.; et al. A gene expression atlas of embryonic neurogenesis in Drosophila reveals complex spatiotemporal regulation of lncRNAs. Development 2019, 146, dev175265. [Google Scholar] [CrossRef] [Green Version]
  85. Rodriques, S.G.; Stickels, R.R.; Goeva, A.; Martin, C.A.; Murray, E.; Vanderburg, C.R.; Welch, J.; Chen, L.M.; Chen, F.; Macosko, E.Z. Slide-seq: A scalable technology for measuring genome-wide expression at high spatial resolution. Science 2019, 363, 1463–1467. [Google Scholar] [CrossRef] [PubMed]
  86. Lubeck, E.; Coskun, A.F.; Zhiyentayev, T.; Ahmad, M.; Cai, L. Single-cell in situ RNA profiling by sequential hybridization. Nat. Methods 2014, 11, 360–361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Chen, K.H.; Boettiger, A.N.; Moffitt, J.R.; Wang, S.; Zhuang, X. Spatially resolved, highly multiplexed RNA profiling in single cells. Science 2015, 348, aaa6090. [Google Scholar] [CrossRef] [Green Version]
  88. Eng, C.L.; Lawson, M.; Zhu, Q.; Dries, R.; Koulena, N.; Takei, Y.; Yun, J.; Cronin, C.; Karp, C.; Yuan, G.; et al. Transcriptome-scale super-resolved imaging in tissues by RNA seqFISH+. Nature 2019, 568, 235–239. [Google Scholar] [CrossRef]
  89. Xia, C.; Fan, J.; Emanuel, G.; Hao, J.; Zhuang, X. Spatial transcriptome profiling by MERFISH reveals subcellular RNA compartmentalization and cell cycle-dependent gene expression. Proc. Natl. Acad. Sci. USA 2019, 116, 19490–19499. [Google Scholar] [CrossRef] [Green Version]
  90. Cusanovich, D.A.; Daza, R.; Adey, A.; Pliner, H.A.; Christiansen, L.; Gunderson, K.L.; Steemers, F.J.; Trapnell, C.; Shendure, J. Multiplex single-cell profiling of chromatin accessibility by combinatorial cellular indexing. Science 2015, 348, 910–914. [Google Scholar] [CrossRef] [Green Version]
  91. Cusanovich, D.A.; Hill, A.J.; Aghamirzaie, D.; Daza, R.M.; Pliner, H.A.; Berletch, J.B.; Filippova, G.N.; Huang, X.; Christiansen, L.; DeWitt, W.S.; et al. A Single-Cell Atlas of In Vivo Mammalian Chromatin Accessibility. Cell 2018, 174, 1309–1324.e18. [Google Scholar] [CrossRef] [Green Version]
  92. Stuart, T.; Satija, R. Integrative single-cell analysis. Nat. Rev. Genet. 2019, 20, 257–272. [Google Scholar] [CrossRef] [PubMed]
  93. Wagner, D.E.; Weinreb, C.; Collins, Z.M.; Briggs, J.A.; Megason, S.G.; Klein, A.M. Single-cell mapping of gene expression landscapes and lineage in the zebrafish embryo. Science 2018, 360, 981–987. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Wagner, D.E.; Klein, A.M. Lineage tracing meets single-cell omics: Opportunities and challenges. Nat. Rev. Genet. 2020, 21, 410–427. [Google Scholar] [CrossRef] [PubMed]
  95. Kalhor, R.; Kalhor, K.; Mejia, L.; Leeper, K.; Graveline, A.; Mali, P.; Church, G.M. Developmental barcoding of whole mouse via homing CRISPR. Science 2018, 361, eaat9804. [Google Scholar] [CrossRef] [Green Version]
  96. McCormack, A.L.; Schieltz, D.M.; Goode, B.; Yang, S.; Barnes, G.; Drubin, D.; Yates, J.R. Direct Analysis and Identification of Proteins in Mixtures by LC/MS/MS and Database Searching at the Low-Femtomole Level. Anal. Chem. 1997, 69, 767–776. [Google Scholar] [CrossRef]
  97. Jensen, O.N. Interpreting the protein language using proteomics. Nat. Rev. Mol. Cell Biol. 2006, 7, 391–403. [Google Scholar] [CrossRef]
  98. Shi, Q.; Qin, L.; Wei, W.; Geng, F.; Fan, R.; Shin, Y.S.; Guo, D.; Hood, L.; Mischel, P.S.; Heath, J.R. Single-cell proteomic chip for profiling intracellular signaling pathways in single tumor cells. Proc. Natl. Acad. Sci. USA 2012, 109, 419–424. [Google Scholar] [CrossRef] [Green Version]
