Skip to main content
  • Minireview
  • Published:

The draft genome sequence of the nematode Caenorhabditis briggsae, a companion to C. elegans

Abstract

The publication of the draft genome sequence of Caenorhabditis briggsae improves the annotation of the genome of its close relative Caenorhabditis elegans and will facilitate comparative genomics and the study of the evolutionary changes during development.

Each genome is fascinating in its own right, but some genomes have been chosen for study because of the added advantages of understanding a closely related genome. Caenorhabditis briggsae (Figure 1a) is just such a companion to Caenorhabditis elegans, a model system that has been widely used to study the genetic basis of animal development, behavior and physiology. The sequencing of the genome of C. elegans [1] has led to rapid advances in our understanding of gene function, particularly through the use of techniques such as RNA interference (RNAi) [2]. Stein et al. [3] now report a draft sequence of the C. briggsae genome.

Figure 1
figure 1

Anatomical comparison of C. briggsae and C. elegans. (a) The C. briggsae hermaphrodite. (b, c)Vulval invagination in L4 larvae of (b) C. elegans and (c) C. briggsae. The vulval morphology of C. briggsae is almost indistinguishable from that of C. elegans; minor differences include a slightly thicker vulval-uterine connection (utse) in C. briggsae.

C. briggsae is closely related to C. elegans (Figure 2) and has almost identical morphology (Figure 1b,c) [4–6]. Given the lack of fossil records, the evolutionary distance between C. briggsae and C. elegans has been estimated using a molecular clock, which gives a divergence time of between 20 and 120 million years ago (Mya) [7–11]. By analyzing the 338 sets of orthologous genes found in the C. briggsae, C. elegans, Anopheles and human genomes and using a molecular clock calibrated by the known date of divergence of nematodes and arthropods, Stein et al. [3] now report a much tighter estimate of the divergence time of C. briggsae and C. elegans, between 80 and 110 Mya. This divergence is slightly greater than the estimate of the human-mouse divergence time (65-75 Mya) [12]. The draft covers 98% of the 104 Mb genome; the slightly larger size of the C. briggsae genome compared with that of C. elegans (100.3 Mb according to the WS108 release of September 2003 [13]) is primarily due to additional repetitive DNA.

Figure 2
figure 2

The phylogenetic relationship of C. briggsae with other nematodes (modified from [3, 6]). Oscheius tipulae (family Rhabditida) and Pristionchus pacificus (Diplogasterida) are two other models used to study evolutionary changes during development [31].

Comparison of genes and non-coding regions

The C. briggsae genome was annotated using various gene-finding programs (such as Genefinder [14]) and by comparison with C. elegans. As different programs often disagree with each other in predicting genes, Stein et al. [3] adopted a 'hybrid' approach by combining the predictions made by multiple gene-finding programs and selecting the consensus. In cases in which a consensus could not be obtained, the authors chose the predictions with best overall similarity with the C. elegans genome. This analysis identified 19,507 genes in C. briggsae. In general, the 'hybrid' approach was twice as accurate as any single gene-prediction program. Conversely, the C. briggsae genome has been extremely useful in the annotation of C. elegans genes: 1,275 new genes were predicted in the C. elegans genome (6% of the new total of 20,621 genes) on the basis of the C. briggsae-C. elegans comparison. Almost 300 of these are confirmed by the open reading frame sequence tag data of Reboul et al. [15]. Gene finding is of supreme importance because essentially all C. elegans genes are being studied, for example by RNAi, and this 6% increase in the number of genes will have enormous impact on the intensive analysis of the organism. Further refinements in C. elegans gene-structure predictions are likely as many more genes are compared carefully in the two species.

Using two different approaches - best reciprocal BLASTP matches and conserved gene order (synteny) - Stein et al. [3] have defined orthologs of C. elegans genes in the C. briggsae genome (62% of predicted C. briggsae genes or 12,155; Figure 3). The identity at the protein level between C. briggsae-C. elegans orthologous pairs (mean of 75%) is comparable to mouse-human orthologs (median 78.5%) [3, 12]. Comparison of the orthologs reveals that about 11% of introns are species-specific, with C. elegans having almost twice as many unique introns (C. elegans has about 4,400 and C. briggsae about 2,200; see also [16]). On average, half of the genes with orthologs differ in the presence of an intron between the two species. Just 4% (807) of the C. briggsae genes do not have significant BLASTP matches in C. elegans (Figure 3); these are likely to be highly divergent and novel (species-specific) genes.

