For skeletal muscle to produce movement, individual myofibers must form stable contacts with tendon cells and then assemble sarcomeres. The myofiber precursor is the nascent myotube, and during myogenesis the myotube completes guided elongation to reach its target tendons. Unlike the well-studied events of myogenesis, such as myoblast specification and myoblast fusion, the molecules that regulate myotube elongation are largely unknown. In Drosophila, hoi polloi (hoip) encodes a highly conserved RNA-binding protein and hoip mutant embryos are largely paralytic due to defects in myotube elongation and sarcomeric protein expression. We used the hoip mutant background as a platform to identify novel regulators of myogenesis, and uncovered surprising developmental functions for the sarcomeric protein Tropomyosin 2 (Tm2). We have identified Hoip-responsive sequences in the coding region of the Tm2 mRNA that are essential for Tm2 protein expression in developing myotubes. Tm2 overexpression rescued the hoip myogenic phenotype by promoting F-actin assembly at the myotube leading edge, by restoring the expression of additional sarcomeric RNAs, and by promoting myoblast fusion. Embryos that lack Tm2 also showed reduced sarcomeric protein expression, and embryos that expressed a gain-of-function Tm2 allele showed both fusion and elongation defects. Tropomyosin therefore dictates fundamental steps of myogenesis prior to regulating contraction in the sarcomere.

Nascent myotubes are faced with two major obstacles before they can form functional, contractile myofibers. First, the myotube must elongate over several cell diameters to identify and contact the appropriate tendon cells (Schnorrer and Dickson, 2004). Second, the myotube must express a wide range of muscle structural proteins and assemble those proteins into contractile sarcomeres (Rui et al., 2010). Although the mechanisms that regulate myotube precursor specification and muscle structural gene transcription have been characterized in detail (Buckingham and Rigby, 2014; Ciglar and Furlong, 2009), the molecules that coordinate myotube elongation are poorly understood.

The genetic tools in Drosophila have identified conserved cellular and molecular processes that direct striated muscle development. Drosophila somatic muscle is analogous to vertebrate skeletal muscle, and somatic muscle development initiates with the specification of myoblasts known as founder cells, that then fuse with neighboring fusion-competent cells to form nascent, multinucleate myotubes (Chen and Olson, 2004; de Joussineau et al., 2012). The nascent myotubes then elongate, activate muscle structural gene expression, attach to tendons, and assemble sarcomeres to form a functional myofiber (Rui et al., 2010; Schejter and Baylies, 2010).

Tropomyosin is a sarcomeric protein that binds thin filament actin to regulate contractions. Outside of the sarcomere, Tropomyosin also directs cytoskeletal dynamics in migratory cells. Membrane protrusions at the leading edge of migrating cells are driven by actin- polymerizing proteins in the lamellipodia, and the physical force for cell movement is derived from lamellar expansion (Ponti et al., 2004). Tropomyosins are essential actin-stabilizing proteins in the lamella that act in concert with the actin-polymerizing proteins at the leading edge, including the Wiskott–Aldrich syndrome proteins (WASP) and the actin-related proteins (ARP)2/3 complex, to drive both normal and metastatic cell migration (Bugyi and Carlier, 2010; Gross, 2014). With respect to myogenesis, the WASP and ARP2/3 protein complexes are essential for myoblast fusion (Baas et al., 2012; Berger et al., 2008; Richardson et al., 2007), and WASP proteins promote myoblast migration (Kawamura et al., 2004). Although the mechanisms that direct myotube elongation are thought to mimic those of migratory cells (Bongiovanni et al., 2012), the role of Tropomyosin during myotube elongation has not been characterized.

We previously identified the RNA-binding protein Hoi polloi (Hoip) in a screen for regulators of Drosophila myotube elongation (Johnson et al., 2013). hoip mutant embryos are largely paralytic due to defects in myotube elongation, as well as defects in myoblast fusion and sarcomeric gene expression. The Drosophila genome encodes two Tropomyosin isoforms, Tm1 and Tm2, and both isoforms were dramatically reduced in hoip embryos. Here, we show Tm2 co-localized with F-actin during myotube elongation, and Hoip-responsive sequences in Tm2 mRNA are required for Tm2 protein expression in somatic muscle. Tm2 overexpression rescued hoip myogenic defects by restoring F-actin during myotube elongation, by promoting myoblast fusion, and by enhancing the expression of additional sarcomeric proteins. Tm2 null embryos showed reduced sarcomeric protein expression, and a gain-of-function Tm2 allele disrupted myotube elongation and myoblast fusion. Tropomyosin is therefore an essential regulator of myogenesis prior to sarcomere assembly.