  99. Love, J.C.; Ronan, J.L.; Grotenbreg, G.M.; van der Veen, A.G.; Ploegh, H.L. A microengraving method for rapid selection of single cells producing antigen-specific antibodies. Nat. Biotechnol. 2006, 24, 703–707. [Google Scholar] [CrossRef]
  100. Su, Y.; Shi, Q.; Wei, W. Single cell proteomics in biomedicine: High-dimensional data acquisition, visualization, and analysis. Proteomics 2017, 17, 1600267. [Google Scholar] [CrossRef]
  101. Herzenberg, L.A.; Parks, D.; Sahaf, B.; Perez, O.; Roederer, M.; Herzenberg, L.A. The History and Future of the Fluorescence Activated Cell Sorter and Flow Cytometry: A View from Stanford. Clin. Chem. 2002, 48, 1819–1827. [Google Scholar] [CrossRef] [Green Version]
  102. Bendall, S.C.; Simonds, E.F.; Qiu, P.; Amir, E.-A.D.; Krutzik, P.O.; Finck, R.; Bruggner, R.V.; Melamed, R.; Trejo, A.; Ornatsky, O.I.; et al. Single-Cell Mass Cytometry of Differential Immune and Drug Responses Across a Human Hematopoietic Continuum. Science 2011, 332, 687–696. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Azizi, E.; Carr, A.J.; Plitas, G.; Cornish, A.E.; Konopacki, C.; Prabhakaran, S.; Nainys, J.; Wu, K.; Kiseliovas, V.; Setty, M.; et al. Single-Cell Map of Diverse Immune Phenotypes in the Breast Tumor Microenvironment. Cell 2018, 174, 1293–1308.e36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Yates, J.R., III. Recent technical advances in proteomics. F1000Research 2019, 8, 351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Specht, H.; Slavov, N. Transformative Opportunities for Single-Cell Proteomics. J. Proteome Res. 2018, 17, 2565–2571. [Google Scholar] [CrossRef] [PubMed]
  106. Virant-Klun, I.; Leicht, S.; Hughes, C.; Krijgsveld, J. Identification of maturation-specific proteins by single-cell proteomics of human oocytes. Mol. Cell. Proteom. 2016, 15, 2616–2627. [Google Scholar] [CrossRef] [Green Version]
  107. Lombard-Banek, C.; Moody, S.A.; Nemes, P. Single-Cell Mass Spectrometry for Discovery Proteomics: Quantifying Translational Cell Heterogeneity in the 16-Cell Frog (Xenopus) Embryo. Angew. Chem. Int. Ed. 2016, 55, 2454–2458. [Google Scholar] [CrossRef]
  108. Li, S.; Plouffe, B.D.; Belov, A.M.; Ray, S.; Wang, X.; Murthy, S.K.; Karger, B.L.; Ivanov, A.R. An Integrated Platform for Isolation, Processing, and Mass Spectrometry-based Proteomic Profiling of Rare Cells in Whole Blood. Mol. Cell. Proteom. 2015, 14, 1672–1683. [Google Scholar] [CrossRef] [Green Version]
  109. Zhu, Y.; Piehowski, P.D.; Zhao, R.; Chen, J.; Shen, Y.; Moore, R.J.; Shukla, A.K.; Petyuk, V.A.; Campbell-Thompson, M.; Mathews, C.E.; et al. Nanodroplet processing platform for deep and quantitative proteome profiling of 10–100 mammalian cells. Nat. Commun. 2018, 9, 882. [Google Scholar] [CrossRef] [Green Version]
  110. Giesen, C.; Wang, H.A.O.; Schapiro, D.; Zivanovic, N.; Jacobs, A.; Hattendorf, B.; Schüffler, P.J.; Grolimund, D.; Buhmann, J.M.; Brandt, S.; et al. Highly multiplexed imaging of tumor tissues with subcellular resolution by mass cytometry. Nat. Methods 2014, 11, 417–422. [Google Scholar] [CrossRef]
  111. Schulz, D.; Zanotelli, V.R.T.; Fischer, J.R.; Schapiro, D.; Engler, S.; Lun, X.; Jackson, H.W.; Bodenmiller, B. Simultaneous Multiplexed Imaging of mRNA and Proteins with Subcellular Resolution in Breast Cancer Tissue Samples by Mass Cytometry. Cell Syst. 2018, 6, 25–36.e5. [Google Scholar] [CrossRef]
