Planarians are an ideal model system to study in vivo the dynamics of adult pluripotent stem cells. However, our knowledge of the factors necessary for regulating the ‘stemness’ of the neoblasts, the adult stem cells of planarians, is sparse. Here, we report on the characterization of the first planarian member of the LSm protein superfamily, Smed-SmB, which is expressed in stem cells and neurons in Schmidtea mediterranea. LSm proteins are highly conserved key players of the splicing machinery. Our study shows that Smed-SmB protein, which is localized in the nucleus and the chromatoid body of stem cells, is required to safeguard the proliferative ability of the neoblasts. The chromatoid body, a cytoplasmatic ribonucleoprotein complex, is an essential regulator of the RNA metabolism required for the maintenance of metazoan germ cells. However, planarian neoblasts and neurons also rely on its functions. Remarkably, Smed-SmB dsRNA-mediated knockdown results in a rapid loss of organization of the chromatoid body, an impairment of the ability to post-transcriptionally process the transcripts of Smed-CycB, and a severe proliferative failure of the neoblasts. This chain of events leads to a quick depletion of the neoblast pool, resulting in a lethal phenotype for both regenerating and intact animals. In summary, our results suggest that Smed-SmB is an essential component of the chromatoid body, crucial to ensure a proper RNA metabolism and essential for stem cell proliferation.

Planarians (phylum Platyhelminthes, class Turbellaria) are bilaterian, triploblastic free-living flatworms best known for their impressive ability to regenerate lost body parts (Morgan, 1898; Saló et al., 2009; Forsthoefel, 2009). This extreme plasticity is due to the presence of a large population (20-30% of the total number of cells) (Hay and Coward, 1975; Baguñà, 1976; Baguñà and Romero, 1981; Newmark and Sánchez Alvarado, 2000; Hayashi et al., 2006) of somatic pluripotent stem cells known as neoblasts (Dubois, 1949; Wolff, 1962). Apart from germ cells, neoblasts are the only cells capable of proliferating in adult planaria (Morita and Best, 1984; Newmark and Sánchez-Alvarado, 2000). They reside in the thickness of the parenchyma (Baguñá, 1973; Newmark and Sánchez Alvarado, 2000; Orii et al., 2005) and are distributed, with few exceptions, throughout the entire body of the animal. Neoblasts maintain tissue homeostasis (Baguñà et al., 1989), generate the germline (Sato et al., 2006) and replace structures lost to damage or amputation through a complex process of pattern re-establishment (re-patterning) (Saló and Baguñà, 1984; Baguñà et al., 1989; Saló, 2006), which begins with the formation of an undifferentiated, unpigmented tissue called blastema (Dubois, 1949; Morita and Best, 1984; Saló and Baguñà, 1984; Newmark and Sánchez Alvarado, 2000). The blastema, however, does not contain proliferating neoblasts. Instead, they are enriched in the region adjacent to the blastema, called post-blastema (PB) (Saló and Baguñà, 1984; Eisenhoffer et al., 2008).

Neoblasts are actively recruited to the PB through signals elicited by the wound response and probably through other subsequent signalling (apoptosis, Wnt-P1) (Pellettieri et al., 2010; Petersen and Reddien, 2009). In the PB, neoblasts actively proliferate to regenerate the lost body part, mainly in two temporally defined waves: the minor one after eight hours, the major one between 48 and 72 hours (Saló and Baguñà, 1984). A similar proliferative response was also found following feeding, especially after a period of starvation (Baguñá, 1976; Newmark and Sánchez Alvarado, 2000). The molecular mechanisms underlying the regulation of neoblast proliferation and differentiation remain, however, largely unknown.

The LSm (like-Sm) RNA-binding protein superfamily is a set of genes involved in many aspects of the RNA metabolism (for a review, see Lührmann et al., 1990). LSm protein products are highly conserved throughout metazoans and relatively well-conserved in bacteria and archaea (Schumacher et al., 2002). LSm proteins are characterized by the presence of the Sm fold, a chaperone-like domain that allows a variety of RNA-RNA and RNA-protein interactions (for a review, see Wilusz and Wilusz, 2005). Complexes of six or seven proteins (the LSm ring) (Khusial et al., 2005) interact with different small nuclear RNAs (snRNAs) to generate small nuclear ribonucleoproteins (snRNPs). LSm proteins take part in several processes such as Cajal body formation, telomere elongation (Fu and Collins, 2006) and ribosomal assembling (Kiss, 2004). LSm proteins are, however, best known for their role in mRNA processing (He and Parker, 2000). They are also involved in the organization of the major and minor spliceosome complexes, which are responsible for the splicing processes of the pre-mRNA (Mayes et al., 1999; Patel et al., 2002). Although these complexes are mainly found in the nucleus, virtually all metazoan germ cells display ribonucleoprotein (RNP) granules in the cytoplasm (Eddy, 1975), like the germ granules in Caenorhabditis elegans and Drosophila melanogaster (reviewed in Strome and Lehmann, 2007) and the nuages or chromatoid bodies (CBs) in mammalian spermatids (reviewed in Parvinen, 2005).

A recent report showed that mammalian cell lines can also be induced to form CB-like structures following the expression of a cytoplasmic form of the prion protein PrP (Beaudoin et al., 2009). Two other studies proposed that LSm proteins regulate cell proliferation in the cytoplasm of mammalian cells through the minor spliceosome (König et al., 2007) and, if perturbed, can interfere with the localization and subcellular distribution of the P-granules of C. elegans (Barbee et al., 2002). Planarian CBs are electron-dense RNP particles characteristic for stem cells (Hay and Coward, 1975) and neurons (Yoshida-Kashikawa et al., 2007). Until now, only two proteins have been described as components of the planarian CB: DjCBC-1 (Yoshida-Kashikawa et al., 2007) and Spoultd-1 (Solana et al., 2009). Their function, not yet fully understood, is related to RNA processing and post-transcriptional modifications (Yoshida-Kashikawa et al., 2007).

In this study, we characterize the first member of the large LSm protein superfamily, Smed-SmB, described in the planarian species Schmidtea mediterranea. Smed-SmB is the ortholog of mammalian SmB/B′/N, and is expressed by stem cells and neurons. Double-stranded RNA (dsRNA)-mediated silencing in intact and regenerating animals resulted in a quick reduction in the number of dividing cells, which compromised both the ability to ensure normal tissue turnover and regeneration after amputation. A massive degenerative process progressed from anterior to posterior in both regenerating and intact animals and caused the animals to die within two to four weeks, respectively.

Our data show that inhibition of Smed-SmB impairs the organization and function of the CB, to which the protein also localizes. Furthermore, Smed-SmB RNAi affects the splicing efficiency of Smed-CycB, which indicates a failure to process pre-mRNAs necessary for cell cycle progression. The degeneration of the CBs and the impairment of the splicing efficiency seem to have a central role in the resulting proliferative failure we observed.

Animals

All planarians used in the present study were generated from a single individual of the asexual strain of Schmidtea mediterranea (clonal line BCN-10), collected from a fountain in Montjuic (Barcelona, Spain). Animals were maintained in a 1:1 mixture of tap water treated with AquaSafe (Tetra Aqua) and MilliQ water (Millipore Iberica, Madrid, Spain) and fed twice a week with bovine liver, as previously described (Molina et al., 2007). Animals used for all the presented experiments were starved for one week. To prepare irradiated controls, planarians were γ-irradiated at 75 Gy (1.56 Gy/minute) with a Gammacell 1000 (Atomic Energy of Canada Limited) (Saló and Baguñà, 1985). The animals were collected at different time points (at 12 hours, daily from days 1-11 and at days 14, 15 and 21) according to downstream applications.

Identification and cloning of Smed-SmB

The peptide sequence of Smed-SmB was identified among putative neoblast-specific proteins (E.F.-T., G. Rodriguez-Esteban, E.S. and J. Abril, unpublished data), and confirmed by BLAST in a local database generated from the S. mediterranea genome traces (sequenced by Washington University Genome Sequencing Center, USA). Primers were designed to amplify a 237 bp fragment by RT-PCR (forward primer, 5′-CAACACTTCAAGATGGTCG; reverse primer, 5′-CCATTAACGCCTACTGAA) from total RNA preparation (TRIzol). 5′- and 3′-ends were obtained by RNA ligase-mediated rapid amplification of cDNA ends (RLM-RACE)-PCR (5′ RACE primer, 5′-GACUGGAGCAGGAGGACACUGACAUGGAGUGAAGGAGUAGAAA; 3′ RACE primer, 5′-GCTGTCAACGATAGGCTACGTAACGGCATGACAGTG[T]18). The complete sequence was confirmed by BLAST to the planarian genome draft (accession number: GU562964).