Tm2 expression and localization suggests a novel function

hoip1 mutant embryos showed two major myogenic phenotypes. A subset of somatic myotubes [including lateral longitudinal (LL1), dorsal oblique (DO3-5), lateral transverse (LT1-4), and lateral oblique (LO1)] failed to elongate (Fig. 1A; see supplementary material Fig. S1A for a diagram of somatic muscle), and sarcomeric RNAs were dramatically downregulated in striated muscles (Johnson et al., 2013). Myoblast fusion was also affected in hoip1 embryos. A mechanism that could explain the hoip mutant phenotype is that Hoip promotes the expression of cytoskeletal regulatory proteins that direct myogenesis. To identify potential Hoip targets, we re-examined our hoip RNA-seq data for misregulated transcripts with gene ontology (GO) terms associated with cytoskeletal regulation. Twenty-five misregulated transcripts were associated with these GO terms, including Tm1 and Tm2. Tropomyosin protein expression initiated in the somatic mesoderm during founder cell specification and is robustly expressed in elongating myotubes (Fig 1B,C). By contrast, Mhc expression did not initiate until myotube elongation was largely complete (Fig. 1D). We characterized a third sarcomeric protein, Z-band alternatively spliced PDZ-motif protein 66 (Zasp66), and found Zasp66 expression did not initiate until St16 (supplementary material Fig. S1B-D). Thus, the Tropomyosin expression pattern is temporally and spatially consistent with a role for Tropomyosin in myotube elongation, and Tropomyosin expression is temporally distinct from other sarcomeric proteins.

Fig. 1.

Tropomyosin expression and localization suggests a novel function. (A) Myotube elongation in Drosophila embryos. Bidirectional myotube elongation (double arrows) initiates during St12, continues through St13, and is largely complete by St15 when elongated muscles identify attachment sites associated with tendon cells. hoip embryos are defective in myotube elongation, in particular the lateral longitudinal 1 (LL1) and dorsal oblique 5 (DO5) muscles fail to elongate. (B-D) WT embryos labeled for Tropomyosin (Tm, green) and Mhc (red). Tm is first detected in the somatic muscle (SM) founder cells during St11 (B). Tm is robustly expressed in SM during myotube elongation (St12, C) whereas SM Mhc is largely undetectable until St13 (D). (B′-D′) Mhc expression. VM, visceral muscle. (E-G) Tm2GFP protein-trap embryos double labeled for GFP (green) and F-actin (phalloidin, red). Tm2GFP co-localized with F-actin at the myotube leading edge (white arrows). (E′-G′) F-actin expression. (E″-G″) Tm2GFP expression. Scale bars: 50 µm in B-D; 5 µm in E-G. Embryos are oriented with anterior to the left and dorsal to the top in this and subsequent figures.

Fig. 1.

Tropomyosin expression and localization suggests a novel function. (A) Myotube elongation in Drosophila embryos. Bidirectional myotube elongation (double arrows) initiates during St12, continues through St13, and is largely complete by St15 when elongated muscles identify attachment sites associated with tendon cells. hoip embryos are defective in myotube elongation, in particular the lateral longitudinal 1 (LL1) and dorsal oblique 5 (DO5) muscles fail to elongate. (B-D) WT embryos labeled for Tropomyosin (Tm, green) and Mhc (red). Tm is first detected in the somatic muscle (SM) founder cells during St11 (B). Tm is robustly expressed in SM during myotube elongation (St12, C) whereas SM Mhc is largely undetectable until St13 (D). (B′-D′) Mhc expression. VM, visceral muscle. (E-G) Tm2GFP protein-trap embryos double labeled for GFP (green) and F-actin (phalloidin, red). Tm2GFP co-localized with F-actin at the myotube leading edge (white arrows). (E′-G′) F-actin expression. (E″-G″) Tm2GFP expression. Scale bars: 50 µm in B-D; 5 µm in E-G. Embryos are oriented with anterior to the left and dorsal to the top in this and subsequent figures.

If Tropomyosin is required for myotube elongation, we reasoned that Tropomyosin would co-localize with F-actin at the myotube leading edge. Embryos harboring an endogenous Tm2 protein trap (hereafter Tm2GFP; Buszczak et al., 2007) were labeled for GFP and F-actin. Tm2GFP co-localized with F-actin at the myotube leading edge throughout elongation and during muscle attachment (Fig. 1E-G). Tm2 subcellular localization further suggested that Tropomyosin is required for myotube elongation, and that Tm2 interacts with F-actin prior to sarcomere assembly.

Tm2 protein expression requires Hoip

hoip expression is not ubiquitous, and within the embryonic mesoderm hoip expression is restricted to the striated muscle lineages and the fat body (Johnson et al., 2013). The human Hoip ortholog NHP2L1 performs multiple cellular functions that include pre-mRNA splicing and ribosomal RNA processing (Schultz et al., 2006). In fact, the crystal structure of NHP2L1 bound to the spliceosomal RNA U4 has been solved (Liu et al., 2007). NHP2L1 also shows sequence similarities to the archaeal ribosomal protein L7Ae (Kuhn et al., 2002), suggesting that Hoip could be a ribosomal component. Our previous studies showed Hoip localizes to both the nucleus and the cytoplasm of elongating myotubes (Johnson et al., 2013). Hoip could thus be required for pre-mRNA splicing, mRNA nuclear export, mRNA stability and localization, or mRNA translation.

To distinguish among these possibilities, we designed an in vivo splicing assay in which a series of Tm2 genomic constructs (Tm2GFP#1-3) were cloned upstream of a C-terminal GFP tag in a UAS transgenic vector (Fig. 2A; supplementary material Fig. S2A). For each construct, GFP expression requires correct splicing of the encoded transcript. We considered the possibility that excessive overexpression of the Tm2 constructs would be sufficient to detect GFP expression in hoip1 embryos. To increase the sensitivity of the assay, we used random transposition to generate both low-level and high-level expressing lines for a comparative analysis.

Fig. 2.