  112. Johnson, C.H.; Ivanisevic, J.; Siuzdak, G. Metabolomics: Beyond biomarkers and towards mechanisms. Nat. Rev. Mol. Cell Biol. 2016, 17, 451–459. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Duncan, K.D.; Fyrestam, J.; Lanekoff, I. Advances in mass spectrometry based single-cell metabolomics. Analyst 2019, 144, 782–793. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Alexandrov, T. Spatial Metabolomics and Imaging Mass Spectrometry in the Age of Artificial Intelligence. Annu. Rev. Biomed. Data Sci. 2020, 3, 61–87. [Google Scholar] [CrossRef] [Green Version]
  115. Rappez, L.; Stadler, M.; Triana Sierra, S.H.; Phapale, P.; Heikenwalder, M.; Alexandrov, T. Spatial single-cell profiling of intracellular metabolomes in situ. bioRxiv 2019, 510222. [Google Scholar] [CrossRef] [Green Version]
  116. Haghverdi, L.; Lun, A.T.L.; Morgan, M.D.; Marioni, J.C. Batch effects in single-cell RNA-sequencing data are corrected by matching mutual nearest neighbors. Nat. Biotechnol. 2018, 36, 421–427. [Google Scholar] [CrossRef] [PubMed]
  117. Stuart, T.; Butler, A.; Hoffman, P.; Hafemeister, C.; Papalexi, E.; Mauck, W.M.; Hao, Y.; Stoeckius, M.; Smibert, P.; Satija, R. Comprehensive Integration of Single-Cell Data. Cell 2019, 177, 1888–1902.e21. [Google Scholar] [CrossRef]
  118. Shafer, M.E.R. Cross-Species Analysis of Single-Cell Transcriptomic Data. Front. Cell Dev. Biol. 2019, 7, 175. [Google Scholar] [CrossRef] [Green Version]
  119. Welch, J.D.; Kozareva, V.; Ferreira, A.; Vanderburg, C.; Martin, C.; Macosko, E.Z. Single-Cell Multi-omic Integration Compares and Contrasts Features of Brain Cell Identity. Cell 2019, 177, 1873–1887.e17. [Google Scholar] [CrossRef]
  120. Tasic, B.; Menon, V.; Nguyen, T.N.T.; Kim, T.T.K.; Jarsky, T.; Yao, Z.; Levi, B.B.; Gray, L.T.; Sorensen, S.A.; Dolbeare, T.; et al. Adult mouse cortical cell taxonomy revealed by single cell transcriptomics. Nat. Neurosci. 2016, 19, 335–346. [Google Scholar] [CrossRef] [Green Version]
  121. Zeisel, A.; Hochgerner, H.; Lönnerberg, P.; Johnsson, A.; Memic, F.; van der Zwan, J.; Häring, M.; Braun, E.; Borm, L.E.; La Manno, G.; et al. Molecular Architecture of the Mouse Nervous System. Cell 2018, 174, 999–1014.e22. [Google Scholar] [CrossRef] [Green Version]
  122. Macosko, E.Z.; Basu, A.; Satija, R.; Nemesh, J.; Shekhar, K.; Goldman, M.; Tirosh, I.; Bialas, A.R.; Kamitaki, N.; Martersteck, E.M.; et al. Highly Parallel Genome-wide Expression Profiling of Individual Cells Using Nanoliter Droplets. Cell 2015, 161, 1202–1214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Neural development in Drosophila and in the mammalian cerebral cortex. (a) During embryogenesis, NBs delaminate from the neuroepithelium and then divide asymetrically (type I division mode) to produce a GMC. The latter divides once more to produce a pair of neurons that will be successively organized in a laminar manner: the early-born neurons will be located in deeper layers (light purple layer), whereas late-born neurons will be positioned closer to the surface (dark purple layer). At the larval stage, the larval brain is divided into three main regions: the central brain, the optic lobe, and the VNC, each containing various types of NBs. (b) During cortical development, excitatory neurons are sequentially generated directly from APs or indirectly from IPs. Neurons born earlier during development will settle in the deep layers of the adult cortex, whereas the upper layers will be composed of later-born neurons. At the end of the neurogenic period, APs undergo a self-consuming division to generate glial cells. Abbreviations: AP, apical progenitors; CP, cortical plate; GMC, ganglion mother cell; IP, intermediate progenitors; NB, neuroblast; VNC, ventral nerve cord; VZ, ventricular zone; SVZ, subventricular zone.