Cell dissociation and FACS

Planarian cell dissociation protocol was modified from González-Estévez (González-Estévez et al., 2007). Briefly, animals were treated with 2% L-cysteine hydrochloride monohydrate (Merck, K23484539), pH 7.0, to remove the mucus, transferred in CMF buffer and mechanically dissociated by gentle rocking. The suspension was then serially filtered through 70 and 30 μm nylon meshes, stained for 2 hours with a mixture of 7.5 μg/ml Hoechst 33342 (Sigma, B2261) and 0.05 μg/ml Calcein AM (Invitrogen, Karlsruhe, Germany), pelleted and resuspended for FACS (FACSARIA, BD, Heidelberg, Germany). Eventually, propidium iodide was added at concentration of 1 μg/ml.

In situ hybridization

Wholemount in situ hybridization (ISH) protocol was modified from Agata (Agata et al., 1998). Briefly, after fixation and rehydration, planarians were permeabilized in 0.1% Triton X-100 in PBS (PBST×0.1) added with 20 μg/ml proteinase K and quenched with 2% glycine. Triethanolamine (Sigma, T9534) treatment was performed as previously described (Nogi and Levin, 2005) before hybridization, which was carried out at 55°C for 18 to 30 hours, using digoxigenin-labelled RNA probes prepared using an in vitro labelling kit (Roche, Mannheim, Germany) at a final concentration of 0.07 ng/μl. ISH on dissociated cells was performed as previously described (González-Estévez et al., 2003).

Immunohistochemistry

Wholemount immunohistochemistry (IHC) was performed as described (Cebriá and Newmark, 2005). Briefly, after fixation in Carnoy and rehydration, animals were blocked in 1% BSA in PBST×0.1 for 2 hours at room temperature. The antibody against proliferating cell nuclear antigen (α-PCNA), kindly provided by Dr Hidefumi Orii (Himeji Institute of Technology, Hyogo, Japan), was used 1:1000 in blocking solution for 20 hours under constant agitation at 4°C. Animals were then labelled for 14-16 hours with FITC-conjugated donkey anti-rabbit (Jackson ImmunoResearch, 1:200). PCNA-positive cells were counted using a Leica SP2 confocal laser-scanning microscope (Leica Microsystems, Barcelona, Spain).

RNA interference (RNAi)

In vitro dsRNA preparation of Smed-SmB 5′-region (500 bp) and injection procedure were previously described (Oviedo and Levin, 2007). A total of three injections, distributed over 3 consecutive days (one injection per day with three pulses of 32 nl of 400 ng/μl dsRNA solution) were performed using a Drummond Scientific Nanoject injector (Broomall, PA, USA) targeting the gastrovascular system. The control group was injected with water. Some animals were pre-pharyngeally amputated 1 or 3 days after the last injection (t-cut).

RT-PCR and quantitative real-time RT-PCR (qRT-PCR)

Total RNA was extracted using TRIzol (Invitrogen, Barcelona, Spain) and reverse-transcribed to cDNA with the High-Capacity cDNA Reverse Transcription Kit (AppliedBiosystems, Darmstadt, Germany). Transcript levels were determined on the ABI PRISM SDS 7900HT (AppliedBiosystems, Darmstadt, Germany) using custom-designed oligonucleotides for the 5′-nuclease assays (see Table S1 in the supplementary material) and normalized to the endogenous control SmEf2 that was chosen among other ubiquitously expressed genes for its stable expression throughout the regenerative process, as well as for not being affected by the RNAi treatment performed (see Fig. S1, Fig. S2 in the supplementary material). Relative quantification of gene expression was calculated using the ΔΔCt method. Three technical replicates were used for each real-time PCR reaction; a reverse transcriptase blank and a no-template blank served as negative controls.

For Smed-CycB RT-PCR of spliced and unspliced forms, the equivalent of 5 ng of reverse-transcribed RNA were amplified using oligonucleotides spanning the second intron (forward primer, 5′-ATGCCGCCGAAACTTTATACCTG; reverse primer, 5′-AACTTCTTCGACTTTTGCTGCAA), according to mk4.001494.00.01. Expected amplicon sizes for the spliced and unspliced forms of the transcript were 136 bp and 219 bp, respectively.

Electron microscopy (EM)

Post-blastema fragments of Smed-SmB dsRNA- and water-injected planarians were fixed in 2.5% glutaraldehyde in 0.1 M cacodylate buffer, pH 7.4, for 2 hours at room temperature. Specimens were post-fixed in 1% OsO4 and embedded after dehydration in epon. Ultrathin sections of 70 nm were cut (Leica-UC6 ultramicrotome, Vienna, Austria) and counterstained with uranyl acetate and lead citrate. Specimens were observed at 80 kV on a FEI-Tecnai 12 electron microscope (FEI, Eindhoven, Netherlands). Pictures were taken using imaging plates (Ditabis, Pforzheim, Germany).

For statistical analysis, neoblasts in the PB area were inspected and number, size and morphology of their chromatoid bodies were accounted.

For immuno-EM, fragments of planarians were fixed with either 4% paraformaldehyde or a mixture of 2% paraformaldehyde and 0.2% glutaraldehyde in 0.1 M phosphate buffer, pH 7.4. The specimen was then cryoprotected in 2.3 M sucrose and frozen in liquid nitrogen. Ultrathin cryosectioning and immunogold-labelling were performed as described (Slot and Geuze, 2007), except a permeabilization treatment with 1 mg/ml saponin in PBS, pH 7.4, for 10 minutes prior to blocking to improve the accessibility of the antibodies to their epitopes.

Western blot

Animals were snap-frozen in liquid nitrogen and ground with a disposable potter. After direct lyses in 50 μl 2× Laemmli buffer, samples were loaded on a 10% acrylamide minigel and run under denaturing SDS-PAGE conditions for 3 hours under constant current. Proteins were transferred on a PVDF Immobilon membrane (Millipore, Schwalbach, Germany) and processed for immunodetection with ECL plus chemistry (GE-Healthcare, Solingen, Germany). Antibodies used were: mouse anti-SmB monoclonal antibody (Sigma S0698) diluted 1:1000; rabbit anti-Gapdh polyclonal antibody (Abcam, ab36840) diluted 1:5000; and rabbit anti-Smed-SmB (residues 46-60) affinity-purified polyclonal antibodies SmB310 and 311, both diluted 1:250,000. Secondary HRP-conjugated antibodies were used at 1:50,000, according to the host species of the primary antibodies. Band intensities were quantified using the Chemi-Doc XRS platform equipped with Quantity One software (BioRad, Munich, Germany).

Statistics

One-way ANOVA with Dunnett's post-test, Student's t-test and Fisher's exact test were performed using GraphPad Prism 4.03 (GraphPad software, La Jolla, CA, USA).

Despite being represented throughout the planarian body and experimentally well accessible, neoblasts are still largely defined by their morphology and by their sensitiveness to γ-irradiation. Therefore, γ-irradiated planaria are instrumental in the discovery of new neoblast-specific markers. Using a proteomic approach, we were able to identify the first planarian member of the LSm superfamily as differentially expressed in wild-type versus γ-irradiated regenerating animals (E.F.-T., G. Rodriguez-Esteban, E.S. and J. Abril, unpublished data). ClustalW alignment showed the conservation of the two Sm fold domains at the N-terminus of the protein (see Fig. S3A in the supplementary material), whereas the bootstrap-based UPGMA method was used to calculate the phylogenetic tree that confirmed Smed-SmB similarity with the SmB proteins of different metazoans (see Fig. S3B in the supplementary material).

Smed-SmB transcripts localize to neoblasts

Neoblasts are evenly dispersed throughout the parenchyma, a scarcely differentiated tissue that lies under the epidermal and muscle layers, and surrounds all the organs (Baguñá, 1973; Saló, 2006). Wholemount in situ hybridization (ISH) against Smed-SmB transcripts showed the typical neoblast distribution in adult intact planarians (Fig. 1A, 1-3). Fluorescent ISH on dissociated cells revealed that Smed-SmB signal could be detected in small (5-10 μm) round cells with a high nucleus-to-cytoplasm (N/C) ratio (Fig. 1B-D), consistent with the morphological definition of neoblasts (Baguñà, 1981). Additionally, wholemount ISH against Smed-SmB on regenerating animals 48-72 hours after amputation produced a strong signal in the post-blastema (PB; Fig. 1E), where proliferating neoblasts consequently accumulate to the major proliferation peak induced by amputation (Baguñà, 1976).

Following γ-irradiation, Smed-SmB expression was significantly diminished, although irradiation-tolerant Smed-SmB-positive cells persisted in the cephalic region (Fig. 2A), as reported for other neoblast markers like Bruli (Guo et al., 2006) and Pumilio (Salvetti et al., 2005). This observation was further substantiated as the relative expression of the transcript varied depending on the portion of the body considered. Fourteen days following γ-irradiation, animals were cut into head, trunk (pharynx and surrounding tissues) and tail fragments (Fig. 2B). The expression of Smed-SmB mRNA in the cephalic portion of irradiated animals was similar to the non-irradiated control (P=0.5667), whereas in the trunk and tail fragments it was 29±3% (P<0.0001) and 33±3% (P=0.0005), respectively (Fig. 2C). Western blot confirmed a similar trend for the expression of Smed-SmB protein (Fig. 2D,E).

Fig. 1.

Smed-SmB expression in asexual S. mediterranea. (A) Wholemount ISH showing the expression pattern of Smed-SmB in intact wild-type animals. Smed-SmB mRNA was found in the parenchyma throughout the animals, with the exception of the pharynx (Ph) and the area anterior to the photoreceptors. This pattern is usually referred to as a neoblast pattern. Sections of the specimen in A showed abundance of Smed-SmB transcripts also around the cephalic ganglia (cg in A1). Signal was otherwise absent from nerve cords (nc in A1), pharynx (Ph in A2) and gastro-vascular system (asterisks in A1-3). (B-D) Fluorescent ISH against Smed-SmB on isolated cells. Fluorescence (B) and Nomarski (C) images of a Smed-SmB-positive cell with neoblast morphology, where a cytoplasmic signal was detected. Differentiated cells showed no signal for Smed-SmB (D). (E) Wholemount ISH against Smed-SmB on regenerating wild-type animals 48 hours after amputation. Signal of Smed-SmB transcripts in the post-blastema region (arrowhead) is stronger than in the rest of the body. Scale bars: 0.5 mm in A and E; 0.1 mm in A1-3; 5 μm in B-D.

Fig. 1.

Smed-SmB expression in asexual S. mediterranea. (A) Wholemount ISH showing the expression pattern of Smed-SmB in intact wild-type animals. Smed-SmB mRNA was found in the parenchyma throughout the animals, with the exception of the pharynx (Ph) and the area anterior to the photoreceptors. This pattern is usually referred to as a neoblast pattern. Sections of the specimen in A showed abundance of Smed-SmB transcripts also around the cephalic ganglia (cg in A1). Signal was otherwise absent from nerve cords (nc in A1), pharynx (Ph in A2) and gastro-vascular system (asterisks in A1-3). (B-D) Fluorescent ISH against Smed-SmB on isolated cells. Fluorescence (B) and Nomarski (C) images of a Smed-SmB-positive cell with neoblast morphology, where a cytoplasmic signal was detected. Differentiated cells showed no signal for Smed-SmB (D). (E) Wholemount ISH against Smed-SmB on regenerating wild-type animals 48 hours after amputation. Signal of Smed-SmB transcripts in the post-blastema region (arrowhead) is stronger than in the rest of the body. Scale bars: 0.5 mm in A and E; 0.1 mm in A1-3; 5 μm in B-D.

Fig. 2.

Effects of γ-irradiation on Smed-SmB expression. (A) Intact wild-type planarians were irradiated and subsequently analyzed for Smed-SmB expression by wholemount ISH. Following γ-irradiation (75 Gys), Smed-SmB expression decreased without disappearing, particularly in the cephalic portion of the body. (B,C) Real-time quantification of Smed-SmB transcripts in different body compartments (B) confirmed that most of the residual expression found in irradiated animals comes from the cephalic area (C, orange line). In trunk and tail portions, the effects of the irradiation were more prominent (C, blue and green lines, respectively). Mean ± standard deviation of three independent experiments is shown. (D,E) Commercially available antibodies against mammalian SmB and Gapdh efficiently cross-reacted with the planarian proteins (D) allowing a reliable quantification. The chart in E shows the quantification of the blot shown in D. Scale bar: 0.5 mm in A. In D, + is irradiated, − is wt control.

Fig. 2.

Effects of γ-irradiation on Smed-SmB expression. (A) Intact wild-type planarians were irradiated and subsequently analyzed for Smed-SmB expression by wholemount ISH. Following γ-irradiation (75 Gys), Smed-SmB expression decreased without disappearing, particularly in the cephalic portion of the body. (B,C) Real-time quantification of Smed-SmB transcripts in different body compartments (B) confirmed that most of the residual expression found in irradiated animals comes from the cephalic area (C, orange line). In trunk and tail portions, the effects of the irradiation were more prominent (C, blue and green lines, respectively). Mean ± standard deviation of three independent experiments is shown. (D,E) Commercially available antibodies against mammalian SmB and Gapdh efficiently cross-reacted with the planarian proteins (D) allowing a reliable quantification. The chart in E shows the quantification of the blot shown in D. Scale bar: 0.5 mm in A. In D, + is irradiated, − is wt control.

Smed-SmB is essential for tissue homeostasis and regeneration

In order to understand the function of Smed-SmB, we performed RNAi experiments on regenerating and intact animals. RNAi efficiency was assessed at both the RNA and protein levels. qRT-PCR showed a sudden downregulation of the transcript (more than 90%), which followed the same kinetics in both intact and regenerating animals (Fig. 3A). Smed-SmB protein expression was also clearly influenced by the RNAi, showing different kinetics between intact and regenerating animals. In regenerating animals, Smed-SmB protein was barely detectable as early as 2 days after amputation. A comparable level of protein expression is attained in intact animals after 9 days from the last dsRNA injection (Fig. 3A).

All Smed-SmB dsRNA-treated animals (n=95), which were amputated 1 or 3 days after the third round of injections (t-cut), failed to regenerate and died within two weeks after amputation. Interestingly, in no cases could we observe blastema formation (Fig. 3B). Within three to four weeks following RNAi, intact animals (n=50) also died, with signs of degeneration becoming evident 10 days after the RNAi treatment (Fig. 3C).

To understand whether Smed-SmB has a neoblast-specific function, we observed the changes induced by Smed-SmB RNAi to the planarian cell populations. Largely based on the morphological features of planarian cells, Hoechst 33342/Calcein AM double-staining is able to separate the irradiation-sensitive cells (X1 and X2) from the differentiated cells (Xin) on a FACS plot (Reddien et al., 2005; Hayashi et al., 2006). In a similar way to γ-irradiation, Smed-SmB RNAi induced a dramatic reduction of irradiation-sensitive fractions X1 and X2 in intact animals (Fig. 4A). Irradiation-sensitive X1 and X2 cells were drastically reduced in number after 11 days (18±8.3% and 27±11.3% of the relative control, respectively; P<0.0001), with the decrease beginning after 7 days (P<0.05; Fig. 4B,C). These data nicely correlate with the availability of Smed-SmB protein shown in Fig. 3A. FACS analysis of PB tissue separated from the rest of the body (RoB) of regenerating RNAi animals showed that the reduction of X1 and X2 cells happened much faster in the PB (day 3, P<0.05), where actively proliferating neoblasts are enriched compared with the RoB (day 6, P<0.05; Fig. 4D-G).

Smed-SmB RNAi affects the expression of planarian cyclinB

We analyzed the expression of neoblast-specific markers in both intact and regenerating animals and we found they were, in both cases, significantly reduced following Smed-SmB RNAi. qRT-PCR of regenerating animals showed that Smed-Bruli (Guo et al., 2006) and Smedwi1 (Reddien et al., 2005; Rossi et al., 2006) expression rapidly disappeared, beginning 3 days after Smed-SmB RNAi (P=0.0002 and P=0.0013, respectively; Fig. 5A). By contrast, the expression of differentiated cell markers, like Smed-Pax6 (Pineda et al., 2002) and Smed-Tmus (Cebriá et al., 1997), was essentially unaltered (P=0.4551 and P=0.0983, respectively; Fig. 5B). The expression of Smedwi2, involved in neoblast differentiation (Reddien et al., 2005), also suffered a significant reduction (P<0.05), although milder and at a later time (day 5) than the other neoblast markers (Fig. 5A).

Fig. 3.

Smed-SmB phenotype. (A) Animals were injected with double-stranded RNA to suppress Smed-SmB expression. Some were subsequently amputated and let regenerate (n=90), whereas others were left intact (n=50). A swift reduction of about 90% at the RNA level was followed by a reduction of about 80% at the protein level, as shown by qRT-PCR and western blotting. Although the dynamics of downregulation of the transcripts is identical in intact and regenerating animals, in intact animals, the time required for an effective protein downregulation was about 5× longer than in regenerating ones. (B) Water- and Smed-SmB dsRNA-injected animals were prepharingeally amputated into two pieces (head and tail) in the days following the third dsRNA injection and let to regenerate. We could not observe blastema formation in any of the amputated Smed-SmB RNAi animals, whereas in control animals, this took place normally (red dashed lines). Additionally, a degenerative process (*) began in dsRNA-injected animals after 10 days, which eventually led to death within 2 weeks. Differences in the phenotypes were not observed between anterior and posterior regeneration, nor among animals cut at different times (1 or 3 days) after dsRNA injection. (C) Intact animals were also treated with dsRNA to inhibit Smed-SmB expression. A degenerative process (arrows) began to affect the head of the animals around day 10, and gradually progressed throughout the rest of the body, leading to death within 3-4 weeks. Scale bars: 0.5 mm.

Fig. 3.

Smed-SmB phenotype. (A) Animals were injected with double-stranded RNA to suppress Smed-SmB expression. Some were subsequently amputated and let regenerate (n=90), whereas others were left intact (n=50). A swift reduction of about 90% at the RNA level was followed by a reduction of about 80% at the protein level, as shown by qRT-PCR and western blotting. Although the dynamics of downregulation of the transcripts is identical in intact and regenerating animals, in intact animals, the time required for an effective protein downregulation was about 5× longer than in regenerating ones. (B) Water- and Smed-SmB dsRNA-injected animals were prepharingeally amputated into two pieces (head and tail) in the days following the third dsRNA injection and let to regenerate. We could not observe blastema formation in any of the amputated Smed-SmB RNAi animals, whereas in control animals, this took place normally (red dashed lines). Additionally, a degenerative process (*) began in dsRNA-injected animals after 10 days, which eventually led to death within 2 weeks. Differences in the phenotypes were not observed between anterior and posterior regeneration, nor among animals cut at different times (1 or 3 days) after dsRNA injection. (C) Intact animals were also treated with dsRNA to inhibit Smed-SmB expression. A degenerative process (arrows) began to affect the head of the animals around day 10, and gradually progressed throughout the rest of the body, leading to death within 3-4 weeks. Scale bars: 0.5 mm.

Given the incapacity shown by Smed-SmB knocked-down animals to form a regenerating blastema, we decided to investigate neoblast proliferative potential. PCNA-positive cells were reduced in number after 10 days from Smed-SmB knockdown in intact animals (P<0.0001; Fig. 5C-E). PCNA is an essential component of the DNA replication machinery that accumulates in cells that are virtually able to proliferate (Bravo et al., 1987; Orii et al., 2005). Conversely, cyclinB is expressed only by actively cycling neoblasts of the X1 subpopulation (Reddien et al., 2005; Eisenhoffer et al., 2008). In wild-type regenerating planarians, its expression increases around day 2-3 of regeneration (P<0.05; Fig. 5F), related to the major peak of mitotic activity (Saló and Baguñà, 1984). As a consequence of Smed-SmB RNAi, Smed-CycB expression was promptly reduced 2 days after amputation (P<0.05), reaching a basal level (more than 90% downregulation; P<0.0001) after 7 days (Fig. 5F).

Using as primer oligonucleotides able to amplify both the spliced and unspliced forms of the Smed-CycB transcript, we observed that the decrease in mature Smed-CycB mRNA expression was accompanied by an increase in the expression of its unspliced form (Fig. 5G). The ratio between the expressions of unspliced and spliced transcripts in Smed-SmB RNAi animals increased up to 5-fold after 3 days of regeneration (Fig. 5H). This suggests that Smed-SmB RNAi does not prevent the expression of cyclinB, but rather, at least in part, prevents the splicing of its transcript.

Loss of Smed-SmB impairs chromatoid body organization

We found that Smed-SmB is important to guarantee the proliferative ability of the neoblasts. We then had a closer look at the ultrastructural level of stem cells in both intact and regenerating Smed-SmB RNAi animals. We noticed that, after 5 days of regeneration, the cytoplasm of the neoblasts was less-densely packed (Fig. 6A) compared with water-injected animals (Fig. 6B). This change was accompanied by an increase in cellular membranes of the endolysosomal system and the Golgi (Fig. 6A, asterisks) — normally not present in neoblasts (Baguña, 1973). CBs are electron-dense granules not delimited by membranes (Fig. 6B, arrows), which resemble the appearance of heterochromatic nuclear spots (Auladell et al., 1993). A loss of integrity in the CB structure was also observed at higher magnification (Fig. 6A,C, white frame). Several small CB-like structures (Fig. 6C, arrowheads) appeared, probably owing to the fragmentation of pre-existing CBs. Occasionally, we also observed the presence of large autophagic vacuoles (Fig. 6D, asterisks), suggesting that the decrease in the X1 and X2 populations might be regulated by this degenerative pathway (González-Estévez et al., 2007). Interestingly, we also found that the degeneration of the CBs follows the Smed-SmB protein availability (Fig. 3A). In intact animals, we occasionally observed degenerated CBs, but only after 11 days from RNAi treatment onwards. In regenerating animals, however, the CB state reflected the neoblast position in the planarian body. CBs observed in neoblasts of the posterior stripe began to degenerate after 5 days from amputation, whereas those found in neoblasts of the PB began to degenerate as early as 1 day after cutting (P<0.05; Fig. 6E). As CBs degenerated, an increasing number of neoblast-like cells without CBs were observed in the PB. The percentage of these cells increased from 11.8±1.8% at t-cut to 50.5±12.5% after 3 days (P<0.05), with a clear trend (R2=0.923; Fig. 6E). Conversely, the percentage of healthy neoblasts with normal CBs dropped from 69.4±8.5% at t-cut to 13.7±11.1% after 3 days (P<0.05; Fig. 6E).

Fig. 4.

Dynamics of the planarian cell populations following Smed-SmB RNAi. (A) Planarians were gently dissociated and their cells were double-stained with Hoechst 33342 and Calcein AM. FACS analysis showed different populations according to size and nucleus-to-cytoplasm ratio. (B,C) Compared with water-injected control animals (B), in intact Smed-SmB RNAi animals, the irradiation-sensitive population X1 (red) and X2 (green) gradually disappeared, beginning around day 7 (C). After 11 days, X1 and X2 populations were reduced by about 81% and 73%, respectively. (D-G) Compared with control (D) or intact Smed-SmB RNAi animals (C), in the post-blastema of Smed-SmB RNAi regenerating animals, neoblast reduction happened much faster, beginning as early as 3 days after treatment (E). By day 7, neoblast populations had decreased to levels similar to intact animals after 11 days (C,E). In the rest of the body, however, the effect of Smed-SmB RNAi progressed more gradually, X1 and X2 populations were reduced by about 36% and 33% after 7 days, respectively (G). Means ± standard deviations in B and C refer to four independent experiments. Means ± standard deviations in D-G refer to three independent experiments.

Fig. 4.

Dynamics of the planarian cell populations following Smed-SmB RNAi. (A) Planarians were gently dissociated and their cells were double-stained with Hoechst 33342 and Calcein AM. FACS analysis showed different populations according to size and nucleus-to-cytoplasm ratio. (B,C) Compared with water-injected control animals (B), in intact Smed-SmB RNAi animals, the irradiation-sensitive population X1 (red) and X2 (green) gradually disappeared, beginning around day 7 (C). After 11 days, X1 and X2 populations were reduced by about 81% and 73%, respectively. (D-G) Compared with control (D) or intact Smed-SmB RNAi animals (C), in the post-blastema of Smed-SmB RNAi regenerating animals, neoblast reduction happened much faster, beginning as early as 3 days after treatment (E). By day 7, neoblast populations had decreased to levels similar to intact animals after 11 days (C,E). In the rest of the body, however, the effect of Smed-SmB RNAi progressed more gradually, X1 and X2 populations were reduced by about 36% and 33% after 7 days, respectively (G). Means ± standard deviations in B and C refer to four independent experiments. Means ± standard deviations in D-G refer to three independent experiments.

The evidence we collected at the ultrastructural level support the idea that knockdown of Smed-SmB impairs the functional organization of the CB, as was also observed for the germ granules of C. elegans after RNAi of either SmE or SmB (Barbee et al., 2002).

Smed-SmB protein localizes to neoblast nuclei and chromatoid bodies

The dynamics of the event characteristics of the Smed-SmB phenotype showed that the degeneration of the CBs is one of the first outcomes of the phenotype, particularly when considering the neoblasts located in the PB (Fig. 6E; Fig. 7A). As several articles proved the localization and functional role of the LSm proteins in CBs and CB-like structures (Parvinen, 2005; Strome and Lehmann, 2007; Beaudoin et al., 2009), we tried to uncover whether Smed-SmB protein also localizes to the CB. Cryoimmuno-EM using two different affinity-purified rabbit polyclonal antibodies (SmB310 and SmB311), specifically raised against residues 46-60 of the Smed-SmB protein, showed gold particle labelling of both the nucleus — except the nucleolus — and the CB of wild-type neoblasts (Fig. 7B). These were mainly labelled at the edge, probably as consequence of steric hindrance of the antigens in the condensed structure. Results were confirmed by using both the antibodies specifically raised. Nuclear Smed-SmB signal was not found in differentiated cells (data not shown). These results indicate with good confidence the presence of the Smed-SmB protein inside as well as outside the nuclear compartment of planarian neoblasts.

Fig. 5.

Smed-SmB RNAi affects the expression of neoblast markers. (A,B) The expression of two panels of genes, either specific to neoblasts (A) or to differentiated cells (B), was assessed by real-time qRT-PCR in regenerating animals. The neoblast markers Smed-Bruli, Smedwi1 and Smedwi2, suffered a consistent downregulation that began 3, 4 or 5 days after treatment, respectively. After 7 days, Smed-Bruli and Smedwi1 expression were reduced by about 90% relative to a water-injected control, whereas Smedwi2 expression was only reduced by about 55% (A). Conversely, Smed-SmB RNAi had no apparent effect on the expression of Smed-Pax6 and Smed-Tmus (B). (C-E) As shown by wholemount IHC of the tail portion of intact animals, the number of PCNA-positive cells in Smed-SmB-inhibited animals was significantly reduced around day 10 (D) compared with control (C). Chart E presents means ± standard deviation of PCNA-positive cells from five independent experiments. A significant reduction took place between 6 and 10 days after RNAi treatment. (F) Smed-CycB expression was quantified by qRT-PCR. In regenerating control animals, cyclinB expression resembles the major mitotic event occurring between 48 and 72 hours after amputation (blue line). This peak of expression is virtually abolished in both γ-irradiated and Smed-SmB RNAi regenerating animals (green and red lines, respectively). (G,H) The presence of the unspliced form of Smed-CycB transcript was analyzed separately by conventional RT-PCR (G, upper panel) and compared with the expression of the mature transcript (G, lower panel). As early as 1 day after Smed-SmB RNAi treatment, we could observe an increased ratio between unspliced and spliced forms of Smed-CycB transcript compared with control, which reached its peak (~5-fold) after 3 days (H). Error bars in A, B, E, F and H represent the standard deviation over a minimum of three independent experiments. PCNA-positive cell number was normalized against a standard volume (1 μm3). In E: *, P<0.05; **, P<0.001. The samples checked by conventional RT-PCR for Smed-CycB (G) are the same as screened by qRT-PCR (F). However, amplification products were checked at end-point on ethidium bromide gel, which means during the linear/plateau amplification phase. Although a decrease of expression of the spliced form of Smed-CycB can still be appreciated (G, lower panel), a precise quantification of the spliced transcript is only possible by real-time RT-PCR. Scale bar: 0.2 mm.

Fig. 5.

Smed-SmB RNAi affects the expression of neoblast markers. (A,B) The expression of two panels of genes, either specific to neoblasts (A) or to differentiated cells (B), was assessed by real-time qRT-PCR in regenerating animals. The neoblast markers Smed-Bruli, Smedwi1 and Smedwi2, suffered a consistent downregulation that began 3, 4 or 5 days after treatment, respectively. After 7 days, Smed-Bruli and Smedwi1 expression were reduced by about 90% relative to a water-injected control, whereas Smedwi2 expression was only reduced by about 55% (A). Conversely, Smed-SmB RNAi had no apparent effect on the expression of Smed-Pax6 and Smed-Tmus (B). (C-E) As shown by wholemount IHC of the tail portion of intact animals, the number of PCNA-positive cells in Smed-SmB-inhibited animals was significantly reduced around day 10 (D) compared with control (C). Chart E presents means ± standard deviation of PCNA-positive cells from five independent experiments. A significant reduction took place between 6 and 10 days after RNAi treatment. (F) Smed-CycB expression was quantified by qRT-PCR. In regenerating control animals, cyclinB expression resembles the major mitotic event occurring between 48 and 72 hours after amputation (blue line). This peak of expression is virtually abolished in both γ-irradiated and Smed-SmB RNAi regenerating animals (green and red lines, respectively). (G,H) The presence of the unspliced form of Smed-CycB transcript was analyzed separately by conventional RT-PCR (G, upper panel) and compared with the expression of the mature transcript (G, lower panel). As early as 1 day after Smed-SmB RNAi treatment, we could observe an increased ratio between unspliced and spliced forms of Smed-CycB transcript compared with control, which reached its peak (~5-fold) after 3 days (H). Error bars in A, B, E, F and H represent the standard deviation over a minimum of three independent experiments. PCNA-positive cell number was normalized against a standard volume (1 μm3). In E: *, P<0.05; **, P<0.001. The samples checked by conventional RT-PCR for Smed-CycB (G) are the same as screened by qRT-PCR (F). However, amplification products were checked at end-point on ethidium bromide gel, which means during the linear/plateau amplification phase. Although a decrease of expression of the spliced form of Smed-CycB can still be appreciated (G, lower panel), a precise quantification of the spliced transcript is only possible by real-time RT-PCR. Scale bar: 0.2 mm.

Fig. 6.

Smed-SmB RNAi produces loss of chromatoid body organization. (A-C) Representative appearance of neoblasts of the posterior stripe belonging either to Smed-SmB dsRNA- or water-injected regenerating animals, 5 days after amputation. In neoblasts from water-injected controls (B), chromatoid bodies (CBs) could be observed (arrows), depending on the sectioning level of the cell, in a highly condensed cytoplasm. On the contrary, some portions of the cytoplasm of treated animals were less-densely packed and cellular membranes began to develop (asterisks in A). In these areas, small granules with the aspect of CBs could be recognized (arrowheads in C) together with translucent vesicular membranes. (D) This micrograph shows a neoblast with unusual autophagic vacuoles (asterisks), in a Smed-SmB RNAi regenerating animal 5 days after amputation. (E) Structural integrity and number of CBs were analyzed on ultrathin sections of post-blastema. We began to observe neoblasts either with destructured CBs or without any CBs starting 1 day after amputation of Smed-SmB RNAi animals. The number of neoblasts with normal morphology, conversely, dropped dramatically. After 3 days of regeneration, only 26% of the neoblasts screened (n=20) showed the presence of CBs, and in only 14% of these cells, CBs had a size comparable with control ones. Bars in E represent means ± s.e. of three independent experiments. Scale bars: 2 μm in A,B,D; 0.5 μm in C.

Fig. 6.

Smed-SmB RNAi produces loss of chromatoid body organization. (A-C) Representative appearance of neoblasts of the posterior stripe belonging either to Smed-SmB dsRNA- or water-injected regenerating animals, 5 days after amputation. In neoblasts from water-injected controls (B), chromatoid bodies (CBs) could be observed (arrows), depending on the sectioning level of the cell, in a highly condensed cytoplasm. On the contrary, some portions of the cytoplasm of treated animals were less-densely packed and cellular membranes began to develop (asterisks in A). In these areas, small granules with the aspect of CBs could be recognized (arrowheads in C) together with translucent vesicular membranes. (D) This micrograph shows a neoblast with unusual autophagic vacuoles (asterisks), in a Smed-SmB RNAi regenerating animal 5 days after amputation. (E) Structural integrity and number of CBs were analyzed on ultrathin sections of post-blastema. We began to observe neoblasts either with destructured CBs or without any CBs starting 1 day after amputation of Smed-SmB RNAi animals. The number of neoblasts with normal morphology, conversely, dropped dramatically. After 3 days of regeneration, only 26% of the neoblasts screened (n=20) showed the presence of CBs, and in only 14% of these cells, CBs had a size comparable with control ones. Bars in E represent means ± s.e. of three independent experiments. Scale bars: 2 μm in A,B,D; 0.5 μm in C.

In this study we characterized the function of the planarian SmB/B′/N ortholog, Smed-SmB, the first member of the conserved LSm superfamily identified in planarians.

The data we collected suggest an essential role for Smed-SmB in sustaining the proliferative potential of the planarian stem cells during regeneration and in tissue homeostasis.

Smed-SmB transcripts were mainly found in the parenchyma and around the cephalic ganglia. Expression of Smed-SmB by postmitotic cells of the cephalic ganglia explains the partial downregulation we found after γ-irradiation, which only affects actively dividing cells.

The silencing of Smed-SmB, efficiently achieved through dsRNA-induced RNAi, produced a late degenerative phenotype in intact animals (Fig. 3C) and an early non-regenerating phenotype in amputated ones (Fig. 3B). The phenotype was accompanied, in both cases, by downregulation of the neoblast markers (Fig. 5B) and reduction in the number of neoblasts (Fig. 4A-G). These effects were also described following the perturbation of other genes — like Pumilio, Smedwi2 and Bruli — involved in neoblast maintenance or differentiation (Salvetti et al., 2005; Reddien et al., 2005; Guo et al., 2006). Compared with the latter, however, the Smed-SmB phenotype showed some peculiarities.

The degenerative effects produced by Smed-SmB knockdown in intact animals progressed at a slower rate compared with the other genes essential for neoblast function. At the beginning of the fourth week, more than 60% of the intact animals were still alive, albeit with a degenerated anterior part (Fig. 3C). This implies that maintenance of the differentiated cells was virtually unaffected (Fig. 4A-C), and that the delay in degeneration observed in intact animals was due to the prevention of tissue turnover.

Also, in Smed-SmB RNAi regenerating animals, the inability to form a blastema was not affected by the length of time between injection and amputation, as observed in the case of Smed-Bruli and Smedwi2 RNAi (Guo et al., 2006; Reddien et al., 2005). Animals cut at different time points after injection (1 or 3 days) invariably failed to form a blastema, suggesting that neoblasts could proliferate, and possibly differentiate, for a short time after either Smed-Bruli or Smedwi2 RNAi, but they were unable to do so following Smed-SmB RNAi. On the contrary, proliferation was severely hampered, as depicted by a decrease in the expression of both PCNA (Fig. 5C-D) and Smed-CycB (Fig. 5E). Although we cannot exclude an effect of Smed-SmB RNAi on the early neoblast progeny, we concluded that the inability to regenerate a blastema as demonstrated by Smed-SmB RNAi animals resulted from the absence of actively proliferating cells.

Another interesting observation was that the expression of a gene, which was otherwise ubiquitously expressed and involved in basal cell functions, was restricted in planaria to stem cells and neurons. Conversely, SmB is known to be a component of RNP granules, also known as CBs, which are present in the germ cells of virtually all metazoans (Eddy, 1975). Recently, their presence was also observed in CB-like structures induced in mammalian cells by expressing the cytoplasmic form of the Prion protein (Beaudoin et al., 2009). The large RNP complexes found in the germ cells share remarkable similarities with planarian CBs, which are a characteristic feature of neoblasts (Hay and Coward, 1975), germ cells (Coward, 1974) and neurons (Yoshida-Kashikawa et al., 2007) and are expected to function in posttranscriptional gene regulation.

Like the germ cell granules, planarian CBs assemble components of the machinery to process mRNA in the cytoplasm, like DjCBC-1 (Yoshida-Kashikawa et al., 2007) and Spoltud-1 (Solana et al., 2009). They might also contain small non-coding RNAs (miRNAs, snRNAs) and members of the LSm protein superfamily. Following neoblast differentiation, CBs gradually disappear, only to persist in the germline (Coward, 1974) and in neurons (Yoshida-Kashikawa et al., 2007).

The Smed-SmB RNAi effects we observed at the CB level showed striking similarities with the knockdown of either SmE or SmB in C. elegans embryos, which alters the sub-cellular distribution of the P-granules and leads to the inaccurate segregation of the germline (Barbee et al., 2002). Our data suggest that Smed-SmB RNAi results in a loss of CB structure and organization, which subsequently translates into a loss of function (Fig. 6C). As generally believed, CBs store the transcripts to provide the cell with a quick response to physiological needs (Kotaja and Sassone-Corsi, 2007). Hence, it seems plausible that the observed degeneration of the CBs triggers the proliferative failure of the neoblasts and the inability of the animals to form a blastema. This is in contrast to the Spoltud-1 phenotype, where the CB is probably unaffected, but the neoblast population is depleted regardless, even though after a much longer time (Solana et al., 2009). The observed increase of unspliced Smed-CycB transcripts suggests that Smed-SmB RNAi additionally leads to a splicing defect of genes essential for neoblast proliferation. The moderate increase in the expression of cyclinB pre-mRNA is probably underestimated owing to its active degradation by nonsense-mediated decay following the introduction of stop codons into the transcript sequence by the unspliced introns (for a review, see Isken and Maquat, 2008). Strikingly, the degeneration of the CBs in neoblasts of the PB occurs in parallel to Smed-SmB protein downregulation. Therefore, we propose that CB degeneration is not the effect of a general splicing defect caused by the loss of Smed-SmB. Such an indirect effect would require the degradation of both the mRNAs and proteins of Smed-SmB splicing targets, a process we would expect to take longer to produce the degeneration and disappearance of the CBs that we observed.

Fig. 7.

Dynamics of the cellular and molecular events leading to the Smed-SmB phenotype in intact and regenerating animals, and subcellular localization of Smed-SmB protein. (A) In both intact (I) and regenerating (R) animals, Smed-SmB transcripts were drastically reduced shortly after Smed-SmB knockdown. Smed-SmB protein, however, showed a small decay in intact animals, which only became prominent after 7 days of treatment. In regenerating animals, by contrast, Smed-SmB protein downregulation shortly followed RNA downregulation, resulting in a severe reduction of protein availability after 2 days. Remarkably, the downregulation of Smed-CycB expression, the reduction of X1 neoblasts and the number of neoblasts with CBs were correlated with — and consequent to — Smed-SmB protein downregulation, not only regarding the timing, but also the dynamics: gradual for intact animals, abrupt for regenerating ones. (B) Electron micrograph of an ultrathin cryosection (55 nm) immunogold-labelled against Smed-SmB protein, showing detail of a neoblast. Owing to EM protocol, membranes appear white. ProteinA-coupled gold particles are 15 nm in size. The nucleus (n) revealed gold labelling mainly localized to uncondensed chromatin. Cytoplasm labelling occurred rarely, whereas CBs (dimension emphasized by arrowheads) were mainly labelled at the rim. Mitochondria (mi) are devoid of labelling. Scale bar: 200 nm.

Fig. 7.

Dynamics of the cellular and molecular events leading to the Smed-SmB phenotype in intact and regenerating animals, and subcellular localization of Smed-SmB protein. (A) In both intact (I) and regenerating (R) animals, Smed-SmB transcripts were drastically reduced shortly after Smed-SmB knockdown. Smed-SmB protein, however, showed a small decay in intact animals, which only became prominent after 7 days of treatment. In regenerating animals, by contrast, Smed-SmB protein downregulation shortly followed RNA downregulation, resulting in a severe reduction of protein availability after 2 days. Remarkably, the downregulation of Smed-CycB expression, the reduction of X1 neoblasts and the number of neoblasts with CBs were correlated with — and consequent to — Smed-SmB protein downregulation, not only regarding the timing, but also the dynamics: gradual for intact animals, abrupt for regenerating ones. (B) Electron micrograph of an ultrathin cryosection (55 nm) immunogold-labelled against Smed-SmB protein, showing detail of a neoblast. Owing to EM protocol, membranes appear white. ProteinA-coupled gold particles are 15 nm in size. The nucleus (n) revealed gold labelling mainly localized to uncondensed chromatin. Cytoplasm labelling occurred rarely, whereas CBs (dimension emphasized by arrowheads) were mainly labelled at the rim. Mitochondria (mi) are devoid of labelling. Scale bar: 200 nm.

Without the possibility to trace planarian cells and look at their fate, we could only speculate about the fate of Smed-SmB RNAi neoblasts. Nonetheless, we frequently observed lysosomes and autophagic vacuoles (Fig. 6A,D) in Smed-SmB RNAi neoblasts, suggesting that these cells might undergo resorption by autophagy (González-Estévez et al., 2007). It is interesting, however, that neoblasts are able to survive longer if not actively cycling, as in the case of intact RNAi animals. Non-regenerating animals rely on their pool of stem cells to guarantee cellular turnover, and this is the reason why the neoblast population of intact animals is virtually unaffected by Smed-SmB knockdown for up to 7 days after the RNAi treatment (Fig. 4A-D). This capacity to survive the Smed-SmB RNAi for an extended period of time is probably owing to the long half-life of the Smed-SmB protein (Fig. 3A; Fig. 7A).

In this work, we have shown that a member of the LSm superfamily, Smed-SmB, is required for the proliferation and maintenance of the adult stem cells of Schmidtea mediterranea. Smed-SmB is expressed by stem cells and neurons. Knockdown of Smed-SmB expression, however, apparently affects only neoblasts, causing the downregulation of the expression of specific neoblast markers like Smedwi1 and Smed-Bruli. Expression of the proliferation markers PCNA and Smed-CycB was also quickly impaired, resulting in a massive proliferative failure and drastic reduction of viable neoblasts. As a direct consequence, regenerating animals do not form blastema and die within 2 weeks. Smed-SmB knockdown also leads to the death of intact animals, presumably as consequence of a lack of cellular turnover. At the subcellular level, we observed a severe degeneration of the CBs. Within 3 days, the number of neoblasts with ‘healthy’ CBs in the PB was reduced to 1 out of 8. By immunogold-labelling, we could localize the Smed-SmB protein to the nucleus of planarian stem cells and to the CB. In addition, we also found an increased expression of unspliced Smed-CycB transcripts. The latter is indicative of a failure of the splicing machinery, which further impairs the proliferative capacity of the stem cells.

We thank H. Orii for providing DjPCNA antibody, C. Ortmeier and G. Verberk for real-time qRT-PCR, K. Mildner for technical assistance with the electron microscopy sample preparation, J. Müller-Keuker for assistance with the figures and S. Kölsch for proofreading. This work was supported by grant BFU-2005-00422 and BFU2008-01544 from the Ministerio de Educación y Ciencia (Spain) and grants 2005SGR00769 and 2009SGR1018 from AGAUR (Generalitat de Catalunya, Spain) to E.S., grant DAAD D/06/12871 to H.R.S., and E.F.T. received an FPI fellowship from the Ministerio de Ciencia y Cultura, Spain.

