Skip to main content

Single-cell RNA sequencing of mid-to-late stage spider embryos: new insights into spider development

Abstract

Background

The common house spider Parasteatoda tepidariorum represents an emerging new model organism of arthropod evolutionary and developmental (EvoDevo) studies. Recent technical advances have resulted in the first single-cell sequencing (SCS) data on this species allowing deeper insights to be gained into its early development, but mid-to-late stage embryos were not included in these pioneering studies.

Results

Therefore, we performed SCS on mid-to-late stage embryos of Parasteatoda and characterized resulting cell clusters by means of in-silico analysis (comparison of key markers of each cluster with previously published information on these genes). In-silico prediction of the nature of each cluster was then tested/verified by means of additional in-situ hybridization experiments with additional markers of each cluster.

Conclusions

Our data show that SCS data reliably group cells with similar genetic fingerprints into more or less distinct clusters, and thus allows identification of developing cell types on a broader level, such as the distinction of ectodermal, mesodermal and endodermal cell lineages, as well as the identification of distinct developing tissues such as subtypes of nervous tissue cells, the developing heart, or the ventral sulcus (VS). In comparison with recent other SCS studies on the same species, our data represent later developmental stages, and thus provide insights into different stages of developing cell types and tissues such as differentiating neurons and the VS that are only present at these later stages.

Peer Review reports

Introduction

Arthropod EvoDevo studies, i.e. the investigation of the development and the evolution of arthropods, still relies mainly on data from a single model organism, the vinegar fly Drosophila melanogaster (reviewed in e.g. [15]. Drosophila development is in many aspects derived however, and the processes and underlying genetic networks that lead to the development from the fertilized egg to the imago (the adult form) cannot easily be compared with the development of other arthropods (e.g. [38]. The introduction of new model organisms, and especially the gaining of deeper insights into existing model organisms that represent less derived modes of development is thus of the greatest interest for arthropod EvoDevo research.

The common house spider Parasteatoda tepidariorum (formerly known as Achaearanea tepidariorum) represents an emerging new model organisms that is often used for comparative developmental studies and the study of arthropod evolution in general (reviewed in [70, 139]. In the last two decades, many standard and advanced molecular biological methods have been established for this species, leading to a large number of comparative studies. Most of these studies, however, have relied on classic candidate gene approaches in which genes that are known to play a certain role during Drosophila development (or other established model organisms) are investigated in non-model organisms or emerging model organisms including Parasteatoda (e.g. [87, 138, 162, 163].

In order to investigate genes (and thus gene regulatory networks) beyond the classic candidate gene approach, genomics and embryonic and tissue-specific transcriptomics have been applied in the recent past, leading to the investigation of complete gene families, orphan genes, fast evolving genes and tissue-specific genes (e.g. [20, 77, 89, 142, 155]. A relatively recent new technical innovation now allows investigation of the transcriptome of every cell of a given organism or a developmental stage of a given organism separately. This so-called single-cell sequencing (SCS) technology can thus give much more detailed insights into gene transcription leading to the identification of genetic fingerprints that are specific for a certain cell type, or the elucidation of trajectories (consecutive developing stages) of differentiating cells (e.g. reviewed in [59, 114].

Two very recent studies applying SCS on embryos of Parasteatoda covered the early stages of spider development. The first paper [2] investigated the so-called germ-disc stage of spider development (stage 5, staging after Mittmann and Wolff [130] (Fig. 1). Early during development, a spherical disc forms that in a subsequent step transforms into a bilaterally-symmetrical germ band (e.g. [130]. The second paper [111] investigated the earliest stages of germ band development (stages 7–9). These stages include the posterior addition of segments, the onset of nervous system development and the beginning of limb bud formation (Fig. 1).

Fig. 1
figure 1

Stages of spider development. Redrawn after Mittmann and Wolff [130]. The 14 previously described developmental stages of the spider Parasteatoda tepidariorum are shown. Stages covered by SCS-analysis of Akiyama-Oda et al. [2] is highlighted in orange. Stages covered by Leite et al. [111] are highlighted in blue, and the stages addressed in our study are highlighted in green

In our study, we investigated developmental stages 10–12. These stages are characterized by for example the end of segment addition, the outgrowth and development of the appendages, nervous system differentiation, heart formation, dorsal closure and ventral splitting of the germ band (Fig. 1). In summary, developmental stages 10–12 likely represent a higher degree of cell-type specification and organogenesis compared with the previous studies on earlier developmental stages. The previous work on the earlier stages of development and our data complement each other and thus each contribute to a larger project driven by the international spider research community to cover all developmental stages of spider development by means of SCS [2, 111]. Our data contribute to this goal by delivering new insights into the differentiation of the spider’s central nervous system (CNS), the patterning of the appendages (including the highly-derived opisthosomal (posterior) appendages), morphogenic movements of cells, and organogenesis. During the course of our investigation, we first characterized the 24 identified cell clusters by means of in-silico analysis (literature analysis) using previously known information on marker genes of each cluster from other arthropods and, whenever available, also spiders including Parasteatoda. In order to test and verify the in-silico based predictions, we then investigated the embryonic expression patterns of 68 genes representing prominent markers of each cell cluster that (for the most) have not been studied previously in Parasteatoda (Fig. 2).

Fig. 2
figure 2

Dotplot showing all genes (Y axis) per cluster (X axis) for which whole mount in situ hybridisation was conducted in this study

Methods

Embryonic tissue dissociation, cell capture, cDNA library preparation, and single-cell RNA sequencing were performed at the Department of Developmental Biology and Gene Core facilities of the European Molecular Biology Laboratory (EMBL), Heidelberg, Germany.

Embryonic tissue dissociation, cell capture, cDNA library preparation, and single-cell mRNA sequencing

In a watch glass, stage 10–12 embryos of Parasteatoda tepidariorum were submerged in phosphate buffered saline (PBS), and the vitelline membranes were removed using tweezers. As much yolk as possible was removed manually with tweezers and a pipette to prevent it from clogging the device used for later cell capture. The tissue-suspension was transferred into 1 ml of PBS and collected in a 1.5 ml low protein binding tube. To reduce the number of fat-droplets, the tissue-suspension was shaken on a shaking-platform at 1100 rpm for 5 min, then 100 µl of surface liquid collecting the fat-droplets was discarded and replaced with fresh PBS. To get rid of any remaining fat-droplets, the suspension was centrifuged at 800 rpm for 4 min, the supernatant was discarded, and the pellet was re-suspended in PBS. The latter centrifugation step was repeated one more time.

Cells were disaggregated (cell dissociation) by intermittent pipetting and swirling of the remaining embryonic tissue in an enzymatic mixture of 2% pronase and 2% sodium thioglycolate in PBS. The dissociation progress was constantly monitored under a microscope. When cell aggregates were no longer visible, the suspension was filtered using a 40 µm cell strainer to remove remaining tissue clumps, yolk granules and debris (e.g. strings of leaked DNA). This cell suspension was then centrifuged at 1100 rpm for 5 min to collect intact cells. The supernatant containing leaked RNA molecules and remaining small pieces of debris was removed. Finally, the cell suspension was filtered through a 20 µm cell strainer twice, and collected into a new 1.5 ml low protein binding tube. Cells were counted using a hemocytometer and cell viability was assayed using Fluorescein Diacetate and Propidium Iodide. A detailed tissue dissociation protocol is available in the supplementary data (Supplementary File 1).

The cell suspension was loaded on a 10X Genomics Chromium Controller™, and cell capture was performed following the Chromium Next GEM Single Cell 3’ version 3.1 protocol, a process by which transcripts from every cell are labelled with a barcoded oligonucleotide. RT-PCR allowed the synthesis of cDNA from each cell’s transcriptome. cDNA quality was checked via electrophoresis on an agarose 2% E-gel precast system (Invitrogen, Cat. No. G402022) (Supplementary Fig. 1). Subsequently, a cDNA library was prepared for sequencing. Paired-end sequencing was performed on an Illumina NextSeq500 platform. Sequenced short reads were de-multiplexed and converted into compressed “(fastq)”-formatted files.

Quality control and mapping of raw data

The bioinformatics tool FastQC (https://www.bioinformatics.babraham.ac.uk/projects/fastqc/) was used to check the quality of the sequencing. A report is available in the supplementary data (Supplementary File 2).

The Parasteatoda genome (version Ptep 3.0) and corresponding annotation were downloaded from NCBI (https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/genome/?term=Parasteatoda+tepidariorum).

High levels of mitochondrial transcripts indicate cell stress, which would render the cells unfit for downstream analysis. To identify and remove such cells from our data, a nearly complete mitochondrial genome was assembled and concatenated to the genome prior to mapping. For its assembly, raw DNA sequencing reads (SRR891584) were used. These reads come from the BioProject PRJNA167405 conducted to assemble the Parasteatoda genome [166]. The mitochondrial genes present in this dataset were searched for and assembled using GetOrganelle v1.7.3.3, a toolkit for de novo assembly of organelle genomes (https://github.com/Kinggerm/GetOrganelle), with kmer sizes 21, 45, 65, 85, 105. The scaffold was annotated using the MITOS server [17], and this information was added to the gene coordinates file (in GTF format) of the genome. The assembled mitochondrial scaffold (“.txt”) is available in the supplementary data (Supplementary File 3).

Mapping of the reads to the reference genome was done using the CellRanger ‘mkref’ and ‘count’ pipelines from the Single Cell Software Suite offered by 10 × Genomics (version 6.0.2). The output included a web summary in “(.html)” format (Supplementary File 3), and a folder containing three compressed files in “(.gz)” format: a list of putative genes and cells (“features.tsv” and “barcodes.tsv”, respectively), and a matrix that displays the number of unique transcripts (UMIs) per cell and per gene (“matrix.mtx”). This information is available in a compressed format “(.zip)” in the supplementary data (Supplementary File 3).

Data processing and downstream analysis

The gene expression matrix was loaded into Seurat ver. 4.1.0 [156] to be processed for downstream analysis (i.e. dimensional reduction, clustering of cells by differential gene expression, identification of cluster markers). In order to filter out low quality cell barcodes (i.e. background RNA encapsulated in droplets, and cells undergoing stress), a subset was made that consisted of cells that express between 250 and 2500 genes, and contained between 1400 and 3500 transcripts, and in which less than 5 percent of transcripts were representing mitochondrial genes. Normalization and variance stabilization of the dataset were performed with the R package SCTransform [30, 61] following the ‘glmGamPoi’ method. A principal component analysis (PCA) was run for linear dimensional reduction. To visualize the data, 50 principal components (PCs) were selected for Uniform Manifold and Projection (UMAP) analysis. To construct a Shared Nearest Neighbours (SNN) graph, 50 PCs were used. Cells were clustered together using the Seurat function “FindClusters” at a resolution of 2. The Seurat function ‘FindAllMarkers’ was used to identify the differentially expressed genes (DEGs) to be used for cluster identification. Only the DEGs that are upregulated (only.pos = TRUE), and those that are expressed in at least 1% of the cells in a cluster (min.pct = 0.01) were considered. For more specific details regarding the parameters followed for cell filtering, clustering, and marker selection, we provide an R notebook file (“.Rmd”) and R object (“.rds”) containing our data analysis, as well as a gene annotation table (“.tsv”). These files are available in the supplementary data (Supplementary File 4). The top most differentially expressed markers per cluster were selected for in silico cell type identification (i.e. by literature review). A spreadsheet containing the markers per cluster is available in the supplementary data (Supplementary File 5). Beyond the in-silico analysis we also conducted additional whole-mount in-situ hybridization (WISH) experiments to a) verify in-silico based cluster-identification of marker genes for which WISH data were not available for Parasteatoda, and b) to investigate the spatial expression of marker genes for which comparative data from spiders and other organisms were not available at all, or not in sufficient extent and quality (and thus were not conclusive in the in-silico analysis).

Gene amplification, probe synthesis, in-situ hybridization, nuclear staining, and data documentation

Total RNA was extracted from a mix of embryos of different developmental stages using TRIZOL (Invitrogen, Cat. No. 15596029). mRNA was isolated from this total RNA using the Dynabeads mRNA Purification Kit (Invitrogen, Cat. No. 61006). cDNA was synthesized from mRNA using RevertAid H Minus First Strand cDNA Synthesis Kit (Thermo scientific, Cat. No. K1631). For fragments of most genes, we performed an initial polymerase chain reaction (PCR), and a subsequent nested (or semi-nested) PCR with a set of internal primers (primer sequences are listed in Supplementary File 5). All backward primers were equipped with a 5´-T7-RNA promotor sequence (gggTAATACGACTCACTATAG) [37]). The three extra guanines serve as protection against degradation of the PCR fragments prior to probe synthesis. PCRs were purified using the QIAGEN PCR purification kit (QIAGEN, Cat. No. 28104). The purified PCRs directly served as templates for subsequent probe synthesis with T7 RNA polymerase (ROCHE, Cat. No. 10881767001). Synthesized probes were purified using the QIAGEN RNeasy Kit (QIAGEN, Cat. No. 74104). Whole mount in-situ hybridization was performed as previously described [85]. Whenever indicated, for better display, embryos were incubated in glycerol and flat-mounted. Appendages were dissected from glycerol-incubated embryos. SYBR Green (incubation of stained embryos in 1:10,000 SYBR Green in phosphate buffered saline with 0.1% Tween-20 (PBST-0.1%) for 20–30 min) was used to better visualize the morphology of embryos. Stained embryos were photographed under a MZ-FLIII Leica dissection microscope equipped with a Leica DC490 digital camera and an external UV-light source. Linear adjustments on colour, contrast and brightness were performed applying the image-processing software Adobe Photoshop 2022.

Results and discussion

Single-cell mRNA sequencing of stage 10–12 spider embryos

Embryonic tissue dissociation and single cell isolation were conducted using a mix of 50 Parasteatoda tepidariorum embryos at stages 10–12. This yielded 70 µl of suspension containing nearly 90% of live cells with a concentration of 920 cells/µl. Single cell mRNA sequencing was performed using equipment and reagents kits provided by the droplet-based technology Chromium Single Cell 3’ Gene Expression™ sold by 10xGenomics®. Following the specifications in the User Guide, 30 000 cells were loaded on the Chromium Controller™. After cell capture, the RT-PCR generated cDNA had a concentration of 1.43 ng/µl. Quality control performed via electrophoresis showed that the cDNA was of good quality, with little degradation of mRNAs (Supplementary Fig. 1). The subsequent cDNA library had a concentration of 37.6 ng/µl. Paired-end sequencing generated 330,047,610 short reads and a sequencing saturation of 11.1%. Quality control via FastQC indicated an overall good base calling process (i.e. high Phred scores for all reads).

Mapping the reads onto the genome, filtering and processing the resulting gene expression matrix using Cell Ranger pipelines and Seurat v4.1.0 yielded a set of 4103 cells expressing 16 669 genes. These cells were grouped into 24 clusters (indicated by Roman Numerals I-XXIV), with 2784 markers in total (Fig. 3; Table 1); note that some genes represent markers for more than one cluster.

Fig. 3
figure 3

Integrated data UMAP showing 24 identified clusters. A short description of the nature of each cluster is given below the UMAP data set

Table 1 Number of markers and cells per cluster. Clusters listed in Roman numerals

Marker genes for each cluster were determined using a “cluster against all-other-clusters” approach, including all genes that were expressed in at least 1% of the cells of their associated cluster and a return threshold p-value of 0.01. “Top-marker genes” are defined as genes with the best in-cluster against all-other-clusters values, i.e. genes with the lowest p-value.

Cluster analysis

Cluster-I: Dorsal tissue, and ectoderm patterning of developing appendages (1)

The top markers of Cluster-I have not been investigated in great detail previously, and it is therefore difficult to determine the nature of Cluster-I cells by in-silico analysis. Somewhat further down the list of markers, however, we find genes that have previously been studied in Parasteatoda and other spiders: Frizzled-4.1 (Fz4.1) (LOC107442148) and Frizzled-4.2 (Fz4.2) (LOC107441380) [82], optomotor-blind (omb)/Tbx3 (LOC107450980) [81], decapentaplegic (dpp) (LOC107441097) [149], and irx2 (LOC107439315) [110]. Fz4.1 and irx2 are strongly expressed in dorsal tissue of the developing Parasteatoda embryo. All of these genes have in common that they are expressed in various patterns in the developing appendages.