Figure 1. Neural development in Drosophila and in the mammalian cerebral cortex. (a) During embryogenesis, NBs delaminate from the neuroepithelium and then divide asymetrically (type I division mode) to produce a GMC. The latter divides once more to produce a pair of neurons that will be successively organized in a laminar manner: the early-born neurons will be located in deeper layers (light purple layer), whereas late-born neurons will be positioned closer to the surface (dark purple layer). At the larval stage, the larval brain is divided into three main regions: the central brain, the optic lobe, and the VNC, each containing various types of NBs. (b) During cortical development, excitatory neurons are sequentially generated directly from APs or indirectly from IPs. Neurons born earlier during development will settle in the deep layers of the adult cortex, whereas the upper layers will be composed of later-born neurons. At the end of the neurogenic period, APs undergo a self-consuming division to generate glial cells. Abbreviations: AP, apical progenitors; CP, cortical plate; GMC, ganglion mother cell; IP, intermediate progenitors; NB, neuroblast; VNC, ventral nerve cord; VZ, ventricular zone; SVZ, subventricular zone.
Ijms 21 07491 g001
Figure 2. Epigenetic and transcriptional regulation of temporal progression. (a) During Drosophila CNS development, temporal progression can be paced by two possible scenarios: the accumulation of a tTF or by the decay of upstream tTF repressors. (b) In the developing Drosophila, gene transcriptional activation (green arrow) can be achieved through the binding of TFs onto CREs. (c) PREs mediate gene silencing (red arrow) by recruiting polycomb complexes. Abbreviations: CRE, cis regulatory elements; TFs, transcription factors; tTFs, temporal transcription factors; PRE, polycomb responsive elements; PhoRC, pleiotropic repressive complex; PRC1, polycomb repressive complex 1; PRC2, polycomb repressive complex 2.
Figure 2. Epigenetic and transcriptional regulation of temporal progression. (a) During Drosophila CNS development, temporal progression can be paced by two possible scenarios: the accumulation of a tTF or by the decay of upstream tTF repressors. (b) In the developing Drosophila, gene transcriptional activation (green arrow) can be achieved through the binding of TFs onto CREs. (c) PREs mediate gene silencing (red arrow) by recruiting polycomb complexes. Abbreviations: CRE, cis regulatory elements; TFs, transcription factors; tTFs, temporal transcription factors; PRE, polycomb responsive elements; PhoRC, pleiotropic repressive complex; PRC1, polycomb repressive complex 1; PRC2, polycomb repressive complex 2.
Ijms 21 07491 g002
Figure 3. Temporal patterning in Drosophila and in the mammalian cerebral cortex. (a) Illustration of the cell-intrinsic and -extrinsic temporal patterning events that take place in the developing Drosophila. During embryogenesis, neuroblasts transition through a cell-autonomous, temporally regulated tTFs cascade to generate distinct neuronal subtypes. In the early larvae, extrinsic cues exert their influence by mediating the exit from quiescence and consequent cell cycle re-entry via nutritional-dependent stimulus. The later temporal switch from early to late tTFs is triggered by extrinsic hormonal cues, in the late larvae stage. (b) Summary of the cell-intrinsic and -extrinsic controls on the temporal progression of mammalian progenitors. During corticogenesis, a six-layer neocortex is generated by the direct or indirect production of neurons from APs or IPs, respectively. The restricted core of TFs drives the deep versus superficial layer identity via cross-regulation. As development proceeds, the transition in neuronal identities is accompanied by the progressive decrease in progenitor competence, hyperpolarization of APs and a switch from a neurogenic to a gliogenic state under the influence of signaling pathways. Abbreviations: AP, apical progenitors; dILPS, Drosophila insulin-like peptides; FBDM, fat body-derived mitogen; IP, intermediate progenitors; NB, neuroblast; RA, retinoic acid; tTF, temporal transcription factor.
