WormBook Header Image Embryo series courtesy of Einhard Schierenberg
Home
About
Contact
Cite WormBook
for Authors
Downloads
Sponsors
Links
 HTML   PDF
or Search with Textpresso

RNA-binding proteins*

Min-Ho Lee § , Tim Schedl§
Department of Genetics, Washington University School of Medicine, St. Louis, MO 63110 USA



View/Add Comments

Table of Contents

1. Overview
2. RBPs function during germline and early embryo development
3. RBPs function in somatic development
4. RNA targets of RBPs
5. RNA binding specificity of RBPs
6. Closing remarks
7. Acknowledgements
8. References

Abstract

The C. elegans genome encodes many RNA-binding proteins (RBPs) with diverse functions in development, indicative of extensive layers of post-transcriptional control of RNA metabolism. A number of C. elegans RBPs have been identified by forward or reverse genetics. They tend to display tissue-specific mutant phenotypes, which underscore their functional importance. In addition, several RBPs that bind regulatory sequences in the 3’untranslated regions of mRNAs have been identified molecularly. Most C. elegans RBPs are conserved throughout evolution, suggesting that their study in C. elegans may uncover new conserved biological functions. In this review, we primarily discuss RBPs that are associated with well-characterized mutant phenotypes in the germ line, the early embryo, or in somatic tissues. We also discuss the identification of RNA targets of RBPs, which is an important first step to understand how an RBP controls C. elegans development. It is likely that most RBPs regulate multiple RNA targets. Once multiple RNA targets are identified, specific features that distinguish target from non-target RNAs and the type(s) of RNA metabolism that each RBP controls can be determined. Furthermore, one can determine whether the RBP regulates all targets by the same mechanism or different targets by distinct mechanisms. Such studies will provide insights into how RBPs exert coordinate control of their RNA targets, thereby affecting development in a concerted fashion.

1. Overview

RNA-binding proteins (RBPs) play key roles in post-transcriptional control of RNAs, which, along with transcriptional regulation, is a major way to regulate patterns of gene expression during development. Post-transcriptional control can occur at many different steps in RNA metabolism, including splicing, polyadenylation, mRNA stability, mRNA localization and translation (Curtis et al., 1995; Wickens et al., 2000; de Moor and Richter, 2001; Johnstone and Lasko, 2001). In the C. elegans genome, approximately 500 genes are annotated to encode RBPs (Wormbase), as they have one or more known RNA binding domains such as the RNA Recognition Motif (RRM, also known as RBD or RNP domain), K Homology (KH) domain, Zinc finger (mainly C-x8-C-x5-C-x3-H type), RGG box, DEAD/DEAH box, Pumilio/FBF (PUF) domain, double-stranded RNA binding domain (DS-RBD), Piwi/Argonaute/Zwille (PAZ) domain, Sm domain, etc. Many RBPs have one or more copies of the same RNA binding domain while others have two or more distinct domains. Several RNA binding domains are suggestive for the molecular function of the RBP; DEAD/DEAH box for RNA helicase activity, PAZ domain for short single-stranded RNA binding in RNAi or microRNAs (miRNA) processes, and Sm domain for snRNA binding in splicing and possibly in tRNA processing. However, other domains only predict RNA binding and do not specifically indicate in which aspect of RNA metabolism they may participate. The known function of RBP homologues in other species can provide insights into RBP function in C. elegans. In addition, functional studies of the C. elegans RBPs may reveal unexpected roles for the conserved RBPs. For example, mammalian Y14 is a component of the exon-junction complex that mediates nonsense-mediated mRNA decay (NMD; Kataoka et al., 2000; Fribourg et al., 2003; Singh and Lykke-Andersen, 2003). The C. elegans homologue of Y14, RPN-4, does not appear to mediate NMD, even though it is still preferentially associated with spliced mRNAs. Interestingly, C. elegans Y14, RPN-4, controls germline sex, suggesting that RPN-4 likely has other functions (Kawano et al., 2004).

In C. elegans, many RBPs have been identified genetically, which have tissue-specific mutant phenotypes caused by mutations in RBP genes. In addition, several RBPs that bind regulatory sequences in the 3’untranslated regions (3’UTRs) of mRNAs have been identified by molecular methods (Zhang et al., 1997; Jan et al., 1999; Marin and Evans, 2003; Mootz et al., 2004). Even though the mechanisms by which RBPs influence protein expression patterns in their respective tissues are still poorly understood, the association of many RBPs with mutant phenotypes underscores their importance in C. elegans development (Table 1). In this review, we will primarily limit our discussion to RBPs that are associated with well-characterized mutant phenotypes.

Many RBPs have been identified as essential factors during germline and early embryo development and RBPs with essential functions in the development of somatic tissues, including neuron, muscle, hypodermis, and excretory cells, as well as in the timing of development have been also identified (see Table 1). These findings indicate that RBPs and post-transcriptional control are employed in most aspects of development. The recent emergence of RBPs implicated in RNAi and miRNA processes, such as DCR-1, ALG-1/-2, PPW-1, RDE-1, and RDE-4, further emphasizing the importance of RBPs and the complexity of post-transcriptional control in C. elegans development (Ambros, 2003; Bartel, 2004).

Even though many RBPs have been identified as having critical roles during the development of various tissues in C. elegans, it is largely unclear how such RBPs control development, primarily due to the difficulty in identifying their RNA targets. Most RBPs likely have multiple RNA targets. Therefore, their mutant phenotypes may result from the mis-regulation of many RNA targets, which makes it difficult to identify individual RNA targets using classical genetic approaches. Nevertheless, it is essential to identify a majority of the RNA targets of each RBP in order to fully understand the function of the RBP. Furthermore, the identification of RBPs that bind to known regulatory sequences, such as those defined by mutations, is crucial to understanding the mechanism of post-transcriptional control mediated by such sequences.

Table 1. RNA-binding proteins associated with characterized mutant phenotypes in C. elegans.

RBP gene RNA binding domain Developmental function Post-transcriptional function RNA targets1 References
RBPs function during germline and early embryo development
gld-1 Maxi-KH Female meiotic prophase progression, germline sex determination, entry of meiotic development, germ cell apoptosis Translational repressor Multiple Francis et al., 1995a; Francis et al., 1995b; Jones and Schedl, 1995; Jan et al., 1999; Clifford et al., 2000; Lee and Schedl, 2001; 2004; Xu et al., 2001; Marin and Evans, 2003; Mootz et al., 2004
gld-3 KH-like Gemline sex determination, germline survival, embryo development Translational activator? Unknown Eckmann et al., 2002; Wang et al., 2002
fbf-1/-22 PUF Gemline sex determination, stem cell maintenance Translational repressor fem-3, gld-1, gld-3s, fbf-1, fbf-2 Zhang et al., 1997; Crittenden et al., 2002; Eckmann et al., 2004; Lamont et al., 2004
puf-3 to puf-102 PUF Meiotic division of primary spermatocytes (puf-8) Unknown Unknown Lee and Schedl, 2001; Wickens et al., 2001; Subramaniam and Seydoux, 2003
nos-1/-2 Nanos RNA binding Germline survival (with nos-3?) Unknown Unknown Kraemer et al., 1999; Subramaniam and Seydoux, 1999
nos-3 Nanos RNA binding Gemline sex determination, entry of meiotic development Translational activator? gld-1? Kraemer et al., 1999; Hansen et al., 2004
daz-1 RRM Female meiotic prophase progression Unknown Unknown Karashima et al., 2000
oma-1/-2 (moe-1/-2) Zn-finger (C-x8-C-x5- C-x3-H type) Oocyte maturation Unknown Unknown Detwiler et al., 2001; Shimada et al., 2002
pgl-1 RGG box P-Granule maintenance, early germ cell proliferation/ differentiation Unknown Unknown Kawasaki et al., 1998
glh-1 to glh-4 DEAD/ DEAH box, Zinc knuckle (CCHC) Oogenesis and spermatogenesis (glh-1 and glh-4 redundantly) RNA helicase Unknown Gruidl et al., 1996; Kuznicki et al., 2000
pie-1 Zn-finger (C-x8-C-x5- C-x3-H type) Embryonic germline specification and maintenance Translational activator? nos-2? Mello et al., 1992; Mello et al., 1996; Tenenhaus et al., 2001
mex-3 KH Specifying the fate of the germ line blastomere and control anterior blastomere in early embryo Translational repressor? pal-1 Draper et al., 1996; Huang et al., 2002; Mootz et al., 2004
mex-5/-6 Zn-finger (C-x8-C-x5- C-x3-H type) Soma/germline asymmetry in early embryo Unknown Unknown Schubert et al., 2000
pos-1 Zn-finger (C-x8-C-x5- C-x3-H type) Specifying the fate of the germline blastomere in early embryo Translational repressor glp-1 Tabara et al., 1999a; Ogura et al., 2003
spn-4 RRM Soma/germline asymmetry in early embryo Translational activator glp-1 Gomes et al., 2001; Ogura et al., 2003
fog-1 (cpb-4) RRM, Zn-finger (C2H2 type) Gemline sex determination Cytoplasmic poly(A) element binding? fog-1? Luitjens et al., 2000; Jin et al., 2001a; 2001b
cpb-1 RRM, Zn-finger (C2H2 type) Spermatogenesis Cytoplasmic poly(A) element binding? Unknown Luitjens et al., 2000
mog-1/-4/-5 DEAD/ DEAH box Gemline sex determination, embryo development RNA helicase Unknown Puoti and Kimble, 1999; Puoti and Kimble, 2000
rnp-4 (Ce-Y14) RRM Gemline sex determination, embryo development Exon-junction complex? Unknown Kawano et al., 2004
RBPs function in RNAi/miRNA processes (germ line and soma)
dcr-1 DEAD/ DEAH box, DS-RBD, PAZ Processing of pre-miRNA and long dsRNA, oogenesis RNase endonuclease Pre-miRNA, dsRNA Grishok et al., 2001; Ketting et al., 2001; Knight and Bass, 2001
alg-1/-2 PAZ miRNA dependent translational repression and RNAi RISC complex Short ssRNA Grishok et al., 2001
rde-1 PAZ miRNA dependent translational repression and RNAi RISC complex Short ssRNA Tabara et al., 1999b; Tabara et al., 2002
ppw-1 PAZ Germline RNAi RISC complex Short ssRNA Tijsterman et al., 2002
RBPs function in somatic development
RBP function in primary sex determination
fox-1 RRM Somatic sex determination, X-chromosome counting Unknown Unknown Nicoll et al., 1997; Skipper et al., 1999
RBPs function in neuromuscular system
unc-75 RRM Fine tuning of synaptic transmission Putative splicing factor Unknown Loria et al., 2003
exc-7 RRM Fine tuning of synaptic transmission Putative splicing factor Unknown Loria et al., 2003
etr-1 RRM Muscle formation and function Unknown Unknown Milne and Hodgkin, 1999
mec-8 RRM Specific alternative splices of unc-52 transcripts in hypodermis Putative splicing factor unc-52? Lundquist et al., 1996; Spike et al., 2002
msi-1 RRM Male mating behavior Unknown Unknown Yoda et al., 2000
RBP function in excretory cells
exc-7 RRM Excretory canals development. Putative splicing factor sma-1 Fujita et al., 2003
RBP function as a heterochronic regulator
lin-28 Cold Shock Domain, Zinc knuckle (CCHC) A regulator of developmental timing in the first larval stage Unknown Unknown Moss et al., 1997; Moss and Tang, 2003
Footnotes
Proteins known to bind RBPs and regulate RNA metabolism such as GLD-2 (Wang et al., 2002; Eckmann et al., 2004) and ATX-2 (Kiehl et al., 2000; Ciosk et al., 2004) are not included in this table.
1 Only characterized RNA targets are listed. However, most RBPs likely have additional RNA targets. RNA targets with “?” indicate that the direct relationship between RBPs and RNA targets has not yet been proven.
2 C. elegans has ~10 puf genes while Drosophila and mammals have fewer, suggesting that an expansion in the puf family has occurred in C. elegans. Each PUF may have unique function while different combinations of PUF proteins have redundant functions, suggesting that each PUF likely has its own RNA targets and some of the targets may be regulated by more than one PUF.

2. RBPs function during germline and early embryo development

Many RBPs have essential functions during late germline and early embryo development (Table 1) because post-transcriptional control of maternal mRNAs is a predominant mechanism for temporal/spatial regulation of gene expression during this period (Wickens et al., 2000; de Moor and Richter, 2001; Kuersten and Goodwin, 2003). In general, RBPs bind regulatory sequences that are usually located in 5’UTR and/or 3’UTR of the mRNAs to exert post-transcriptional regulation. During meiotic prophase progression, chromosomes become condensed at diakinesis, and the genome becomes transcriptionally silent (Gibert et al., 1984; Schisa et al., 2001; Kelly et al., 2002). Therefore, during the process of late oogenesis (and late spermatogenesis), fertilization/ meiotic divisions, and early embryogenesis, post-transcriptional control of pre-existing mRNAs, mainly through localization and translational regulation, is the predominant mechanism regulating protein expression. However, post-transcriptional control in the germline also occurs during periods of active transcription (distal mitotic region through the pachytene stage; Crittenden et al., 2002; Marin and Evans, 2003; Hansen et al., 2004; Lee and Schedl, 2001, 2004).

As in other systems, RBP regulatory networks are beginning to be uncovered. For example, FBF-1/-2 are repressors while NOS-3 functions redundantly with a poly-A polymerase, GLD-2, as a putative direct activator of gld-1 mRNA translation (Crittenden et al., 2002; Hansen et al., 2004). GLD-1 in turn is a putative repressor of mex-3 mRNA translation and MEX-3 and GLD-1 are spatially non-overlapping repressors of the translation of the pal-1 mRNA (Mootz et al., 2004). Interestingly, a number of RBPs and some specific mRNAs are localized, at least transiently, to germline specific granules, P granules (Schisa et al., 2001; Barbee et al., 2002 and references therein). However, the function of these associations, particularly in regard to post-transcriptional control, remains to be understood.

Germline sex determination and the proliferation vs. meiotic development decision (stem cell maintenance) rely heavily on post-transcriptional mechanisms in the control of these processes (Table 1). Interestingly, the RBPs, GLD-1, GLD-3, FBF-1/-2, and NOS-3 function in both processes (Francis et al., 1995a; Francis et al., 1995b; Jones and Schedl, 1995; Zhang et al., 1997; Kraemer et al., 1999; Crittenden et al., 2002; Eckmann et al., 2002; Hansen et al., 2004). It is unclear at this time whether the observation that these RBPs function in both processes is due to (1) a temporal/spatial regulatory coordination between the sex determination and the proliferation vs. meiosis decisions or (2) the various RBPs having numerous mRNA targets, and the sex determination and the proliferation vs. meiosis phenotypes are just the processes that have been uncovered genetically so far. Other RBPs are known to function in regulating the sex determination decision, such as FOG-1, MOG-1/-4/-5, and RNP-4 (Ce-Y14) (Puoti et al, 1999, 2000; Jin et al., 2001a; Kawano et al., 2004). It is currently unknown if they also function in the proliferation vs. meiotic development decision.

Several RBPs have been shown to function in meiotic prophase progression, gametogenesis and oocyte maturation such as GLD-1, GLD-3, DAZ-1, OMA-1/-2, PGL-1, and GLH-1/-2/-3/-4. Others (MEX-3, PIE-1, POS-1, GLD-1 and SPN-4) have also been implicated in regulating aspects of early embryo development. These RBPs likely regulate the temporal and spatial pattern of protein expression of subsets of maternal mRNAs, which is crucial in the absence of de novo mRNA synthesis.

3. RBPs function in somatic development

Post-transcriptional control is also important in somatic development as a number of RBPs that have somatic tissue-specific mutant phenotypes have been identified genetically (Table 1). Unlike RBPs that function in germline and early embryo development, several RBPs essential for somatic development are proposed to act as splicing factors, presumably regulating tissue-specific alternative splicing of their mRNA targets. For example, two RRM domain-containing RBPs, MEC-8 and UNC-75, localize to nuclear speckles in the hypodermis and nervous system, respectively (Spike et al., 2002; Loria et al., 2003), consistent with a role in pre-mRNA splicing. Furthermore, MEC-8 regulates the alternative splicing of unc-52 pre-mRNA primarily in the hyperdermis (Spike et al., 2002). Although this regulation is likely direct, it has not yet been proved. In addition, other mec-8 loss-of-function phenotypes are likely independent of unc-52 function, suggesting that MEC-8 regulates additional pre-mRNAs. Another RRM domain-containing RBP, EXC-7, is nuclear localized in embryonic excretory canal cells and throughout the nervous system and is proposed to regulate its mRNA targets through the control of splicing and/or stability (Fujita et al., 2003; Loria et al., 2003).

Somatic RBPs are not limited to regulating RNA processing in the nucleus, as two other somatic RBPs, MSI-1 and LIN-28, are enriched in the cytoplasm (Moss et al., 1997; Yoda et al., 2000). RNAi and miRNA dependent regulation occur during both germline and somatic development and recent discoveries of RBPs functioning in RNAi and miRNA processes, such as DCR-1, ALG-1/-2, PPW-1, RDE-1, and RDE-4 (Table 1), has provided a glimpse into understanding this new mode of post-transcriptional control in C. elegans development.

4. RNA targets of RBPs

Several approaches have been taken to identify RNA targets of RBPs or to identify RBPs that bind to known regulatory sequences (Lundquist et al., 1996; Zhang et al., 1997; Jan et al., 1999; Lee and Schedl, 2001; Xu et al., 2001; Fujita et al., 2003; Marin and Evans, 2003; Mootz et al., 2004). In general, genetic analysis indicates that the RBPs must regulate multiple RNA targets because loss of an RBP has more pleiotropic mutant phenotypes than mis-regulations of the known mRNA targets. For example, fbf-1/-2 loss-of-function phenotype suggests that FBF-1/-2 must bind to other mRNAs to maintain germline stem cells in addition to binding fem-3 mRNA to regulate germline sex. Indeed, FBF-1/-2 have been shown to bind and repress the translation of gld-1 mRNA to regulate germline stem cell proliferation (Crittenden et al., 2002). These results support the view that a comprehensive understanding of how such RBPs control development requires the identification of many or all of their mRNA targets. The identification of multiple mRNA targets will allow one to identify sequences and/or structures that distinguish target from non-target RNAs, as well to determine the type of RNA metabolism the RBP controls.

RNA targets of RBPs have been identified by candidate gene approaches (“educated guess”), as genes that encode putative targets that have similar (or opposite) mutant phenotypes to that of RBP, and/or by studying the expression patterns of the RBPs and the proteins of their putative mRNA targets. For example, 1) Mutations in mec-8 strongly enhance the phenotype of specific mutations in unc-52 and MEC-8 controls alternative splicing of unc-52 transcripts (Lundquist et al., 1996; Spike et al., 2002). 2) exc-7 mutants exhibit synergistic excretory canal defects with mutations in sma-1 and EXC-7 binds to sma-1 mRNA (Fujita et al., 2003). 3) GLP-1 is mis-expressed in pos-1 mutant embryos and POS-1 binds to the glp-1 3’UTR (Ogura et al., 2003). 4) GLD-1 is a cytoplasmic RNA binding protein that represses the translation of several mRNA targets in the distal germline. It was also known that mes-3 and pal-1 mRNAs are translationally repressed in the distal germ line. Therefore it was directly tested and confirmed that GLD-1 binds and represses the translation of mes-3 and pal-1 mRNAs (Xu et al., 2001; Mootz et al., 2004). Because indirect regulation by the RBP also fits the criteria used to predict the candidate mRNA target, these approaches require the experimental test of whether the RBP binds to the candidate mRNA target directly.

Gain-of-function mutations in the fem-3 and the tra-2 3’UTRs (Ahringer and Kimble, 1991; Goodwin et al., 1993), which result in deregulated expression, provided a starting point for molecular identification of RBPs that function in translational repression of these mRNAs. For example, FBF-1/-2 were identified by the yeast three-hybrid system using the 3’UTR of fem-3 mRNA as a bait. FBF-1/-2 bind to the 3’UTR of fem-3 to repress the male sexual fate in the C. elegans hermaphrodite germline (Zhang et al., 1997). GLD-1 was found to bind to the tra-2 3’UTR, also by using the yeast three-hybrid system (Jan et al., 1999). GLD-1 was also identified as binding to the glp-1 3’UTR biochemically (Marin and Evans, 2003).

More direct methods to identify multiple endogenous RNAs associated with RBPs using high-throughput gene array technologies have been described (Tenenbaum et al., 2000, 2002; Brown et al., 2001). They are based on the isolation of endogenous RNA-protein complexes under optimized conditions mostly by immunoprecipitation (IP). The isolated endogenous RNAs are then identified by microarray analysis. This approach has the potential to identify most RNA targets of an RPB without prior knowledge of mutant phenotypes or expression patterns. Mis-regulations of such RNA targets may have small effects or contributions to phenotypes of the RBP. This approach can also uncover mRNA targets from functionally redundant paralagous genes that are co-regulated by an RBP (Lee and Schedl, 2001, 2004).

Sixteen mRNA targets of GLD-1 have been identified by a similar approach (Lee and Schedl, 2001, 2004). Functional GLD-1 was immunoprecipitated from cytosol extracts. The isolated mRNAs were identified after subtractive hybridization/cloning/sequencing. GLD-1 represses the translation of several targets through direct binding (Lee and Schedl, 2001, 2004). Recently, an essentially identical experiment was performed with microarray analysis to detect specifically enriched mRNAs in the GLD-1 IP. This resulted in the identification of significantly more mRNAs (129), which are enriched more than two fold (p < 0.05) in the GLD-1 IP over the control IP (Lee, M.-H., Reinke, V., and Schedl, T., unpubl. data). Essentially, all targets identified previously were identified again. These and other results indicate that most of the 129 mRNA targets are likely to interact with GLD-1 in vivo and demonstrate that there is a significant increase in mRNA target detection power with the microarray analysis compared with the subtractive hybridization/cloning/sequencing strategy.

Several efforts are currently undertaken to identify in vivo mRNA targets of RBPs, PIE-1 (D'Agostino, I., Reinke, V., and Seydoux, G., pers. comm.), DAZ-1 (Karashima, T., Sugioka, K., Uesugi, K, Kohara, Y., and Yamamoto, M., pers. comm.), and LIN-28 (Moss, E. and Kemper, K., pers. comm.), or to identify in vivo mRNA targets of a putative cytoplasmic poly-A polymerase, GLD-2, that interacts with RBPs (Suh, N. and Kimble, J, pers. comm.).

5. RNA binding specificity of RBPs

One of the important questions to understand the function of RBPs in C. elegans development is how RBPs distinguish their targets from non-targets in vivo. In other words, how RBPs specifically recognize their RNA targets. RBPs may recognize specific sequences, structures, or both, which are present in their RNA targets. Understanding of RNA binding specificity of an RBP can be another way to identify unknown targets that contain similar features, if they have enough information to distinguish targets from non-targets computationally. It will also provide important tools to find the molecular mechanism of the post-transcriptional process that the RBP controls. RNA binding specificity of most RBPs is unknown because their RNA targets are not identified yet. However, RNA binding specificity of GLD-1 and FBF-1/-2 in their mRNA targets are beginning to emerge.

GLD-1 has been shown to form a homodimer to bind to two close sites in a single TGE (tra-2/GLI element) of tra-2 3’UTR in vitro where one site has a hexanucleotide consensus (Ryder et al., 2004). Among other mRNA targets of GLD-1 identified in the GLD-1 IP, thirteen independent GLD-1 binding regions in seven targets have been found where GLD-1 binds 5’UTR, 3’UTR or in the open reading frame (ORF) depending on the target (Lee and Schedl, 2001, 2004, unpubl. data). Interestingly, many GLD-1 binding regions have the hexanucleotide consensus, suggesting it is likely important for GLD-1 binding. However, several GLD-1 binding regions do not contain the hexanucleotide consensus, suggesting that other distinct features that dictate GLD-1 binding likely exist (Lee, M.-H., and Schedl, T., unpubl. data).

FBF-1/-2 have been shown to bind to sites that carry crucial UGU(G/A) motif in the 3’UTRs of their mRNA targets (Zhang et al., 1997; Crittenden et al., 2002; Eckmann et al., 2004; Lamont et al., 2004). It is interesting to note that FBF-1/-2 do not bind to all potential sites that contain UGU(G/A) motif in the 3’UTRs of their mRNA targets (Eckmann et al., 2004; Lamont et al., 2004). This suggests that this motif alone is not sufficient to be recognized by FBF-1/-2 and other distinct features should exist.

6. Closing remarks

The C. elegans genome encodes many RBPs with diverse functions in development, revealing a large layer of post-transcriptional control of RNA metabolism, which had not been previously appreciated in the control of gene expression. Most RBPs with tissue-specific functions are conserved throughout evolution, suggesting that their studies in C. elegans may uncover new conserved biological functions.

In general, RBPs likely regulate multiple RNA targets and the identification of many or all RNA targets of RBPs will be an important first step in a comprehensive understanding of how RBPs control C. elegans development. With multiple RNA targets, specific features that distinguish target from non-target RNAs, as well as the type(s) of RNA metabolism that each RBP controls can be determined. Furthermore, with the identity of multiple RNA targets, one can determine whether the RBP regulates all targets by the same mechanism or different targets by distinct mechanisms. Such studies will provide insights into how RBPs exert coordinate control of their RNA targets, thereby affecting development in a concerted fashion.

7. Acknowledgements

We are grateful to Sudhir Nayak for providing information about genes that are annotated to encode RNA-binding proteins in Wormbase and Dave Hansen and members of the Schedl lab for comments on this review. TS is supported by NIH grant GM63310.

8. References

Ahringer, J., and Kimble, J. (1991). Control of the sperm-oocyte switch in Caenorhabditis elegans hermaphrodites by the fem-3 3' untranslated region. Nature 349, 346–348. Abstract Article

Ambros, V. (2003). MicroRNA pathways in flies and worms: growth, death, fat, stress, and timing. Cell 113, 673–676. Abstract Article

Barbee, S.A., Lublin, A.L., and Evans, T.C. (2002). A novel function for the Sm proteins in germ granule localization during C. elegans embryogenesis. Curr. Biol. 12, 1502–1506. Abstract Article

Bartel, D.P. (2004). MicroRNAs: genomics, biogenesis, mechanism, and function. Cell 116, 281–297. Abstract Article

Brown, V., Jin, P., Ceman, S., Darnell, J.C., O'Donnell, W.T., Tenenbaum, S.A., Jin, X., Feng, Y., Wilkinson, K.D., Keene, J.D., et al. (2001). Microarray identification of FMRP-associated brain mRNAs and altered mRNA translational profiles in fragile X syndrome. Cell 107, 477–487. Abstract Article

Ciosk, R., DePalma, M., and Priess, J.R. (2004). ATX-2, the C. elegans ortholog of ataxin 2, functions in translational regulation in the germline. Development 131, 4831–4841. Abstract Article

Clifford, R., Lee, M.H., Nayak, S., Ohmachi, M., Giorgini, F., and Schedl, T. (2000). FOG-2, a novel F-box containing protein, associates with the GLD-1 RNA binding protein and directs male sex determination in the C. elegans hermaphrodite germline. Development 127, 5265–5276. Abstract

Crittenden, S.L., Bernstein, D.S., Bachorik, J.L., Thompson, B.E., Gallegos, M., Petcherski, A.G., Moulder, G., Barstead, R., Wickens, M., and Kimble, J. (2002). A conserved RNA-binding protein controls germline stem cells in Caenorhabditis elegans. Nature 417, 660–663. Abstract Article

Curtis, D., Lehmann, R., and Zamore, P.D. (1995). Translational regulation in development. Cell 81, 171–178. Abstract Article

de Moor, C.T., and Richter, J.D. (2001). Translational control in vertbrate development. In Cell lineage specification and patterning of the embryo, L.D. Etkin, and K.W. Jeon, eds. (San Diego, Academic Press), pp. 567–608. Abstract

Detwiler, M.R., Reuben, M., Li, X., Rogers, E., and Lin, R. (2001). Two zinc finger proteins, OMA-1 and OMA-2, are redundantly required for oocyte maturation in C. elegans. Dev. Cell 1, 187–199. Abstract Article

Draper, B.W., Mello, C.C., Bowerman, B., Hardin, J., and Priess, J.R. (1996). MEX-3 is a KH domain protein that regulates blastomere identity in early C. elegans embryos. Cell 87, 205–216. Abstract Article

Eckmann, C.R., Crittenden, S.L., Suh, N., and Kimble, J. (2004). GLD-3 and control of the mitosis/meiosis decision in the germline of Caenorhabditis elegans. Genetics 168, 147–160. Abstract Article

Eckmann, C.R., Kraemer, B., Wickens, M., and Kimble, J. (2002). GLD-3, a bicaudal-C homolog that inhibits FBF to control germline sex determination in C. elegans. Dev. Cell 3, 697–710. Abstract Article

Francis, R., Barton, M.K., Kimble, J., and Schedl, T. (1995a). gld-1, a tumor suppressor gene required for oocyte development in Caenorhabditis elegans. Genetics 139, 579–606. Abstract

Francis, R., Maine, E., and Schedl, T. (1995b). Analysis of the multiple roles of gld-1 in germline development: interactions with the sex determination cascade and the glp-1 signaling pathway. Genetics 139, 607–630. Abstract

Fribourg, S., Gatfield, D., Izaurralde, E., and Conti, E. (2003). A novel mode of RBD-protein recognition in the Y14-Mago complex. Nat. Struct. Biol. 10, 433–439. Abstract Article

Fujita, M., Hawkinson, D., King, K.V., Hall, D.H., Sakamoto, H., and Buechner, M. (2003). The role of the ELAV homologue EXC-7 in the development of the Caenorhabditis elegans excretory canals. Dev. Biol. 256, 290–301. Abstract Article

Gibert, M.A., Starck, J., and Beguet, B. (1984). Role of the gonad cytoplasmic core during oogenesis of the nematode Caenorhabditis elegans. Biol. Cell 50, 77–85. Abstract

Gomes, J.E., Encalada, S.E., Swan, K.A., Shelton, C.A., Carter, J.C., and Bowerman, B. (2001). The maternal gene spn-4 encodes a predicted RRM protein required for mitotic spindle orientation and cell fate patterning in early C. elegans embryos. Development 128, 4301–4314. Abstract

Goodwin, E.B., Okkema, P.G., Evans, T.C., and Kimble, J. (1993). Translational regulation of tra-2 by its 3' untranslated region controls sexual identity in C. elegans. Cell 75, 329–339. Abstract Article

Grishok, A., Pasquinelli, A.E., Conte, D., Li, N., Parrish, S., Ha, I., Baillie, D.L., Fire, A., Ruvkun, G., and Mello, C.C. (2001). Genes and mechanisms related to RNA interference regulate expression of the small temporal RNAs that control C. elegans developmental timing. Cell 106, 23–34. Abstract Article

Gruidl, M.E., Smith, P.A., Kuznicki, K.A., McCrone, J.S., Kirchner, J., Roussell, D.L., Strome, S., and Bennett, K.L. (1996). Multiple potential germ-line helicases are components of the germ-line-specific P granules of Caenorhabditis elegans. Proc. Natl. Acad. Sci. USA 93, 13837–13842. Abstract Article

Hansen, D., Wilson-Berry, L., Dang, T., and Schedl, T. (2004). Control of the proliferation versus meiotic development decision in the C. elegans germline through regulation of GLD-1 protein accumulation. Development 131, 93–104. Abstract Article

Huang, N.N., Mootz, D.E., Walhout, A.J., Vidal, M., and Hunter, C.P. (2002). MEX-3 interacting proteins link cell polarity to asymmetric gene expression in Caenorhabditis elegans. Development 129, 747–759. Abstract

Jan, E., Motzny, C.K., Graves, L.E., and Goodwin, E.B. (1999). The STAR protein, GLD-1, is a translational regulator of sexual identity in C. elegans. EMBO J. 18, 258–269. Abstract Article

Jin, S.W., Arno, N., Cohen, A., Shah, A., Xu, Q., Chen, N., and Ellis, R.E. (2001a). In Caenorhabditis elegans, the RNA-binding domains of the cytoplasmic polyadenylation element binding protein FOG-1 are needed to regulate germ cell fates. Genetics 159, 1617–1630. Abstract Article

Jin, S.W., Kimble, J., and Ellis, R.E. (2001b). Regulation of cell fate in Caenorhabditis elegans by a novel cytoplasmic polyadenylation element binding protein. Dev. Biol. 229, 537–553. Abstract Article

Johnstone, O., and Lasko, P. (2001). Translational regulation and RNA localization in Drosophila oocytes and embryos. Annu. Rev. Genet. 35, 365–406. Abstract Article

Jones, A.R., and Schedl, T. (1995). Mutations in gld-1, a female germ cell-specific tumor suppressor gene in Caenorhabditis elegans, affect a conserved domain also found in Src-associated protein Sam68. Genes Dev. 9, 1491–1504. Abstract

Karashima, T., Sugimoto, A., and Yamamoto, M. (2000). Caenorhabditis elegans homologue of the human azoospermia factor DAZ is required for oogenesis but not for spermatogenesis. Development 127, 1069–1079. Abstract

Kataoka, N., Yong, J., Kim, V.N., Velazquez, F., Perkinson, R.A., Wang, F., and Dreyfuss, G. (2000). Pre-mRNA splicing imprints mRNA in the nucleus with a novel RNA-binding protein that persists in the cytoplasm. Mol. Cell 6, 673–682. Abstract Article

Kawano, T., Kataoka, N., Dreyfuss, G., and Sakamoto, H. (2004). Ce-Y14 and MAG-1, components of the exon-exon junction complex, are required for embryogenesis and germline sexual switching in Caenorhabditis elegans. Mech. Dev. 121, 27–35. Abstract Article

Kawasaki, I., Shim, Y.H., Kirchner, J., Kaminker, J., Wood, W.B., and Strome, S. (1998). PGL-1, a predicted RNA-binding component of germ granules, is essential for fertility in C. elegans. Cell 94, 635–645. Abstract Article

Kelly, W.G., Schaner, C.E., Dernburg, A.F., Lee, M.H., Kim, S.K., Villeneuve, A.M., and Reinke, V. (2002). X-chromosome silencing in the germline of C. elegans. Development 129, 479–492. Abstract

Ketting, R.F., Fischer, S.E., Bernstein, E., Sijen, T., Hannon, G.J., and Plasterk, R.H. (2001). Dicer functions in RNA interference and in synthesis of small RNA involved in developmental timing in C. elegans. Genes Dev. 15, 2654–2659. Abstract Article

Kiehl, T.R., Shibata, H., and Pulst, S.M. (2000). The ortholog of human ataxin-2 is essential for early embryonic patterning in C. elegans. J. Mol. Neurosci. 15, 231–241. Abstract Article

Knight, S.W., and Bass, B.L. (2001). A role for the RNase III enzyme DCR-1 in RNA interference and germ line development in Caenorhabditis elegans. Science. 293, 2269–2271. Abstract Article

Kraemer, B., Crittenden, S., Gallegos, M., Moulder, G., Barstead, R., Kimble, J., and Wickens, M. (1999). NANOS-3 and FBF proteins physically interact to control the sperm- oocyte switch in Caenorhabditis elegans. Curr. Biol. 9, 1009–1018. Abstract Article

Kuersten, S., and Goodwin, E.B. (2003). The power of the 3' UTR: translational control and development. Nat. Rev. Genet. 4, 626–637. Abstract Article

Kuznicki, K.A., Smith, P.A., Leung-Chiu, W.M., Estevez, A.O., Scott, H.C., and Bennett, K.L. (2000). Combinatorial RNA interference indicates GLH-4 can compensate for GLH-1; these two P granule components are critical for fertility in C. elegans. Development 127, 2907–2916. Abstract

Lamont, L.B., Crittenden, S.L., Bernstein, D., Wickens, M., and Kimble, J. (2004). FBF-1 and FBF-2 regulate the size of the mitotic region in the C. elegans germline. Dev. Cell 7, 697–707. Abstract Article

Lee, M.H., and Schedl, T. (2001). Identification of in vivo mRNA targets of GLD-1, a maxi-KH motif containing protein required for C. elegans germ cell development. Genes Dev. 15, 2408–2420. Abstract Article

Lee, M.H., and Schedl, T. (2004). Translation repression by GLD-1 protects its mRNA targets from nonsense-mediated mRNA decay in C. elegans. Genes Dev. 18, 1047–1059. Abstract Article

Loria, P.M., Duke, A., Rand, J.B., and Hobert, O. (2003). Two neuronal, nuclear-localized RNA binding proteins involved in synaptic transmission. Curr. Biol. 13, 1317–1323. Abstract Article

Luitjens, C., Gallegos, M., Kraemer, B., Kimble, J., and Wickens, M. (2000). CPEB proteins control two key steps in spermatogenesis in C. elegans. Genes Dev. 14, 2596–2609. Abstract Article

Lundquist, E.A., Herman, R.K., Rogalski, T.M., Mullen, G.P., Moerman, D.G., and Shaw, J.E. (1996). The mec-8 gene of C. elegans encodes a protein with two RNA recognition motifs and regulates alternative splicing of unc-52 transcripts. Development 122, 1601–1610. Abstract

Marin, V.A., and Evans, T.C. (2003). Translational repression of a C. elegans Notch mRNA by the STAR/KH domain protein GLD-1. Development 130, 2623–2632. Abstract Article

Mello, C.C., Draper, B.W., Krause, M., Weintraub, H., and Priess, J.R. (1992). The pie-1 and mex-1 genes and maternal control of blastomere identity in early C. elegans embryos. Cell 70, 163–176. Abstract Article

Mello, C.C., Schubert, C., Draper, B., Zhang, W., Lobel, R., and Priess, J.R. (1996). The PIE-1 protein and germline specification in C. elegans embryos. Nature 382, 710–712. Abstract Article

Milne, C.A., and Hodgkin, J. (1999). ETR-1, a homologue of a protein linked to myotonic dystrophy, is essential for muscle development in Caenorhabditis elegans. Curr. Biol. 9, 1243–1246. Abstract Article

Mootz, D., Ho, D.M., and Hunter, C.P. (2004). The STAR/Maxi-KH domain protein GLD-1 mediates a developmental switch in the translational control of C. elegans PAL-1. Development 131, 3263–3272. Abstract Article

Moss, E.G., Lee, R.C., and Ambros, V. (1997). The cold shock domain protein LIN-28 controls developmental timing in C. elegans and is regulated by the lin-4 RNA. Cell 88, 637–646. Abstract

Moss, E.G., and Tang, L. (2003). Conservation of the heterochronic regulator Lin-28, its developmental expression and microRNA complementary sites. Dev. Biol. 258, 432–442. Abstract Article

Nicoll, M., Akerib, C.C., and Meyer, B.J. (1997). X-chromosome-counting mechanisms that determine nematode sex. Nature 388, 200–204. Abstract Article

Ogura, K., Kishimoto, N., Mitani, S., Gengyo-Ando, K., and Kohara, Y. (2003). Translational control of maternal glp-1 mRNA by POS-1 and its interacting protein SPN-4 in Caenorhabditis elegans. Development 130, 2495–2503. Abstract Article

Puoti, A., and Kimble, J. (1999). The Caenorhabditis elegans sex determination gene mog-1 encodes a member of the DEAH-Box protein family. Mol. Cell Biol. 19, 2189–2197. Abstract

Puoti, A., and Kimble, J. (2000). The hermaphrodite sperm/oocyte switch requires the Caenorhabditis elegans homologs of PRP2 and PRP22. Proc. Natl. Acad. Sci. USA 97, 3276–3281. Abstract Article

Ryder, S.P., Frater, L.A., Abramovitz, D.L., Goodwin, E.B., and Williamson, J.R. (2004). RNA target specificity of the STAR/GSG domain post-transcriptional regulatory protein GLD-1. Nat. Struct. Mol. Biol. 11, 20–28. Abstract Article

Schisa, J.A., Pitt, J.N., and Priess, J.R. (2001). Analysis of RNA associated with P granules in germ cells of C. elegans adults. Development 128, 1287–1298. Abstract

Schubert, C.M., Lin, R., de Vries, C.J., Plasterk, R.H., and Priess, J.R. (2000). MEX-5 and MEX-6 function to establish soma/germline asymmetry in early C. elegans embryos. Mol. Cell 5, 671–682. Abstract Article

Shimada, M., Kawahara, H., and Doi, H. (2002). Novel family of CCCH-type zinc-finger proteins, MOE-1, -2 and -3, participates in C. elegans oocyte maturation. Genes Cells 7, 933–947. Abstract Article

Singh, G., and Lykke-Andersen, J. (2003). New insights into the formation of active nonsense-mediated decay complexes. Trends Biochem. Sci. 28, 464–466. Abstract Article

Skipper, M., Milne, C.A., and Hodgkin, J. (1999). Genetic and molecular analysis of fox-1, a numerator element involved in Caenorhabditis elegans primary sex determination. Genetics 151, 617–631. Abstract

Spike, C.A., Davies, A.G., Shaw, J.E., and Herman, R.K. (2002). MEC-8 regulates alternative splicing of unc-52 transcripts in C. elegans hypodermal cells. Development 129, 4999–5008. Abstract

Subramaniam, K., and Seydoux, G. (1999). nos-1 and nos-2, two genes related to Drosophila nanos, regulate primordial germ cell development and survival in Caenorhabditis elegans. Development 126, 4861–4871. Abstract

Subramaniam, K., and Seydoux, G. (2003). Dedifferentiation of primary spermatocytes into germ cell tumors in C. elegans lacking the pumilio-like protein PUF-8. Curr. Biol. 13, 134–139. Abstract Article

Tabara, H., Hill, R.J., Mello, C.C., Priess, J.R., and Kohara, Y. (1999a). pos-1 encodes a cytoplasmic zinc-finger protein essential for germline specification in C. elegans. Development 126, 1–11. Abstract

Tabara, H., Sarkissian, M., Kelly, W.G., Fleenor, J., Grishok, A., Timmons, L., Fire, A., and Mello, C.C. (1999b). The rde-1 gene, RNA interference, and transposon silencing in C. elegans. Cell 99, 123–132. Abstract Article

Tabara, H., Yigit, E., Siomi, H., and Mello, C.C. (2002). The dsRNA binding protein RDE-4 interacts with RDE-1, DCR-1, and a DExH-box helicase to direct RNAi in C. elegans. Cell 109, 861–871. Abstract Article

Tenenbaum, S.A., Carson, C.C., Lager, P.J., and Keene, J.D. (2000). Identifying mRNA subsets in messenger ribonucleoprotein complexes by using cDNA arrays. Proc. Natl. Acad. Sci. USA 97, 14085–14090. Abstract Article

Tenenbaum, S.A., Lager, P.J., Carson, C.C., and Keene, J.D. (2002). Ribonomics: identifying mRNA subsets in mRNP complexes using antibodies to RNA-binding proteins and genomic arrays. Methods 26, 191–198. Abstract Article

Tenenhaus, C., Subramaniam, K., Dunn, M.A., and Seydoux, G. (2001). PIE-1 is a bifunctional protein that regulates maternal and zygotic gene expression in the embryonic germ line of Caenorhabditis elegans. Genes Dev. 15, 1031–1040. Abstract Article

Tijsterman, M., Okihara, K.L., Thijssen, K., and Plasterk, R.H. (2002). PPW-1, a PAZ/PIWI protein required for efficient germline RNAi, is defective in a natural isolate of C. elegans. Curr. Biol. 12, 1535–1540. Abstract Article

Wang, L., Eckmann, C.R., Kadyk, L.C., Wickens, M., and Kimble, J. (2002). A regulatory cytoplasmic poly(A) polymerase in Caenorhabditis elegans. Nature 419, 312–316. Abstract Article

Wickens, M., Bernstein, D., Crittenden, S., Luitjens, C., and Kimble, J. (2001). PUF proteins and 3'UTR regulation in the Caenorhabditis elegans germ line. Cold Spring Harb. Symp. Quant. Biol. 66, 337–343. Abstract Article

Wickens, M., Goodwin, E.B., Kimble, J., Strickland, S., and Hentze, M.W. (2000). Translational control of developmental decisions. In Translational control of gene expression, N. Sorenberg, J.W.B. Hershey, and M.B. Mathews, eds. (Cold Spring Harbor: Cold Spring Harbor Laboratory Press), pp. 295–370.

Xu, L., Paulsen, J., Yoo, Y., Goodwin, E.B., and Strome, S. (2001). Caenorhabditis elegans MES-3 is a target of GLD-1 and functions epigenetically in germline development. Genetics 159, 1007–1017. Abstract

Yoda, A., Sawa, H., and Okano, H. (2000). MSI-1, a neural RNA-binding protein, is involved in male mating behaviour in Caenorhabditis elegans. Genes Cells 5, 885–895. Abstract Article

Zhang, B., Gallegos, M., Puoti, A., Durkin, E., Fields, S., Kimble, J., and Wickens, M.P. (1997). A conserved RNA-binding protein that regulates sexual fates in the C. elegans hermaphrodite germ line. Nature 390, 477–484. Abstract Article



*Edited by Thomas Blumenthal. Last revised February 8, 2005. Published April 18, 2006. This chapter should be cited as: Lee, M.-H. and Schedl, T. RNA-binding proteins (April 18, 2006), WormBook, ed. The C. elegans Research Community, WormBook, doi/10.1895/wormbook.1.79.1, http://www.wormbook.org.

Copyright: © 2006 Min-Ho Lee and Tim Schedl. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.

§To whom correspondence should be addressed. E-mail: mhlee@albany.edu or ts@genetics.wustl.edu

Current address: Department of Biological Sciences, University at Albany, SUNY, Albany, NY 12222 USA

Creative Commons License All WormBook content, except where otherwise noted, is licensed under a Creative Commons Attribution License.