We tried to substantiate the suggestion that Cluster-I cells could represent dorsal derivatives of the spider embryo and/or the developing appendages by additional whole-mount in-situ hybridization (WISH) experiments choosing markers that are high up in the list, and that are expressed as tissue (cluster) specific as possible. Therefore, we chose hexosaminidase-1 (hex1) (LOC107442242) and elongation of very long chain fatty acids protein 7 (elovl7) (LOC107446178). These two genes first are expressed in complementing leg gap-gene like patterns proximally and distally respectively and later in rings in the appendages suggesting a function in joint formation. Importantly, however, both genes are also dominantly expressed in dorsal tissue of both the pro- and opisthosoma, while ventral tissue does not express these genes (Fig. 4A-H and Supplementary Figs. 2 and 3). This pattern is very similar to the patterns of irx2 and Fz4.1 in dorsal tissue supporting the suggestion that the Cluster-I represents dorsal tissue.

Fig. 4
figure 4

Gene expression of Cluster-I, -II, and -IV markers. In all panels, anterior is to the left, except panels D, H, L, P, T and X where anterior is up. Panels A, I, K, M and R represent lateral views. Panels B, C, E, F, G, J, N, Q, S, V and W represent ventral views. Panels D, H, L, P, T and X represent anterior views. Schematic drawings in the right column summarize the main gene expression characteristic for a given cell cluster (blue tissue). Arrows and arrowheads in schematic drawings point to comparable expression as marked by the same symbols in the original photographs. Abbreviations of gene names is indicated in the upper left corner. Developmental stages are indicated in the bottom right corner. For further information, see Supplementary Figs. 24 and 911, and Supplementary File 6. Abbreviations used in schematic drawings: aSP, anterior pair of spinnerets; bl, book lung; ch, chelicera; CNS, central nervous system; h, heart; L1-L4, first to fourth leg; lr, labrum; m, mouth; pp, pedipalp; pSP, posterior pair of spinnerets; tr, trachea; vs, ventral sulcus

With respect to the somewhat later expression of Cluster-I genes in what we believe are the developing joints, it is worth mentioning that also the Cluster-I marker dpp is expressed in a comparative pattern in the developing joints [149]. It is thus likely that Cluster-I indeed also harbours cells that are involved in spider joint formation.

Cluster-II: Ectoderm patterning of developing appendages (2)

Most top markers of Cluster-II represent uncharacterized genes. Among the characterized markers, however, range a number of genes that have previously been described for Parasteatoda, other spiders and/or other arthropods. Notably, a large number of these markers are expressed in the ectoderm of developing appendages, and are involved in appendage development, such as disconnected (disco) (aka basonuclin) (LOC107447159) (Fig. 4I-L and Supplementary Fig. 4), spineless2 (ss2) (LOC107457395) (Fig. 10F-H, cf. Cluster-XVI), aristaless (al) (LOC107448374), AP2.2 (LOC107443623) (Fig. 9U-X, cf. Cluster-XV) (note that Leite et al. [111] investigated another paralog of this gene, AP2 (LOC107452006)), Fz4.1 (LOC107442148), Fz4.2 (LOC107441380), and unchoordinated-5.1 (unc5.1) (LOC107445619) (e.g. [4, 6, 45, 64, 82, 83, 88, 109, 111, 132, 134, 141, 150, 160, 167].

We therefore believe that Cluster-II cells generally contribute to the developing appendages. At least some of the investigated genes are expressed in the form of rings that could indicate a function in proximal–distal axis patterning and thus joint formation.

Cluster-III: The developing central nervous system (CNS) (1) – Differentiating and differentiated neurons (1)

In-silico analysis of Cluster-III gene markers strongly suggests that these cells represent part of the developing nervous system. Best markers are sax3 (LOC107455609), and two paralogs of neural-cadherin (Ncad1 (LOC107454545) and Ncad2 (LOC107451854)). Other high-ranked markers are follistatin-related protein 5 (fstl5) (LOC107441591), Dscam-2 (LOC107456604), pikachurin (pika) (LOC107439395), and ELAV-like protein 3 (Elavl3) (LOC107436216). Of these markers, sax3, pika, fstl5, and Elavl3 are unique markers for this cell cluster (see Supplementary Fig. 5). Ncad1, Ncad2 and Dscam2 are markers of both Cluster-III and the closely related Cluster-XII (cf. chapter on Cluster-XII). These genes are all involved in neuron development, neuron differentiation, and neuronal pathfinding, and thus also in differentiated neurons (e.g. [36, 43, 66, 75, 92, 140, 151, 191, 203]. Beyond that, many of the lesser markers of this cluster indeed also represent genes typically expressed in neurons, such as the Dscams (e.g. LOC107456604, LOC107444681, LOC107450135, LOC107445852, LOC107452025, LOC107448201, LOC107443039, LOC107455980) (e.g. reviewed in [171], and synaptotagmins (LOC107455395, LOC107449342) (e.g. reviewed in [195]. It is therefore reasonable to suggest that Cluster-III cells represent differentiating and mature neurons.

The expression and function of most of these genes has not been studied in great detail in arthropods other than Drosophila, and thus data from spiders are not available. In-situ hybridization of some of these genes show, as expected, that they are expressed in the developing central nervous system (CNS), and some are expressed in the peripheral nervous system (PNS) too (Fig. 5A-L and Supplementary Figs. 58).

Fig. 5
figure 5

Gene expression of Cluster-III, XII, and -XVI markers. In all panels, anterior is to the left. All panels represent ventral views. Panels D, H, L, P, T, X, b, and f show flat-mounted embryos. Schematic drawings in the right column summarize the main gene expression characteristic for a given cell cluster (blue tissue). Arrow in schematic drawings point to comparable expression as marked by the same symbol in the original photographs. Abbreviations of gene names is indicated in the upper left corner. Developmental stages are indicated in the bottom right corner. For further information, see Supplementary Figures B5-B8, B28-B31 and B46-B48, and Supplementary File 6. For abbreviations used in schematic drawings see legend of Fig. 4

Cluster-IV: Ectoderm patterning of developing appendages (3)

Most of the top markers of Cluster-IV have not been studied in great detail, or not at all. Two of these markers, cuticular protein 47Ef-like (cp47EF-like) (LOC107436879) and chitinase-7 (cht7) (LOC107437310), are exclusively expressed in the developing joints (Fig. 4M-T and Supplementary Figs. 9 and 10). Another marker, the protease notopleural (np) (LOC107439948), is expressed in dorsal tissue and what we believe are the developing joints of the appendages, very similar like the markers of the closely-related Cluster-I and the aforementioned markers of Cluster-IV (Fig. 4U-X and Supplementary Fig. 11). Interestingly, another marker of Cluster-IV is trachealess (trh) (LOC110283103) which has been studied in Parasteatoda and other chelicerates such as a harvestman and a scorpion, is also expressed in the form of rings in the developing appendages suggesting a possible role in joint formation [168]. Likewise, AP2.2 (LOC107443623) (cf. Cluster-XV) is expressed in the form of rings in the appendages (Fig. 9U-X). Other previously investigated and less prominent markers of this cluster (e.g. the two Fz4 ohnologs (LOC107442148, LOC107441380) [82], Wnt6, Wnt5, and Wnt1 (LOC107438387, LOC107445649, and LOC107438386) [87], and unc5 [88] are also expressed in the developing appendages, albeit not in the form of rings. Gene expression analysis of the high-ranked markers of this cluster, however, suggests that these cells could contribute to the development of the joints of the spider appendages.

Cluster-V: The developing central nervous system (CNS) (2) – EMT-like processes in neural precursor determination?

Top markers of this cluster are neurotactin (nrt) (LOC107441543), magu (LOC107453771), zinc finger protein 395-like (zfp395-like) (LOC122271089), fasciclin-2 (fas2) (LOC107436358), noggin (nog) (LOC107444265) and otopetrin-2-like (otop2) (LOC107451015).

In Drosophila, nrt is involved in morphogenic movements and is expressed in dynamic patterns in both mesodermal and ectodermal tissues including the developing nervous system where it is expressed strongly [39]. The expression of Drosophila magu is not known, but over-expression of magu in Drosophila causes an elongated life span, especially when it is overexpressed in the nervous system [112]. To our knowledge, there are no data on zfp395-like. In Drosophila, the cell adhesion molecule fas2 is an important factor of synaptic growth and maintenance [57, 165]. In vertebrates, nog1 induces neural tissue development (reviewed in e.g. [98]. These genes have not been studied in spiders prior to this study, except for the recent data on nog showing expression in earlier developmental stages [111].

Some other significant markers of Cluster-V, however, have been investigated previously in Parasteatoda or other species of spiders. netrin-2 (net2) (LOC107450632) for example is prominently expressed in the developing nervous system, the heart, and the appendages [88, 118], and snail (sna) (LOC107443696) is prominently expressed in the developing nervous system of the American wandering spider Cupiennius salei [177, 193]. In summary, these data suggest that Cluster-V cells are involved in the development of the nervous system. Our gene expression analysis supports this as all investigated genes of this cluster are expressed inter alia in the developing CNS (Fig. 6A-F and Supplementary Figs. 1215).

Fig. 6
figure 6

Gene expression of Cluster-V, -XIV, -XVIII and -XXII markers. In all panels, anterior is to the left. All panels represent ventral views, except panels K (lateral view), N (dorsal view), and V/V´ (anterior view). T´ and V´ represent SYBR-green staining of the embryos shown in panels T and V. Schematic drawings in the right column summarize the main gene expression characteristic for a given cell cluster (blue tissue). Arrows and arrowheads in schematic drawings point to comparable expression as marked by the same symbol in the original photographs. For Cluster-XVIII, it was not possible to determine a specific tissue marked by these genes. Abbreviations of gene names is indicated in the upper left corner. Developmental stages are indicated in the bottom right corner. For further information, see Supplementary Figs. 1215, 3537, 4951, and 6063, and Supplementary File 6. For abbreviations used in schematic drawings see legend of Fig. 4

Unlike the expression of other CNS-Cluster genes (e.g. clusters VI, X, XII, and XVII), markers of Cluster-V are also expressed in other tissues such as the visceral mesoderm, the heart and the dorsal field. These aspects of expression are very similar to the pattern of Cluster-VII (discussed below) and may hint to a function in visceral mesoderm development and epithelial-to-mesenchymal transition (EMT). Although Cluster-V cells still show a neural fingerprint, this fingerprint hints to a “near” transition of these cells into a possible mesenchymal nature. Alternatively, Cluster-V cells represent another developmental state of CNS development in which future neural precursors are selected and delaminate from the “neural epithelium”, in an EMT-related process (e.g. [7].

Cluster-VI: The developing central nervous system (CNS) (3) – Early differentiating neural cells (1)

Top markers of Cluster-VI are delta-2 (LOC107438011), and the uncharacterized gene unc1852 (LOC107451852). delta is a well-known neurogenic gene (e.g. [60, 190]. Two other genes that are markers of Cluster-VI, cyclin D (cycD) (LOC107457517) and cyclin dependent kinase 1 (CDK1) (LOC107436497) are known factors of cell cycle regulation and have been reported to be highly expressed in the developing nervous system (e.g. [58, 71, 86]. Further down the list of markers, we also find the genes dachshund (dac) (LOC107453438) and Nkx6.2 (LOC107450777) (the latter has recently been studied by [111]. We confirm in this study that Nkx6.2 (Fig. 10Q-T, cf. Cluster-XIX) and delta-2 (Fig. 7A-D and Supplementary Fig. 16) are expressed exclusively in the developing nervous system of the spider. Previous studies have shown that dac is expressed in the developing nervous system of Parasteatoda and other spiders [144, 149, 187]. These data and the fact that Cluster-VI is encircled by other cell clusters that likely represent cells of the developing spider nervous system (cf. clusters V, VIII, X, XI, and XIX), strongly suggest that Cluster-VI represents a subtype of cells of the developing nervous system. Notably, none of the detected Achaete-Scute complex (ASH-C) genes is expressed in Cluster-VI, but are expressed in the neighbouring clusters XIII, X and XVII (Supplementary File 5), suggesting that Cluster-VI cells do not represent the earliest steps of CNS development.

Fig. 7
figure 7

Gene expression of Cluster-VI, -VIII, and -X markers. In all panels, anterior is to the left. All panels represent ventral views. Panels D, H, L, P, T and X represent flat-mounted embryos. Schematic drawings in the right column summarize the main gene expression characteristic for a given cell cluster (blue tissue). The arrow in the schematic drawing points to comparable expression as marked by the same symbol in the original photographs. Abbreviations of gene names is indicated in the upper left corner. Developmental stages are indicated in the bottom right corner. For further information, see Supplementary Figs. 16, 2023, 26, and 27, and Supplementary File 6. For abbreviations used in schematic drawings see legend of Fig. 4

Cluster-VII: The developing mesoderm (1) – EMT-like processes in visceral mesoderm development?

In-silico analysis of Cluster-VII markers gives some insight into the possible nature of the cells represented by this cluster. Top markers are integrin alpha-PS2 (PS2) (LOC107444850), collagen alpha-1(I) chain (LOC107442626), its paralog (LOC107443413), several other collagen genes (LOC107442286, LOC107441250, LOC107442289), several laminin genes (LOC107439460, LOC107452433, LOC107448892), papilin (LOC107447945), and a fibrinogen-like gene (LOC107449293).

Many of these genes represent cell adhesion molecules needed for cell migration as present during epithelial-to-mesenchymal transition (EMT) (reviewed in e.g. [14, 73]. These and other genes such as ECM protein 3-like (LOC122271235) and ECM organizing protein Fras1 (LOC107444379) are components of the extracellular matrix (ECM) that is used as a substrate for cell migration (reviewed in e.g. [181]. Interestingly, in Drosophila, the visceral mesoderm serves as a substrate for many populations of migrating cells (e.g. [22, 153], thus connecting cell migration, visceral mesoderm and the ECM. Among the integrins, PS2, the top marker of this cluster, is specifically expressed in the visceral mesoderm (e.g. [21, 128]. Another group of genes that are among the top markers of this cluster are the fibroblast growth factor receptors (FGFRs) (LOC107437526, LOC107438841, LOC107445728) which are involved in the development of the visceral mesoderm, somatic muscles and the heart (e.g. [122, 172]. papilin is an important factor of basement membrane development of somatic and visceral muscles, interacts with collagen genes, and is generally involved in cell movement [28, 97], reviewed in [48]. Finally, we also found FoxF1 (LOC107456534) to be a marker of this cell cluster, a gene that is known to be involved in the development of the visceral mesoderm in Drosophila (e.g. [78, 146]. A recent study has shown that in Parasteatoda, FoxF1 is almost exclusively expressed in the visceral mesoderm of the opisthosoma [89].

We performed in-situ hybridization experiments with top markers of this cluster and show that they all are expressed inter alia in the visceral mesoderm of the trunk, the mesoderm of the appendages, and the developing heart (Fig. 8A-F and Supplementary Figs. 1719). It is thus likely that Cluster-VII genes represent cells that are in the process of EMT and that are developing towards becoming cells of the visceral mesoderm (cf. Cluster-V).

Fig. 8
figure 8

Gene expression of Cluster-VII, -XX, -XXI and -XXIII markers. In all panels, anterior is to the left. Panels A, B, E, F, U and V represent ventral views. Panels C, D, M, P, R, T and X represent lateral views. Panels G-L, N, O, Q, S and W represent dorsal views. Schematic drawings in the right column summarize the main gene expression characteristic for a given cell cluster (blue tissue). Arrows and asterisks in schematic drawings point to comparable expression as marked by the same symbol in the original photographs. Abbreviations of gene names is indicated in the upper left corner. Developmental stages are indicated in the bottom right corner. For further information, see Supplementary Figs. 17, 18, 5456, 5759, and 6466, and Supplementary File 6. For abbreviations used in schematic drawings see legend of Fig. 4

Cluster-VIII: The developing central nervous system (CNS) (4) – Early differentiating neuronal cells

The top marker of this cluster is couch potato (cpo) (LOC107453124), a gene that in Drosophila is involved in the development of neuronal precursors, and the sensory nervous system (e.g. [16, 54]. Other top markers are Chronophage (Cph) (LOC107438166), the two ohnologs of delta (LOC107456525, LOC107438011), CD109 antigen (CD109) (LOC107436484) and epidermal growth factor receptor (egfr) (LOC107446048). Cph has recently been identified as a temporal switch of nerve cell subtype specification [51]. One of the two paralogs of Parasteatoda delta (LOC107456525) has been described previously, and it has been shown to be expressed in the developing nervous system around stage 9 (later stages were not presented in this study) [138]. In another spider, the American wandering spider Cupiennius salei, both ohnologs of delta have been investigated, and it has been shown that both are expressed early in the developing nervous system [176]. In Drosophila, delta represents a neurogenic gene, and is thus involved in cell fate determination within the developing nervous system (reviewed in e.g. [136]. CD109 is involved in Drosophila septate junction formation, but to our knowledge there is no data about its potential function during nervous system development [10]. Finally, egfr represents an important factor of nervous system development as EGFR-signalling interacts with TOR-signalling and thereby contributes to neuronal differentiation [93, 94]. In-silico analysis thus suggests that Cluster-VIII cells represent nerve cells that are in the (relatively) early process of differentiation. We investigated the embryonic gene expression profiles of cpo, delta, delta-2, CD109, and an uncharacterized marker of this cluster (unc6289) (LOC107436289). These genes all are expressed almost exclusively in the developing nervous system, thereby supporting the conclusion drawn from in-silico analysis (Fig. 7E-P and Supplementary Figs. 2023).