Hoip directs Tm2 protein expression but not Tm2 pre-mRNA splicing. (A) Diagram of the Tm2 locus and the Tm2-GFP constructs. Boxes represent exons, shaded regions denote the coding region. Tm2 encodes two alternative final exons (5a/b). Both low-level and high-level expressing insertions were recovered for each construct. (B-O) Live St16 embryos. The somatic muscle driver RP298.gal4 was used to co-express CD8.mCherry and τGFP (B,C) or Tm2-GFPs (D-O). (B,C) τGFP and CD8.mCherry were expressed at comparable levels in WT and hoip1 embryos. (C′) τGFP expression. Low-level expressing Tm2-GFP#1-3 and Tm2-cDNA.GFP produced Tm2-GFP in WT embryos (D-G), but Tm2-GFP was largely undetectable in hoip1 embryos (H-K). RP298.gal4 also directs transgene expression in the salivary gland (arrows) and the PNS (arrowheads). Non-muscle tissues showed robust Tm2-GFP in hoip1 embryos. (H′-K′) Tm2-GFP expression. (L-O) Tm2-GFP from high-level expressing Tm2-GFP constructs was detectable in hoip1 embryos. (P) Quantification of GFP fluorescence in hoip1 DO2 muscles relative to WT. DO2 muscles elongate in hoip1 embryos and were used to assay transgene expression in this and subsequent figures. ***P<0.001, Student's t-test; n≥36 per genotype. (Q) Western blots from S2 cells co-transfected with Hoip-GFP and the Tm2 constructs shown in (A) with a C-terminal Flag tag. Hoip promoted Tm2 protein expression from both cDNA and genomic constructs. (R) Tm2-GFP#2 splicing in S2 cells transfected with Hoip or control vector. qPCR results show the abundance of intron-free transcripts compared with intron-containing transcripts relative to control transfected cells. Error bars in P and R represent s.e.m. Scale bars: 50 µm.

Fig. 2.

Hoip directs Tm2 protein expression but not Tm2 pre-mRNA splicing. (A) Diagram of the Tm2 locus and the Tm2-GFP constructs. Boxes represent exons, shaded regions denote the coding region. Tm2 encodes two alternative final exons (5a/b). Both low-level and high-level expressing insertions were recovered for each construct. (B-O) Live St16 embryos. The somatic muscle driver RP298.gal4 was used to co-express CD8.mCherry and τGFP (B,C) or Tm2-GFPs (D-O). (B,C) τGFP and CD8.mCherry were expressed at comparable levels in WT and hoip1 embryos. (C′) τGFP expression. Low-level expressing Tm2-GFP#1-3 and Tm2-cDNA.GFP produced Tm2-GFP in WT embryos (D-G), but Tm2-GFP was largely undetectable in hoip1 embryos (H-K). RP298.gal4 also directs transgene expression in the salivary gland (arrows) and the PNS (arrowheads). Non-muscle tissues showed robust Tm2-GFP in hoip1 embryos. (H′-K′) Tm2-GFP expression. (L-O) Tm2-GFP from high-level expressing Tm2-GFP constructs was detectable in hoip1 embryos. (P) Quantification of GFP fluorescence in hoip1 DO2 muscles relative to WT. DO2 muscles elongate in hoip1 embryos and were used to assay transgene expression in this and subsequent figures. ***P<0.001, Student's t-test; n≥36 per genotype. (Q) Western blots from S2 cells co-transfected with Hoip-GFP and the Tm2 constructs shown in (A) with a C-terminal Flag tag. Hoip promoted Tm2 protein expression from both cDNA and genomic constructs. (R) Tm2-GFP#2 splicing in S2 cells transfected with Hoip or control vector. qPCR results show the abundance of intron-free transcripts compared with intron-containing transcripts relative to control transfected cells. Error bars in P and R represent s.e.m. Scale bars: 50 µm.

The Tm2 genomic constructs were expressed in developing somatic muscles with RP298.gal4. As a control for RP298.gal4 activity in hoip1 embryos, we assayed somatic muscle GFP expression from UAS.τGFP. RP298.gal4.  As UAS.τGFP. RP298.gal4 is also active in the salivary gland, salivary gland fluorescence was used to normalize somatic muscle transgene expression. Wild-type (WT) and hoip1 muscles produced comparable levels of τGFP in somatic muscles (Fig. 2B,C; hoip1 τGFP fluorescence=117.9% of WT, n=42). However, hoip1 embryos that expressed the low-level Tm2-GFP constructs produced only a fraction of WT GFP fluorescence in somatic muscles (Fig 2D-J,P; Tm2-GFP#1: 18.0%, n=36; Tm2-GFP#2: 24.0%, n=48; Tm2-GFP#3: 24.7%, n=48). By contrast, hoip1 embryos that expressed the high-level Tm2-GFP constructs in somatic muscles showed GFP fluorescence that was comparable to WT embryos (Fig. 2L-N; supplementary material Fig. S2EB-G; Tm2-GFP#1: 89.3%, n=66; Tm2-GFP#2: 167.6%, n=48; Tm2-GFP#3: 137.9%, n=36). In fact, some hoip muscles showed higher Tm2-GFP fluorescence than WT muscles. We attribute this to the fact that hoip muscles are smaller than WT muscles, and the GFP signal is concentrated over a smaller area. Our splicing assays showed Hoip regulates Tm2 protein expression during myogenesis, but surprisingly suggest that Hoip is not necessary for pre-mRNA splicing.

Hoip promotes Tm2 protein expression through a splicing-independent mechanism

To test the possibility that Hoip facilitates Tm2 protein expression after pre-mRNA splicing, we generated stable insertions of a C-terminal GFP-tagged Tm2-cDNA under UAS control. Similar to the intron-containing constructs, hoip1 embryos that expressed a low-level Tm2-cDNA.GFP showed reduced Tm2-GFP expression compared with WT embryos (Fig. 2K,P; 52.1%, n=48), and hoip1 embryos that expressed a high-level Tm2-cDNA.GFP construct showed WT levels of Tm2-GFP expression (Fig. 2O,P; 125.0%, n=42). Thus, Tm2-GFP expression from a cDNA also required Hoip, which further argues that Hoip regulates Tm2 expression through a splicing-independent mechanism.

To confirm these observations, we generated Flag-tagged Tm2 constructs for expression in Drosophila S2 cells (Fig. 2Q). S2 cells co-transfected with Hoip and either the Tm2 genomic constructs or the Tm2 cDNA construct produced more Tm2 protein than cells co-transfected with the Tm2 constructs and a control vector (Fig. 2Q). Importantly, cells transfected with Hoip did not show enhanced expression of β-tubulin, which argues that Hoip is not a global regulator of protein expression. We could not detect Tm2-Flag protein expression from Tm2#3, which suggests that alternative exon 5a is exclusively selected in S2 cells (data not shown). The Tm2#2 genomic construct contains three exons and two introns, and cells transfected with Hoip did not show enhanced splicing of Tm2#2 transcripts compared with control transfected cells by quantitative real-time PCR (qPCR; Fig. 2R). These in vitro experiments confirmed that Hoip regulates Tm2 protein expression through a splicing-independent mechanism.

Hoip promotes protein expression after Tm2 mRNA nuclear export

To understand if Hoip is required for Tm2 mRNA nuclear export, we assayed Tm2 mRNA localization by in situ hybridization. WT embryos showed robust Tm2 expression in all somatic muscles, whereas hoip1 embryos expressed Tm2 at low levels in the dorsal and ventral muscle groups, but not in the lateral muscles (Fig. 3A,B). hoip1 embryos that expressed the low-level Tm2-GFP#3 construct showed significant Tm2 mRNA in the somatic muscle cytoplasm, but Tm2-GFP protein was largely undetectable (Fig. 3C). However, hoip1 embryos that expressed the high-level Tm2-GFP#3 showed robust Tm2-GFP protein expression in somatic muscles (Fig. 3D). Tm2 mRNAs are therefore exported from the nucleus in hoip1 embryos, but the transcript is not robustly translated when expressed at low levels.

Fig. 3.

Tm2 coding region sequences direct somatic muscle protein expression. (A-D) St16 embryos that expressed τGFP (A,B) or Tm2-GFP#3 (C,D) under the control of RP298.gal4 double labeled for Tm2 mRNA (green) and GFP (red). WT embryos expressed Tm2 in all SMs (A), whereas hoip1 embryos expressed only weak Tm2 in only the dorsal and ventral muscles (B, arrows). (C) hoip1 embryos that expressed low-level Tm2-GFP#3 accumulated Tm2 RNA in the cytoplasm but did not express significant Tm2-GFP protein. (D) hoip1 embryos that expressed high-level Tm2-GFP#3 expressed more Tm2-GFP protein than low-level expressing Tm2-GFP#3 embryos. (E) Western blots from S2 cells co-transfected with Hoip.GFP and Tm2 constructs with a C-terminal Flag tag. S2 cells transfected with Hoip and the Tm2-ΔcDNA construct, which lacks the first 260 bp downstream of the start codon, produced less Tm2 protein than cells transfected with the full-length construct. (F,G) Live St16 embryos that co-expressed CD8.mCherry and Tm2-cDNA.GFPs under the control of RP298.gal4. Tm2-ΔGFP protein is expressed in non-muscle tissues (arrowheads) but not SM. (H) Normalized Tm2-GFP expression in DO2 muscles. ***P<0.001, Student's t-test; n≥38 per genotype; error bars represent s.e.m. (I,J) St16 embryos that expressed Tm2-cDNA.GFPs under the control of RP298.gal4 double labeled for GFP mRNA (green) and GFP protein (red). Tm2-cDNA.GFP and Tm2-ΔcDNA.GFP RNAs accumulated at equal levels in the cytoplasm, but only the full-length Tm2-GFP protein was expressed at high levels in SM. Scale bars: 10 µm in A-D; 50 µm in F-J. SM, somatic muscle.

Fig. 3.

Tm2 coding region sequences direct somatic muscle protein expression. (A-D) St16 embryos that expressed τGFP (A,B) or Tm2-GFP#3 (C,D) under the control of RP298.gal4 double labeled for Tm2 mRNA (green) and GFP (red). WT embryos expressed Tm2 in all SMs (A), whereas hoip1 embryos expressed only weak Tm2 in only the dorsal and ventral muscles (B, arrows). (C) hoip1 embryos that expressed low-level Tm2-GFP#3 accumulated Tm2 RNA in the cytoplasm but did not express significant Tm2-GFP protein. (D) hoip1 embryos that expressed high-level Tm2-GFP#3 expressed more Tm2-GFP protein than low-level expressing Tm2-GFP#3 embryos. (E) Western blots from S2 cells co-transfected with Hoip.GFP and Tm2 constructs with a C-terminal Flag tag. S2 cells transfected with Hoip and the Tm2-ΔcDNA construct, which lacks the first 260 bp downstream of the start codon, produced less Tm2 protein than cells transfected with the full-length construct. (F,G) Live St16 embryos that co-expressed CD8.mCherry and Tm2-cDNA.GFPs under the control of RP298.gal4. Tm2-ΔGFP protein is expressed in non-muscle tissues (arrowheads) but not SM. (H) Normalized Tm2-GFP expression in DO2 muscles. ***P<0.001, Student's t-test; n≥38 per genotype; error bars represent s.e.m. (I,J) St16 embryos that expressed Tm2-cDNA.GFPs under the control of RP298.gal4 double labeled for GFP mRNA (green) and GFP protein (red). Tm2-cDNA.GFP and Tm2-ΔcDNA.GFP RNAs accumulated at equal levels in the cytoplasm, but only the full-length Tm2-GFP protein was expressed at high levels in SM. Scale bars: 10 µm in A-D; 50 µm in F-J. SM, somatic muscle.

Hoip acts on the Tm2 coding region to promote protein expression

RNA-binding proteins typically act outside of the coding region to regulate mRNA translation and stability. However, our Tm2 transgenic constructs contained exogenous 5′ and 3′ untranslated regions (UTRs). We suspected that Hoip must act within the coding region to regulate Tm2 protein expression. To identify Hoip- responsive sequences, we co-transfected S2 cells with Hoip and a series of Tm2 coding region fragments. These assays showed that the 5′ 260 bp of the Tm2 mRNA are Hoip responsive (supplementary material Fig. S3). We deleted these sequences from the Tm2 cDNA (Tm2-ΔcDNA), and found S2 cells transfected with Tm2-ΔcDNA alone did not produce as much Tm2 protein as cells transfected with full-length Tm2-cDNA (Fig. 3E). In addition, cells co-transfected with Hoip and Tm2-ΔcDNA did not express robust Tm2Δ protein (Fig. 3E). To confirm these observations in vivo, we expressed Tm2-ΔcDNA.GFP with RP298.gal4. Although Tm2Δ.GFP protein was clearly visible in the salivary gland, Tm2Δ.GFP protein was largely undetectable in somatic muscle (Fig. 3F-H). In addition, Tm2-ΔcDNA.GFP transcripts were present in the cytoplasm of somatic muscles at levels comparable to full-length Tm2-cDNA.GFP transcripts (Fig. 3I,J). Hoip therefore acts within the Tm2 coding region to direct Tm2 protein expression in somatic muscles.

Tm2 rescues myotube elongation defects in hoip embryos

Control hoip1 embryos that expressed τGFP in the somatic musculature showed a dramatic reduction in the number of completely elongated LL1 (16.1%, n=57) and LT1-3 (3.9%, n=147) muscles compared with WT embryos (Fig. 4A,B,O). hoip1 embryos that expressed low-level Tm2-GFP constructs in the somatic musculature showed myotube elongation defects similar to hoip1 embryos that expressed τGFP (Fig. 4C-H; Tm2-GFP#1: LL1=16.1%, LT1-3=15.3%; Tm2-GFP#2: LL1=26.7%, LT1-3=28.2%; Tm2-GFP#3: LL1=11.1%, LT1-3=6.8%). Remarkably, hoip1 embryos that expressed high-level Tm2-GFP constructs in the somatic musculature showed a significant recovery in the number of elongated myotubes compared with control hoip1 embryos (Fig. 4I-N; Tm2-GFP#1: LL1=50.4%, n=71, LT1-3=39.1%, n=213; Tm2-GFP#2: LL1=59.8%, n=67, LT1-3=55.4%, n=204; Tm2-GFP#3: LL1=69.5%, n=53, LT1-3=14.6%, n=159).

Fig. 4.

Tm2 rescues myotube elongation defects in hoip1 embryos. Live St16 embryos that co-expressed CD8.mCherry and τ.GFP (A,B), low-level Tm2-GFPs (C-H), or high-level Tm2-GFPs (I-N) under the control of RP298.gal4. White arrows denote properly elongated LL1 muscles in WT embryos. (A,B) τ.GFP and CD8.mCherry were expressed at comparable levels in WT and hoip1 embryos. LL1 muscles failed to elongate (red arrows) in hoip1 embryos that expressed τ.GFP. (C-H) Compared with WT embryos, hoip1 embryos showed reduced Tm2-GFP fluorescence from low-level expressing lines. A majority of LL1 muscles failed to elongate in hoip1 embryos that expressed low-level Tm2-GFPs. (I-N) Tm2-GFP fluorescence from high-level expressing lines was comparable between WT and hoip1 embryos. LL1 muscles showed improved elongation and morphology in hoip1 embryos that expressed high-level Tm2-GFPs (green arrows). LT1-3, DO3-4, and LO1 muscles also elongated in the presence of Tm2-GFP. mCherry was used to assess elongation in muscles that did not express Tm2-GFP. (B′-N′) GFP expression. (O) Quantification of GFP fluorescence and muscle morphology in hoip1 embryos that expressed GFP transgenes. Significance was determined between hoip1, τ.GFP embryos and hoip1, Tm2-GFP embryos. *P<0.05, **P<0.01, ***P<0.001, ns, not significant, Student's t-test; n≥53 per genotype; error bars represent s.e.m. Scale bars: 20 µm.

Fig. 4.

Tm2 rescues myotube elongation defects in hoip1 embryos. Live St16 embryos that co-expressed CD8.mCherry and τ.GFP (A,B), low-level Tm2-GFPs (C-H), or high-level Tm2-GFPs (I-N) under the control of RP298.gal4. White arrows denote properly elongated LL1 muscles in WT embryos. (A,B) τ.GFP and CD8.mCherry were expressed at comparable levels in WT and hoip1 embryos. LL1 muscles failed to elongate (red arrows) in hoip1 embryos that expressed τ.GFP. (C-H) Compared with WT embryos, hoip1 embryos showed reduced Tm2-GFP fluorescence from low-level expressing lines. A majority of LL1 muscles failed to elongate in hoip1 embryos that expressed low-level Tm2-GFPs. (I-N) Tm2-GFP fluorescence from high-level expressing lines was comparable between WT and hoip1 embryos. LL1 muscles showed improved elongation and morphology in hoip1 embryos that expressed high-level Tm2-GFPs (green arrows). LT1-3, DO3-4, and LO1 muscles also elongated in the presence of Tm2-GFP. mCherry was used to assess elongation in muscles that did not express Tm2-GFP. (B′-N′) GFP expression. (O) Quantification of GFP fluorescence and muscle morphology in hoip1 embryos that expressed GFP transgenes. Significance was determined between hoip1, τ.GFP embryos and hoip1, Tm2-GFP embryos. *P<0.05, **P<0.01, ***P<0.001, ns, not significant, Student's t-test; n≥53 per genotype; error bars represent s.e.m. Scale bars: 20 µm.

To extend this observation, we assayed Tm2-GFP expression and myotube elongation in embryos with two copies of the low-level Tm2-GFP insertions. hoip1 embryos with two copies of the low-level expressing Tm2-GFP genomic constructs produced more Tm2-GFP than hoip1 embryos with just a single copy (supplementary material Fig. S4A-I; Tm2-GFP#1: 497.6%, n=48; Tm2-GFP#2: 703.8%, n=48; Tm2-GFP#3: 599.4%, n=66; Tm2-cDNA.GFP: 161.3%, n=66; percent relative to single copy). In addition, hoip1 embryos with two copies of each Tm2-GFP showed a significant recovery in the number of elongated myotubes compared with hoip1 embryos with just a single copy (supplementary material Fig. S4I; Tm2-GFP#1: LL1=336.5%, n=62, LT1-3=290.0%, n=186; Tm2-GFP#2: LL1=304.3%, n=66, LT1-3=155.9%, n=171; Tm2-GFP#3: LL1=598.0%, n=74, LT1-3=169.3%, n=222; Tm2-cDNA.GFP: LL1=251.7%, n=73, LT1-3=326.3%, n=219). Tm2 therefore promotes myotube elongation in hoip1 embryos.

Tm2 regulates F-actin during myotube elongation

Tropomyosin localizes to F-actin in the lamella of migratory cells, and the contractile F-actin/myosin lamellar network provides the physical force to drive cell movement (Ponti et al., 2004; Ridley et al., 2003). As Tm2 co-localized with F-actin during myotube elongation (Fig. 1E,F), we predicted that Tm2 regulates F-actin at the myotube leading edge. Compared with WT embryos (Fig. 5A-C,P; St12 DO5=9.5; St13 DO5=14.9; St15 DO5=9.9; arbitrary units, n≥12 myotubes per stage), hoip1 embryos showed a significant decrease in leading edge F-actin (Fig. 5D-F,P; St12 DO5=3.9; St13 DO5=3.5; St15 DO5=4.5; n≥18). However, hoip1 embryos that expressed the high-level Tm2-GFP#3 assembled significantly more leading edge F-actin than hoip1 embryos (Fig. 5G-I,P; St12 DO5=5.0; St13 DO5=7.2; St15 DO5=9.0; n≥21). Thus, Tm2 promotes F-actin assembly at the myotube leading edge.

Fig. 5.

Tm2 restores actin expression in hoip1 embryos. (A-I) Embryos that expressed τGFP (A-F) or one copy of Tm2-GFP#3 (G-I) under the control of RP298.gal4 co-labeled for F-actin (phalloidin, red). (A-C) WT embryos showed robust F-actin accumulation at the leading edge of elongating myotubes (St12-13; A,B) and during target site recognition (St15; C). The DO5 muscle is outlined in white or blue throughout the figure. Arrows denote one leading-edge. (D-F) hoip1 embryos showed reduced F-actin accumulation at DO5 myotube leading edges throughout myogenesis. (G-I) hoip1 embryos that expressed high-level Tm2-GFP#3 showed improved F-actin accumulation at myotube leading edges compared with control hoip1 embryos. (J-L) St13 embryos that expressed τ.GFP (J,K) or one copy of Tm2-GFP#3 (L) labeled for Act57B mRNA. hoip1 embryos showed reduced Act57B. hoip1 embryos that expressed Tm2-GFP#3 showed improved Act57B expression. (M-O) St13 Act57B.-593/+2.nlacZ embryos that expressed τ.GFP (M,N) or one copy of Tm2-GFP#3 (O). hoip1 embryos showed reduced β-gal expression. hoip1 embryos that expressed Tm2-GFP#3 showed improved β-gal expression. Arrowheads show a subset of somatic muscle nuclei. (P) Quantification of leading-edge F-actin and Act57B.-593/+2.nlacZ expression. (Q) qPCR of mRNA isolated from St17 embryos. Sarcomeric transcripts are downregulated in hoip1 embryos and largely restored in hoip1 embryos that expressed Tm2-GFP#3. The enrichment of Tm1 in hoip1 embryos suggests a compensatory mechanism might regulate Tropomyosin levels. *P<0.05, **P<0.01, ***P<0.001, Student's t-test; error bars represent s.e.m. Scale bars: 5 µm in A-I; 50 µm (top three rows) or 10 µm (bottom row) in J-O.

Fig. 5.

Tm2 restores actin expression in hoip1 embryos. (A-I) Embryos that expressed τGFP (A-F) or one copy of Tm2-GFP#3 (G-I) under the control of RP298.gal4 co-labeled for F-actin (phalloidin, red). (A-C) WT embryos showed robust F-actin accumulation at the leading edge of elongating myotubes (St12-13; A,B) and during target site recognition (St15; C). The DO5 muscle is outlined in white or blue throughout the figure. Arrows denote one leading-edge. (D-F) hoip1 embryos showed reduced F-actin accumulation at DO5 myotube leading edges throughout myogenesis. (G-I) hoip1 embryos that expressed high-level Tm2-GFP#3 showed improved F-actin accumulation at myotube leading edges compared with control hoip1 embryos. (J-L) St13 embryos that expressed τ.GFP (J,K) or one copy of Tm2-GFP#3 (L) labeled for Act57B mRNA. hoip1 embryos showed reduced Act57B. hoip1 embryos that expressed Tm2-GFP#3 showed improved Act57B expression. (M-O) St13 Act57B.-593/+2.nlacZ embryos that expressed τ.GFP (M,N) or one copy of Tm2-GFP#3 (O). hoip1 embryos showed reduced β-gal expression. hoip1 embryos that expressed Tm2-GFP#3 showed improved β-gal expression. Arrowheads show a subset of somatic muscle nuclei. (P) Quantification of leading-edge F-actin and Act57B.-593/+2.nlacZ expression. (Q) qPCR of mRNA isolated from St17 embryos. Sarcomeric transcripts are downregulated in hoip1 embryos and largely restored in hoip1 embryos that expressed Tm2-GFP#3. The enrichment of Tm1 in hoip1 embryos suggests a compensatory mechanism might regulate Tropomyosin levels. *P<0.05, **P<0.01, ***P<0.001, Student's t-test; error bars represent s.e.m. Scale bars: 5 µm in A-I; 50 µm (top three rows) or 10 µm (bottom row) in J-O.

Tm2 regulates actin expression

Similar to vertebrates, Drosophila embryos express muscle-specific actin isoforms. Actin 57B (Act57B) is a component of somatic muscle thin filaments, and Act57B expression initiates as early as St11 (Kelly et al., 2002). Compared with WT embryos, Act57B mRNA levels were dramatically reduced in hoip1 embryos (Fig. 5J,K; supplementary material Fig. S5A,B), and hoip1 embryos that expressed Tm2-GFP#3 showed improved Act57B expression in somatic muscles (Fig. 5L; supplementary material Fig. S5C). Importantly, the probe used to detect Act57B binds to the highly divergent 3′UTR that distinguishes it from other actin isoforms (Kelly et al., 2002). To understand how Hoip and Tm2 regulate Act57B, we used the minimal reporter gene Act57B.-593/+2.nlacZ to assay transcriptional activity (Kelly et al., 2002). Compared with WT embryos (Fig. 5M,Q; supplementary material Fig. S5D; St13=6.7; St16=8.4; arbitrary units, n≥150 nuclei per stage), Act57B.-593/+2 reporter expression was reduced in hoip1 embryos (Fig. 5N,P; supplementary material Fig. S5E; St13=2.7; St16=5.7; n≥150). hoip1 embryos that expressed Tm2-GFP#3 showed a slight, yet significant, increase in reporter gene expression compared with hoip1 embryos (Fig. 5O,P; supplementary material Fig. S5F; St13=4.6; St16=6.9; n≥150).

Tm2 promotes sarcomere assembly

F-actin performs multiple functions during myogenesis beyond myotube elongation. For example, F-actin stress fibers are thought to provide the template for sarcomere assembly (Friedrich et al., 2012). As Tm2 restored Act57B transcription and leading edge F-actin in hoip1 embryos, we suspected that Tm2 might also regulate F-actin during sarcomere assembly. By St17, WT somatic muscles have assembled thin filament F-actin. However, F-actin was absent from the somatic muscles of hoip1 embryos and hoip1 embryos that expressed the low-level Tm2-GFP#3 (supplementary material Fig. S5H,N). Strikingly, somatic muscles of hoip1 embryos that expressed the high-level Tm2-GFP#3 assembled thin filament F-actin (supplementary material Fig. S5I). Tm2-GFP also localized to sarcomeres in WT St17 embryos (supplementary material Fig. S5J,L). Consistent with our F-actin results, Tm2-GFP from the high-level expressing line localized to sarcomeres in hoip1 embryos whereas Tm2-GFP from the low-level expressing line did not (supplementary material Fig. S5K,M). It is possible that the low-level expressing line did not produce enough Tm2-GFP for us to detect sarcomeric localization in hoip1 embryos. However, these embryos also lacked thin filaments so the minimal Tm2-GFP produced likely lacked a substrate for localization. In either case, high levels of Tm2 restored sarcomere assembly in hoip1 embryos.