Figure 3
figure 3

A pie chart showing the relationship between C. briggsae and C. elegans genes. BLASTP matches and conserved gene order (synteny) have revealed 62% of C. briggsae genes to have orthologs in C. elegans. About 27% of C. briggsae genes have multiple matches in C. elegans genome and represent various gene families. Of the remaining 11% of genes, 7% have very weak similarity to C. elegans (with a BLASTP E value of over 10-5), whereas 4% appear to be unique.

The C. briggsae genome contains 5,211 genes that have multiple matches in the C. elegans genome (Figure 3) and that correspond to various gene families. Although in most cases such families contain comparable numbers of proteins in the two species (for example, there are 376 protein kinases in C. elegans and 399 in C. briggsae), the olfactory-type chemosensory receptors and the cyclin-like F-box proteins have significantly more members in C. elegans (718 and 243 members, respectively) than in C. briggsae (464 and 98, respectively) [3]. For the chemoreceptors, gene expansion is likely to have occurred in the C. elegans-specific families. The functional significance of such divergence might indicate physiological or ecological differences between these species.

In addition to finding coding exons, blocks of alignable sequence were found using the WABA algorithm [16]. About 1.3 million such blocks were identified, covering coding exons (32.2%), introns (35.3%) and other regions (untranslated regions, intergenic regions, and so on). These conserved regions are a rich resource for further investigation, but it would be useful to have additional genome sequences in order to help discriminate between regions that are conserved because of selection and those that are conserved because they have not yet faded away during neutral evolution. The WABA algorithm distinguishes different codon positions and thus reveals many of the presumed regions of intergenic conservation to have the signature of coding exons. The C. briggsae genome sequence has been a tool to help analyze cis-regulatory sequences for some time already. In general, the C. briggsae-C. elegans comparison has been helpful in locating functionally critical regions of non-coding DNA [17, 18] and may be a good filter for eliminating false positives for some types of bioinformatic searches.

By aligning clear orthologs, Stein et al. [3] were able to evaluate the syntenic relationships for much of the two genomes. Alignment of ortholog pairs reveals over 3,000 rearrangements, and these are ten times as common within a chromosome as between chromosomes. A higher-resolution view of the syntenic relationships between the genomes will require additional physical mapping, or genetic mapping, for example using single-nucleotide polymorphisms. Comparison of the genomic rearrangements between C. briggsae-C. elegans and Drosophila species reveals a breakpoint rate roughly five times higher in nematodes.

One striking feature of nematode genome organization is the existence of trans-spliced operons, in which a primary transcript is processed to give multiple protein-coding transcripts by a trans-splicing reaction with the SL2 spliced leader sequence. There are estimated to be about 1,000 operons in C. elegans (WS108 release) [19]. Of the 800 well-characterized C. elegans operons (WS77 release), 32 are broken in C. briggsae, either by insertion, transposition or rearrangement.

Developmental differences between C. briggsae and C. elegans

As we have seen, despite the morphological similarity, there are significant molecular differences between the C. briggsae and C. elegans genomes. The C. briggsae genome contains about 800 genes with no apparent match in C. elegans [3]. Together with the divergent genes and gene families (Figure 3), nearly one third of the genome is arguably different from C. elegans. Are these changes reflected in significant biological differences? Careful examination has revealed a number of subtle differences between the two species. For example, the excretory system in C. elegans plays a critical role in osmoregulation [20], and comprises three cells including a single duct cell [21]. Although the excretory system looks morphologically identical in the two species, C. briggsae animals have a more anterior duct opening than C. elegans [22]. This difference is the result of the altered expression of lin-48, which encodes a member of the Ovo family of transcription factors. Expression of lin-48 is observed in the excretory duct cell in C. elegans but not in C. briggsae [22, 23]. This is the first example of a morphological difference between the two nematode species with a known molecular basis.

Studies of vulval development in C. briggsae have revealed that, although the overall vulval morphology is similar to that in C. elegans (Figure 1b,c), there are some differences in the underlying mechanisms. In C. elegans, six vulval precursor cells are competent to respond to a gonad-derived inductive signal and to produce the vulval progeny cells [24]. For example, ablation of all vulval precursor cells but the anterior-most one, known as P3.p, allows the latter to generate vulval cells, even though it does not do so in an intact animal. By contrast, P3.p in C. briggsae is not competent in this assay [25]. On the other hand, in certain multivulva mutants of C. briggsae, P3.p is induced and makes vulval progeny cells (B.P.G., unpublished observations). Thus, there is a subtle difference in the competence of the P3.p cell between the two species. Another example of the differences in vulval development comes from the study of the glp-1 gene function in three Caenorhabditis species: C. briggsae, C. elegans and C. remanei. In C. elegans, GLP-1, a receptor of the LIN-12/Notch family, mediates cell-cell communication during development [26]. Loss-of-function mutations in glp-1 do not affect C. elegans vulval development, but in C. briggsae, injection of glp-1 double-stranded RNA causes a multivulva phenotype, suggesting that glp-1 inhibits vulval development in C. briggsae [27].

Nematodes are morphologically diverse and have two different modes of reproduction (hermaphroditic, such as in C. briggsae and C. elegans, and male-female, such as in C. remanei and Caenorhabditis species CB5161); they thus provide an excellent opportunity to study the molecular basis of sexual differentiation and evolution. Several of the known sex-determination genes are remarkably diverged between C. briggsae and C. elegans [28]; for example, FEM-3 and TRA-2, which physically interact with each other, have coevolved and function strictly in a species-specific manner [29]. RNAi targeting the fem-1 and fem-2 genes in C. briggsae have revealed functional differences from C. elegans ([28] and references therein).

A need for more sequences and functional tests

An immediate reward of the C. briggsae genome sequence has been to increase our understanding of C. elegans gene structure. The C. briggsae-C. elegans comparison will inspire many new projects to study gene function and evolutionary changes. Although sequence comparison has revealed a significant number of gene sequences conserved between C. briggsae and C. elegans (62% orthologs; see Figure 3), it is not clear whether all these genes are functionally conserved. Furthermore, conservation in the coding sequences might be misleading if the regulatory elements have evolved differentially (see, for example, [17]). In addition, there are significant differences in some gene families (see above). An understanding of the evolutionary conservation and divergence between the two genomes will require systematic study of the function of C. briggsae genes. The recent success of large-scale RNAi screens in C. elegans raises the promise of a similar approach in C. briggsae. If there is a difference in the RNAi phenotype of a particular gene, however, it will be unclear whether it is a quantitative difference in the use of that gene or whether the gene differs in its susceptibility to RNAi. Thus, classical genetics or targeted gene knockouts are likely to be valuable. To facilitate such experiments, a classical genetic linkage map (B.P.G., R. Johnsen, T. Inoue, A. Mah, G. Jo, D. Baillie and P.W.S., unpublished results) and a single-nucleotide polymorphism map (R. Miller, S. Baird, L. Fulton and R. Waterston, personal communication) are being developed for C. briggsae. These maps will help order contigs in the genome and allow researchers to carry out genetic analysis, as well as helping clone C. briggsae genes with novel mutant phenotypes and study their biological functions.

The C. briggsae genome sequence will probably also help in understanding parasitic nematode genomes such as that of the filarial parasite Brugia malayi. Additional genomes could bridge the gap from C. briggsae and C. elegans to other species (Figure 2). Intensive analysis of transcriptional regulatory networks, in particular cis-regulatory elements, has clearly been helped by the availability of C. briggsae sequence. There is significant value for computational analysis in having additional close nematode genomes (E. Schwarz, J. DeModena, E. Moon, H. Shizuya, B. Wold and P.W.S., unpublished observations). The biological differences between the other nematodes might make the sequence comparisons slightly less informative, but it is possible to test the function of sequences by reciprocal transformation experiments [30], providing a way to know whether there is indeed conservation in gene function. We look forward to the new insights into nematode biology that more genome sequences and further analysis of the C. briggsae genome will bring.

References

  1. The C. elegans sequencing consortium: Genome sequence of the nematode C. elegans : a platform for investigating biology. Science. 1998, 282: 2012-2018. 10.1126/science.282.5396.2012.

    Article  Google Scholar 

  2. Fire A, Xu S, Montgomery MK, Kostas SA, Driver SE, Mello CC: Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature. 1998, 391: 806-811. 10.1038/35888.

    Article  PubMed  CAS  Google Scholar 

  3. Stein LD, Bao Z, Blasiar D, Blumenthal T, Brent M, Chen NS, Chinwalla A, Clarke L, Clee C, Coghlan A, et al: The genome sequence of Caenorhabditis briggsae: a platform for comparative genomics. PLoS Biology. 2003, 1: 166-192. 10.1371/journal.pbio.0000045.

    Article  CAS  Google Scholar 

  4. Nigon V, Dougherty EC: Reproductive patterns and attempts at receiprocal crossing of Rhabditis elegans Maupas, 1900, and Rhabditis briggsae Dougherty & Nigon, 1949 (Nematoda: Rhabditidae). J Exp Zool. 1949, 112: 485-503.

    Article  PubMed  CAS  Google Scholar 

  5. Fodor A, Riddle DL, Nelson FK, Golden JW: Comparison of a new wild-type Caenorhabditis briggsae with laboratory strains of C. briggsae and C. elegans. Nematol. 1983, 29: 203-217.

    Article  Google Scholar 

  6. Blaxter ML, Ley PD, Garey JR, Liu LX, Scheldeman P, Vierstraete A, Vanfleteren JR, Mackey LY, Dorris M, Frisse LM, et al: A molecular evolutionary framework for the phylum Nematoda. Nature. 1998, 392: 71-75. 10.1038/32160.

    Article  PubMed  CAS  Google Scholar 

  7. Coghlan A, Wolfe KH: Fourfold faster rate of genome rearrangement in nematodes than in Drosophila. Genome Res. 2002, 12: 857-867. 10.1101/gr.172702.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  8. Kennedy BP, Aamodt EJ, Allen FL, Chung MA, Heschl MF, McGhee JD: The gut esterase gene (ges-1) from the nematodes Caenorhabditis elegans and Caenorhabditis briggsae. J Mol Biol. 1993, 229: 890-908. 10.1006/jmbi.1993.1094.

    Article  PubMed  CAS  Google Scholar 

  9. Lee YH, Huang XY, Hirsh D, Fox GE, Hecht RM: Conservation of gene organization and trans-splicing in the glyceraldehydes-3-phosphate dehydrogenase-encoding genes of Caenorhabditis briggsae. Gene. 1992, 121: 227-235. 10.1016/0378-1119(92)90126-A.

    Article  PubMed  CAS  Google Scholar 

  10. Heschl MF, Baillie DL: Functional elements and domains inferred from sequence comparisons of a heat shock gene in two nematodes. J Mol Evol. 1990, 31: 3-9.

    Article  PubMed  CAS  Google Scholar 

  11. Prasad SS, Baillie DL: Evolutionarily conserved coding sequences in the dpy-20-unc-22 region of Caenorhabditis elegans. Genomics. 1989, 5: 185-198.

    Article  PubMed  CAS  Google Scholar 

  12. Waterston RH, Lindblad-Toh K, Birney E, Rogers J, Abril JF, Agarwal P, Agarwala R, Ainscough R, Alexandersson M, An P, et al: Mouse genome sequencing consortium. Initial sequencing and comparative analysis of the mouse genome. Nature. 2002, 420: 520-562. 10.1038/nature01262.

    Article  PubMed  CAS  Google Scholar 

  13. Wormbase. [http://www.wormbase.org]

  14. Genefinder. [http://repeatmasker.genome.washington.edu/cgi-bin/Genefinder]

  15. Reboul J, Vaglio P, Rual JF, Lamesch P, Martinez M, Armstrong CM, Li S, Jacotot L, Bertin N, Janky R, et al: C. elegans ORFeome version 1.1: experimental verification of the genome annotation and resource for proteome-scale protein expression. Nat Genet. 2003, 34: 35-41. 10.1038/ng1140.

    Article  PubMed  Google Scholar 

  16. Kent WJ, Zahler AM: Conservation, regulation, synteny, and introns in a large-scale C. briggsae-C. elegans genomic alignment. Genome Res. 2000, 10: 1115-1125. 10.1101/gr.10.8.1115.

    Article  PubMed  CAS  Google Scholar 

  17. Kirouac M, Sternberg PW: cis-regulatory control of three cell fate-specific genes in vulval organogenesis of Caenorhabditis elegans and C. briggsae. Dev Biol. 2003, 257: 85-103. 10.1016/S0012-1606(03)00032-0.

    Article  PubMed  CAS  Google Scholar 

  18. Gower NJD, Temple GR, Schein JE, Marra M, Walker DS, Baylis HA: Dissection of the promoter region of the inositol 1,4,5-trisphosphate receptor gene, itr-1, in C. elegans: A molecular basis for cell-specific expression of IP3R isoforms. J Mol. 2001, 306: 145-157. 10.1006/jmbi.2000.4388.

    Article  CAS  Google Scholar 

  19. Blumenthal T, Gleason KS: Caenorhabditis elegans operons: forms and function. Nat Rev Genet. 2003, 4: 110-118. 10.1038/nrg995.

    Article  Google Scholar 

  20. Nelson FK, Riddle DL: Functional study of the Caenorhabditis elegans secretory-excretory system. J Exp Zool. 1984, 231: 45-56.

    Article  PubMed  CAS  Google Scholar 

  21. Nelson FK, Albert PS, Riddle DL: Fine structure of the Caenorhabditis elegans secretory-excretory system. J Ultrastruct Res. 1983, 82: 156-171.

    Article  PubMed  CAS  Google Scholar 

  22. Wang X, Chamberlin HM: Multiple regulatory changes contribute to the evolution of the Caenorhabditis lin-48 ovo gene. Genes Dev. 2002, 16: 2345-2349. 10.1101/gad.996302.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  23. Johnson AD, Fitzsimmons D, Hagman J, Chamberlin HM: EGL-38 Pax regulates the ovo-related gene lin-48 during Caenorhabditis elegans organ development. Development. 2001, 128: 2857-2865.

    PubMed  CAS  Google Scholar 

  24. Horvitz HR, Sternberg PW: Multiple intercellular signalling systems control the development of the C. elegans vulva. Nature. 1991, 351: 535-541. 10.1038/351535a0.

    Article  PubMed  CAS  Google Scholar 

  25. Delattre M, Felix M-A: Polymorphism and evolution of vulval precursor cell lineages within two nematode genera, Caenorhabditis and Oscheius. Curr Biol. 2001, 11: 631-643. 10.1016/S0960-9822(01)00202-0.

    Article  PubMed  CAS  Google Scholar 

  26. Kimble J, Simpson P: The LIN-12/Notch signaling pathway and its regulation. Annu Rev Cell Dev Biol. 1997, 13: 333-361. 10.1146/annurev.cellbio.13.1.333.

    Article  PubMed  CAS  Google Scholar 

  27. Rudel D, Kimble J: Conservation of glp-1 regulation and function in nematodes. Genetics. 2001, 157: 639-654.

    PubMed  CAS  PubMed Central  Google Scholar 

  28. Stothard P, Pilgrim D: Sex-determination gene and pathway evolution in nematodes. BioEssays. 2003, 25: 221-231. 10.1002/bies.10239.

    Article  PubMed  CAS  Google Scholar 

  29. Haag ES, Wang S, Kimble J: Rapid coevolution of the nematode sex-determining genes fem-3 and tra-2. Curr Biol. 2002, 12: 2035-2041. 10.1016/S0960-9822(02)01333-7.

    Article  PubMed  CAS  Google Scholar 

  30. Ruvinsky I, Ruvkun G: Functional tests of enhancer conservation between distantly related species. Development. 2003, 130: 5133-5142. 10.1242/dev.00711.

    Article  PubMed  CAS  Google Scholar 

  31. Sommer RJ: Comparative genetics: A third model nematode species. Curr Biol. 2000, 10: R879-R881. 10.1016/S0960-9822(00)00815-0.

    Article  PubMed  CAS  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Paul W Sternberg.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Gupta, B.P., Sternberg, P.W. The draft genome sequence of the nematode Caenorhabditis briggsae, a companion to C. elegans. Genome Biol 4, 238 (2003). https://doi.org/10.1186/gb-2003-4-12-238

Download citation

  • Published:

  • DOI: https://doi.org/10.1186/gb-2003-4-12-238

Keywords