Cluster-IX: Opisthosomal appendages

The two top markers of this cluster represent the uncharacterized genes unc0558 (LOC107440558) and unc2247 (LOC107452247). These genes, and other markers of this cluster do not allow any in-silico prediction. These two genes are expressed in the developing spinnerets and book lungs (Fig. 9A-H and Supplementary Figs. 24 and 25). Cluster-IX could thus represent cells of the highly-modified opisthosomal appendages. Another previously studied gene of this cluster is one of the two ohnologs of trachealess (trh) (LOC107455153) which is expressed more specifically in the developing tracheae of spiders than its paralog (LOC110283103) which itself is expressed in the developing tracheae, but also strongly in legs and dorsal tissue (a marker of clusters IV, XIII, XV, and XVIII) (Zhang (née Turetzek) 2016).

Fig. 9
figure 9

Gene expression of Cluster-IX, -XIII, and -XV markers. In all panels, anterior is to the left. Panels A, C, E, F, H, I-K, M, O, Q-S, V, W, Y-a, and c-e represent ventral views. Panels B, G, and N represent lateral views. Panel D represents dorsal view. Panels L, P, T, X, b, and f represent anterior views on flat-mounted head regions. Schematic drawings in the right column summarize the main gene expression characteristic for a given cell cluster (blue tissue). Arrows, arrowheads and asterisks in schematic drawings point to comparable expression as marked by the same symbol in the original photographs. Abbreviations of gene names is indicated in the upper left corner. Developmental stages are indicated in the bottom right corner. For further information, see Supplementary Figs. 24, 25, 3234, and 3841, and Supplementary File 6. For abbreviations used in schematic drawings see legend of Fig. 4

Cluster-X: The developing central nervous system (CNS) (5) – Early differentiating neural cells (2)

One of the two ohnologs of the neurogenic gene delta, delta-2 (LOC107438011) represents the strongest marker of this cluster, followed by CD109 (OC107436484) (cf. Cluster-VI). Other high-ranked markers are the uncharacterized genes unc1847 (LOC107451847) (cf. clusters VI and XIII) and unc3246 (LOC110283246) (cf. clusters VI, VIII and XI), and the histone-lysine N-methyltransferase MECOM (LOC107436282) (cf. clusters V, VI, VIII, XI, XVII, and XIX). Cluster-X cells thus clearly share a large number of markers with its neighbouring clusters VI, VIII and XI. One potentially crucial difference between Cluster-X and clusters VI and VIII could be the presence of cyclin-dependent kinase inhibitor 1 (CKI1) (LOC107443706) which likely interacts with cycD (another high-ranked marker of Cluster-X) (LOC107457517) (cf. clusters VI and VIII) [170]. Clusters VI, VIII and X could thus represent cells in a different stage of cell cycle (see distribution of cycD, CDK1, CKI1 genes) in the developing nervous system (reviewed in [117]. Interestingly, CKI1 is also involved in neural stem cell differentiation and neuron induction by silencing Sox2 (= SoxN), a top marker of cluster-XI [113, 126], reviewed in [117]. Unfortunately, we do not have gene expression data on cycD, CDK1 and CKI1, but it has recently been proposed that cycD and CDK1 (and other cycs and CDKs) are expressed in waves of expression in the nervous system of the onychophoran Euperipatoides kanangrensis [86]. Cluster-X or clusters VI, X and VIII could thus represent early differentiating neural cells, and Cluster-XI, in which SoxN (LOC107457313) is highly expressed, could represent neural precursors (discussed below). In line with this assumption is the fact that the proneural gene achaete scute homolog 1 (ash1) (LOC107451231) is also among the higher-ranked markers of Cluster-X. In Drosophila, achaete-scute complex genes are expressed in segregating neuroblasts (neuronal precursors), but not in neurons [25, 174, 175]. Interestingly, however, studies in insects other than Drosophila, such as the beetle Tribolium castaneum, and in crustaceans, have shown that the function of ash as the first gene to be expressed in developing neuroblasts is not conserved, but instead there seems to be more flexibility in the patterning of the early nervous system (e.g. [188, 194]. What all these studies show, however, is that ash is an early factor of nervous system development.

In-situ hybridization analysis of previously unstudied genes that are among the top markers of this cluster reveals that they, as expected, are expressed predominantly in the developing nervous system (Fig. 7Q-X and Supplementary Figs. 26 and 27) (cf. Cluster-VIII for CD109 expression and Cluster-VI for delta-2 expression).

Cluster-XI: The developing central nervous system (CNS) (6) – Neural precursors

Two highly-ranked markers of Cluster-XI are SoxN (LOC107457313), a well-known factor of stem cell maintenance in the developing nervous system [113, 126], reviewed in [47], and snail (sna) (LOC107443696), an important factor of stem cell maintenance in the developing nervous system (e.g. [9, 27]. Both genes are prominently expressed in the developing CNS of spiders [20, 177, 193]. Most interestingly, both genes represent proneural genes that link neural progenitor (neuroblasts in Drosophila) selection with epithelial-to-mesenchymal transition (EMT) [7] (cf. sna as a marker of Cluster-V). Another highly-ranked marker of this cluster is another potential proneural gene in Drosophila, goosecoid (gsc) (LOC107439785) [62]. gsc is involved in neuroblast delamination in vertebrates (e.g. [96]. The homeobox gene Dbx (LOC122269629) is expressed in the developing nervous system of Parasteatoda [110], and in vertebrates, its two paralogs are expressed in neural progenitors (e.g. [148]. In Drosophila, the single Dbx gene is expressed in interneurons that form from a subset of neural progenitor cells (neuroblasts) [106]. Finally, the trunk-determining zinc finger transcription factor tiptop/teashirt (tio/tsh) (LOC107441771) is expressed in the CNS of the trunk of Parasteatoda [125, 129]. In Drosophila, tio and tsh are expressed in the trunk CNS as well, and it has been shown that both genes can mediate the survival of neurons [107, 135]. Given the genetic fingerprint of Cluster-XI it is possible that these cells represent neural precursor cells.

Cluster-XII: The developing central nervous system (CNS) (7) – Differentiating and differentiated neurons (2)

Cluster-XII is closely related to Cluster-III. Top markers of Cluster-XII are the uncharacterized gene unc3142 (LOC107443142), defective proboscis extension response 6 (dpr6) (LOC107436606), neural-cadherin 2 (Ncad2) (LOC107451854), collier (col) (aka knot) (LOC107437107) and a scratch-like (scrtl) gene (LOC107456893).

col has previously been studied in spiders, and indeed a wide range of arthropods, showing that it has a conserved role in nervous system development and neuronal differentiation [13, 35, 41, 158]. Among arthropods, scrt has only been investigated in Drosophila (CG1130) where it is expressed in most (or all) neural precursors promoting neuronal development [154]. Information on Drosophila scrtl (CG12650) is restricted to gene expression data, showing that it is expressed in the developing CNS (BDGP in situ homepage: https://insitu.fruitfly.org/cgi-bin/ex/report.pl?ftype=3&ftext=LP01683). dpr6 and Ncad2 are also key-factors of Drosophila nervous system development that are involved in the self-organization of neurons and the specification of neuronal subtypes (e.g. [3, 19, 34, 186]. Together, this strongly suggests that Cluster-XII represents cells of the developing nervous system, and indeed, in-situ hybridization experiments confirm this suggestion: In Parasteatoda, unc3142, scrtl, dpr6 and Ncad2 all are expressed exclusively in the CNS (Fig. 5M-X and Supplementary Figs. 2831). Beyond that, Cluster-XII is like its neighbouring cluster (Cluster-III) characterized by a significant number of Dscams (LOC107450135, LOC107445852, LOC107444681, LOC107456604). In summary, it is therefore likely that Cluster-XII cells represent differentiating and differentiated neurons (cf. Cluster-III).

Cluster-XIII—Ectoderm patterning of developing appendages (4)

Among the markers of Cluster-XIII are very few previously investigated genes. Top markers are elongation of very long chain fatty acids protein 7.2 (elovl7.2) (LOC107441276) and endochitinase A (endoA) (LOC107442511). We investigated the in-situ hybridization patterns of these two markers and another top marker, the uncharacterized gene unc9645 (LOC107439645). All three marker genes have in common that they are expressed in the form of rings in the developing appendages, indicating a function in appendage ectoderm patterning and thus likely also joint formation (F9g. 9I-T and Supplementary Figs. 3234).

Cluster-XIV: The ventral sulcus (VS)

The top markers of Cluster-XIV represent two Tenectin-like genes, Tenectin1 (Tnc1) (LOC107436695) and Tenectin1 (Tnc2) (LOC122273808). Additional top markers of this cluster are a sidestep (side)/hemicentin-2 like gene hereafter called sidestep VIII (sideVIII) (LOC107456176), netrin-1 (LOC107455212) and slit (LOC107443293 & LOC122270376). netrin-1 has previously been investigated in spiders. It is expressed inter alia in the ventral sulcus (VS) (aka the ventral midline epithelium [119] where it is likely involved in axon guidance [88, 118]. slit genes encode ligands of the Roundabout receptors and are thus important key players in axon guidance as well (reviewed in [199]. Likewise, hemicentin-2/sidestep functions as an important guidance cue for growing axons (e.g. [173]. Finally, also Tnc is involved in nervous system development in Drosophila where it is expressed in both longitudinal and commissural axon tracts that span the ventral midline [52, 192]. We therefore assume that Cluster-XIV cells represent cells of the VS, a suggestion that is backed-up by in-situ hybridization that confirmed that top markers of this cluster (Tnc1, Tnc2, and sideVIII) indeed are expressed in either the complete VS, or, like netrin-1, in transverse stripes spanning the VS (Fig. 6G-L and Supplementary Figs. 3537). Another high-ranked marker of this cluster, vitamin K-dependent protein C (vitK-C) (LOC107451660) has recently been studied by Leite et al. [111] showing that it is exclusively expressed in the ventral midline region prior to ventral splitting,data on developmental stages that possess the split ventral midline and thus the VS are unfortunately not shown in their paper.

Cluster-XV: Ectoderm patterning of developing appendages (5)

Top markers of Cluster-XV are the uncharacterized gene unc4096 (LOC107444096), clotting factor G subunit (cfGbs) (LOC107446841), the transcription factor AP2.2 (LOC107443623) and the homeobox containing gene unc-4 like (LOC107444630) (Drosophila DPHD-1). Information about the two top markers unc4096 and cfGbs is not available. AP2, however, has been investigated in some detail. In Drosophila, this gene is expressed inter alia in the leg and antennal discs [12, 99]. A previous study in another spider, Cupiennius salei, has shown that AP2 is expressed in the appendages and functions during leg development [150]. The unc-4 like gene has been studied in Drosophila where it is expressed in the developing nervous system, in a segment-polarity like pattern in the epidermis, and in the eye-antennal disc [179].

A closer look at the Cluster-XV markers also reveals a large number of genes that are known factors of arthropod (and indeed spider) appendage development, such as clawless/C15 (LOC107451627) Zhang [201] 2016, dally (LOC107446074) [72], distal-less (LOC107450100) [1, 145, 164], dachsous (LOC107449611) [124], the Hox gene Deformed (LOC107444120) [166], unc5.1 (LOC107445619) [88], optomotor-blind (LOC107450980) [81], SP6/9 (LOC107448645) [102] and Fz4.2 (LOC107441380) [82]. In-situ hybridization of the top markers in this cluster revealed that these genes all are expressed strongly in the ectoderm of the development appendages, many of them in the form of rings, indicating a role in proximal–distal axis patterning and joint formation (Fig. 9U-f and Supplementary Figs. 3841).

Cluster-XVI: The peripheral nervous system (PNS)

Two top markers of this cluster represent ohnologs of the aryl hydrocarbon receptor spineless (ss) (LOC107454134, LOC107457395). In Drosophila, ss causes the transformation of the distal antenna into a distal leg [24, 44], but beyond that ss also is involved in sensory neuron development [40, 100, 116]. Both functions appear to be conserved in other insects [182]. Other top markers of this cluster are a VEGF receptor like gene (LOC107454058), eagle (LOC107450205), sevenless (sev) (LOC107441450) and Pax2 (LOC107444558). In Drosophila, VEGF receptor-like is involved in mechanical nociception and normal axon branching, is expressed in sensory neurons, and overexpression leads to mechanical hypersensitivity [108, 120]. eagle represents a steroid receptor that is inter alia involved in nervous system development [69, 121]. Drosophila sev is a well-known factor of photoreceptor specification (reviewed in e.g. [183]. Finally, Pax2 has been shown to be expressed in the developing peripheral nervous system of Parasteatoda [79]. In summary, this suggests that cells of Cluster-XVI represent the developing peripheral nervous system. In-situ hybridization experiments of the aforementioned genes supports this suggestion as they are all expressed in distinct cells or groups of cells in the periphery of the developing embryo such as the appendages that display a large number of sensory structures (e.g. [11, 55, 79, 178] (Fig. 10A-L and Supplementary Figs. 4245). This is further supported by two genes that have been investigated as top markers of other clusters, unc-4 like (LOC107444630) (Fig. 9c-f, cf. Cluster-XV) and the uncharacterized gene unc6289 (LOC107436289) (Supplementary Fig. 23, cf. Cluster-VIII). Both are expressed in similar patterns in the appendages suggesting a function during peripheral/sensory nervous system development.

Fig. 10
figure 10

Gene expression of Cluster-XVI, -XIX, and -XXIV markers. In all panels, anterior is to the left. Panels A-M, P, R-U, and W represent ventral views. Panel N represents dorsal view. Panel O represents anterior view. Panel Q represents lateral view. Panels V and X show isolated walking legs (left), pedipalps (middle) and chelicerae (right), lateral views. Schematic drawings in the right column summarize the main gene expression characteristic for a given cell cluster (blue tissue). Arrows, arrowheads and asterisks in schematic drawings point to comparable expression as marked by the same symbol in the original photographs. Black arrowheads in panel O point to expression in the developing eyes. Abbreviations of gene names is indicated in the upper left corner. Developmental stages are indicated in the bottom right corner. For further information, see Supplementary Figs. 4245, 52, 53, and 6769, and Supplementary File 6. For abbreviations used in schematic drawings see legend of Fig. 4

Cluster-XVII: The developing central nervous system (CNS) (8) – Early differentiating neurons

The top markers of Cluster-XVII are inscuteable (insc) (LOC107457283 & LOC122271737), nervous fingers (nerfin) (LOC107446363), prospero (pros) (LOC107448306), and (less high-ranked) brain tumor (brat) (LOC107438139). insc is a neural precursor gene in Drosophila that is involved in defining neuroblast identity [5, 104]. Beyond that, insc is involved in controlling asymmetric cell division during nervous system development where it acts via miranda (mira) upstream of pros [76, 169]. Interestingly, pros, another confirmed factor of spider and arthropod nervous system development [23, 101, 185, 193], is one of the highest-ranking markers of Cluster-XVII as well. In Drosophila, pros functions as a factor of early neural differentiation by repressing neural stem cell markers such as snail (sna) (e.g. [200]. brat is involved in the regulation and distinction of intermediate progenitor cells from mitotically active neuroblasts, again interacting with the scaffolding protein mira [18]. Finally, nerfin prevents reversion and dedifferentiation of neurons into neural stem cells [53, 196]. In-situ hybridization of insc, nerfin, pros (in Cupiennius salei), and brat reveals relatively early expression in the developing CNS suggesting that the cells represented by Cluster-XVII are early differentiating neurons (Fig. 5Y-f and Supplementary Figs. 4648).

Cluster-XVIII: Ectoderm patterning of developing appendages (6) – EMT-like processes in the appendage epithelium

Among the top markers are papilin-like (LOC122270011), a paralog of the aforementioned papilin (LOC107447945) (cf. Cluster-VII), integrin alpha-PS1 (PS1) (LOC107454271) (cf. Cluster-V), and nord (LOC107451942).

papilin-like is an extracellular matrix protein encoding gene that is a likely factor of basement membrane development and cell movement (reviewed in [48]. Comparing the expression patterns of papilin-like and papilin in Parasteatoda (cf. Cluster-VII) reveals only little similarity (Figs. 6M-O and 8D-F). This suggests that these two paralogs have undergone neo- or sub-functionalization, and thus represent markers of different but related cell clusters (clusters XVIII and VII respectively). Unlike PS2 that is expressed in the visceral mesoderm (discussed above,cf. Cluster-VII), PS1 is expressed in epithelial cells and mediates cell migration [189]. Notably, a number of laminin-subunit encoding genes (LOC107448892, LOC107452433), the putative ligands of the integrins [67, 127], and a collagen gene (LOC107442289), an interaction partner of papillin [97], are also among the strongest markers of Cluster-XVIII, and so is the extracellular matrix (ECM) protein encoding gene nord (LOC107451942) [198]. The genetic fingerprint of Cluster-XVIII is thus very similar to that of Cluster-VII, suggesting that its cells are actively involved in cell migration and thus likely in EMT-related processes.

In-situ hybridization of Cluster-XVIII marker genes, however, is not conclusive. All investigated top markers appear to have quite complex and in many aspects different expression patterns, and also the markers of Cluster-XVIII that have been studied previously in Parasteatoda such as Wnt16 (LOC107457243) [87], dally (LOC107446074) [72], FoxB (LOC107443349) [65], and hex1 (cf. Cluster-I) show a variety of different expression pattern (Fig. 6M-R and Supplementary Figs. 4951). What these genes have in common, however, is expression in the developing appendages. It is thus likely that Cluster-XVIII represents cells that undergo an EMT-like process in the developing appendages.

Cluster-XIX: The developing central nervous system (CNS) (9) – The sensory nervous system of the head

Many of the top markers of Cluster-XIX have previously been studied in Parasteatoda. Among those are irx-4 (LOC107456088) [110], six3 (LOC107436457) [111, 157], orthodenticle (otd/otx) (LOC107457564) [163] and eyes-absent (eya) (LOC107452693) [163]. Although these genes are predominantly expressed in the developing nervous system of the head, including the developing eyes, they are also expressed (albeit at a lower level) in the developing CNS of the trunk [110, 157, 163]. Expression data of irx-4 in later developmental stages are not shown in Leite et al. [110], but the putative homologs of irx-4 in Drosophila, araucan (ara) and caupolican (caup) both are expressed in the sensory nervous system of the head including the developing eyes (e.g. [56, 147]. The second highest-ranked marker of this cluster is a scavenger receptor class B member 1-like gene (LOC122271437) that shows weak sequence similarity with Drosophila croquemort and Sensory neuron membrane protein 1 (Snmp-1), a gene that is involved in the function of sensory pheromone receptors [91]. We investigated the embryonic expression of the uncharacterized gene unc1341 (LOC122271341) and the homeobox gene Nkx6.2 (LOC107450777) (cf. Cluster-VI) and show that both genes are expressed in the CNS including part of the most anterior region of the CNS (Fig. 10M-T and Supplementary Figs. 52 and 53). Notably, within the developing brain, unc1341 is specifically expressed in the developing eyes (Fig. 10M-P). Cells representing Cluster-XIX thus clearly contribute to the developing CNS, possibly representing developing sensory structures in the head.

Cluster-XX: The developing heart

The top marker of this cluster is a ryanodine receptor (ryr) (LOC107450496) encoding gene that is typically found in muscle tissue (e.g. [63, 180]. The second-best marker is a lethal 2 like gene (LOC107453784) and the third-best marker is rho-associated protein kinase 2 (rapk2) (LOC107453369). Other high-ranked markers are two titin/twitchin/bent ohnologs (LOC107453137, LOC107453633), filamin (LOC107438849), myosin heavy chain (LOC107457063), myocardin-related transcription factor (LOC107443971) and kon-tiki (kon) (LOC107436488), all of which are involved in muscle and heart development (e.g. [105, 123, 133, 161, 202]. A lower-ranked marker, myocyte enhancer factor 2 (Mef2) (LOC107445920) is also expressed in the developing heart of Parasteatoda [111] and another spider [80]. In-silico analysis thus clearly suggests that Cluster-XX cells represent developing muscle tissue, including the developing heart. We show by in-situ hybridization analysis that all three previously unstudied top markers also are expressed almost exclusively in the heart, strongly suggesting that Cluster-XX cells indeed represent the developing heart (Fig. 8G-L and Supplementary Figs. 5456).

Cluster-XXI: Midgut development, yolk metabolism, hematopoiesis, and immune response

Among the top markers of Cluster-XXI are few previously investigated genes. The top markers are an aquaporin-7-like (aqp7l) gene (LOC107457171), the uncharacterized gene unc8180 (LOC107448180), and a Na+/K+ transporting ATPase subunit (NaK-t-ATPase) (LOC107456567). The majority of markers appear to represent enzymes (e.g. alkaline phosphatase (LOC107438694) and snake venom 5’-nucleotidase (LOC107439410)), transporters (e.g. the aforementioned NaK-t-ATPase and organic cation transporter protein (LOC107457207)), and channel-forming proteins (e.g. the aforementioned aquaporin gene and apolipophorin (LOC107450568)), suggesting that these cells are involved in yolk metabolism and uptake. Another group of Cluster-XXI marker genes is involved in the vertebrate immune response (e.g. the aforementioned snake venom 5’-nucleotidase, equilibrative nucleoside transporter 1 (LOC107445032) and venom phosphodiesterase 2 (LOC107442458)) suggesting that this may be another function of Cluster-XXI cells (e.g. [74, 95, 159]. Indeed, in arthropods, the so-called extraembryonic tissues (sensu lato) are involved in providing immune response (recently reviewed in e.g. [143, 184]. In-situ hybridization shows that the three top markers of this cluster all exclusively are expressed in cells underlying the germ band (and possibly also the dorsal field and ventral sulcus), and thus the cells/tissue that connects the developing embryo (sensu stricto) with the yolk (Fig. 8M-R and Supplementary Fig. 5759). Interestingly, we also find a midgut marker of Parasteatoda in this cluster, the GATA transcription factor serpent (srp) (LOC107456523), supporting that these cells may also contribute to midgut development that appears to go hand in hand with nutrition uptake in spiders [46]. Beyond that, hematopoiesis, a process that is closely linked with innate immune defence, depends on the interplay of the aforementioned GATA transcription factor srp with the ‘friend of GATA’ factor U-shaped (Ush) (e.g. [49, 50, 152], another unique marker of this cell cluster (LOC107440842).

Cluster-XXII: The stomodaeum and the ventral midline

The top gene markers of this cluster are a number of uncharacterized genes (e.g. unc 5848 (LOC107445848), unc3843 (LOC107453843), unc5174 (LOC107455174) and atrial natriuretic peptide-converting enzyme (anapece) (LOC107455346). The only previously investigated spider genes that represent markers of Cluster-XXII are the two forkhead domain transcription factors FoxC/crocodile (croc) (LOC107456536) and FoxA/forkhead (fkh) (LOC107452746), six3 (LOC107450741), and visual system homeobox (vsx) (LOC122268388). In Parasteatoda and other spiders, these genes are expressed in the stomodaeum and along the split ventral midline [84, 89, 111, 163]. Expression of unc5848, unc3843, and anapece, however, shows that these top markers are expressed exclusively in the stomodaeum, or strongly in the stomodaeum and faintly at the edges of the split ventral midline (Fig. 6S-V and Supplementary Figs. 6063). It appears thus that Cluster-XXII predominantly represents cells of the stomodaeum and possibly also cells along the ventral midline.

Cluster-XXIII: Midgut development and yolk metabolism

Cluster-XXIII forms a well-separated small group of cells. The best marker of this cluster is the uncharacterized gene unc7981 (LOC107437981). The second-best marker is the Na+/H+-exchanger beta (nhe2) (LOC107452408). This gene is expressed and functions in the epithelium of the mammalian gut (e.g. [33, 42]. To the best of our knowledge, there are no data on the expression or function of this gene in any arthropod species. Other markers of this cluster are the Na-dependent phosphate transporter 1A (PiT1-like) (LOC107453365) and the multidrug resistance-associated protein 1 (mrp1) (LOC107440407). The former gene is involved in the uptake of inorganic phosphate, displays kidney functions, and is often expressed in the intestine (e.g. [68, 131]. Mrp1 represents a universal transporter, including the transport of lipid derivatives (reviewed in [31, 32]. In-silico analysis thus suggests that these cells are involved in metabolic processes, possibly including the uptake of nutrition (yolk metabolism), and the development of the midgut, similar to the predicted functions of Cluster-XXI cells. Detection of gene expression supports this assumption as all of these genes are expressed exclusively in cells of (or underneath) the dorsal field, underneath the embryo proper, and in the tail region that appears to be a key-connective tissue between the yolk and the embryo proper (Fig. 8S-X and Supplementary Fig. 6466) (cf. Cluster-XXI). The fact that Cluster-XXI and -XXIII cells are separated into two clusters suggest that they represent different cell populations of this tissue. In this context, it is interesting to note that one endodermal midgut gene, serpent (srp) (LOC107456523), is expressed in Cluster-XXI cells, but another midgut marker, hepatocyte nuclear factor 4 (hnf4) (LOC107439273) is expressed in Cluster-XXIII cells [46].

Cluster-XXIV: A mini-cluster consisting of only 5 cells that suggests a function in leg-innervation

Some of the top markers of this cluster are RYamide receptor-like (LOC107443440), elongation of very long chain fatty acids protein 4-like (elovl4-l) (LOC122269995), nose resistant to fluoxetine protein 6-like (nrf6-l) (LOC107448146) and prostaglandin reductase-3 (ptgr3) (LOC107457101).

In-silico analysis of these genes is difficult because there is little or no information about the function of these genes in Drosophila and other arthropods. Only the RYamide receptor-like gene hints to a function in nervous system development. RYamide peptides have been identified in at least arthropods and tardigrades [26, 103], and represent the orthologs of lophotrochozoan Luqin genes [90]. RYamides/Luqins are involved in a wide range of physiological function, including feeding behaviour and locomotion (e.g. [115]. We failed to PCR-amplify RYamide receptor-like, but the other aforementioned markers all are expressed exclusively inside the developing legs, and with lower intensity the developing pedipalps (Fig. 10U-X and Supplementary Figs. 6769), suggesting a function in locomotion, possibly innervation of these appendages.

Mega-clusters and sequencing depth

Overall, the achieved SCS data appear to be of acceptable quality, i.e. most of the predicted cell clusters indeed represent specific tissues of the developing spider embryos such as the developing heart, the ventral sulcus or the peripheral nervous system (PNS). These are also the clusters that are most clearly separated from other clusters, and thus are most well-defined. Other clusters appear to be sub-clusters of larger clusters, often part of what we would like to call mega-clusters (Mega-Cluster-A (Ectoderm patterning of developing appendages; clusters I, II, IV, XIII, XV, and XVIII), Mega-Cluster-B (The developing central nervous system (CNS); clusters III, V, VI, VIII, X, XI, XII, XIV, XVI, XVII, XIX, and XXIV), and Mega-Cluster-C (Epithelial-to-mesenchymal transition (EMT)); clusters V, VII, XIV, and XVIII); note that some clusters such as Cluster-XVIII appear to be part of two mega clusters (summarized in Table 2). Naturally, the underlying data that subdivide those closely-related and physically connected (in the UMAP) clusters is less strong (i.e. there are less many markers that are specifically expressed in cells of each given cluster, and that are expressed in more vs less cells of a given cluster). The mega-clusters could represent the actual situation in developing embryos in terms that many related cell types or developmental stages of the same cell type indeed express very similar sets of marker genes. Since our SCS data stem from embryos and thus developing tissues and differentiating cell types, this assumption is not unlikely.

Table 2 Cluster overview

Alternatively, however, mega-clusters may represent “artefacts” as a result of the shortcomings of our analysis. These cells may indeed be much more different from each another, and thus the corresponding clusters could be much more defined than shown in our analysis. A possible reason for such a scenario could be the relatively low sequencing saturation in our analysis (ca. 11%). Low sequencing depth could have let to the lack of detection of more specific, but less numerous transcripts.

A bias towards ectodermal tissues?

We recognized that the majority of all detected tissues, and thus dissociated and captured cells represent ectodermal derivatives, while less mesodermal, and few endodermal cells were detected in our analysis. This could either represent the natural correlation of ectodermal vs mesodermal vs endodermal cells in the developing spider embryo at the investigated stages, or it could be a bias introduced by our methodology, i.e. dissociation, cell capture and downstream analysis. Additionally, we likely removed a fraction of the cells that cover the yolk outside the embryo proper (i.e. cells of the ventral sulcus and the dorsal field) prior to dissociation and cell capture.

We assume, however, that ectodermal cells represent the majority of cells in embryos of the stages 10–12, followed by mesodermal cells and finally endodermal cells. We base this assumption on the fact that most tissues indeed are ectodermal, followed by mesodermal tissues/cells. This has also been shown previously by the expression profiles of typical ectodermal, mesodermal and endodermal markers (e.g. [46, 87, 88, 110, 111]. The study by Feitosa et al. [46], for example, has shown that only relatively few cells of the so called dorsal extraembryonic tissue (the dorsal field in this study) likely contribute to the later (endodermal) midgut [46]. Therefore, we believe that the distribution of captured ectodermal, mesodermal and endodermal cells likely represents the true distribution of these cell types in developing spider embryos of stage 10–12. The recent SCS-study by Leite et al. [111] on earlier stages of Parasteatoda further support this assumption.

Whole genome duplication in the spider: How reliable is “in-silico” analysis?

One part of the analysis of cell markers is what we call “in-silico” analysis, the comparison of known data from other arthropods, usually the vinegar fly Drosophila melanogaster. One potential problem of this kind of comparative analysis, however, is the fact that many genes are duplicated in Parasteatoda (e.g. [87, 110, 166], which likely represents the result of a whole genome duplication (WGD) in this species that dates back to a WGD in the last common ancestor of Arachnopulmonata (e.g. [137, 166]. In-silico analysis therefore often compares one (of two) paralogs/ohnologs of a given marker gene of a particular cell cluster in Parasteatoda with the one ortholog in Drosophila (or other previously investigated organisms). Only in a few cases, both Parasteatoda ohnologs of a given gene (if present as two ohnologs) represent markers of the same cell cluster and are ranked similarly as markers of this cluster. Examples are the two spineless ohnologs in Cluster-XVI, or the two delta ohnologs in Cluster-VIII.

For all Parasteatoda genes, however, that are present in the form of two ohnologs and that are not both markers of a given cell cluster, this may cause a problem. We know that many ohnologs have undergone sub-functionalization and neo-functionalization. It is thus possible that a marker, if this particular Parasteatoda ohnolog has been the subject of neo-functionalization, may have a completely different function than its ortholog in other organisms. Likewise, spatial sub-functionalization (expression in different tissues) and temporal sub-functionalization (expression in the same tissue/cell type at different time points during development) may lead to misinterpretation of the function of a given marker gene. This problem can only be addressed beyond doubt by comprehensive WISH analysis of every considered marker (as partially done in this study).

Cell-type vs developmental stage of a cell

Interpretation of the obtained data must consider that an identified cell cluster may either represent a certain cell type (as defined by a cell-type specific genetic fingerprint), or a developmental stage of a certain cell type that therefore may also be represented by other, typically accompanying cell cluster(s). Possible examples are represented by the mega-clusters. Some clusters of Mega-Cluster B (e.g. clusters VI, VIII, X and XI) all harbour cells of the early developing CNS, representing either different (differentiating) cell types, different developmental stages of the same general cell type (that at all times during their development are represented by a more or less stable and specific genetic fingerprint), or cells that represent the same future cell type and/or developmental stage (of that cell type), but show variation (heterogeneity) in their expression profile (e.g. [29]. Indeed, suggestions have been made how to overcome this apparent problem [8], but we currently do not have the tools to apply them to our current data set. Thus, although we can make predictions on this matter, we do not believe that it is currently possible to distinguish the one from the other beyond considerable doubt.

Our data compared to previously published SCS-projects in Parasteatoda tepidariorum

Two previous studies on SCS in Parasteatoda embryos covered the developmental stages 5 [2] and stages 7–9 [111]. At stage 5, a germ disc and a cumulus have formed. The cumulus represents later endomesodermal cells and the disc represents the ectodermal tissues of the developing germ band. Additional endodermal and mesodermal tissue come from the rim of the disc (e.g. [138, 197]. The centre of the disc represents the later posterior of the embryo and its segment addition zone (SAZ), and the periphery of the disc represents the anterior of the later germ band (e.g. [2]. Cells from outside the disc represent the yolk and part of the later-forming dorsal field (DF). At this point, the embryo consists of approximately 2000 cells representing mainly undifferentiated cells of the three germ layers [2]. Already at this stage of development, the genetic fingerprint of endodermal cells appears to be quite different from the ectoderm and mesoderm, the latter two which form clusters that are in close proximity to each another (i.e. having related genetic fingerprints) [2]. Interestingly, the study by Akiyama-Oda and colleagues presents two endodermal cell clusters, one single mesodermal cell cluster, and a much larger number of ectodermal cell clusters that represent different regions along the centre-to-periphery-axis of the disc [2].

The study by Leite et al. [111] targets the stages 7–9 each represented by separated dissociation and SCS experiments that were then combined into a single SCS-atlas. Given the more advanced developmental stages that are addressed in this study compared to the study by Akiyama-Oda et al. [2], there are considerably more and more distinct cell clusters [111]. Like in our study, Leite et al. [111] also identified 23 distinct cell clusters, but the nature of these clusters is in many cases quite different from the 23 cell clusters identified in our study. Therefore, it is not possible to directly compare the clusters recovered in Leite et al. [111] with the clusters that we identified in our study. Indeed, there are obvious explanations for these differences. One main difference between the two studies is the higher number of CNS-representing clusters in our analysis which may easiest be explained by the more differentiated (and differentiating) state of the central nervous system in stage 10–12 vs stage 7–9 embryos, where in the latter, development and differentiation of the CNS may only just have begun [111]. Some clusters identified by Leite et al. [111] like the segment addition zone (SAZ) and the segment maturation zone (SMZ) have not been identified in our study, most easily explained by the lack or miniaturization of such tissues in later stage embryos (e.g. [130]. On the other hand, Leite et al. [111] did not detect cells specific for the ventral sulcus (VS), simply because this structure has not yet formed in the developmental stages studied by them. Another eye-striking difference in the outcome of the previous studies compared with our study is the apparent lack of EMT-related cell types in their studies. This, so we believe, can also be explained by the stages of the sequenced embryos because main morphogenetic events may take place at later developmental stages and after the development of the complete AP axis, rather than early during the processes of axis formation, germ band formation, and segment addition. Finally, Leite et al. [111] also identify a cluster representing putative stem cells, but we do not find a corresponding cell population in our study, most likely because the number of stem cells is low in the later and thus more differentiated developmental stages that we investigated. The paper by Leite et al. [111] uses Hox genes as regional markers and defines some of the recovered clusters according to the expression/distribution of Hox genes. In this way, they for example discriminate between the first two pairs of legs (L1 and L2) and the remaining legs, L3 and L4. Likewise they identify a cluster representing the pedipalp-bearing segment (or merely the developing pedipalp). We did not do this in our analysis because we believe that regional markers are potentially misguiding. It is unlikely, for example, that the cells that build L1 and L2 are fundamentally different than those involved in the development of L3 and L4, especially because these segments and their appendages all form from the early germ disc, and do not represent segments/appendages that are partially from the germ disc, and partially added sequentially from the SAZ. Using regional markers such as labial (lab) to discriminate the pedipalp-bearing segments from the leg-bearing segments, however, makes sense if the genes involved in pedipalp development indeed differ somewhat from those in leg development.

Mega-Cluster-A: Ectoderm patterning of developing appendages

According to our data and subsequent analysis, six clusters clearly are associated with the ectoderm of developing appendages (i.e. clusters I, II, IV, IX, XIII, and XV). Although both in-silico analysis of known genes representing markers of these clusters and accompanying in-situ hybridization analysis show that these cells come from the ectoderm of developing appendages, characterization of each cluster turned out to be rather complicated. Firstly, the detected patterns in the appendages are complex and dynamic making it difficult to find common patterns of cells of a given cluster. Secondly, even top markers of certain clusters such as Cluster-XIII showed to be expressed in several tissues further complicating analysis. It is obviously necessary to further investigate these cell clusters, either by applying an improved SCS data set, or by more detailed in-situ hybridization studies either comparing marker gene expression in the appendages of exactly staged embryos, or applying double (or multi) staining. We assume that such data could reveal the interplay(s) of genes involved in appendage patterning.

An interesting finding of our study is the presence of multiple genes that are expressed like genes that are involved in joint formation, i.e. in neatly spaced rings along the proximal–distal (PD) axis of the appendages. The number of cell clusters, and the accompanying large number of markers, that are involved in PD-appendage patterning and (likely) also joint formation is remarkably high, possibly reflecting the complexity of the crucial process of leg-patterning and joint development. Our data thus provide further insight and the possibility to study multiple new genes that likely are involved in these processes.

Another interesting outcome of our study is the finding that Cluster-IX cells are apparently specific for the heavily modified opisthosomal appendages of spiders, the breathing organs (tracheae and book lungs) and the possibly most characteristic feature of spiders, the spinnerets. This may allow a comparison between the genetic networks involved in the developing of the more basal locomotory prosomal legs and the derived appendages of the opisthosoma.

Mega-Cluster-B: The developing CNS

Several of the identified cell clusters represent cells of the developing CNS. Interestingly, the genetic fingerprint of these clusters also enables us to predict their specific identity beyond their mere involvement in nervous system development. We identify proneural genes predominantly in Cluster-XI, suggesting that these cells represent neural progenitors. The adjacent clusters VI and X both likely represent early differentiating neural cells based on the presence of the neurogenic delta genes, the proneural gene ash that in Drosophila is involved in neuroblast segregation (Cluster-X), and a cassette of nervous system cell cycle controlling factors that also block the action of the proneural gene SoxN found in Cluster-XI (neural progenitors). Adjacent to clusters VI and X lies Cluster-VIII which possibly represents early differentiating neuronal cells including neurons. Although the cluster is marked by a number of the aforementioned markers of clusters VI and X it also expresses markers such as cph that is involved in the early differentiation of neural cell types, and cpo that inter alia is involved in neuronal precursor development. Cells of Cluster-VIII thus likely represent a later step in nervous system development and differentiation than cells of clusters VI, X and XI. Closely connected to Cluster-VIII lies Cluster-XVII which, based on its genetic fingerprint, appears to represent early differentiating neurons: insc is involved in neuronal progenitor determination, pros inhibits neural stem cell development and initiates neuronal differentiation, brat defines intermediate neuronal progenitors from primary neuronal progenitors, and nerfin blocks reversion and dedifferentiation of neurons into neural precursors. In the periphery of the CNS mega-cluster we find two closely-connected clusters, Cluster-III and Cluster-XII, that very likely represent developing and/or mature neurons because they both express genes that are key-factors of neuron development, neuron differentiation, and neuronal pathfinding. Compared to Cluster-XII, Cluster-III expresses more neuron-specific markers such as the Dscams. It may thus be that Cluster-XII indeed represents mostly late developing neurons, and Cluster-III mostly mature neurons.

Remarkably, we can follow CNS development from neural progenitor determination (Cluster-XI), over the early differentiation of these cells (clusters VI, X and VIII) towards the determination of neuronal precursors and early developing neurons (Cluster-XVII) up to the late stages of neuron development and finally mature neurons (clusters III and XII).

A second branch of neuronal cells is represented by clusters XVI, XIX, XXIV, and possibly even XXII. Closest to clusters VI and X, the early differentiating neural cells, lies Cluster-XIX which we believe represents mainly developing sensory nerve cells of the head. Close by lies the mini-cluster Cluster-XXIV that only contains 5 cells. These cells, however, appear to be part of the appendage-innervation system. Cluster-XVI is represented by cells that express typical markers of the peripheral nervous system such as Pax2, sev and eagle. Finally, Cluster-XXII that represents cells of the developing stomodaeum (and ventral midline?) may also represent a neuronal cell type. While the developmental trajectory of Cluster-XI to Cluster-III appears to represent the development from neuronal precursors towards mature neurons, the possible trajectory of Cluster-XI to Cluster-XXII appears less straightforward and thus may need further investigation, probably with the help of an improved data set.

A third short trajectory of CNS-related clusters appears to be the connection of Cluster-XI to its neighbouring Cluster-V. Both clusters express sna, a gene that in Drosophila regulates the delamination of neuroblasts (neuronal precursors) and that is thus involved in EMT-like processes during neurogenesis [7]. Other genes of Cluster-V indeed are involved in EMT and the expression patterns of many markers of Cluster-V shows that they are not only expressed in the CNS but also various other tissues, likely as a result of the delamination and morphogenic movements of these cells. The trajectory from Cluster-XI to Cluster-V and the involvement of EMT-like processes in CNS development is thus closely related to other cell types such as mesodermal cells (Cluster-VII) or the ectoderm of the developing appendages (Cluster-XVIII) that share an EMT-related genetic fingerprint (discussed below).

Mega-Cluster-C: Epithelial-to-mesenchymal transitions (EMTs)

Several of the identified gene clusters are marked by an EMT-related genetic fingerprint. The main difference between these cells is the underlying second genetic fingerprint that is for example neuronal (clusters V and XIV) (discussed above), mesodermal (Cluster-VII), or appendage-ectoderm specific (Cluster-XVIII). We believe that the reason for detecting a large number of cells that show an EMT-related genetic fingerprint is likely correlated with the relatively late developmental stages we investigated in this study. This goes in line with the other Parasteatoda SCS studies that investigate earlier developmental stages and that did not specifically identify cells that are undergoing EMT [2, 111]. Indeed, processes like CNS differentiation, the formation of neuronal networks, and the formation of the mesoderm and mesodermal organs such as the heart most certainly require EMT and morphogenic movement.

Future perspectives

The recent SCS-based papers published on the development of Parasteatoda clearly highlight the interest in this model system and the interest in SCS data [2, 111], this study). In the future, the existing data including our own will, we hope, be supplemented and extended towards earlier as well as later developmental stages, including nymphs, juveniles and adults of both sexes, to develop a comprehensive overview over cell types and their developmental trajectories in this model spider. Detailed knowledge about the changes of the genetic fingerprints of developing cell types and the identification of definite cell types will significantly improve our knowledge about the development of spiders. This will include the identification of novel genes, and the untangling of new and conserved gene regulatory networks (GRNs). These data can then be used to study newly identified marker genes (or complete GRNs) in classic candidate gene approaches beyond spiders, possibly enabling us to reconstruct the development and evolution of Arthropoda as a whole. The current study aims to contribute to this goal.

Availability of data and materials

The data generated and analysed during this study are included in this published article and its supplementary information files. Raw sequencing reads in fastq.gz format are available from the corresponding authors on reasonable request.

References

  1. Abzhanov A, Kaufman TC. Homologs of Drosophila appendage genes in the patterning of arthropod limbs. Dev Biol. 2000;227(2):673–89.

    Article  CAS  PubMed  Google Scholar 

  2. Akiyama-Oda Y, Akaiwa T, Oda H. Reconstruction of the global polarity of an early spider embryo by single-cell and single-nucleus transcriptome analysis. Front Cell Dev Biol. 2022;10:933220.

    Article  PubMed  PubMed Central  Google Scholar 

  3. Aksu M, Seiradake E. DIPping into the Fly visual system. Neuron. 2018;100(6):1270–2.

    Article  CAS  PubMed  Google Scholar 

  4. Ahn Y, Zou J, Mitchell PJ. Segment-specific regulation of the Drosophila AP-2 gene during leg and antennal development. Dev Biol. 2011;355(2):336–48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. An H, Ge W, Xi Y, Yang X. Inscuteable maintains type I neuroblast lineage identity via numb/notch signaling in the Drosophila larval brain. J Genet Genomics. 2017;44(3):151–62.

    Article  CAS  PubMed  Google Scholar 

  6. Angelini DR, Kikuchi M, Jockusch EL. Genetic patterning in the adult capitate antenna of the beetle Tribolium castaneum. Dev Biol. 2009;327(1):240–51.

    Article  CAS  PubMed  Google Scholar 

  7. Arefin B, Parvin F, Bahrampour S, Stadler CB, Thor S. Drosophila neuroblast selection is gated by notch, snail, SoxB, and EMT gene interplay. Cell Rep. 2019;29(11):3636–3651.e3.

    Article  CAS  PubMed  Google Scholar 

  8. Arendt D, Bertucci PY, Achim K, Musser JM. Evolution of neuronal types and families. Curr Opin Neurobiol. 2019;56:144–52.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Ashraf SI, Ip YT. The snail protein family regulates neuroblast expression of inscuteable and string, genes involved in asymmetry and cell division in Drosophila. Development. 2001;128(23):4757–67.

    Article  CAS  PubMed  Google Scholar 

  10. Bätz T, Förster D, Luschnig S. The transmembrane protein macroglobulin complement-related is essential for septate junction formation and epithelial barrier function in Drosophila. Development. 2014;141(4):899–908.

    Article  PubMed  Google Scholar 

  11. Barth FG. Spider mechanoreceptors. Curr Opin Neurobiol. 2004;14(4):415–22.

    Article  CAS  PubMed  Google Scholar 

  12. Bauer R, McGuffin ME, Mattox W, Tainsky MA. Cloning and characterization of the Drosophila homologue of the AP-2 transcription factor. Oncogene. 1998;17(15):1911–22.

    Article  CAS  PubMed  Google Scholar 

  13. Baumgardt M, Miguel-Aliaga I, Karlsson D, Ekman H, Thor S. Specification of neuronal identities by feedforward combinatorial coding. PLoS Biol. 2007;5(2):e37.

    Article  PubMed  PubMed Central  Google Scholar 

  14. Barczyk M, Carracedo S, Gullberg D. Integrins. Cell Tissue Res. 2010;339(1):269–80.

    Article  CAS  PubMed  Google Scholar 

  15. Beckingham KM, Armstrong JD, Texada MJ, Munjaal R, Baker DA. Drosophila melanogaster–the model organism of choice for the complex biology of multi-cellular organisms. Gravit Space Biol Bull. 2005;18(2):17–29.

    PubMed  Google Scholar 

  16. Bellen HJ, Kooyer S, D’Evelyn D, Pearlman J. The Drosophila couch potato protein is expressed in nuclei of peripheral neuronal precursors and shows homology to RNA-binding proteins. Genes Dev. 1992;6(11):2125–36.

    Article  CAS  PubMed  Google Scholar 

  17. Bernt M, Donath A, Jühling F, Externbrink F, Florentz C, Fritzsch G, Pütz J, Middendorf M, Stadler PF. MITOS: improved de novo metazoan mitochondrial genome annotation. Mol Phylogenet Evol. 2013;69(2):313–9.

    Article  PubMed  Google Scholar 

  18. Betschinger J, Mechtler K, Knoblich JA. Asymmetric segregation of the tumor suppressor brat regulates self-renewal in Drosophila neural stem cells. Cell. 2006;124(6):1241–53.

    Article  CAS  PubMed  Google Scholar 

  19. Biersmith B, Liu ZC, Bauman K, Geisbrecht ER. The DOCK protein sponge binds to ELMO and functions in Drosophila embryonic CNS development. PLoS ONE. 2011;6(1):e16120.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  20. Bonatto Paese CL, Leite DJ, Schönauer A, McGregor AP, Russell S. Duplication and expression of sox genes in spiders. BMC Evol Biol. 2018;18(1):205.

    Article  PubMed  PubMed Central  Google Scholar 

  21. Boube M, Martin-Bermudo MD, Brown NH, Casanova J. Specific tracheal migration is mediated by complementary expression of cell surface proteins. Genes Dev. 2001;15(12):1554–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Bradley PL, Myat MM, Comeaux CA, Andrew DJ. Posterior migration of the salivary gland requires an intact visceral mesoderm and integrin function. Dev Biol. 2003;257(2):249–62.

    Article  CAS  PubMed  Google Scholar 

  23. Broadus J, Doe CQ. Evolution of neuroblast identity: seven-up and prospero expression reveal homologous and divergent neuroblast fates in Drosophila and Schistocerca. Development. 1995;121(12):3989–96.

    Article  CAS  PubMed  Google Scholar 

  24. Burgess EA, Duncan I. Direct control of antennal identity by the spineless-aristapedia gene of Drosophila. Mol Gen Genet. 1990;221(3):347–57.

    Article  CAS  PubMed  Google Scholar 

  25. Cabrera CV, Martinez-Arias A, Bate M. The expression of three members of the achaete-scute gene complex correlates with neuroblast segregation in Drosophila. Cell. 1987;50(3):425–33.

    Article  CAS  PubMed  Google Scholar 

  26. Caers J, Verlinden H, Zels S, Vandersmissen HP, Vuerinckx K, Schoofs L. More than two decades of research on insect neuropeptide GPCRs: an overview. Front Endocrinol (Lausanne). 2012;3:151.

    Article  CAS  PubMed  Google Scholar 

  27. Cai Y, Chia W, Yang X. A family of snail-related zinc finger proteins regulates two distinct and parallel mechanisms that mediate Drosophila neuroblast asymmetric divisions. EMBO J. 2001;20(7):1704–14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Campbell AG, Fessler LI, Salo T, Fessler JH. Papilin: a Drosophila proteoglycan-like sulfated glycoprotein from basement membranes. J Biol Chem. 1987;262(36):17605–12.

    Article  CAS  PubMed  Google Scholar 

  29. Cembrowski MS, Menon V. Continuous variation within cell types of the nervous system. Trends Neurosci. 2018;41(6):337–48.

    Article  CAS  PubMed  Google Scholar 

  30. Choudhary S, Satija R. Comparison and evaluation of statistical error models for scRNA-seq. Genome Biol. 2022;23(1):27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Cole SP. Targeting multidrug resistance protein 1 (MRP1, ABCC1): past, present, and future. Annu Rev Pharmacol Toxicol. 2014;54:95–117.

    Article  CAS  PubMed  Google Scholar 

  32. Cole SP. Multidrug resistance protein 1 (MRP1, ABCC1), a “multitasking” ATP-binding cassette (ABC) transporter. J Biol Chem. 2014;289(45):30880–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Collins JF, Honda T, Knobel S, Bulus NM, Conary J, DuBois R, Ghishan FK. Molecular cloning, sequencing, tissue distribution, and functional expression of a Na+/H+ exchanger (NHE-2). Proc Natl Acad Sci U S A. 1993;90(9):3938–42.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  34. Cosmanescu F, Katsamba PS, Sergeeva AP, Ahlsen G, Patel SD, Brewer JJ, Tan L, Xu S, Xiao Q, Nagarkar-Jaiswal S, Nern A, Bellen HJ, Zipursky SL, Honig B, Shapiro L. Neuron-subtype-specific expression, interaction affinities, and specificity determinants of DIP/DPR cell recognition proteins. Neuron. 2018;100(6):1385–1400.e6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Crozatier M, Vincent A. Control of multidendritic neuron differentiation in Drosophila: the role of Collier. Dev Biol. 2008;315(1):232–42.

    Article  CAS  PubMed  Google Scholar 

  36. Daiber T, VanderZwan-Butler CJ, Bashaw GJ, Evans TA. Conserved and divergent aspects of Robo receptor signaling and regulation between Drosophila Robo1 and C. elegans SAX-3. Genetics. 2021;217(3):iya018.

    Article  Google Scholar 

  37. David R, Wedlich D. PCR-based RNA probes: a quick and sensitive method to improve whole mount embryo in situ hybridizations. Biotechniques. 2001;30(4):769–72 (74).

    Article  CAS  PubMed  Google Scholar 

  38. Davis GK, Patel NH. Short, long, and beyond: molecular and embryological approaches to insect segmentation. Annu Rev Entomol. 2002;47:669–99.

    Article  CAS  PubMed  Google Scholar 

  39. de la Escalera S, Bockamp EO, Moya F, Piovant M, Jiménez F. Characterization and gene cloning of neurotactin, a Drosophila transmembrane protein related to cholinesterases. EMBO J. 1990;9(11):3593–601.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Deak II. Demonstration of sensory neurones in the ectopic cuticle of spineless-aristapedia, a homoeotic mutant of Drosophila. Nature. 1976;260(5548):252–4.

    Article  ADS  CAS  PubMed  Google Scholar 

  41. Demilly A, Simionato E, Ohayon D, Kerner P, Garcès A, Vervoort M. Coe genes are expressed in differentiating neurons in the central nervous system of protostomes. PLoS ONE. 2011;6(6):e21213.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  42. Dudeja PK, Rao DD, Syed I, Joshi V, Dahdal RY, Gardner C, Risk MC, Schmidt L, Bavishi D, Kim KE, Harig JM, Goldstein JL, Layden TJ, Ramaswamy K. Intestinal distribution of human Na+/H+ exchanger isoforms NHE-1, NHE-2, and NHE-3 mRNA. Am J Physiol. 1996;271(3 Pt 1):G483–93.

    CAS  PubMed  Google Scholar 

  43. Dumstrei K, Wang F, Shy D, Tepass U, Hartenstein V. Interaction between EGFR signaling and DE-cadherin during nervous system morphogenesis. Development. 2002;129(17):3983–94.

    Article  CAS  PubMed  Google Scholar 

  44. Duncan DM, Burgess EA, Duncan I. Control of distal antennal identity and tarsal development in Drosophila by spineless-aristapedia, a homolog of the mammalian dioxin receptor. Genes Dev. 1998;12(9):1290–303.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Emmons RB, Duncan D, Duncan I. Regulation of the Drosophila distal antennal determinant spineless. Dev Biol. 2007;302(2):412–26.

    Article  CAS  PubMed  Google Scholar 

  46. Feitosa NM, Pechmann M, Schwager EE, Tobias-Santos V, McGregor AP, Damen WGM, Nunes da Fonseca R. Molecular control of gut formation in the spider Parasteatoda tepidariorum. Genesis. 2017;55(5). https://0-doi-org.brum.beds.ac.uk/10.1002/dvg.23033.

  47. Feng R, Wen J. Overview of the roles of Sox2 in stem cell and development. Biol Chem. 2015;396(8):883–91.

    Article  CAS  PubMed  Google Scholar 

  48. Fessler JH, Kramerova I, Kramerov A, Chen Y, Fessler LI. Papilin, a novel component of basement membranes, in relation to ADAMTS metalloproteases and ECM development. Int J Biochem Cell Biol. 2004;36(6):1079–84.

    Article  CAS  PubMed  Google Scholar 

  49. Fossett N, Tevosian SG, Gajewski K, Zhang Q, Orkin SH, Schulz RA. The friend of GATA proteins U-shaped, FOG-1, and FOG-2 function as negative regulators of blood, heart, and eye development in Drosophila. Proc Natl Acad Sci U S A. 2001;98(13):7342–7.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  50. Fossett N, Hyman K, Gajewski K, Orkin SH, Schulz RA. Combinatorial interactions of serpent, lozenge, and U-shaped regulate crystal cell lineage commitment during drosophila hematopoiesis. Proc Natl Acad Sci U S A. 2003;100(20):11451–6.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  51. Fox PM, Tang JLY, Brand AH. The Drosophila homologue of CTIP1 (Bcl11a) and CTIP2 (Bcl11b) regulates neural stem cell temporal patterning. Development. 2022;149(17):dev200677.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Fraichard S, Bouge AL, Chauvel I, Bouhin H. Tenectin, a novel extracellular matrix protein expressed during Drosophila melanogaster embryonic development. Gene Expr Patterns. 2006;6(8):772–6.

    Article  CAS  PubMed  Google Scholar 

  53. Froldi F, Szuperak M, Weng CF, Shi W, Papenfuss AT, Cheng LY. The transcription factor Nerfin-1 prevents reversion of neurons into neural stem cells. Genes Dev. 2015;29(2):129–43.

    Article  PubMed  PubMed Central  Google Scholar 

  54. Glasscock E, Tanouye MA. Drosophila couch potato mutants exhibit complex neurological abnormalities including epilepsy phenotypes. Genetics. 2005;169(4):2137–49.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Gold K, Cotton JA, Stollewerk A. The role of notch signalling and numb function in mechanosensory organ formation in the spider Cupiennius salei. Dev Biol. 2009;327(1):121–31.

    Article  CAS  PubMed  Google Scholar 

  56. Gomez-Skarmeta JL, Diez del Corral R, De la Calle-Mustienes E, Ferré-Marcó D, Modolell J. Araucan and caupolican, two members of the novel iroquois complex, encode homeoproteins that control proneural and vein-forming genes. Cell. 1996;85(1):95–105.

    Article  CAS  PubMed  Google Scholar 

  57. Grenningloh G, Rehm EJ, Goodman CS. Genetic analysis of growth cone guidance in Drosophila: fasciclin II functions as a neuronal recognition molecule. Cell. 1991;67(1):45–57.

    Article  CAS  PubMed  Google Scholar 

  58. Grison A, Atanasoski S. Cyclins, cyclin-dependent kinases, and cyclin-dependent kinase inhibitors in the mouse nervous system. Mol Neurobiol. 2020;57(7):3206–18.

    Article  CAS  PubMed  Google Scholar 

  59. Grün D, van Oudenaarden A. Design and analysis of single-cell sequencing experiments. Cell. 2015;163(4):799–810.

    Article  PubMed  Google Scholar 

  60. Haenlin M, Kunisch M, Kramatschek B, Campos-Ortega JA. Genomic regions regulating early embryonic expression of the drosophila neurogenic gene Delta. Mech Dev. 1994;47(1):99–110.

    Article  CAS  PubMed  Google Scholar 

  61. Hafemeister C, Satija R. Normalization and variance stabilization of single-cell RNA-seq data using regularized negative binomial regression. Genome Biol. 2019;20(1):296.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Hahn M, Jäckle H. Drosophila goosecoid participates in neural development but not in body axis formation. EMBO J. 1996;15(12):3077–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Hasan G, Rosbash M. Drosophila homologs of two mammalian intracellular Ca(2+)-release channels: identification and expression patterns of the inositol 1,4,5-triphosphate and the ryanodine receptor genes. Development. 1992;116(4):967–75.

    Article  CAS  PubMed  Google Scholar 

  64. Heilig JS, Freeman M, Laverty T, Lee KJ, Campos AR, Rubin GM, Steller H. Isolation and characterization of the disconnected gene of Drosophila melanogaster. EMBO J. 1991;10(4):809–15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Heingård M, Turetzek N, Prpic NM, Janssen R. FoxB, a new and highly conserved key factor in arthropod dorsal-ventral (DV) limb patterning. EvoDevo. 2019;10:28.

    Article  PubMed  PubMed Central  Google Scholar 

  66. Hemmati-Brivanlou A, Kelly OG, Melton DA. Follistatin, an antagonist of activin, is expressed in the Spemann organizer and displays direct neuralizing activity. Cell. 1994;77(2):283–95.

    Article  CAS  PubMed  Google Scholar 

  67. Henchcliffe C, García-Alonso L, Tang J, Goodman CS. Genetic analysis of laminin a reveals diverse functions during morphogenesis in drosophila. Development. 1993;118(2):325–37.

    Article  CAS  PubMed  Google Scholar 

  68. Hernando N, Wagner CA. Mechanisms and regulation of intestinal phosphate absorption. Compr Physiol. 2018;8(3):1065–90.

    Article  PubMed  Google Scholar 

  69. Higashijima S, Shishido E, Matsuzaki M, Saigo K. eagle, a member of the steroid receptor gene superfamily, is expressed in a subset of neuroblasts and regulates the fate of their putative progeny in the Drosophila CNS. Development. 1996;122(2):527–36.

    Article  CAS  PubMed  Google Scholar 

  70. Hilbrant M, Damen WG, McGregor AP. Evolutionary crossroads in developmental biology: the spider Parasteatoda tepidariorum. Development. 2012;139(15):2655–62.

    Article  CAS  PubMed  Google Scholar 

  71. Hindley C, Philpott A. Co-ordination of cell cycle and differentiation in the developing nervous system. Biochem J. 2012;444(3):375–82.

    Article  CAS  PubMed  Google Scholar 

  72. Hogvall M, Budd GE, Janssen R. Gene expression analysis of potential morphogen signalling modifying factors in Panarthropoda. EvoDevo. 2018;9:20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Huang H, Wright S, Zhang J, Brekken RA. Getting a grip on adhesion: cadherin switching and collagen signaling. Biochim Biophys Acta Mol Cell Res. 2019;1866(11):118472.

    Article  CAS  PubMed  Google Scholar 

  74. Huang YS, Tseng WY, Clanchy FIL, Topping LM, Ogbechi J, McNamee K, Perocheau D, Chiang NY, Ericsson P, Sundstedt A, Xue ZT, Salford LG, Sjögren HO, Stone TW, Lin HH, Luo SF, Williams RO. Pharmacological modulation of T cell immunity results in long-term remission of autoimmune arthritis. Proc Natl Acad Sci U S A. 2021;118(19):e2100939118.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Hunter DD, Manglapus MK, Bachay G, Claudepierre T, Dolan MW, Gesuelli KA, Brunken WJ. CNS synapses are stabilized trans-synaptically by laminins and laminin-interacting proteins. J Comp Neurol. 2019;527(1):67–86.

    Article  CAS  PubMed  Google Scholar 

  76. Ikeshima-Kataoka H, Skeath JB, Nabeshima Y, Doe CQ, Matsuzaki F. Miranda directs prospero to a daughter cell during Drosophila asymmetric divisions. Nature. 1997;390(6660):625–9.

    Article  ADS  CAS  PubMed  Google Scholar 

  77. Iwasaki-Yokozawa S, Nanjo R, Akiyama-Oda Y, Oda H. Lineage-specific, fast-evolving GATA-like gene regulates zygotic gene activation to promote endoderm specification and pattern formation in the Theridiidae spider. BMC Biol. 2022;20(1):223.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Jakobsen JS, Braun M, Astorga J, Gustafson EH, Sandmann T, Karzynski M, et al. Temporal ChIP-on-chip reveals Biniou as a universal regulator of the visceral muscle transcriptional network. Genes Dev. 2007;21:2448–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Janeschik M, Schacht MI, Platten F, Turetzek N. It takes two: discovery of spider Pax2 duplicates indicates prominent role in chelicerate central nervous system, eye, as well as external sense organ precursor formation and diversification after neo-and subfunctionalization. Front Ecol Evol. 2022;10. https://0-doi-org.brum.beds.ac.uk/10.3389/fevo.2022.810077.

  80. Janssen R, Damen WG. Diverged and conserved aspects of heart formation in a spider. Evol Dev. 2008;10(2):155–65.

    Article  CAS  PubMed  Google Scholar 

  81. Janssen R, Feitosa NM, Damen WG, Prpic NM. The T-box genes H15 and optomotor-blind in the spiders Cupiennius salei, Tegenaria atrica and Achaearanea tepidariorum and the dorsoventral axis of arthropod appendages. Evol Dev. 2008;10(2):143–54.

    Article  CAS  PubMed  Google Scholar 

  82. Janssen R, Schönauer A, Weber M, Turetzek N, Hogvall M, Goss GE, Patel NH, McGregor AP, Hilbrant M. The evolution and expression of panarthropod frizzled genes. Front Ecol Evol. 2015;3:96.

    Article  Google Scholar 

  83. Janssen R. Gene expression reveals evidence for EGFR-dependent proximal-distal limb patterning in a myriapod. Evol Dev. 2017;19(3):124–35.

    Article  MathSciNet  CAS  PubMed  Google Scholar 

  84. Janssen R, Budd GE. Investigation of endoderm marker-genes during gastrulation and gut-development in the velvet worm Euperipatoides kanangrensis. Dev Biol. 2017;427(1):155–64.

    Article  CAS  PubMed  Google Scholar 

  85. Janssen R, Andersson E, Betnér E, Bijl S, Fowler W, Höök L, Leyhr J, Mannelqvist A, Panara V, Smith K, Tiemann S. Embryonic expression patterns and phylogenetic analysis of panarthropod sox genes: insight into nervous system development, segmentation and gonadogenesis. BMC Evol Biol. 2018;18(1):88.

    Article  PubMed  PubMed Central  Google Scholar 

  86. Janssen R, Budd GE. Oscillating waves of fox, cyclin and CDK gene expression indicate unique spatiotemporal control of cell cycling during nervous system development in onychophorans. Arthropod Struct Dev. 2021;62:101042.

    Article  PubMed  Google Scholar 

  87. Janssen R, Pechmann M, Turetzek N. A chelicerate Wnt gene expression atlas: novel insights into the complexity of arthropod Wnt-patterning. EvoDevo. 2021;12(1):12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Janssen R, Budd GE. Expression of netrin and its receptors uncoordinated-5 and frazzled in arthropods and onychophorans suggests conserved and diverged functions in neuronal pathfinding and synaptogenesis. Dev Dyn. 2022. https://0-doi-org.brum.beds.ac.uk/10.1002/dvdy.459.

    Article  PubMed  Google Scholar 

  89. Janssen R, Schomburg C, Prpic NM, Budd GE. A comprehensive study of arthropod and onychophoran Fox gene expression patterns. PLoS ONE. 2022;17(7):e0270790.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Jékely G. Global view of the evolution and diversity of metazoan neuropeptide signaling. Proc Natl Acad Sci U S A. 2013;110(21):8702–7.

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  91. Jin X, Ha TS, Smith DP. SNMP is a signaling component required for pheromone sensitivity in drosophila. Proc Natl Acad Sci U S A. 2008;105(31):10996–1001.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  92. Jontes JD. The cadherin superfamily in neural circuit assembly. Cold Spring Harb Perspect Biol. 2018;10(7):a029306.

    Article  PubMed  PubMed Central  Google Scholar 

  93. Jussen D, von Hilchen J, Urbach R. Genetic regulation and function of epidermal growth factor receptor signalling in patterning of the embryonic drosophila brain. Open Biol. 2016;6(12):160202.