Figure 3. Temporal patterning in Drosophila and in the mammalian cerebral cortex. (a) Illustration of the cell-intrinsic and -extrinsic temporal patterning events that take place in the developing Drosophila. During embryogenesis, neuroblasts transition through a cell-autonomous, temporally regulated tTFs cascade to generate distinct neuronal subtypes. In the early larvae, extrinsic cues exert their influence by mediating the exit from quiescence and consequent cell cycle re-entry via nutritional-dependent stimulus. The later temporal switch from early to late tTFs is triggered by extrinsic hormonal cues, in the late larvae stage. (b) Summary of the cell-intrinsic and -extrinsic controls on the temporal progression of mammalian progenitors. During corticogenesis, a six-layer neocortex is generated by the direct or indirect production of neurons from APs or IPs, respectively. The restricted core of TFs drives the deep versus superficial layer identity via cross-regulation. As development proceeds, the transition in neuronal identities is accompanied by the progressive decrease in progenitor competence, hyperpolarization of APs and a switch from a neurogenic to a gliogenic state under the influence of signaling pathways. Abbreviations: AP, apical progenitors; dILPS, Drosophila insulin-like peptides; FBDM, fat body-derived mitogen; IP, intermediate progenitors; NB, neuroblast; RA, retinoic acid; tTF, temporal transcription factor.
Ijms 21 07491 g003
Figure 4. Multi-omics data integration for in silico reconstruction of temporal progression in cortical development. Samples are gathered at multiple time points during cortical development. Independent experiments are performed accessing information from spatial transcriptomics (e.g., FISH-seq), deep single-cell transcriptomic (e.g., scRNA-seq), single cell chromatin accessibility (e.g., scATAC-seq), lineage-seq (e.g., Marc1 mouse), proteomics and metabolomics and ultimately cross referenced with literature data and functional validation. Finally, the cell identities observed can be aligned across different methods, reconstructing emblematic characteristics of single-cells across temporal progression.
Figure 4. Multi-omics data integration for in silico reconstruction of temporal progression in cortical development. Samples are gathered at multiple time points during cortical development. Independent experiments are performed accessing information from spatial transcriptomics (e.g., FISH-seq), deep single-cell transcriptomic (e.g., scRNA-seq), single cell chromatin accessibility (e.g., scATAC-seq), lineage-seq (e.g., Marc1 mouse), proteomics and metabolomics and ultimately cross referenced with literature data and functional validation. Finally, the cell identities observed can be aligned across different methods, reconstructing emblematic characteristics of single-cells across temporal progression.
Ijms 21 07491 g004
Table 1. Main advantages and disadvantages of the various omics techniques.
Table 1. Main advantages and disadvantages of the various omics techniques.
MethodsAdvantagesDisadvantagesReferences
scRNA-seqMeasures gene expression at the single-cell resolution; Straightfoward protocol; Available benchmark analysis; Compatible with scATAC-seqStatic snapshot of cell state; Loss of spatial information[19,80,82,83,84,120]
Spatial transcriptomicsSpatially resolved gene expressionStatic snapshot of cell state; Restricted target number of genes; Tedious protocol and analysis[85,86,87,88,89]
scATAC-seqMapping of the accessible chromatin regions at the single-cell resolution; Straightfoward protocol; Available benchmark analysis; Compatible with scRNA-seqStatic snapshot of the current state; Loss of spatial information; Limited starting material[90,91]
Lineage reconstructionMapping of cellular ontogenyDetectability; Benchmark analysis not available[93,95,121]
scProteomicsMore accurate reflection of a cell stateStatic snapshot of the current state; Unstable starting material; Benchmark analysis not available; Limited access to the technology[98,101,105,110]
scMetabolomicsDirect readout of ongoing biological processes; Closest to the functional phenotypeStatic snapshot of the current state; Lack of standardized data repository; Benchmark analysis not available; Limited access to the technology[112,113]
Spatial proteomics/metabolomicsSpatially resolved proteomic and metabolomic expressionLimited access to the technology[111,115]

Share and Cite

MDPI and ACS Style

Lopes, A.; Magrinelli, E.; Telley, L. Emerging Roles of Single-Cell Multi-Omics in Studying Developmental Temporal Patterning. Int. J. Mol. Sci. 2020, 21, 7491. https://doi.org/10.3390/ijms21207491

AMA Style

Lopes A, Magrinelli E, Telley L. Emerging Roles of Single-Cell Multi-Omics in Studying Developmental Temporal Patterning. International Journal of Molecular Sciences. 2020; 21(20):7491. https://doi.org/10.3390/ijms21207491

Chicago/Turabian Style

Lopes, Andrea, Elia Magrinelli, and Ludovic Telley. 2020. "Emerging Roles of Single-Cell Multi-Omics in Studying Developmental Temporal Patterning" International Journal of Molecular Sciences 21, no. 20: 7491. https://doi.org/10.3390/ijms21207491

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop