A central challenge of developmental and evolutionary biology is to understand how anatomy is encoded in the genome. Elucidating the genetic mechanisms that control the development of specific anatomical features will require the analysis of model morphogenetic processes and an integration of biological information at genomic, cellular and tissue levels. The formation of the endoskeleton of the sea urchin embryo is a powerful experimental system for developing such an integrated view of the genomic regulatory control of morphogenesis. The dynamic cellular behaviors that underlie skeletogenesis are well understood and a complex transcriptional gene regulatory network (GRN) that underlies the specification of embryonic skeletogenic cells (primary mesenchyme cells, PMCs) has recently been elucidated. Here, we link the PMC specification GRN to genes that directly control skeletal morphogenesis. We identify new gene products that play a proximate role in skeletal morphogenesis and uncover transcriptional regulatory inputs into many of these genes. Our work extends the importance of the PMC GRN as a model developmental GRN and establishes a unique picture of the genomic regulatory control of a major morphogenetic process. Furthermore, because echinoderms exhibit diverse programs of skeletal development, the newly expanded sea urchin skeletogenic GRN will provide a foundation for comparative studies that explore the relationship between GRN evolution and morphological evolution.

The dynamic anatomical changes that characterize embryogenesis are encoded in the genome. The genomic regulatory control of development can be understood in terms of transcriptional gene regulatory networks (GRNs), which can be defined as dynamic networks of interacting genes that encode transcription factors (i.e. regulatory genes). Such networks are being used to analyze cell specification in diverse organisms (Stathopoulos and Levine, 2005; Ettensohn, 2009; Nikitina et al., 2009; Peter and Davidson, 2009; Davidson, 2010). There has been less progress, however, in using GRNs to explain the complex cell and tissue behaviors that drive changes in embryonic form. Insights in this area will emerge first from model systems in which there is a detailed understanding of both (1) cell specification at the GRN level and (2) morphogenetic mechanisms at the molecular, cellular and tissue levels. Establishing linkages between early cell specification networks and specific morphogenetic processes is crucially important, not just for understanding embryogenesis per se, but also in an evolutionary context, i.e. for understanding the ways in which evolutionary modifications to genetic networks have led to changes in morphological features.

The formation of the embryonic endoskeleton of sea urchins is a powerful experimental model for developing an integrated view of the genomic regulatory control of morphogenesis. The skeleton is the primary determinant of the distinctive, angular shape of the larva and influences its orientation, swimming and feeding (Pennington and Strathmann, 1990; Hart and Strathmann, 1994). The skeleton is a biomineral that consists of calcite and small amounts of occluded proteins. It is secreted by primary mesenchyme cells (PMCs), a specialized population of skeletogenic cells that have a well-defined embryonic lineage. During gastrulation, PMCs undergo a striking sequence of morphogenetic behaviors that includes epithelial-mesenchymal transition, directional cell migration and cell-cell fusion (Wilt and Ettensohn, 2007; Ettensohn, 2009). These cellular behaviors have been analyzed in remarkable detail, largely owing to the optical transparency of the sea urchin embryo and its suitability for in vivo imaging (Gustafson and Wolpert, 1967; Ettensohn and Malinda, 1993; Malinda et al., 1995; Miller et al., 1995; Peterson and McClay, 2003; Hodor and Ettensohn, 2008).

Recently, a GRN that underlies PMC specification has been described (Oliveri et al., 2008; Ettensohn, 2009). This network is initially deployed through the activity of polarized, maternal inputs that activate a small set of early zygotic regulatory genes selectively in the large micromere-PMC lineage. The transcription factors encoded by these genes engage additional layers of regulatory genes; various feedback and feed-forward interactions subsequently stabilize the transcriptional network and drive it forward (Oliveri et al., 2008). Although much information is available concerning the interactions among regulatory genes that are deployed in the network, little is known concerning the downstream circuitry of the network, i.e. the current network model includes few connections to downstream genes that play a direct role in the morphogenetic program of these cells. Some of these downstream morphogenetic effector (‘morphoregulatory’) genes are known, although many important genes in this class remain to be identified.

In previous work, we generated an arrayed cDNA library from PMCs at the mid-gastrula stage, when these cells are engaged in their most prominent morphogenetic behaviors. Initial analysis of this library allowed us to identify several components of the PMC GRN, including delta (Sweet et al., 2002), the transcription factors alx1 and erg (Zhu et al., 2001; Ettensohn et al., 2003), and several biomineralization-related genes (Illies et al., 2002; Cheers and Ettensohn, 2005; Livingston et al., 2006). In the present study, we greatly expand the current PMC GRN model by identifying many additional morphoregulatory genes and by uncovering regulatory inputs into these genes. Our work extends the value of the PMC GRN as a model developmental GRN and establishes a framework for understanding the genomic circuitry that encodes a major anatomical feature.

Embryo culture

Strongylocentrotus purpuratus embryos were obtained and cultured at 15°C as described previously (Zhu et al., 2001).

The PMC cDNA library and expressed sequence tag (EST) collection

The construction and arraying of the PMC cDNA library has been described (Zhu et al., 2001; Livingston et al., 2006). Briefly, the library was generated from polyA(+) RNA that was isolated from micromeres (presumptive PMCs) that were cultured until sibling control embryos reached the mid-gastrula stage, ∼36 hours post-fertilization. The cDNA library was not normalized or subtracted in any way. Reverse transcription was primed using oligo(dT) and the resulting cDNAs were directionally cloned into pSPORT with an average insert size of 1.5-2 kb. Clones were robotically arrayed in 384-well plates and subjected to Sanger-based sequencing. A total of 51,097 cDNA sequence reads (ESTs) were obtained from randomly selected clones (Genbank entries BG780044-BG789446, DN560823-DN586179 and DN781759-DN810571). In most cases, single 5′-reads were obtained from each clone; in some cases, both 5′ and 3′ sequence reads were obtained.

Whole-mount in situ hybridization (WMISH)

WMISH was performed as described previously (Zhu et al., 2001).

Morpholino (MO) injections

MOs were injected into fertilized eggs as described by Cheers and Ettensohn (Cheers and Ettensohn, 2004), with the modification that eggs were fertilized in the presence of 0.1% (wt/vol) para-aminobenzoic acid to prevent hardening of the fertilization envelope. MO sequences were: SpAlx1: 5′-TATTGAGTTAAGTCTCGGCACGACA-3′; SpEts1: 5′-GAACAGTGCATAGACGCCATGATTG-3′; SpTbr: 5′-TGTAATTCTTCTCCCATCATGTCTC-3′.

All three MOs were translation-blocking MOs. The SpAlx1 MO has been shown to specifically and effectively block SpAlx1 expression (Ettensohn et al., 2003). The SpEts1 and SpTbr MOs were also described previously (Oliveri et al., 2008). The SpEts1 MO phenocopies the overexpression of a dominant-negative form of Ets1, which blocks PMC specification and ingression (Sharma and Ettensohn, 2010). The SpTbr MO produces a selective effect on skeletogenesis without affecting PMC ingression or migration (Oliveri et al., 2008). MOs were injected at working concentrations of 4 mM (Alx1) or 2 mM (Ets1, Tbr).

Quantitative polymerase chain reaction (QPCR)

QPCR was carried out as described previously (Stamateris et al., 2010). PCR primers that were used in this study are shown in supplementary material Table S4. Technical duplicates were included for each sample and the average of the two Ct values was used to calculate ΔCt. For temporal QPCR profiles, the absolute numbers of mRNA molecules per embryo at various developmental stages were calculated using Sp-z12 as a standard. The numbers of Sp-z12 transcripts at various stages have been determined by RNA titration (Wang et al., 1995). To determine regulatory inputs into genes, MOs were used to block the expression of ets1, alx1 or tbr and the effects on the expression of downstream skeletogenic genes were analyzed at 28-30 hours post-fertilization by QPCR.

We assessed the expression of each gene in morphant embryos compared with control siblings in three batches of embryos that were derived from independent matings (three biological replicates). We calculated ΔCt values with respect to Sp-z12 and used Student’s t-test to compare the levels of expression of each gene in control and morphant embryos (n=3). A P value of ≤0.05 (i.e. a 95% confidence value) was taken to be significant and such genes were considered to have an input from the gene that was targeted by the MO. Of the genes that satisfied this criterion, most showed an average level of expression that was <25% of the control level. We also considered an effect on gene expression to be meaningful if, in all three biological replicates, the fold-difference in expression in control embryos relative to morphants was at least threefold and always in the same direction, a criterion applied by Oliveri and co-workers (Oliveri et al., 2008). Applying this criterion caused us to score as regulatory targets a small number of genes (three cases in total) that consistently showed pronounced effects, but which also showed a relatively high level of variability across biological replicates. Lastly, in three cases, genes that showed expression of <33% of the control level in two of the three biological replicates were analyzed independently by WMISH. We considered a gene to have an input if its level of expression was clearly and consistently reduced by MO knockdown, as assessed by WMISH (all three cases).

Analysis of genes expressed by PMCs during gastrulation

Sanger-based sequence reads (51,097; average read length >900 nts) from randomly selected PMC cDNAs were mapped to the complete set of GLEAN3 gene predictions (Sea Urchin Genome Sequencing Consortium, 2006). We expanded the GLEAN3 models to include 3′-UTR sequences that were identified through a genome-wide tiling analysis of gene expression (Samanta et al., 2006). Of the 51,097 ESTs, about half (24,238) aligned to GLEAN3 gene models; these ESTs were mapped to 7415 different GLEAN3 genes (supplementary material Table S1). The number of EST matches per gene varied from one to 669 (supplementary material Fig. S1). Approximately one-third of the 7415 GLEAN3 models had a single EST match, approximately one-third had two to four matches, and approximately one-third had five or more matches. The number of EST hits per gene corresponded very roughly to the abundances of the corresponding transcripts at the mid-gastrula stage (36 hours post-fertilization) as measured by QPCR (see below). Manual curation of a randomly selected sample of 100 ESTs that did not match GLEAN3 models showed that ∼50% could be attributed to overly conservative predictions of 3′-UTR sequences, ∼25% represented rRNA sequences or previously unidentified exons, and the remaining 25% represented simple repetitive sequences or other sequences that could not be mapped to the current sea urchin nuclear or mitochondrial genome assemblies.

As one means of assessing the completeness of the collection of PMC-expressed genes, we investigated whether it included all genes that had been shown in previous studies to be expressed selectively by PMCs. Of the 52 such genes that had been identified prior to this study (supplementary material Table S2), only three were not represented in our collection of PMC-expressed genes. In two cases (foxN2/3 and snail), this was attributable to the temporal expression patterns of the genes; in S. purpuratus, foxN2/3 expression is extinguished in the micromere lineage by the gastrula stage (Tu et al., 2006), whereas snail is not expressed until late in gastrulation, and then only at extremely low levels (Oliveri et al., 2008). In the case of fgfr2, closer inspection showed that the collection of PMC ESTs included sequences that corresponded to the 3′UTR of the fgfr2 mRNA, which was incorrectly predicted in the gene model.

Fig. 1.

WMISH analysis of mRNAs that are restricted to PMCs during gastrulation. BL, blastula; LG, late gastrula; MB, mesenchyme blastula; MG, mid-gastrula; PL, pluteus larva; PR, prism. Most embryos are viewed laterally; some (e.g. alx4 MG and fos LG, panels at far right) are also viewed along the animal-vegetal axis.

Fig. 1.

WMISH analysis of mRNAs that are restricted to PMCs during gastrulation. BL, blastula; LG, late gastrula; MB, mesenchyme blastula; MG, mid-gastrula; PL, pluteus larva; PR, prism. Most embryos are viewed laterally; some (e.g. alx4 MG and fos LG, panels at far right) are also viewed along the animal-vegetal axis.

Although the catalog of genes that are expressed by PMCs during gastrulation (supplementary material Table S1) is likely to be nearly complete, this list also contains some genes that are not expressed by PMCs but are represented in the catalog because the micromere preparation that was used to generate the original cDNA library was not completely pure and contained small numbers of other cell types (∼5%) (Zhu et al., 2001). For example, we identified several ESTs that corresponded to spec1, a highly abundant, aboral ectoderm-specific transcript. It is significant, however, that many transcripts that are expressed at moderate to high levels at the gastrula stage in various non-skeletogenic lineages were completely absent from the PMC mRNA catalog, or were extremely rare. For example, ESTs that corresponded to the endoderm marker endo16 (GLEAN3_11038/9), the pigment cell marker pks (GLEAN3_02895) and the apical ectoderm marker nk2-1 (GLEAN3_00757) were absent from the PMC EST collection, as expected. Based on these considerations, it seems prudent to conclude that the presence of any specific gene in the PMC gene catalog, particularly if the gene has very few EST matches, is only suggestive of expression in PMCs and confirmation by independent methods is required.

Identification of new morphoregulatory genes expressed selectively by PMCs

The spatial patterns of expression of 180 genes in the PMC gene catalog were analyzed by whole-mount in situ hybridization (WMISH). These genes were chosen by focusing on relatively abundant transcripts (based on the number of EST matches/gene) and by excluding obvious housekeeping genes or genes that had been previously analyzed by WMISH. Embryonic stages from the fertilized egg to the early pluteus larva were examined. Many of the genes we analyzed by WMISH showed rather general patterns of expression during development; however, we also identified 33 mRNAs that were restricted to the large micromere-PMC lineage or were highly enriched in these cells during gastrulation, when the PMCs are actively engaged in their salient morphogenetic activities (Figs 1, 2). Most of these mRNAs encode proteins that are expected to function in morphogenetic processes and not in cell specification; e.g. cytoskeletal proteins and their regulators, putative cell adhesion and extracellular matrix (ECM) proteins, new biomineralization proteins and novel proteins. We did, however, also identify three new PMC-enriched transcription factors (Fos, Smad1/5/8 and Alx4). The functions of some of the gene products that emerged from the WMISH analysis have been well characterized in other organisms (e.g. Wasp, Arp proteins, Cdc42, etc.) but others are novel proteins or have cellular functions that are less well understood. Information concerning the newly identified PMC-specific and PMC-enriched proteins that emerged from the WMISH analysis is summarized in supplementary material Table S3.

Fig. 2.

WMISH analysis of mRNAs that are enriched in PMCs and NSM cells during gastrulation. Many of these mRNAs encode cytoskeleton-related proteins. BL, blastula; EB, early blastula; EG, early gastrula; LG, late gastrula; MB, mesenchyme blastula. Most embryos are viewed laterally; some (e.g. arp1 LG and arp3 LG, panels at far right) are also viewed along the animal-vegetal axis.

Fig. 2.

WMISH analysis of mRNAs that are enriched in PMCs and NSM cells during gastrulation. Many of these mRNAs encode cytoskeleton-related proteins. BL, blastula; EB, early blastula; EG, early gastrula; LG, late gastrula; MB, mesenchyme blastula. Most embryos are viewed laterally; some (e.g. arp1 LG and arp3 LG, panels at far right) are also viewed along the animal-vegetal axis.

Quantitative polymerase chain reaction (QPCR) was used to determine high-resolution temporal expression profiles of many of the newly identified mRNAs (Fig. 3). Several PMC-specific mRNAs that had been identified in previous studies, but for which detailed expression profiles were not available, were also included in the analysis. Only four of the genes that we examined (casc1, cyp2, smad1/5/8 and jun) were expressed maternally at levels of >1000 transcripts per egg. Zygotic transcription of all of the genes was activated prior to the onset of PMC ingression (24 hours post-fertilization), as shown by an increase in steady-state transcript levels. The timing of zygotic activation, however, appeared to vary considerably. The first genes to be expressed (casc1, cdi, fos, p19, egf-rich/p41, pks2 and stomatin) were activated by 12 hours post-fertilization (early blastula stage) whereas transcripts encoded by late genes (can1 and p16rel1) did not begin to accumulate until shortly before PMC ingression. Maximal levels of expression for most genes were observed during gastrulation (24-48 hours post-fertilization), which includes the developmental stage from which the PMC cDNA library was prepared (mid-gastrula stage; ∼36 hours post-fertilization). More than half of the genes (13/25) showed maximal expression at 42 hours post-fertilization; two exceptions were sm29 and Clectin, which showed maximal transcript levels at 72 hours post-fertilization (the latest stage examined). Calculated on a per-cell basis (32 PMCs per embryo), peak expression levels varied from 30 transcripts per cell (can1) to 600 transcripts per cell (p19).

Fig. 3.

High-resolution, temporal expression profiles of newly identified components of the PMC GRN, as determined by quantitative polymerase chain reaction (QPCR). Genes are ordered on the vertical axis according to the approximate time of activation. The levels of expression for each gene are given as the number of transcripts per embryo (see Materials and methods) and colored according to the key below. Samples were collected every 3 hours from the unfertilized egg (0 hour) to the late gastrula stage (48 hours), then at 54 hours (prism), 60 hours and 72 hours (early pluteus larva). For reference, other developmental stages included early cleavage (3-6 hours), early blastula (12 hours), mesenchyme blastula (24 hours), early gastrula (30 hours) and mid-gastrula (36 hours).

Fig. 3.

High-resolution, temporal expression profiles of newly identified components of the PMC GRN, as determined by quantitative polymerase chain reaction (QPCR). Genes are ordered on the vertical axis according to the approximate time of activation. The levels of expression for each gene are given as the number of transcripts per embryo (see Materials and methods) and colored according to the key below. Samples were collected every 3 hours from the unfertilized egg (0 hour) to the late gastrula stage (48 hours), then at 54 hours (prism), 60 hours and 72 hours (early pluteus larva). For reference, other developmental stages included early cleavage (3-6 hours), early blastula (12 hours), mesenchyme blastula (24 hours), early gastrula (30 hours) and mid-gastrula (36 hours).

Transcriptional inputs into morphoregulatory genes

We used morpholino (MO) knockdowns and QPCR to identify regulatory inputs into many of the newly identified genes. We focused on Ets1, Alx1 and Tbr, three well-characterized transcription factors that provide some of the earliest inputs into the network. For QPCR studies, we restricted our analysis to downstream genes that were expressed exclusively by PMCs, so that potentially subtle changes in transcript levels in PMCs would not be obscured by the presence of the same mRNA in other cell types. We examined the effects of gene knockdowns at 28-30 hours of development (i.e. at the early gastrula stage), when all morphoregulatory genes that we tested were ordinarily expressed at high levels (see Fig. 3).

The effects of Ets1, Alx1 and Tbr knockdowns on the levels of expression of PMC-specific mRNAs are shown in Fig. 4. The regulatory inputs of Ets1, Alx1, and Tbr into their morphoregulatory gene targets are invariably positive inputs. We did, however, detect auto-repression by two of the regulatory genes, alx1 and tbr, as was reported previously (Ettensohn et al., 2003; Oliveri et al., 2008). We also used WMISH to confirm our QPCR results; in every case that we examined, WMISH data supported the assignments of regulatory linkages that were based on QPCR (Fig. 5). We also used WMISH to assess certain inputs into morphoregulatory genes that showed substantial changes in expression in more than one biological replicate but failed to meet the statistical criteria that we established (see Materials and methods). In all three such cases, WMISH analysis supported the assignment of a positive regulatory input (see Fig. 5; effects of Ets1 knockdown on the expression of p58B, sm49 and Clectin).

Figure 6 summarizes the new network connections that were identified in this study. Our findings reveal that: (1) tbr has very few inputs into morphoregulatory genes; (2) ets1 and alx1 provide inputs into the majority (approximately two-thirds) of the morphoregulatory genes, including all of the known biomineralization-related genes that we tested; and (3) the targets of ets1 and alx1 are nearly identical. We detected no essential inputs from ets1, alx1 or tbr into six of the morphoregulatory genes that we tested (casc1, cdi, cyp2, egf-rich/p41, fgfr2 and stomatin). By combining this new network circuitry with previously published work, it is now possible to construct an expanded model of the PMC GRN that bridges upstream regulatory genes to many of the effector genes in the network (Fig. 7).

The formation of the complex and precisely patterned embryonic skeleton is the culmination of a sequence of morphogenetic activities on the part of the biomineral-forming PMCs (Wilt and Ettensohn, 2007). At present, there is a relatively detailed picture of the gene network that underlies early PMC specification; indeed, the micromere-PMC GRN is arguably the most complete developmental GRN in any experimental model (Oliveri et al., 2008; reviewed by Ettensohn, 2009). It is evident that this early specification network impinges on genes that control morphogenesis, as the behaviors of PMCs during gastrulation are dependent upon zygotic transcriptional inputs (Kurokawa et al., 1999; Ettensohn et al., 2003; Wu and McClay, 2007). The connections between the transcriptional GRN and the various cellular activities that underlie skeletogenesis, however, are not well understood. It is precisely this set of regulatory linkages that is most directly related to the assembly and patterning of the larval skeleton.

Fig. 4.

Transcriptional inputs into morphoregulatory genes. (A-C) Three early regulatory genes, ets1 (A), alx1 (B) and tbr (C), were knocked down individually using morpholinos (MOs) and the effect on levels of mRNAs encoded by morphoregulatory genes was analyzed by QPCR at 28-30 hours post-fertilization. The effect of each MO on each mRNA was assessed in three independent biological replicates. Vertical bars indicate the ratio of expression in morphant embryos relative to sibling controls (equal expression=1, as indicated by the dotted lines; values <1 indicate reduced expression in morphant embryos). All genes in the shaded gray areas were assigned a regulatory input based on QPCR. Blue boxes indicate genes that failed the initial t-test criterion (see Materials and methods) but were assigned a regulatory input because all three biological replicates showed levels of expression in morphant embryos that were one-third or less of control levels (Oliveri and Davidson, 2008). Red boxes indicate genes that showed variable results in the QPCR analysis but were assigned regulatory inputs based on WMISH analysis (Fig. 5). Asterisks indicate other regulatory interactions that were examined by WMISH; in all cases, WMISH analysis confirmed findings that were based on QPCR. The double asterisk in B indicates a single biological replicate in which the fold-increase in alx1 expression in alx1 morphant embryos was off-scale (>2.5).

Fig. 4.

Transcriptional inputs into morphoregulatory genes. (A-C) Three early regulatory genes, ets1 (A), alx1 (B) and tbr (C), were knocked down individually using morpholinos (MOs) and the effect on levels of mRNAs encoded by morphoregulatory genes was analyzed by QPCR at 28-30 hours post-fertilization. The effect of each MO on each mRNA was assessed in three independent biological replicates. Vertical bars indicate the ratio of expression in morphant embryos relative to sibling controls (equal expression=1, as indicated by the dotted lines; values <1 indicate reduced expression in morphant embryos). All genes in the shaded gray areas were assigned a regulatory input based on QPCR. Blue boxes indicate genes that failed the initial t-test criterion (see Materials and methods) but were assigned a regulatory input because all three biological replicates showed levels of expression in morphant embryos that were one-third or less of control levels (Oliveri and Davidson, 2008). Red boxes indicate genes that showed variable results in the QPCR analysis but were assigned regulatory inputs based on WMISH analysis (Fig. 5). Asterisks indicate other regulatory interactions that were examined by WMISH; in all cases, WMISH analysis confirmed findings that were based on QPCR. The double asterisk in B indicates a single biological replicate in which the fold-increase in alx1 expression in alx1 morphant embryos was off-scale (>2.5).

Fig. 5.

WMISH analysis of gene expression in morphant embryos. The expression of various genes was compared in morpholino-injected and sibling control embryos at 28-30 hours of development (early gastrula stage).

Fig. 5.

WMISH analysis of gene expression in morphant embryos. The expression of various genes was compared in morpholino-injected and sibling control embryos at 28-30 hours of development (early gastrula stage).

The identification of morphoregulatory genes

Our EST analysis and WMISH screen led to the identification of many PMC-specific and PMC-enriched mRNAs that encode cytoskeletal proteins and their regulators, putative adhesion proteins, extracellular matrix proteins and biomineralization proteins (supplementary material Table S3). The developmental functions of many of these proteins have yet to be analyzed in detail, but in other cases their functions are relatively well understood. For example, spicule matrix proteins are a family of secreted proteins that are occluded within the biomineral and influence its growth and physical properties (Wilt and Ettensohn, 2007). A recent proteomic analysis of proteins in the embryonic spicules of S. purpuratus (Mann et al., 2010) has converged on essentially the same suite of C-type lectin domain (CTLD)-containing proteins that has emerged from our analysis of the PMC transcriptome (Illies et al., 2002; Livingston et al., 2006) (this study). It is therefore likely that all of the canonical (CTLD-containing) spicule matrix proteins that are expressed in the embryo have now been identified (supplementary material Fig. S2). Another protein that has an essential function in biomineralization is the novel, PMC-specific transmembrane protein P16 (Cheers and Ettensohn, 2005). P16rel1 and P16rel2, which are similar in structure to P16 and are encoded by closely linked genes, probably play a similar role in biomineralization. A recent functional analysis of P58A and P58B, two related Type I transmembrane proteins that were identified in the present study, has shown that both proteins are required for biomineral formation but not for PMC specification, migration or fusion (Adomako-Ankomah and Ettensohn, 2011). The precise biochemical functions of the P16 and P58 proteins are unknown, although it was recently shown that P16 is phosphorylated and binds to hydroxyapatite (Alvares et al., 2009).

Fig. 6.

A summary of the regulatory inputs into downstream effector genes, based on QPCR and WMISH analysis ofalx1,ets1andtbrmorphant embryos at 28-30 hours of development. Note that the interactions that are shown here among the four regulatory genes at the top of the GRN were identified in previous studies and were confirmed in the course of this work.

Fig. 6.

A summary of the regulatory inputs into downstream effector genes, based on QPCR and WMISH analysis ofalx1,ets1andtbrmorphant embryos at 28-30 hours of development. Note that the interactions that are shown here among the four regulatory genes at the top of the GRN were identified in previous studies and were confirmed in the course of this work.

In addition to biomineralization-related proteins, we identified several other secreted or transmembrane proteins that are expressed by PMCs (e.g. P11, EGF-rich/P41, P133, Stomatin and Tsp). P11 and Tsp are both small, secreted proteins; P11 lacks identifiable motifs but Tsp contains a single thrombospondin type I repeat (TSR). In vertebrates, TSR domain-containing proteins, which are typically larger and have more complex domain organizations, are constituents of the ECM and function in cell adhesion, signaling and motility (Adams and Tucker, 2000). Stomatin is a membrane protein; vertebrate stomatins regulate the function of other membrane-associated proteins, including ion channels, by mechanisms that are poorly understood (Lapatsina et al., 2011). P133 is a single-pass transmembrane protein with multiple, interspersed EGF and Laminin G domains, an overall organization that is strikingly similar to that of members of the neurexin family of cell adhesion receptors (Lisé and El-Husseini, 2006). These features suggest that P133 might mediate adhesion between PMCs and substrate molecules in the basal lamina (Hodor et al., 2000), or PMC-PMC adhesion as part of the process of cell fusion.

Fig. 7.

A time-averaged overview of the PMC GRN. Interactions among intermediate regulatory genes are based primarily on data from Oliveri and Davidson (Oliveri and Davidson, 2008) and inputs into late morphoregulatory genes are based primarily on the data provided in the present study. See the Endomesoderm Gene Regulatory Network at the Sea Urchin Genome Project website (http://sugp.caltech.edu/endomes/) for additional references and for a dynamic representation of changes in the state of the PMC GRN through time, including the data that are presented in this paper. Note that only a small fraction of the regulatory interactions indicated by the arrows have been shown to be direct, by means of biochemical studies and/or cis-regulatory analysis. An input from Ets1 into cyp1 is shown, based on the data of Amore and Davidson (Amore and Davidson, 2006), although this interaction narrowly failed to meet the threshold that we set in our study. Dashed boxes indicate genes that are not expressed by PMCs, owing to repressive inputs.

Fig. 7.

A time-averaged overview of the PMC GRN. Interactions among intermediate regulatory genes are based primarily on data from Oliveri and Davidson (Oliveri and Davidson, 2008) and inputs into late morphoregulatory genes are based primarily on the data provided in the present study. See the Endomesoderm Gene Regulatory Network at the Sea Urchin Genome Project website (http://sugp.caltech.edu/endomes/) for additional references and for a dynamic representation of changes in the state of the PMC GRN through time, including the data that are presented in this paper. Note that only a small fraction of the regulatory interactions indicated by the arrows have been shown to be direct, by means of biochemical studies and/or cis-regulatory analysis. An input from Ets1 into cyp1 is shown, based on the data of Amore and Davidson (Amore and Davidson, 2006), although this interaction narrowly failed to meet the threshold that we set in our study. Dashed boxes indicate genes that are not expressed by PMCs, owing to repressive inputs.

Our analysis has also identified several PMC-enriched mRNAs that encode cytoskeletal proteins. Almost all of these mRNAs are co-expressed by PMCs and non-skeletogenic mesoderm (NSM) cells, two kinds of mesenchymal cells that exhibit similar kinds of invasive, motile behaviors (see below). Many of these proteins have known biochemical functions and act cooperatively in cells. For example, we identified four components or regulators of the Arp2/3 complex (Arp3 and p21arc/ARPC3, Cdc42 and Wasp) in the set of PMC-enriched mRNAs. Given the well-known role of the Arp2/3 complex in regulating actin dynamics and cell protrusive activity, including the formation of filopodia, it seems likely that this collection of proteins plays a role in regulating the filopodial motility of sea urchin mesenchyme cells (Pollard, 2007; Faix et al., 2009; Mellor, 2010). Cdc42 has been specifically implicated in the formation of filopodia in several cell types (Ahmed et al., 2010). We also identified Lasp1 as a PMC-specific transcript. In mammalian cells, this actin-binding phosphoprotein is enriched in cell protrusions, and Lasp1 has recently been shown to mediate the directional migration of leucocytes by binding to the chemotactic receptor CXCR2 (Raman et al., 2010). Lasp1 is therefore a strong candidate for a motility-related protein in PMCs.

In other experimental systems, emphasis has been placed on the post-translational regulation of cytoskeletal machinery in regulating early morphogenetic processes. For example, studies in Drosophila indicate that spatially controlled, zygotic transcriptional activation of a small number of key proteins that regulate Rho- or G protein-mediated signaling results in the local activation of ubiquitous cytoskeletal machinery, thereby leading to region-specific cell shape changes and cell movements during gastrulation (Dawes-Hoang et al., 2005; Kölsch et al., 2007). Although local, post-translational regulation of the cytoskeleton undoubtedly also contributes to region-specific cell behaviors during sea urchin gastrulation, we also find evidence of a widespread transcriptional upregulation of many cytoskeletal genes in PMCs. In other embryonic and metastatic cell types, invasive cell behavior has been shown to be under transcriptional control. It is of interest that jun and fos have been demonstrated to regulate invasiveness in several cellular contexts (Sherwood et al., 2005; Ozanne et al., 2007) and that both genes are selectively upregulated in PMCs.

Linking the PMC specification network to morphoregulatory genes

Our findings reveal several important features of the regulatory inputs that are provided by alx1, ets1 and tbr into the set of PMC-expressed morphoregulatory genes. (Note, however, that our gene knockdown analysis does not allow us to determine whether these inputs are direct or indirect.) alx1 and ets1 provide positive regulatory inputs into most (approximately two-thirds) of the morphoregulatory genes that we tested. By contrast, tbr provides inputs into a much smaller subset of genes (only lasp1, p16rel1, p16rel2 and sm49), even though tbr is activated early in development; i.e. several hours before PMC ingression and prior to the activation of most of the downstream genes that we analyzed here. The paucity of morphoregulatory genes that receive inputs from tbr is consistent with the finding that tbr also provides few inputs into regulatory genes in the GRN (Oliveri et al., 2008) (SpBase Endomesoderm Gene Regulatory Network). It is significant that tbr has undergone significant, recent evolutionary modifications with respect to its network connectivity (Hinman et al., 2007) and the lack of tbr expression in the skeletogenic centers of adult echinoids and sea stars suggests that this gene was recently co-opted into the micromere-PMC GRN of echinoids (Gao and Davidson, 2008). The limited connectivity of tbr in the PMC GRN probably reflects the recent recruitment of this gene into the network.