    Article  PubMed  PubMed Central  Google Scholar 

  94. McNeill H, Craig GM, Bateman JM. Regulation of neurogenesis and epidermal growth factor receptor signaling by the insulin receptor/target of rapamycin pathway in Drosophila. Genetics. 2008;179(2):843–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Kabashima K, Nakashima C, Nonomura Y, Otsuka A, Cardamone C, Parente R, De Feo G, Triggiani M. Biomarkers for evaluation of mast cell and basophil activation. Immunol Rev. 2018;282(1):114–20.

    Article  CAS  PubMed  Google Scholar 

  96. Kantarci H, Gerberding A, Riley BB. Spemann organizer gene Goosecoid promotes delamination of neuroblasts from the otic vesicle. Proc Natl Acad Sci U S A. 2016;113(44):E6840–8.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  97. Kawano T, Zheng H, Merz DC, Kohara Y, Tamai KK, Nishiwaki K, Culotti JGC. elegans mig-6 encodes papilin isoforms that affect distinct aspects of DTC migration, and interacts genetically with mig-17 and collagen IV. Development. 2009;136(9):1433–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Kelly OG, Melton DA. Induction and patterning of the vertebrate nervous system. Trends Genet. 1995;11(7):273–8.

    Article  CAS  PubMed  Google Scholar 

  99. Kerber B, Monge I, Mueller M, Mitchell PJ, Cohen SM. The AP-2 transcription factor is required for joint formation and cell survival in Drosophila leg development. Development. 2001;128(8):1231–8.

    Article  CAS  PubMed  Google Scholar 

  100. Kim MD, Jan LY, Jan YN. The bHLH-PAS protein spineless is necessary for the diversification of dendrite morphology of drosophila dendritic arborization neurons. Genes Dev. 2006;20(20):2806–19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Klann M, Stollewerk A. Evolutionary variation in neural gene expression in the developing sense organs of the crustacean Daphnia magna. Dev Biol. 2017;424(1):50–61.

    Article  CAS  PubMed  Google Scholar 

  102. Königsmann T, Turetzek N, Pechmann M, Prpic NM. Expression and function of the zinc finger transcription factor Sp6-9 in the spider Parasteatoda tepidariorum. Dev Genes Evol. 2017;227(6):389–400.

    Article  PubMed  Google Scholar 

  103. Koziol U. Precursors of neuropeptides and peptide hormones in the genomes of tardigrades. Gen Comp Endocrinol. 2018;267:116–27.

    Article  CAS  PubMed  Google Scholar 

  104. Kraut R, Campos-Ortega JA. inscuteable, a neural precursor gene of drosophila, encodes a candidate for a cytoskeleton adaptor protein. Dev Biol. 1996;174(1):65–81.

    Article  CAS  PubMed  Google Scholar 

  105. Kuwahara K, Barrientos T, Pipes GC, Li S, Olson EN. Muscle-specific signaling mechanism that links actin dynamics to serum response factor. Mol Cell Biol. 2005;25(8):3173–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Lacin H, Zhu Y, Wilson BA, Skeath JB. dbx mediates neuronal specification and differentiation through cross-repressive, lineage-specific interactions with eve and hb9. Development. 2009;136(19):3257–66. https://0-doi-org.brum.beds.ac.uk/10.1242/dev.037242.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Laugier E, Yang Z, Fasano L, Kerridge S, Vola C. A critical role of teashirt for patterning the ventral epidermis is masked by ectopic expression of tiptop, a paralog of teashirt in Drosophila. Dev Biol. 2005;283(2):446–58.

    Article  CAS  PubMed  Google Scholar 

  108. Learte AR, Forero MG, Hidalgo A. Gliatrophic and gliatropic roles of PVF/PVR signaling during axon guidance. Glia. 2008;56(2):164–76.

    Article  CAS  PubMed  Google Scholar 

  109. Lee KJ, Freeman M, Steller H. Expression of the disconnected gene during development of Drosophila melanogaster. EMBO J. 1991;10(4):817–26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Leite DJ, Baudouin-Gonzalez L, Iwasaki-Yokozawa S, Lozano-Fernandez J, Turetzek N, Akiyama-Oda Y, Prpic NM, Pisani D, Oda H, Sharma PP, McGregor AP. Homeobox gene duplication and divergence in arachnids. Mol Biol Evol. 2018;35(9):2240–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Leite DJ, Schönauer A, Blakeley G, Harper A, Garcia-Castro H, Baudouin-Gonzalez L, Wang R, Sarkis N, Nikola AG, Koka VS, Kenny NJ. An atlas of spider development at single-cell resolution provides new insights into arthropod embryogenesis. bioRxiv. 2022.

  112. Li Y, Tower J. Adult-specific over-expression of the Drosophila genes magu and hebe increases life span and modulates late-age female fecundity. Mol Genet Genomics. 2009;281(2):147–62.

    Article  CAS  PubMed  Google Scholar 

  113. Li H, Collado M, Villasante A, Matheu A, Lynch CJ, Cañamero M, Rizzoti K, Carneiro C, Martínez G, Vidal A, Lovell-Badge R, Serrano M. p27(Kip1) directly represses Sox2 during embryonic stem cell differentiation. Cell Stem Cell. 2012;11(6):845–52.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Li H. Single-cell RNA sequencing in Drosophila: technologies and applications. Wiley Interdiscip Rev Dev Biol. 2021;10(5):e396.

    Article  CAS  PubMed  Google Scholar 

  115. Li C, Zheng Y, Cong X, Liu H, Storey KB, Chen M. Molecular and functional characterization of the luqin-type neuropeptide signaling system in the sea cucumber Apostichopus japonicus. Peptides. 2022;155:170839.

    Article  CAS  PubMed  Google Scholar 

  116. Lienhard MC, Stocker RF. The development of the sensory neuron pattern in the antennal disc of wild-type and mutant (lz3, ssa) Drosophila melanogaster. Development. 1991;112(4):1063–75.

    Article  CAS  PubMed  Google Scholar 

  117. Lim S, Kaldis P. Cdks, cyclins and CKIs: roles beyond cell cycle regulation. Development. 2013;140(15):3079–93.

    Article  CAS  PubMed  Google Scholar 

  118. Linne V, Stollewerk A. Conserved and novel functions for netrin in the formation of the axonal scaffold and glial sheath cells in spiders. Dev Biol. 2011;353(1):134–46.

    Article  CAS  PubMed  Google Scholar 

  119. Linne V, Eriksson BJ, Stollewerk A. Single-minded and the evolution of the ventral midline in arthropods. Dev Biol. 2012;364(1):66–76.

    Article  CAS  PubMed  Google Scholar 

  120. Lopez-Bellido R, Puig S, Huang PJ, Tsai CR, Turner HN, Galko MJ, Gutstein HB. Growth factor signaling regulates mechanical nociception in flies and vertebrates. J Neurosci. 2019;39(30):6012–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Lundell MJ, Hirsh J. eagle is required for the specification of serotonin neurons and other neuroblast 7–3 progeny in the Drosophila CNS. Development. 1998;125(3):463–72.

    Article  CAS  PubMed  Google Scholar 

  122. Mandal L, Dumstrei K, Hartenstein V. Role of FGFR signaling in the morphogenesis of the Drosophila visceral musculature. Dev Dyn. 2004;231(2):342–8.

    Article  CAS  PubMed  Google Scholar 

  123. Manieu C, Olivares GH, Vega-Macaya F, Valdivia M, Olguín P. Jitterbug/Filamin and myosin-II form a complex in tendon cells required to maintain epithelial shape and polarity during musculoskeletal system development. Mech Dev. 2018;154:309–14.

    Article  CAS  PubMed  Google Scholar 

  124. Mao Y, Rauskolb C, Cho E, Hu WL, Hayter H, Minihan G, Katz FN, Irvine KD. Dachs: an unconventional myosin that functions downstream of fat to regulate growth, affinity and gene expression in Drosophila. Development. 2006;133(13):2539–51.

    Article  CAS  PubMed  Google Scholar 

  125. March LE, Smaby RM, Setton EVW, Sharma PP. The evolution of selector gene function: Expression dynamics and regulatory interactions of tiptop/teashirt across Arthropoda. Evol Dev. 2018;20(6):219–32.

    Article  CAS  PubMed  Google Scholar 

  126. Marqués-Torrejón MÁ, Porlan E, Banito A, Gómez-Ibarlucea E, Lopez-Contreras AJ, Fernández-Capetillo O, Vidal A, Gil J, Torres J, Fariñas I. Cyclin-dependent kinase inhibitor p21 controls adult neural stem cell expansion by regulating Sox2 gene expression. Cell Stem Cell. 2013;12(1):88–100.

    Article  PubMed  Google Scholar 

  127. Martin D, Zusman S, Li X, Williams EL, Khare N, DaRocha S, Chiquet-Ehrismann R, Baumgartner S. wing blister, a new Drosophila laminin alpha chain required for cell adhesion and migration during embryonic and imaginal development. J Cell Biol. 1999;145(1):191–201.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Martin-Bermudo MD, Alvarez-Garcia I, Brown NH. Migration of the Drosophila primordial midgut cells requires coordination of diverse PS integrin functions. Development. 1999;126(22):5161–9.

    Article  CAS  PubMed  Google Scholar 

  129. Medina-Jiménez BI, Budd GE, Janssen R. Panarthropod tiptop/teashirt and Spalt orthologs and their potential role as “trunk”-selector genes. EvoDevo. 2021;12(1):7.

    Article  PubMed  PubMed Central  Google Scholar 

  130. Mittmann B, Wolff C. Embryonic development and staging of the cobweb spider Parasteatoda tepidariorum C. L. Koch, 1841 (syn.: Achaearanea tepidariorum; Araneomorphae; Theridiidae). Dev Genes Evol. 2012;222(4):189–216.

    Article  PubMed  Google Scholar 

  131. Miyaji T, Kawasaki T, Togawa N, Omote H, Moriyama Y. Type 1 sodium-dependent phosphate transporter acts as a membrane potential-driven urate exporter. Curr Mol Pharmacol. 2013;6(2):88–94.

    Article  CAS  PubMed  Google Scholar 

  132. Miyawaki K, Inoue Y, Mito T, Fujimoto T, Matsushima K, Shinmyo Y, Ohuchi H, Noji S. Expression patterns of aristaless in developing appendages of Gryllus bimaculatus (cricket). Mech Dev. 2002;113(2):181–4.

    Article  CAS  PubMed  Google Scholar 

  133. Mogami K, O’Donnell PT, Bernstein SI, Wright TR, Emerson CP Jr. Mutations of the Drosophila myosin heavy-chain gene: effects on transcription, myosin accumulation, and muscle function. Proc Natl Acad Sci U S A. 1986;83(5):1393–7.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  134. Monge I, Mitchell PJ. DAP-2, the Drosophila homolog of transcription factor AP-2. Mech Dev. 1998;76(1–2):191-doe5.

    Google Scholar 

  135. Moran MT, Tare M, Kango-Singh M, Singh A. Homeotic gene teashirt (tsh) has a neuroprotective function in amyloid-beta 42 mediated neurodegeneration. PLoS ONE. 2013;8(11):e80829.

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  136. Muskavitch MA. Delta-notch signaling and Drosophila cell fate choice. Dev Biol. 1994;166(2):415–30.

    Article  CAS  PubMed  Google Scholar 

  137. Nolan ED, Santibáñez-López CE, Sharma PP. Developmental gene expression as a phylogenetic data class: support for the monophyly of Arachnopulmonata. Dev Genes Evol. 2020;230(2):137–53.

    Article  CAS  PubMed  Google Scholar 

  138. Oda H, Nishimura O, Hirao Y, Tarui H, Agata K, Akiyama-Oda Y. Progressive activation of Delta-notch signaling from around the blastopore is required to set up a functional caudal lobe in the spider Achaearanea tepidariorum. Development. 2007;134(12):2195–205.

    Article  CAS  PubMed  Google Scholar 

  139. Oda H, Akiyama-Oda Y. The common house spider Parasteatoda tepidariorum. EvoDevo. 2020;11:6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Ogawa Y, Kakumoto K, Yoshida T, Kuwako KI, Miyazaki T, Yamaguchi J, Konno A, Hata J, Uchiyama Y, Hirai H, Watanabe M, Darnell RB, Okano H, Okano HJ. Elavl3 is essential for the maintenance of Purkinje neuron axons. Sci Rep. 2018;8(1):2722.

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  141. Oliveira MB, Liedholm SE, Lopez JE, Lochte AA, Pazio M, Martin JP, Mörch PR, Salakka S, York J, Yoshimoto A, Janssen R. Expression of arthropod distal limb-patterning genes in the onychophoran Euperipatoides kanangrensis. Dev Genes Evol. 2014;224(2):87–96.

    Article  PubMed  Google Scholar 

  142. Panara V, Budd GE, Janssen R. Phylogenetic analysis and embryonic expression of panarthropod Dmrt genes. Front Zool. 2019;16:23.

    Article  PubMed  PubMed Central  Google Scholar 

  143. Panfilio KA, Chuva de Sousa Lopes SM. The extended analogy of extraembryonic development in insects and amniotes. Phil Trans R Soc. 2022;B377:20210268. https://0-doi-org.brum.beds.ac.uk/10.1098/rstb.2021.0268.

    Article  Google Scholar 

  144. Pechmann M, Prpic NM. Appendage patterning in the South American bird spider Acanthoscurria geniculata (araneae: mygalomorphae). Dev Genes Evol. 2009;219(4):189–98.

    Article  PubMed  Google Scholar 

  145. Pechmann M, Khadjeh S, Turetzek N, McGregor AP, Damen WG, Prpic NM. Novel function of distal-less as a gap gene during spider segmentation. PLoS Genet. 2011;7(10):e1002342.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Perez Sanchez C, Casas-Tinto S, Sanchez L, Rey-Campos J, Granadino B. Dm fox F, a novel drosophila fork head factor expressed in visceral mesoderm. Mech Dev. 2002;111:163–6.

    Article  CAS  PubMed  Google Scholar 

  147. Pichaud F, Casares F. homothorax and Iroquois-C genes are required for the establishment of territories within the developing eye disc. Mech Dev. 2000;96(1):15–25.

    Article  CAS  PubMed  Google Scholar 

  148. Pierani A, Moran-Rivard L, Sunshine MJ, Littman DR, Goulding M, Jessell TM. Control of interneuron fate in the developing spinal cord by the progenitor homeodomain protein Dbx1. Neuron. 2001;29(2):367–84.

    Article  CAS  PubMed  Google Scholar 

  149. Prpic NM, Janssen R, Wigand B, Klingler M, Damen WG. Gene expression in spider appendages reveals reversal of exd/hth spatial specificity, altered leg gap gene dynamics, and suggests divergent distal morphogen signaling. Dev Biol. 2003;264(1):119–40.

    Article  CAS  PubMed  Google Scholar 

  150. Prpic NM, Damen WG. Notch-mediated segmentation of the appendages is a molecular phylotypic trait of the arthropods. Dev Biol. 2009;326(1):262–71.

    Article  CAS  PubMed  Google Scholar 

  151. Punovuori K, Migueles RP, Malaguti M, Blin G, Macleod KG, Carragher NO, Pieters T, van Roy F, Stemmler MP, Lowell S. N-cadherin stabilises neural identity by dampening anti-neural signals. Development. 2019;146(21):dev183269.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Rehorn KP, Thelen H, Michelson AM, Reuter R. A molecular aspect of hematopoiesis and endoderm development common to vertebrates and Drosophila. Development. 1996;122(12):4023–31.

    Article  CAS  PubMed  Google Scholar 

  153. Reuter R, Grunewald B, Leptin M. A role for the mesoderm in endodermal migration and morphogenesis in Drosophila. Development. 1993;119(4):1135–45.

    Article  CAS  PubMed  Google Scholar 

  154. Roark M, Sturtevant MA, Emery J, Vaessin H, Grell E, Bier E. scratch, a pan-neural gene encoding a zinc finger protein related to snail, promotes neuronal development. Genes Dev. 1995;9(19):2384–98.

    Article  CAS  PubMed  Google Scholar 

  155. Sasaki M, Akiyama-Oda Y, Oda H. Evolutionary origin of type IV classical cadherins in arthropods. BMC Evol Biol. 2017;17(1):142.

    Article  PubMed  PubMed Central  Google Scholar 

  156. Satija R, Farrell JA, Gennert D, Schier AF, Regev A. Spatial reconstruction of single-cell gene expression data. Nat Biotechnol. 2015;33(5):495–502.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Schacht MI, Schomburg C, Bucher G. six3 acts upstream of foxQ2 in labrum and neural development in the spider Parasteatoda tepidariorum. Dev Genes Evol. 2020;230(2):95–104.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Schaeper ND, Pechmann M, Damen WG, Prpic NM, Wimmer EA. Evolutionary plasticity of collier function in head development of diverse arthropods. Dev Biol. 2010;344(1):363–76.