Agata
K.
,
Soejima
Y.
,
Kato
K.
,
Kobayashi
C.
,
Umesono
Y.
,
Watanabe
K.
(
1998
).
Structure of the planarian central nervous system (CNS) revealed by neuronal cell markers
.
Zool. Sci.
15
,
433
-
440
.
Auladell
C.
,
Garcia-Valero
J.
,
Baguñà
J.
(
1993
).
Ultrastructural localization of RNA in chromatoid bodies of undifferentiated stem cells (neoblasts) in planarians by the RNase-Gold complex technique
.
J. Morphol.
216
,
319
-
326
.
Baguñà
J.
(
1973
).
Estudios citotaxonomicos, ecologicos, e histofisiologia de la regulacion morfogenetica durante el crecimiento y la regeneracion de la raza asexuada de la planaria Dugesia mediterranea n. sp. (Turbellaria, Trioladida, Paludicola)
.
Genetics
,
Universidad de Barcelona
.
PhD thesis
.
Baguñà
J.
(
1976
).
Mitosis in the intact and regenerating planarian Dugesia mediterranea n. sp. II. Mitotic studies during regeneration and a possible mechanism of blastema formation
.
J. Exp. Zool.
195
,
65
-
80
.
Baguñà
J.
(
1981
).
Planarian neoblasts
.
Nature
290
,
14
-
15
.
Baguñà
J.
,
Romero
R.
(
1981
).
Quantitative analysis of cell types during growth, degrowth and regeneration in the planarians Dugesia mediterranea and Dugesia tigrina
.
Hydrobiologia
84
,
181
-
194
.
Baguñà
J.
,
Saló
E.
,
Auladell
C.
(
1989
).
Regeneration and pattern formation in planarians III. Evidence that neoblasts are totipotent stem cells and the source of blastema cells
.
Development
107
,
77
-
86
.
Barbee
S. A.
,
Lublin
A. L.
,
Evans
T. C.
(
2002
).
A novel function for the Sm proteins in germ granule localization during C. elegans embryogenesis
.
Curr. Biol.
12
,
1502
-
1506
.
Beaudoin
S.
,
Vanderperre
B.
,
Grenier
C.
,
Tremblay
I.
,
Leduc
F.
,
Roucou
X.
(
2009
).
A large ribonucleoprotein particle induced by cytoplasmic PrP shares striking similarities with the chromatoid body, an RNA granule predicted to function in posttranscriptional gene regulation
.
Biochim. Biophys. Acta
1793
,
335
-
345
.
Bravo
R.
,
Frank
R.
,
Blundell
P. A.
,
Macdonald-Bravo
H.
(
1987
).
Cyclin/PCNA is the auxiliary protein of DNA polymerase-delta
.
Nature
326
,
515
-
517
.
Cebriá
F.
,
Newmark
P. A.
(
2005
).
Planarian homologs of netrin and netrin receptor are required for proper regeneration of the central nervous system and the maintenance of nervous system architecture
.
Development
132
,
3691
-
3703
.
Cebriá
F.
,
Vispo
M.
,
Newmark
P.
,
Bueno
D.
,
Romero
R.
(
1997
).
Myocite differentiation and body wall muscle regeneration in the planarian Girardia tigrina
.
Dev. Genes Evol.
207
,
306
-
316
.
Coward
S. J.
(
1974
).
Chromatoid bodies in somatic cells of the planarian: observations on their behavior during mitosis
.
Anat. Rec.
180
,
533
-
534
.
Dubois
F.
(
1949
).
Contribution á l'etude de la migration des cellules de règènèration chez les planaires dulcicoles
.
Bull. Biol. Fr. Belg.
83
,
213
-
283
.
Eddy
E.M.
(
1975
).
Germ plasm and the differentiation of the germ cell line
.
Int. Rev. Cytol.
43
,
229
-
280
.
Eisenhoffer
G. T.
,
Kang
H.
,
Sánchez-Alvarado
A.
(
2008
).
Molecular analysis of stem cells and their descendants during cell turnover and regeneration in the planarian Schmidtea mediterranea
.
Cell Stem Cell
3
,
327
-
339
.
Forsthoefel
D. J.
,
Newmark
P. A.
(
2009
).
Emerging patterns in planarian regeneration
.
Curr. Opin. Genet. Dev.
194
,
12
-
20
.
Fu
D.
,
Collins
K.
(
2006
).
Human telomerase and Cajal body ribonucleoproteins share a unique specificity of Sm protein association
.
Genes Dev.
20
,
531
-
536
.
González-Estévez
C.
,
Momose
T.
,
Gehring
W. J.
,
Saló
E.
(
2003
).
Transgenic planarian lines obtained by electroporation using transposon-derived vectors and an eye-specific GFP marker
.
Proc. Natl. Acad. Sci. USA
100
,
14046
-
14051
.
González-Estévez
C.
,
Felix
D. A.
,
Aboobaker
A. A.
,
Saló
E.
(
2007
).
Gtdap-1 promotes autophagy and is required for planarian remodeling during regeneration and starvation
.
Proc. Natl. Acad. Sci. USA
104
,
13373
-
13378
.
Guo
T.
,
Peters
A. H.
,
Newmark
P. A.
(
2006
).
A Bruno-like gene is required for stem cell maintenance in planarians
.
Dev. Cell
11
,
159
-
169
.
Hay
E.D.
,
Coward
S.J.
(
1975
).
Fine structure studies on the planarian, Dugesia. I. Nature of the “neoblast” and other cell types in noninjured worms
.
J. Ultrastruct. Res.
50
,
1
-
21
.
Hayashi
T.
,
Asami
M.
,
Higuchi
S.
,
Shibata
N.
,
Agata
K.
(
2006
).
Isolation of planarian X-ray-sensitive stem cells by fluorescence-activated cell sorting
.
Dev. Growth Differ.
48
,
371
-
380
.
He
W.
,
Parker
R.
(
2000
).
Functions of Lsm proteins in mRNA degradation and splicing
.
Curr. Opin. Cell Biol.
12
,
346
-
350
.
Isken
O.
,
Maquat
L. E.
(
2008
).
The multiple lives of NMD factors: balancing roles in gene and genome regulation
.
Nat. Rev. Genet.
9
,
699
-
712
.
Khusial
P.
,
Plaag
R.
,
Zieve
G. W.
(
2005
).
LSm proteins form heptameric rings that bind to RNA via repeating motifs
.
Trends Biochem. Sci.
30
,
522
-
528
.
Kiss
T.
(
2004
).
Biogenesis of small nuclear RNPs
.
J. Cell Sci.
117
,
5949
-
5951
.
König
H.
,
Matter
N.
,
Bader
R.
,
Thiele
W.
,
Müller
F.
(
2007
).
Splicing segregation: the minor spliceosome acts outside the nucleus and controls cell proliferation
.
Cell
131
,
718
-
729
.
Kotaja
N.
,
Sassone-Corsi
P.
(
2007
).
The chromatoid body: a germ-cell-specific RNA-processing centre
.
Nat. Rev. Mol. Cell. Biol.
8
,
85
-
90
.
Lührmann
R.
,
Kastner
B.
,
Bach
M.
(
1990
).
Structure of spliceosomal snRNPs and their role in pre-mRNA splicing
.
Biochim. Biophys. Acta
1087
,
265
-
292
.
Mayes
A. E.
,
Verdone
L.
,
Legrain
P.
,
Beggs
J. D.
(
1999
).
Characterization of Sm-like proteins in yeast and their association with U6 snRNA
.
EMBO J.
18
,
4321
-
4331
.
Molina
M. D.
,
Saló
E.
,
Cebriá
F.
(
2007
).
The BMP pathway is essential for re-specification and maintenance of the dorsoventral axis in regenerating and intact planarians
.
Dev. Biol.
311
,
79
-
94
.
Morgan
T. H.
(
1898
).
Experimental studies of the regeneration of Planaria maculata
,
Arch. Entwicklungsmech. Org.
7
,
364
-
397
.
Morita
M.
,
Best
J. B.
(
1984
).
Electron microscopic studies of planarian regeneration. IV. Cell division of neoblasts in Dugesia dorotochepala
.
J. Exp. Zool.
229
,
425
-
436
.
Newmark
P. A.
,
Sánchez Alvarado
A.
(
2000
).
Bromodeoxyuridine specifically labels the regenerative stem cells of planarians
.
Dev. Biol.
220
,
142
-
153
.
Newmark
P. A.
,
Reddien
P. W.
,
Cebriá
F.
,
Sánchez Alvarado
A.
(
2003
).
Ingestion of bacterially expressed double-stranded RNA inhibits gene expression in planarians
.
Proc. Natl. Acad. Sci. USA
100
,
11861
-
11865
.
Nogi
T.
,
Levin
M.
(
2005
).
Characterization of innexin gene expression and functional roles of gap-junctional communication in planarian regeneration
.
Dev. Biol.
287
,
314
-
335
.
Orii
H.
,
Sakurai
T.
,
Watanabe
K.
(
2005
).
Distribution of the stem cells (neoblasts) in the planarian Dugesia japonica
.
Dev. Genes Evol.
215
,
143
-
157
.
Oviedo
N. J.
,
Levin
M.
(
2007
).
Smedinx-11 is a planarian stem cell gap junction gene required for regeneration and homeostasis
.
Development
134
,
3121
-
3131
.
Parvinen
M.
(
2005
).
The chromatoid body in spermatogenesis
.
Int. J. Androl.
28
,
189
-
201
.
Patel
A. A.
,
McCarthy
M.
,
Steiz
J. A.
(
2002
).
The splicing of U12-type introns can be a rate-limiting step in gene expression
.
EMBO J.
21
,
3804
-
3815
.
Pellettieri
J.
,
Fitzgerald
P.
,
Watanabe
S.
,
Mancuso
J.
,
Green
D. R.
,
Alvarado
A. S.
(
2010
).
Cell death and tissue remodeling in planarian regeneration
.
Dev. Biol.
338
,
76
-
85
.
Petersen
C. P.
,
Reddien
P. W.
(
2009
).
Wnt signaling and the polarity of the primary body axis
.
Cell
139
,
1056
-
1068
.
Pineda
D.
,
Rossi
L.
,
Batistoni
R.
,
Salvetti
A.
,
Marsal
M.
,
Gremigni
V.
,
Falleni
A.
,
Gonzalez-Linares
J.
,
Deri
P.
,
Saló
E.
(
2002
).
The genetic network of prototypic planarian eye regeneration is Pax6 independent
.
Development
129
,
1423
-
1434
.
Reddien
P. W.
,
Oviedo
N. J.
,
Jennings
J. R.
,
Jenkin
J. C.
,
Sánchez-Alvarado
A.
(
2005
).
SMEDWI-2 is a PIWI-like protein that regulates planarian stem cells
.
Science
310
,
1327
-
1330
.
Rossi
L.
,
Salvetti
A.
,
Lena
A.
,
Batistoni
R.
,
Deri
P.
,
Pugliesi
C.
,
Loreti
E.
,
Gremigni
V.
(
2006
).
DjPiwi-1, a member of the PAZ-Piwi gene family, defines a subpopulation of planarian stem cells
.
Dev. Genes Evol.
216
,
335
-
346
.
Saló
E.
(
2006
).
The power of regeneration and the stem-cell kingdom: freshwater planarians (Platyhelminthes)
.
BioEssays
28
,
546
-
559
.
Saló
E.
,
Baguñà
J.
(
1984
).
Regeneration and pattern formation in planarians. I. The pattern of mitosis in anterior and posterior regeneration in Dugesia (G) tigrina, and a new proposal for blastema formation
.
J. Embryol. Exp. Morphol.
83
,
63
-
80
.
Saló
E.
,
Baguñà
J.
(
1985
).
Cell movement in intact and regenerating planarians. Quantitation using chromosomal, nuclear and cytoplasmic markers
.
J. Embryol. Exp. Morphol.
89
,
57
-
70
.
Saló
E.
,
Abril
J. F.
,
Adell
T.
,
Cebrià
F.
,
Eckelt
K.
,
Fernandez-Taboada
E.
,
Handberg-Thorsager
M.
,
Iglesias
M.
,
Molina
M. D.
,
Rodríguez-Esteban
G.
(
2009
).
Planarian regeneration: achievements and future directions after 20 years of research
.
Int. J. Dev. Biol.
53
,
1317
-
1327
.
Salvetti
A.
,
Rossi
L.
,
Lena
A.
,
Batistoni
R.
,
Deri
P.
,
Rainaldi
G.
,
Locci
M. T.
,
Evangelista
M.
,
Gremigni
V.
(
2005
).
DjPum, a homologue of Drosophila pumilio, is essential to planarian stem cell maintenance
.
Development
132
,
1863
-
1874
.
Sato
K.
,
Shibata
N.
,
Orii
H.
,
Amikura
R.
,
Sakurai
T.
,
Agata
K.
,
Kobayashi
S.
,
Watanabe
K.
(
2006
).
Identification and origin of the germline stem cells as revealed by the expression of nanos-related gene in planarians
.
Dev. Growth Differ.
48
,
615
-
628
.
Schumacher
M. A.
,
Pearson
R. F.
,
Moller
T.
,
Valentin-Hansen
P.
,
Brennan
R. G.
(
2002
).
Structures of the pleiotropic translational regulator Hfq and an Hfq-RNA complex: a bacterial Sm-like protein
.
EMBO J.
21
,
3546
-
3556
.
Slot
J. W.
,
Geuze
H. J.
(
2007
).
Cryosectioning and immunolabeling
.
Nat. Protoc.
2
,
2480
-
2491
.
Solana
J.
,
Lasko
P.
,
Romero
R.
(
2009
).
Spoltud-1 is a chromatoid body component required for planarian long-term stem cell self-renewal
.
Dev. Biol.
328
,
410
-
421
.
Strome
S.
,
Lehmann
R.
(
2007
).
Germ versus soma decisions: lessons from flies and worms
.
Science
316
,
392
-
393
.
Wilusz
C. J.
,
Wilusz
J.
(
2005
).
Eukaryotic Lsm proteins: lessons from bacteria
.
Nat. Struct. Mol. Biol.
12
,
1031
-
1036
.
Yoshida-Kashikawa
M.
,
Shibata
N.
,
Takechi
K.
,
Agata
K.
(
2007
).
DjCBC-1, a conserved DEAD box RNA helicase of the RCK/p54/Me31B family, is a component of RNA-protein complexes in planarian stem cells and neurons
.
Dev. Dyn.
236
,
3436
-
3450
.

Competing interests statement

The authors declare no competing financial interests.

Supplementary information