One of the most striking findings from our gene knockdown studies is that the targets of alx1 and ets1 are almost identical. The single exception that we identified is p58B, which receives a positive input from ets1 but not from alx1. Even in this case, the data are equivocal; our QPCR measurements suggest that alx1 knockdown might have a modest effect on p58B expression that did not meet the significance criteria that we established. This case aside, we note that all the biomineralization genes that were examined (spicule matrix genes, can1 and members of the p16 and p58 gene families) are co-regulated by ets1 and alx1. Parallel inputs from ets1 and alx1 also regulate vegfr-Ig-10, which encodes a receptor tyrosine kinase with a crucial role in PMC guidance and differentiation (Duloquin et al., 2007).

A variety of mechanisms might explain the parallel connectivity of ets1 and alx1. alx1 knockdown has little, if any, effect on ets1 expression (Ettensohn et al., 2003; Oliveri et al., 2008). By contrast, perturbation of ets1 expression or function, although it does not affect the early phase of alx1 expression, suppresses the later phase (Oliveri et al., 2008; Sharma and Ettensohn, 2010). It seems likely, therefore, that the parallel connectivity of alx1 and ets1 can be attributed partly to the influence of ets1 on alx1 expression. Other evidence indicates, however, that ets1 regulates terminal genes in the network by mechanisms that are independent of alx1 expression. For example, ets1 provides positive inputs into other regulatory genes, including erg and hex, that are required for the expression of differentiation genes but that do not provide regulatory inputs into alx1 (Oliveri et al., 2008) (SpBase Endomesoderm Gene Regulatory Network). Given the evidence that ets1 regulates terminal genes in the network by alx1-independent mechanisms, it is striking that we found no example of a gene that receives an essential input from ets1 but not from alx1, save the one questionable exception noted above. Oliveri and co-workers (Oliveri et al., 2008) proposed that ets1 might cooperate with various regulatory genes in feed-forward loops (A>B, B>C, A>C) that impinge on terminal genes in the network, and experimental evidence for a feed-forward loop that involves ets1 and alx1 has come from analysis of the cis-regulatory control of the cyp1 gene, which receives direct inputs from Dri (a target of alx1) and Ets1 (Amore and Davidson, 2006). If feed-forward regulatory interactions are a common feature of the wiring of the skeletogenic GRN, then one interpretation of our findings is that almost every morphoregulatory gene that receives an input from ets1 is wired with a feed-forward loop that includes alx1.

Although most morphoregulatory genes in the GRN are regulated by ets1 and alx1, six of the genes that we identified (casc1, cdi, cyp2, egf-rich/p41, fgfr2 and stomatin) do not appear to require inputs from any of the three transcription factors that we analyzed. One of the important morphoregulatory genes that is presently unconnected to the GRN is fgfr2, which functions in PMC migration and differentiation (Röttinger et al., 2008). Because the expression of every regulatory gene in the current model of the GRN, with the exception of tel, is influenced, directly or indirectly, by ets1, alx1 or tbr, our findings suggest that these morphoregulatory genes might receive inputs from as yet undiscovered subcircuits within the PMC GRN. These genes might receive inputs from tel, from one or more of the new PMC-enriched transcription factors that were identified in the present study (fos, alx4 and smad1/5/8) or from other regulatory genes in the network that have not yet been subjected to gene knockdown analysis, such as jun.

Regulation of the PMC GRN by extrinsic signals

Analysis of gene expression in cultures of micromeres or dissociated embryonic cells indicates that the initial deployment of the skeletogenic GRN in the micromere-PMC lineage is largely independent of signals from other cell types (Harkey and Whiteley, 1983; Stephens et al., 1989; Page and Benson, 1992). Much evidence, however, demonstrates that, later in development, skeletal growth and patterning is regulated by local, ectoderm-derived cues (Ettensohn, 1990; Kiyomoto and Tsukahara, 1991; Armstrong et al., 1993; Guss and Ettensohn, 1997; Duloquin et al., 2007; Röttinger et al., 2008). Many genes, including many of the downstream effector genes that were identified in the present study, show non-uniform patterns of expression within the PMC syncytium, with high levels of transcripts accumulating at sites of skeletal rod growth (Harkey et al., 1992; Guss et al., 1997). There is evidence from the expression of transgenes within the PMC syncytium that mRNAs remain localized predominantly in the PMC cell body in which they were synthesized (Harkey et al., 1995; Makabe et al., 1995; Wilt et al., 2008), a finding which suggests that local variations in mRNA abundance within the syncytium arise through the local regulation of gene transcription, rather than by an indirect mechanism (e.g. the diffusion and selective trapping of mRNAs). One model, which we favor, is that local, ectoderm-derived signals directly modulate the skeletogenic GRN and control the local availability of gene products (e.g. biomineralization proteins) that control the rate of biomineral deposition. The ectoderm has other important, and probably more subtle, influences on skeletal morphogenesis (for example, on the branching of skeletal rods), which remain mysterious. It will be important in the future to elucidate the various mechanisms by which the ectoderm influences skeletal growth and patterning.

Evolution of the PMC GRN

PMCs are a relatively recent evolutionary invention. The appearance of this cell population probably involved at least two steps: (1) the importation of a late larval/adult program of biomineralization into the late (gastrula stage) embryo, and (2) a second heterochronic shift, which shifted the skeletogenic program into the pre-gastrula stage embryo and which was associated with the formation of micromeres and an early-ingressing skeletogenic mesenchyme (Ettensohn, 2009). Our findings shed additional light on the mechanisms that might have accompanied the evolutionary origin of PMCs.

In euechinoid sea urchins (subclass Euechinoidea), a group that includes most modern sea urchins, several regulatory and signaling proteins that are expressed in the micromere-PMC lineage during early development (ets1, erg, foxN2/3, snail, wnt8 and delta) are also deployed in the adjacent NSM territory later in development, during gastrulation. In the present study, we have identified numerous non-regulatory genes (e.g. cytoskeletal proteins and secreted factors) that are expressed selectively in these two territories (Fig. 2). Thus, our current findings provide additional evidence of a striking similarity in the genomic regulatory states of PMCs and NSM cells. These observations support the view that the invention of the micromere-PMC lineage was associated with a heterochronic shift in the deployment of an ancestral GRN that was operative in the late-ingressing skeletogenic mesenchyme of ancestral echinoids. This mode of skeletogenesis is still seen in modern cidaroid urchins, which closely resemble the ancestral stock that gave rise to all extant sea urchins (Wray and McClay, 1989). It seems likely that one vestige of this heterochronic shift that has been retained in modern euechinoid sea urchins is the biphasic temporal expression (i.e. first in PMCs then in NSM cells) of many genes that are co-expressed in these two embryonic territories.

Gene duplication has played a major role in the evolution of the PMC GRN. Many of the genes that are expressed selectively in the micromere-PMC lineage are found in multiple, tandem copies. The stomatin, pmar1 and sm30 loci, for example, consist of five to ten tandem copies of these genes. All msp130 family members are found in small clusters of two to three genes, as are most of the spicule matrix genes (Livingston et al., 2006). The biomineralization gene p16 is adjacent to several related genes, at least one of which (p16rel2) also bears some similarity to spicule matrix genes. In addition, tsp, lasp1, p58 and alx1 are found as pairs of distinct but closely related genes. There is evidence that higher order clustering of these and other PMC-expressed genes is obscured by the incomplete nature of the current S. purpuratus genome assembly. For example, one spicule matrix gene that is expressed in the adult (GLEAN3_13825) (Mann et al., 2008) is closely linked to the msp130-msp130rel1-msp130rel3 gene cluster, indicating that at least some members of these two major gene families are clustered in the genome. These observations point to the active role of gene duplication in expanding the set of terminal morphoregulatory genes in the PMC GRN. Most of the duplicated genes contain introns (the stomatin genes are an exception) and are tandemly arranged, indicating that their duplication has occurred by DNA-mediated processes, such as unequal crossing over, rather than by retrotransposition.

Gene duplications are remarkably common events on an evolutionary time scale, but the usual fate of a duplicate gene pair is the silencing of one copy within a few million years (Hancock, 2005). Several of the gene duplications described above, however, occurred >100 million years ago and the duplicate copies have remained functional (Adomako-Ankomah and Ettensohn, 2011). An adaptive expansion of biomineralization-related genes might have occurred if natural selection favored an increase in their dosage or if a rapid ‘microfunctionalization’ of biomineralization proteins occurred (Hancock, 2005). Extensive duplication of genes that encode secreted, biomineralization proteins has occurred independently in vertebrates (Kawasaki et al., 2009). The duplication of other functional classes of genes might have facilitated the evolution of novel, cell type-specific protein functions or new gene expression patterns. In this regard, we note that the PMC-specific representatives of some protein families, such as the receptor tyrosine kinase family, have atypical structures that might be associated with unusual biochemical properties (Duloquin et al., 2007; Röttinger et al., 2008). Of particular interest is the duplication of transcription factor-encoding genes such as alx1 and pmar1, which might have allowed these genes to acquire new upstream regulators (through changes in cis-regulatory sequences) or new downstream targets (through changes in the biochemical properties of the proteins). Such gene duplication events might have been intimately involved in the heterochronic shifts in the deployment of the skeletogenic GRN that occurred during echinoderm evolution.

The authors thank Angela Bozak, Rachel Schuster and Zhongling Sun (Carnegie Mellon University) for assistance with WMISH analysis; and Melissa Landrum (National Center for Biotechnology Information) for mapping ESTs to the genome assembly.

Funding

This work was supported by National Science Foundation grants [IOS-0745875 and IOS-1021805].