    Article  CAS  PubMed  Google Scholar 

  159. Schneider E, Winzer R, Rissiek A, Ricklefs I, Meyer-Schwesinger C, Ricklefs FL, Bauche A, Behrends J, Reimer R, Brenna S, Wasielewski H, Lauten M, Rissiek B, Puig B, Cortesi F, Magnus T, Fliegert R, Müller CE, Gagliani N, Tolosa E. CD73-mediated adenosine production by CD8 T cell-derived extracellular vesicles constitutes an intrinsic mechanism of immune suppression. Nat Commun. 2021;12(1):5911.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  160. Schneitz K, Spielmann P, Noll M. Molecular genetics of aristaless, a prd-type homeo box gene involved in the morphogenesis of proximal and distal pattern elements in a subset of appendages in Drosophila. Genes Dev. 1993;7(1):114–29.

    Article  CAS  PubMed  Google Scholar 

  161. Schnorrer F, Kalchhauser I, Dickson BJ. The transmembrane protein Kon-tiki couples to Dgrip to mediate myotube targeting in Drosophila. Dev Cell. 2007;12(5):751–66.

    Article  CAS  PubMed  Google Scholar 

  162. Schönauer A, Paese CL, Hilbrant M, Leite DJ, Schwager EE, Feitosa NM, Eibner C, Damen WG, McGregor AP. The Wnt and Delta-Notch signalling pathways interact to direct pair-rule gene expression via caudal during segment addition in the spider Parasteatoda tepidariorum. Development. 2016;143(13):2455–63.

    PubMed  Google Scholar 

  163. Schomburg C, Turetzek N, Schacht MI, Schneider J, Kirfel P, Prpic NM, Posnien N. Molecular characterization and embryonic origin of the eyes in the common house spider Parasteatoda tepidariorum. EvoDevo. 2015;6:15.

    Article  PubMed  PubMed Central  Google Scholar 

  164. Schoppmeier M, Damen WG. Double-stranded RNA interference in the spider Cupiennius salei: the role of Distal-less is evolutionarily conserved in arthropod appendage formation. Dev Genes Evol. 2001;211(2):76–82.

    Article  CAS  PubMed  Google Scholar 

  165. Schuster CM, Davis GW, Fetter RD, Goodman CS. Genetic dissection of structural and functional components of synaptic plasticity. I. Fasciclin II controls synaptic stabilization and growth. Neuron. 1996;17(4):641–54.

    Article  CAS  PubMed  Google Scholar 

  166. Schwager EE, Sharma PP, Clarke T, Leite DJ, Wierschin T, Pechmann M, Akiyama-Oda Y, Esposito L, Bechsgaard J, Bilde T, Buffry AD, Chao H, Dinh H, Doddapaneni H, Dugan S, Eibner C, Extavour CG, Funch P, Garb J, Gonzalez LB, Gonzalez VL, Griffiths-Jones S, Han Y, Hayashi C, Hilbrant M, Hughes DST, Janssen R, Lee SL, Maeso I, Murali SC, Muzny DM, Nunes da Fonseca R, Paese CLB, Qu J, Ronshaugen M, Schomburg C, Schönauer A, Stollewerk A, Torres-Oliva M, Turetzek N, Vanthournout B, Werren JH, Wolff C, Worley KC, Bucher G, Gibbs RA, Coddington J, Oda H, Stanke M, Ayoub NA, Prpic NM, Flot JF, Posnien N, Richards S, McGregor AP. The house spider genome reveals an ancient whole-genome duplication during arachnid evolution. BMC Biol. 2017;15(1):62.

    Article  PubMed  PubMed Central  Google Scholar 

  167. Setton EVW, March LE, Nolan ED, Jones TE, Cho H, Wheeler WC, Extavour CG, Sharma PP. Expression and function of spineless orthologs correlate with distal deutocerebral appendage morphology across Arthropoda. Dev Biol. 2017;430(1):224–36.

    Article  CAS  PubMed  Google Scholar 

  168. Sharma PP. Chelicerates and the conquest of land: a view of arachnid origins through an evo-devo spyglass. Integr Comp Biol. 2017;57(3):510–22.

    Article  PubMed  Google Scholar 

  169. Shen CP, Jan LY, Jan YN. Miranda is required for the asymmetric localization of Prospero during mitosis in Drosophila. Cell. 1997;90(3):449–58.

    Article  CAS  PubMed  Google Scholar 

  170. Sherr CJ, Roberts JM. CDK inhibitors: positive and negative regulators of G1-phase progression. Genes Dev. 1999;13(12):1501–12.

    Article  CAS  PubMed  Google Scholar 

  171. Shi L, Lee T. Molecular diversity of Dscam and self-recognition. Adv Exp Med Biol. 2012;739:262–75.

    Article  CAS  PubMed  Google Scholar 

  172. Shishido E, Ono N, Kojima T, Saigo K. Requirements of DFR1/Heartless, a mesoderm-specific Drosophila FGF-receptor, for the formation of heart, visceral and somatic muscles, and ensheathing of longitudinal axon tracts in CNS. Development. 1997;124(11):2119–28.

    Article  CAS  PubMed  Google Scholar 

  173. Sink H, Rehm EJ, Richstone L, Bulls YM, Goodman CS. sidestep encodes a target-derived attractant essential for motor axon guidance in Drosophila. Cell. 2001;105(1):57–67.

    Article  CAS  PubMed  Google Scholar 

  174. Skeath JB, Carroll SB. The achaete-scute complex: generation of cellular pattern and fate within the Drosophila nervous system. FASEB J. 1994;8(10):714–21.

    Article  CAS  PubMed  Google Scholar 

  175. Skeath JB, Doe CQ. The achaete-scute complex proneural genes contribute to neural precursor specification in the Drosophila CNS. Curr Biol. 1996;6(9):1146–52.

    Article  CAS  PubMed  Google Scholar 

  176. Stollewerk A. Recruitment of cell groups through Delta/Notch signalling during spider neurogenesis. Development. 2002;129(23):5339–48.

    Article  CAS  PubMed  Google Scholar 

  177. Stollewerk A, Tautz D, Weller M. Neurogenesis in the spider: new insights from comparative analysis of morphological processes and gene expression patterns. Arthropod Struct Dev. 2003;32(1):5–16.

    Article  PubMed  Google Scholar 

  178. Stollewerk A, Seyfarth EA. Evolutionary changes in sensory precursor formation in arthropods: embryonic development of leg sensilla in the spider Cupiennius salei. Dev Biol. 2008;313(2):659–73.

    Article  CAS  PubMed  Google Scholar 

  179. Tabuchi K, Yoshikawa S, Yuasa Y, Sawamoto K, Okano H. A novel Drosophila paired-like homeobox gene related to Caenorhabditis elegans unc-4 is expressed in subsets of postmitotic neurons and epidermal cells. Neurosci Lett. 1998;257(1):49–52.

    Article  CAS  PubMed  Google Scholar 

  180. Takeshima H, Nishi M, Iwabe N, Miyata T, Hosoya T, Masai I, Hotta Y. Isolation and characterization of a gene for a ryanodine receptor/calcium release channel in Drosophila melanogaster. FEBS Lett. 1994;337(1):81–7.

    Article  CAS  PubMed  Google Scholar 

  181. Theocharis AD, Skandalis SS, Gialeli C, Karamanos NK. Extracellular matrix structure. Adv Drug Deliv Rev. 2016;97:4–27.

    Article  CAS  PubMed  Google Scholar 

  182. Toegel JP, Wimmer EA, Prpic NM. Loss of spineless function transforms the Tribolium antenna into a thoracic leg with pretarsal, tibiotarsal, and femoral identity. Dev Genes Evol. 2009;219(1):53–8.

    Article  PubMed  Google Scholar 

  183. Tomlinson A, Mavromatakis YE, Arias R. The role of sevenless in Drosophila R7 photoreceptor specification. Dev Biol. 2019;454(2):181–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  184. Treffkorn S, Mayer G, Janssen R. Review of extra-embryonic tissues in the closest arthropod relatives, onychophorans and tardigrades. Phil Trans R Soc B. 2022;377:20210270. https://0-doi-org.brum.beds.ac.uk/10.1098/rstb.2021.0270.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Tripathi BK, Das R, Mukherjee A, Mutsuddi M. Interaction of spoonbill with prospero in Drosophila: Implications in neuroblast development. Genesis. 2017;55(9). https://0-doi-org.brum.beds.ac.uk/10.1002/dvg.23049.

  186. Trush O, Liu C, Han X, Nakai Y, Takayama R, Murakawa H, Carrillo JA, Takechi H, Hakeda-Suzuki S, Suzuki T, Sato M. N-cadherin orchestrates self-Organization of Neurons within a columnar unit in the Drosophila medulla. J Neurosci. 2019;39(30):5861–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Turetzek N, Pechmann M, Schomburg C, Schneider J, Prpic NM. Neofunctionalization of a duplicate dachshund gene underlies the evolution of a novel leg segment in Arachnids. Mol Biol Evol. 2016;33(1):109–21.

    Article  CAS  PubMed  Google Scholar 

  188. Ungerer P, Eriksson BJ, Stollewerk A. Neurogenesis in the water flea Daphnia magna (Crustacea, Branchiopoda) suggests different mechanisms of neuroblast formation in insects and crustaceans. Dev Biol. 2011;357(1):42–52.

    Article  CAS  PubMed  Google Scholar 

  189. Urbano JM, Domínguez-Giménez P, Estrada B, Martín-Bermudo MD. PS integrins and laminins: key regulators of cell migration during Drosophila embryogenesis. PLoS ONE. 2011;6(9):e23893.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  190. Vässin H, Bremer KA, Knust E, Campos-Ortega JA. The neurogenic gene Delta of Drosophila melanogaster is expressed in neurogenic territories and encodes a putative transmembrane protein with EGF-like repeats. EMBO J. 1987;6(11):3431–40.

    Article  PubMed  PubMed Central  Google Scholar 

  191. Wang J, Ding M. Robo and Ror function in a common receptor complex to regulate Wnt-mediated neurite outgrowth in Caenorhabditis elegans. Proc Natl Acad Sci U S A. 2018;115(10):E2254–63.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  192. Wang Q, Han TH, Nguyen P, Jarnik M, Serpe M. Tenectin recruits integrin to stabilize Bouton architecture and regulate vesicle release at the Drosophila neuromuscular junction. Elife. 2018;7:e35518.

    Article  PubMed  PubMed Central  Google Scholar 

  193. Weller M, Tautz D. Prospero and snail expression during spider neurogenesis. Dev Genes Evol. 2003;213(11):554–66.

    Article  CAS  PubMed  Google Scholar 

  194. Wheeler SR, Carrico ML, Wilson BA, Brown SJ, Skeath JB. The expression and function of the achaete-scute genes in Tribolium castaneum reveals conservation and variation in neural pattern formation and cell fate specification. Development. 2003;130(18):4373–81.

    Article  CAS  PubMed  Google Scholar 

  195. Wu X, Hu S, Kang X, Wang C. Synaptotagmins: beyond presynaptic neurotransmitter release. Neuroscientist. 2020;26(1):9–15.

    Article  PubMed  Google Scholar 

  196. Xu J, Hao X, Yin MX, Lu Y, Jin Y, Xu J, Ge L, Wu W, Ho M, Yang Y, Zhao Y, Zhang L. Prevention of medulla neuron dedifferentiation by Nerfin-1 requires inhibition of notch activity. Development. 2017;144(8):1510–7.

    CAS  PubMed  Google Scholar 

  197. Yamazaki K, Akiyama-Oda Y, Oda H. Expression patterns of a twist-related gene in embryos of the spider Achaearanea tepidariorum reveal divergent aspects of mesoderm development in the fly and spider. Zoolog Sci. 2005;22(2):177–85.

    Article  CAS  PubMed  Google Scholar 

  198. Yang S, Wu X, Daoutidou EI, Zhang Y, Shimell M, Chuang KH, Peterson AJ, O’Connor MB, Zheng X. The NDNF-like factor Nord is a hedgehog-induced extracellular BMP modulator that regulates Drosophila wing patterning and growth. Elife. 2022;11:e73357.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  199. Ypsilanti AR, Chedotal A. Roundabout receptors. Adv Neurobiol. 2014;8:133–64.

    Article  PubMed  Google Scholar 

  200. Zander MA, Cancino GI, Gridley T, Kaplan DR, Miller FD. The snail transcription factor regulates the numbers of neural precursor cells and newborn neurons throughout mammalian life. PLoS ONE. 2014;9(8):e104767.

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  201. Zhang N (née Turetzek). Developmental Studies of Appendage Patterning and Formation in Spiders. PhD thesis. 2016. Georg-August University Göttingen, Germany. https://0-doi-org.brum.beds.ac.uk/10.53846/goediss-5988

  202. Zhang Y, Featherstone D, Davis W, Rushton E, Broadie K. Drosophila D-titin is required for myoblast fusion and skeletal muscle striation. J Cell Sci. 2000;113(Pt 17):3103–15.

    Article  CAS  PubMed  Google Scholar 

  203. Zipursky SL, Sanes JR. Chemoaffinity revisited: dscams, protocadherins, and neural circuit assembly. Cell. 2010;143(3):343–53.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We are grateful to the members of the Developmental Biology department (Arendt Lab) and GeneCore at the European Molecular Biology Laboratory (EMBL) in Heidelberg, Germany for providing facilities, equipment and assistance during the single cell experiments. Special thanks to Jacob M. Musser (Yale University), Daniel Leite, Laura Piovani (University College London), Alba Almazán (Institut de Génomique Fonctionnelle de Lyon), Periklis Paganos (Stazione Zoologica Anton Dohrn) and Åsa Björklund (National Bioinformatics Infrastructure Sweden) for their guidance during the process of data analysis.

The data handling for this project was enabled by resources provided by the Swedish National Infrastructure for Computing (SNIC) at UPPMAX, partially funded by the Swedish Research Council under grant agreement No. 2018-05973.

Funding

Open access funding provided by Uppsala University. This project has received funding from the European Union’s Horizon 2020 Research and Innovation programme under grant agreement No. 766053.

Author information

Authors and Affiliations

Authors

Contributions

BIMJ, RJ and GEB designed the outline of the project. BIMJ performed all steps of the single-cell RNA sequencing experiments and the subsequent in-silico analysis of the data. BIMJ performed most of the in-situ hybridization experiments. RJ validated the in-situ hybridization experiments and performed additional in-situ hybridization experiments. BIMJ and RJ analysed and discussed the results of the experiments and wrote the first draft version of the manuscript. RJ created the figures. All authors worked on the final version of the manuscript. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Brenda I. Medina-Jiménez or Ralf Janssen.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1.

Additional file 2.

Additional file 3.

Additional file 4.

Additional file 5.

Additional file 6.

Additional file 7.

Additional file 8.

Additional file 9.

Additional file 10.

Additional file 11.

Additional file 12.

Additional file 13.

Additional file 14.

Additional file 15.

Additional file 16.

Additional file 17.

Additional file 18.

Additional file 19.

Additional file 20.

Additional file 21.

Additional file 22.

Additional file 23.

Additional file 24.

Additional file 25.

Additional file 26.

Additional file 27.

Additional file 28.

Additional file 29.

Additional file 30.

Additional file 31.

Additional file 32.

Additional file 33.

Additional file 34.

Additional file 35.

Additional file 36.

Additional file 37.

Additional file 38.

Additional file 39.

Additional file 40.

Additional file 41.

Additional file 42.

Additional file 43.

Additional file 44.

Additional file 45.

Additional file 46.

Additional file 47.

Additional file 48.

Additional file 49.

Additional file 50.

Additional file 51.

Additional file 52.

Additional file 53.

Additional file 54.

Additional file 55.

Additional file 56.

Additional file 57.

Additional file 58.

Additional file 59.

Additional file 60.

Additional file 61.

Additional file 62.

Additional file 63.

Additional file 64.

Additional file 65.

Additional file 66.

Additional file 67.

Additional file 68.

Additional file 69.

Additional file 70.

Additional file 71.

Additional file 72.

Additional file 73.

Additional file 74.

Additional file 75.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Medina-Jiménez, B.I., Budd, G.E. & Janssen, R. Single-cell RNA sequencing of mid-to-late stage spider embryos: new insights into spider development. BMC Genomics 25, 150 (2024). https://0-doi-org.brum.beds.ac.uk/10.1186/s12864-023-09898-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/s12864-023-09898-x

Keywords