Adams
J. C.
,
Tucker
R. P.
(
2000
).
The thrombospondin type 1 repeat (TSR) superfamily: diverse proteins with related roles in neuronal development
.
Dev. Dyn.
218
,
280
299
.
Adomako-Ankomah
A.
,
Ettensohn
C. A.
(
2011
).
P58-A and P58-B: Novel proteins that mediate skeletogenesis in the sea urchin embryo
.
Dev. Biol.
353
,
81
93
.
Ahmed
S.
,
Goh
W. I.
,
Bu
W.
(
2010
).
I-BAR domains, IRSp53 and filopodium formation
.
Semin. Cell Dev. Biol.
21
,
350
356
.
Akasaka
K.
,
Frudakis
T. N.
,
Killian
C. E.
,
George
N. C.
,
Yamasu
K.
,
Khaner
O.
,
Wilt
F. H.
(
1994
).
Genomic organization of a gene encoding the spicule matrix protein SM30 in the sea urchin Strongylocentrotus purpuratus
.
J. Biol. Chem.
269
,
20592
20598
.
Alvares
K.
,
Dixit
S. N.
,
Lux
E.
,
Veis
A.
(
2009
).
Echinoderm phosphorylated matrix proteins UTMP16 and UTMP19 have different functions in sea urchin tooth mineralization
.
J. Biol. Chem.
284
,
26149
26160
.
Amore
G.
,
Davidson
E. H.
(
2006
).
cis-Regulatory control of cyclophilin, a member of the ETS-DRI skeletogenic gene battery in the sea urchin embryo
.
Dev. Biol.
293
,
555
564
.
Amore
G.
,
Yavrouian
R. G.
,
Peterson
K. J.
,
Ransick
A.
,
McClay
D. R.
,
Davidson
E. H.
(
2003
).
Spdeadringer, a sea urchin embryo gene required separately in skeletogenic and oral ectoderm gene regulatory networks
.
Dev. Biol.
261
,
55
81
.
Angerer
L. M.
,
Chambers
S. A.
,
Yang
Q.
,
Venkatesan
M.
,
Angerer
R. C.
,
Simpson
R. T.
(
1988
).
Expression of a collagen gene in mesenchyme lineages of the Strongylocentrotus purpuratus embryo
.
Genes Dev.
2
,
239
246
.
Armstrong
N.
,
Hardin
J.
,
McClay
D. R.
(
1993
).
Cell-cell interactions regulate skeleton formation in the sea urchin embryo
.
Development
119
,
833
840
.
Benson
S.
,
Sucov
H.
,
Stephens
L.
,
Davidson
E.
,
Wilt
F.
(
1987
).
A lineage-specific gene encoding a major matrix protein of the sea urchin embryo spicule. I. Authentication of the cloned gene and its developmental expression
.
Dev. Biol.
120
,
499
506
.
Castoe
T. A.
,
Stephens
T.
,
Noonan
B. P.
,
Calestani
C.
(
2007
).
A novel group of type I polyketide synthases (PKS) in animals and the complex phylogenomics of PKSs
.
Gene
392
,
47
58
.
Cheers
M. S.
,
Ettensohn
C. A.
(
2004
).
Rapid microinjection of fertilized eggs
.
Methods Cell Biol.
74
,
287
310
.
Cheers
M. S.
,
Ettensohn
C. A.
(
2005
).
P16 is an essential regulator of skeletogenesis in the sea urchin embryo
.
Dev. Biol.
283
,
384
396
.
Davidson
E. H.
(
2010
).
Emerging properties of animal gene regulatory networks
.
Nature
468
,
911
920
.
Dawes-Hoang
R. E.
,
Parmar
K. M.
,
Christiansen
A. E.
,
Phelps
C. B.
,
Brand
A. H.
,
Wieschaus
E. F.
(
2005
).
Folded gastrulation, cell shape change and the control of myosin localization
.
Development
132
,
4165
4178
.
Duloquin
L.
,
Lhomond
G.
,
Gache
C.
(
2007
).
Localized VEGF signaling from ectoderm to mesenchyme cells controls morphogenesis of the sea urchin embryo skeleton
.
Development
134
,
2293
2302
.
Ettensohn
C. A.
(
1990
).
The regulation of primary mesenchyme cell patterning
.
Dev. Biol.
140
,
261
271
.
Ettensohn
C. A.
(
2009
).
Lessons from a gene regulatory network: echinoderm skeletogenesis provides insights into evolution, plasticity and morphogenesis
.
Development
136
,
11
21
.
Ettensohn
C. A.
,
Malinda
K. M.
(
1993
).
Size regulation and morphogenesis: a cellular analysis of skeletogenesis in the sea urchin embryo
.
Development
119
,
155
167
.
Ettensohn
C. A.
,
Illies
M. R.
,
Oliveri
P.
,
De Jong
D. L.
(
2003
).
Alx1, a member of the Cart1/Alx3/Alx4 subfamily of Paired-class homeodomain proteins, is an essential component of the gene network controlling skeletogenic fate specification in the sea urchin embryo
.
Development
130
,
2917
2928
.
Faix
J
,
Breitsprecher
D.
,
Stradal
T. E.
,
Rottner
K.
(
2009
).
Filopodia: Complex models for simple rods
.
Int. J. Biochem. Cell Biol.
41
,
1656
1664
.
Fuchikami
T.
,
Mitsunaga-Nakatsubo
K.
,
Amemiya
S.
,
Hosomi
T.
,
Watanabe
T.
,
Kurokawa
D.
,
Kataoka
M.
,
Harada
Y.
,
Satoh
N.
,
Kusunoki
S.
, et al. 
. (
2002
).
T-brain homologue (HpTb) is involved in the archenteron induction signals of micromere descendant cells in the sea urchin embryo
.
Development
129
,
5205
5216
.
Gao
F.
,
Davidson
E. H.
(
2008
).
Transfer of a large gene regulatory apparatus to a new developmental address in echinoid evolution
.
Proc. Natl. Acad. Sci. USA
105
,
6091
6096
.
George
N. C.
,
Killian
C. E.
,
Wilt
F. H.
(
1991
).
Characterization and expression of a gene encoding a 30.6-kDa Strongylocentrotus purpuratus spicule matrix protein
.
Dev. Biol.
147
,
334
342
.
Guss
K. A.
,
Ettensohn
C. A.
(
1997
).
Skeletal morphogenesis in the sea urchin embryo: regulation of primary mesenchyme gene expression and skeletal rod growth by ectoderm-derived cues
.
Development
124
,
1899
1908
.
Gustafson
T.
,
Wolpert
L.
(
1967
).
Cellular movement and contact in sea urchin morphogenesis
.
Biol. Rev.
42
,
442
498
.
Hancock
J. M.
(
2005
).
Gene factories, microfunctionalization and the evolution of gene families
.
Trends Genet.
21
,
591
595
.
Harkey
M. A.
,
Whiteley
A. H.
(
1983
).
The program of protein synthesis during the development of the micromere-primary mesenchyme cell line in the sea urchin embryo
.
Dev. Biol.
100
,
12
28
.
Harkey
M. A.
,
Whiteley
H. R.
,
Whiteley
A. H.
(
1992
).
Differential expression of the msp130 gene among skeletal lineage cells in the sea urchin embryo: a three-dimensional in situ hybridization analysis
.
Mech. Dev.
37
,
173
184
.
Harkey
M. A.
,
Klueg
K.
,
Sheppard
P.
,
Raff
R. A.
(
1995
).
Structure, expression, and extracellular targeting of PM27, a skeletal protein associated specifically with growth of the sea urchin larval spicule
.
Dev. Biol.
168
,
549
566
.
Hart
M. W.
,
Strathmann
R. R.
(
1994
).
Functional consequences of phenotypic plasticity in echinoid larvae
.
Biol. Bull.
186
,
291
299
.
Hinman
V. F.
,
Nguyen
A.
,
Davidson
E. H.
(
2007
).
Caught in the evolutionary act: precise cis-regulatory basis of difference in the organization of gene networks of sea stars and sea urchins
.
Dev. Biol.
312
,
584
595
.
Hodor
P. G.
,
Ettensohn
C. A.
(
2008
).
Mesenchymal cell fusion in the sea urchin embryo
.
Methods Mol. Biol.
475
,
315
334
.
Hodor
P. G.
,
Illies
M. R.
,
Broadley
S.
,
Ettensohn
C. A.
(
2000
).
Cell-substrate interactions during sea urchin gastrulation: migrating primary mesenchyme cells interact with and align extracellular matrix fibers that contain ECM3, a molecule with NG2-like and multiple calcium-binding domains
.
Dev. Biol.
222
,
181
194
.
Honoré
B.
(
2009
).
The rapidly expanding CREC protein family: members, localization, function, and role in disease
.
BioEssays
31
,
262
277
.
Howard-Ashby
M.
,
Materna
S. C.
,
Brown
C. T.
,
Chen
L.
,
Cameron
R. A.
,
Davidson
E. H.
(
2006a
).
Identification and characterization of homeobox transcription factor genes in Strongylocentrotus purpuratus, and their expression in embryonic development
.
Dev. Biol.
300
,
74
89
.
Howard-Ashby
M.
,
Materna
S. C.
,
Brown
C. T.
,
Chen
L.
,
Cameron
R. A.
,
Davidson
E. H.
(
2006b
).
Gene families encoding transcription factors expressed in early development of Strongylocentrotus purpuratus
.
Dev. Biol.
300
,
90
107
.
Illies
M. R.
,
Peeler
M. T.
,
Dechtiaruk
A. M.
,
Ettensohn
C. A.
(
2002
).
Identification and developmental expression of new biomineralization proteins in the sea urchin Strongylocentrotus purpuratus
.
Dev. Genes Evol.
212
,
419
431
.
Kadurin
I.
,
Huber
S.
,
Gründer
S.
(
2009
).
A single conserved proline residue determines the membrane topology of stomatin
.
Biochem. J.
418
,
587
594
.
Kawasaki
K
,
Buchanan
A. V.
,
Weiss
K. M.
(
2009
).
Biomineralization in humans: making the hard choices in life
.
Annu. Rev. Genet.
43
,
119
142
.
Killian
C. E.
,
Croker
L.
,
Wilt
F. H.
(
2010
).
SpSM30 gene family expression patterns in embryonic and adult biomineralized tissues of the sea urchin, Strongylocentrotus purpuratus
.
Gene Expr. Patterns
10
,
135
139
.
Kiyomoto
M.
,
Tsukahara
J.
(
1991
).
Spicule formation-inducing substance in the sea urchin embryo
.
Dev. Growth Differ.
33
,
443
450
.
Kölsch
V.
,
Seher
T.
,
Fernandez-Ballester
G. J.
,
Serrano
L.
,
Leptin
M.
(
2007
).
Control of Drosophila gastrulation by apical localization of adherens junctions and RhoGEF2
.
Science
315
,
384
386
.
Kurokawa
D.
,
Kitajima
T.
,
Mitsunaga-Nakatsubo
K.
,
Amemiya
S.
,
Shimada
H.
,
Akasaka
K.
(
1999
).
HpEts, an ets-related transcription factor implicated in primary mesenchyme cell differentiation in the sea urchin embryo
.
Mech. Dev.
80
,
41
52
.
Lapatsina
L.
,
Brand
J.
,
Poole
K.
,
Daumke
O.
,
Lewin
G. R.
(
2011
).
Stomatin-domain proteins
.
Eur. J. Cell Biol.
doi:10.1016/j.ejcb.2011.01.018
.
Leaf
D. S.
,
Anstrom
J. A.
,
Chin
J. E.
,
Harkey
M. A.
,
Showman
R. M.
,
Raff
R. A.
(
1987
).
Antibodies to a fusion protein identify a cDNA clone encoding msp130, a primary mesenchyme-specific cell surface protein of the sea urchin embryo
.
Dev. Biol.
121
,
29
40
.
Lee
Y. H.
,
Britten
R. J.
,
Davidson
E. H.
(
1999
).
SM37, a skeletogenic gene of the sea urchin embryo linked to the SM50 gene
.
Dev. Growth Differ.
41
,
303
312
.
Lisé
M. F.
,
El-Husseini
A.
(
2006
).
The neuroligin and neurexin families: from structure to function at the synapse
.
Cell Mol. Life Sci.
63
,
1833
1849
.
Livingston
B. T.
,
Killian
C. E.
,
Wilt
F.
,
Cameron
A.
,
Landrum
M. J.
,
Ermolaeva
O.
,
Sapojnikov
V.
,
Maglott
D. R.
,
Buchanan
A. M.
,
Ettensohn
C. A.
(
2006
).
A genome-wide analysis of biomineralization-related proteins in the sea urchin Strongylocentrotus purpuratus
.
Dev. Biol.
300
,
335
348
.
Love
A. C.
,
Andrews
M. E.
,
Raff
R. A.
(
2007
).
Gene expression patterns in a novel animal appendage: the sea urchin pluteus arm
.
Evol. Dev.
9
,
51
68
.
Makabe
K. W.
,
Kirchhamer
C. V.
,
Britten
R. J.
,
Davidson
E. H.
(
1995
).
Cis-regulatory control of the SM50 gene, an early marker of skeletogenic lineage specification in the sea urchin embryo
.
Development
121
,
1957
1970
.
Malinda
K. M.
,
Fisher
G. W.
,
Ettensohn
C. A.
(
1995
).
Four-dimensional microscopic analysis of the filopodial behavior of primary mesenchyme cells during gastrulation in the sea urchin embryo
.
Dev. Biol.
172
,
552
566
.
Mann
K.
,
Poustka
A. J.
,
Mann
M.
(
2008
).
The sea urchin (Strongylocentrotus purpuratus) test and spine proteomes
.
Proteome Sci.
6
,
22
.
Mann
K.
,
Wilt
F. H.
,
Poustka
A. J.
(
2010
).
Proteomic analysis of sea urchin (Strongylocentrotus purpuratus) spicule matrix
.
Proteome Sci.
8
,
33
.
Mellor
H.
(
2010
).
The role of formins in filopodia formation
.
Biochim. Biophys. Acta
1803
,
191
200
.
Miller
J.
,
Fraser
S. E.
,
McClay
D.
(
1995
).
Dynamics of thin filopodia during sea urchin gastrulation
.
Development
121
,
2501
2511
.
Minokawa
T.
,
Rast
J. P.
,
Arenas-Mena
C.
,
Franco
C. B.
,
Davidson
E. H.
(
2004
).
Expression patterns of four different regulatory genes that function during sea urchin development
.
Gene Expr. Patterns
4
,
449
456
.
Nikitina
N.
,
Sauka-Spengler
T.
,
Bronner-Fraser
M.
(
2009
).
Chapter 1. Gene regulatory networks in neural crest development and evolution
.
Curr. Top. Dev. Biol.
86
,
1
14
.
Oliveri
P.
,
Carrick
D. M.
,
Davidson
E. H.
(
2002
).
A regulatory gene network that directs micromere specification in the sea urchin embryo
.
Dev. Biol.
246
,
209
228
.
Oliveri
P.
,
Tu
Q.
,
Davidson
E. H.
(
2008
).
Global regulatory logic for specification of an embryonic cell lineage
.
Proc. Natl. Acad. Sci. USA
105
,
5955
5962
.
Ozanne
B. W.
,
Spence
H. J.
,
McGarry
L. C.
,
Hennigan
R. F.
(
2007
).
Transcription factors control invasion: AP-1 the first among equals
.
Oncogene
26
,
1
10
.
Page
L.
,
Benson
S.
(
1992
).
Analysis of competence in cultured sea urchin micromeres
.
Exp. Cell Res.
203
,
305
311
.
Pennington
J. T.
,
Strathmann
R. R.
(
1990
).
Consequences of the calcite skeletons of planktonic echinoderm larvae for orientation, swimming, and shape
.
Biol. Bull.
179
,
121
133
.
Peter
I. S.
,
Davidson
E. H.
(
2009
).
Modularity and design principles in the sea urchin embryo gene regulatory network
.
FEBS Lett.
583
,
3948
3958
.
Peterson
R. E.
,
McClay
D. R.
(
2003
).
Primary mesenchyme cell patterning during the early stages following ingression
.
Dev. Biol.
254
,
68
78
.
Pollard
T. D.
(
2007
).
Regulation of actin filament assembly by Arp2/3 complex and formins
.
Annu. Rev. Biophys. Biomol. Struct.
36
,
451
477
.
Raman
D.
,
Sai
J.
,
Neel
N. F.
,
Chew
C. S.
,
Richmond
A.
(
2010
).
LIM and SH3 protein-1 modulates CXCR2-mediated cell migration
.
PLoS ONE
5
,
e10050
.
Rizzo
F.
,
Fernandez-Serra
M.
,
Squarzoni
P.
,
Archimandritis
A.
,
Arnone
M. I.
(
2006
).
Identification and developmental expression of the ets gene family in the sea urchin (Strongylocentrotus purpuratus)
.
Dev. Biol.
300
,
35
48
.
Röttinger
E.
,
Saudemont
A.
,
Duboc
V.
,
Besnardeau
L.
,
McClay
D.
,
Lepage
T.
(
2008
).
FGF signals guide migration of mesenchymal cells, control skeletal morphogenesis and regulate gastrulation during sea urchin development
.
Development
135
,
353
365
.
Samanta
M. P.
,
Tongprasit
W.
,
Istrail
S.
,
Cameron
R. A.
,
Tu
Q.
,
Davidson
E. H.
,
Stolc
V.
(
2006
).
The transcriptome of the sea urchin embryo
.
Science
314
,
960
962
.
Schroer
T. A.
(
2004
).
Dynactin
.
Annu. Rev. Cell Dev. Biol.
20
,
759
779
.
Sea Urchin Genome Sequencing Consortium et al
. (
2006
).
The genome of the sea urchin Strongylocentrotus purpuratus
.
Science
314
,
941
952
.
Sharma
T.
,
Ettensohn
C. A.
(
2010
).
Activation of the skeletogenic gene regulatory network in the early sea urchin embryo
.
Development
137
,
1149
1157
.
Sherwood
D. R.
(
2006
).
Cell invasion through basement membranes: an anchor of understanding
.
Trends Cell Biol.
16
,
250
256
.
Sherwood
D. R.
,
Butler
J. A.
,
Kramer
J. M.
,
Sternberg
P. W.
(
2005
).
FOS-1 promotes basement-membrane removal during anchor-cell invasion in C. elegans
.
Cell
121
,
951
962
.
Smith
L. C.
,
Harrington
M. G.
,
Britten
R. J.
,
Davidson
E. H.
(
1994
).
The sea urchin profilin gene is specifically expressed in mesenchyme cells during gastrulation
.
Dev. Biol.
164
,
463
474
.
Stamateris
R. E.
,
Rafiq
K.
,
Ettensohn
C. A.
(
2010
).
The expression and distribution of Wnt and Wnt receptor mRNAs during early sea urchin development
.
Gene Expr. Patterns
10
,
60
64
.
Stathopoulos
A.
,
Levine
M.
(
2005
).
Genomic regulatory networks and animal development
.
Dev. Cell
9
,
449
462
.
Stephens
L.
,
Kitajima
T.
,
Wilt
F.
(
1989
).
Autonomous expression of tissue-specific genes in dissociated sea urchin embryos
.
Development
107
,
299
307
.
Suzuki
H. R.
,
Reiter
R. S.
,
D’Alessio
M.
,
Di Liberto
M.
,
Ramirez
F.
,
Exposito
J. Y.
,
Gambino
R.
,
Solursh
M.
(
1997
).
Comparative analysis of fibrillar and basement membrane collagen expression in embryos of the sea urchin, Strongylocentrotus purpuratus
.
Zoolog. Sci.
14
,
449
454
.
Sweet
H. C.
,
Gehring
M.
,
Ettensohn
C. A.
(
2002
).
LvDelta is a mesoderm-inducing signal in the sea urchin embryo and can endow blastomeres with organizer-like properties
.
Development
129
,
1945
1955
.
Tu
Q.
,
Brown
C. T.
,
Davidson
E. H.
,
Oliveri
P.
(
2006
).
Sea urchin Forkhead gene family: phylogeny and embryonic expression
.
Dev. Biol.
300
,
49
62
.
Wang
D.
,
Britten
R. J.
,
Davidson
E. H.
(
1995
).
Maternal and embryonic provenance of a sea urchin embryo transcription factor, SpZ12-1
.
Mol. Mar. Biol. Biotech.
4
,
148
153
.
Wessel
G. M.
,
Chen
S. W.
(
1993
).
Transient, localized accumulation of alpha-spectrin during sea urchin morphogenesis
.
Dev. Biol.
155
,
161
171
.
Whittaker
C. A.
,
Bergeron
K. F.
,
Whittle
J.
,
Brandhorst
B. P.
,
Burke
R. D.
,
Hynes
R. O.
(
2006
).
The echinoderm adhesome
.
Dev Biol.
300
,
252
266
.
Wikramanayake
A. H.
,
Peterson
R.
,
Chen
J.
,
Huang
L.
,
Bince
J. M.
,
McClay
D. R.
,
Klein
W. H.
(
2004
).
Nuclear beta-catenin-dependent Wnt8 signaling in vegetal cells of the early sea urchin embryo regulates gastrulation and differentiation of endoderm and mesodermal cell lineages
.
Genesis
39
,
194
205
.
Wilt
F. H.
,
Ettensohn
C. A.
(
2007
).
The morphogenesis and biomineralization of the sea urchin larval skeleton
. In
Handbook of Biomineralization: Biological Aspects and Structure Formation
(ed.
Bauerlein
E.
), pp.
183
210
.
Weinheim
:
Wiley-VCH
.
Wilt
F. H.
,
Killian
C. E.
,
Hamilton
P.
,
Croker
L.
(
2008
).
Dynamics of secretion during sea urchin embryonic skeleton formation
.
Exp. Cell Res.
314
,
1744
1752
.
Wirschell
M.
,
Yang
C.
,
Yang
P.
,
Fox
L.
,
Yanagisawa
H. A.
,
Kamiya
R.
,
Witman
G. B.
,
Porter
M. E.
,
Sale
W. S.
(
2009
).
IC97 is a novel intermediate chain of I1 dynein that interacts with tubulin and regulates interdoublet sliding
.
Mol. Biol. Cell
20
,
3044
3054
.
Wray
G. A.
,
McClay
D. R.
(
1989
).
Molecular heterochronies and heterotopies in early echinoid development
.
Evolution
43
,
803
813
.
Wu
S. Y.
,
McClay
D. R.
(
2007
).
The Snail repressor is required for PMC ingression in the sea urchin embryo
.
Development
134
,
1061
1070
.
Zhou
J.
,
Wang
Q.
,
Chen
L. L.
,
Carmichael
G. G.
(
2008
).
On the mechanism of induction of heterochromatin by the RNA-binding protein vigilin
.
RNA
14
,
1773
1781
.
Zhu
X.
,
Mahairas
G.
,
Illies
M.
,
Cameron
R. A.
,
Davidson
E. H.
,
Ettensohn
C. A.
(
2001
).
A large-scale analysis of mRNAs expressed by primary mesenchyme cells of the sea urchin embryo
.
Development
128
,
2615
2627
.

Competing interests statement

The authors declare no competing financial interests.

Supplementary information