Tm2 regulates sarcomeric protein expression

Knockdown of individual sarcomeric proteins can disrupt sarcomere assembly (Rui et al., 2010). We had previously shown that sarcomeric RNAs were downregulated in hoip1 embryos at St12-13, and that Mhc protein expression was greatly reduced in St16 hoip1 embryos (Johnson et al., 2013). As Tm2 restored sarcomere assembly in hoip1 embryos, we expected that Tm2 would also promote the expression of sarcomeric RNAs. Compared to WT embryos, the average ratio of sarcomeric mRNA levels decreased in St17 hoip1 embryos [Act57B, 0.44; α-actinin (Actn), 0.49 Myosin heavy chain (Mhc), 0.38; Myosin light chain 1 (Mlc1), 0.52; Mlc2, 0.16; Troponin C at 47D (TpnC47D), 0.42 and Tm2, 0.61; Fig. 5Q]. However, the average ratio of sarcomeric mRNA levels in hoip1 embryos that expressed Tm2-GFP#3 was comparable to WT embryos (Act57B, 1.2; Actn, 1.5; Mhc, 1.7; Mlc1, 2.4; Mlc2, 1.9; TpnC47D, 1.9; Tm2, 3.4; Fig. 5Q). Consistent with these qPCR results, Mhc protein expression, Mhc mRNA expression, and TpnC47D mRNA expression were also restored in hoip1 somatic muscles that expressed Tm2-GFP#3 (Fig. 6A-I).

Fig. 6.

Tm2 restores sarcomeric gene expression and promotes myoblast fusion in hoip1 embryos. (A-I) St16 embryos that expressed τGFP or one copy of Tm2-GFP#3 under the control of RP298.gal4 co-labeled for (A-C) Tropomyosin (Tm, green) and Mhc (red), (D-F) Mhc mRNA (green) and GFP (red), or (G-I) TpnC47D mRNA (green) and GFP (red). Tm and Mhc protein expression is dramatically reduced in hoip1 embryos (A,B). Tm2-GFP#3 partially restored Mhc protein expression in hoip1 embryos (C). Mhc and TpnC47D mRNAs were also reduced in hoip1 embryos (D,E,G,H), and restored in hoip1 embryos that expressed Tm2-GFP#3 (F,I). (J) Quantification of DO2 muscle size, myoblast fusion, and Mhc and Tm protein expression in St16 embryos. The fusion index represents the percentage of RP298.nlacZ positive-nuclei (per hemisegment) versus control embryos. The expression index represents protein expression normalized to muscle area versus control embryos. ***P<0.001, Student's t-test; error bars represent s.e.m. (K-M) DO2 muscles from embryos that co-expressed RP298.lacZ and either τGFP or one copy of Tm2-GFP#3 under the control of RP298.gal4. hoip1 DO2 muscles had fewer nuclei compared with WT, but did elongate. Tm2-GFP#3 increased the number of nuclei in hoip1 DO2 muscles. Scale bars represent 50 µm in K-M and low magnification (left) in A-I; 10 µm in high magnification (right) in A-I.

Fig. 6.

Tm2 restores sarcomeric gene expression and promotes myoblast fusion in hoip1 embryos. (A-I) St16 embryos that expressed τGFP or one copy of Tm2-GFP#3 under the control of RP298.gal4 co-labeled for (A-C) Tropomyosin (Tm, green) and Mhc (red), (D-F) Mhc mRNA (green) and GFP (red), or (G-I) TpnC47D mRNA (green) and GFP (red). Tm and Mhc protein expression is dramatically reduced in hoip1 embryos (A,B). Tm2-GFP#3 partially restored Mhc protein expression in hoip1 embryos (C). Mhc and TpnC47D mRNAs were also reduced in hoip1 embryos (D,E,G,H), and restored in hoip1 embryos that expressed Tm2-GFP#3 (F,I). (J) Quantification of DO2 muscle size, myoblast fusion, and Mhc and Tm protein expression in St16 embryos. The fusion index represents the percentage of RP298.nlacZ positive-nuclei (per hemisegment) versus control embryos. The expression index represents protein expression normalized to muscle area versus control embryos. ***P<0.001, Student's t-test; error bars represent s.e.m. (K-M) DO2 muscles from embryos that co-expressed RP298.lacZ and either τGFP or one copy of Tm2-GFP#3 under the control of RP298.gal4. hoip1 DO2 muscles had fewer nuclei compared with WT, but did elongate. Tm2-GFP#3 increased the number of nuclei in hoip1 DO2 muscles. Scale bars represent 50 µm in K-M and low magnification (left) in A-I; 10 µm in high magnification (right) in A-I.

Tm2 promotes myoblast fusion in hoip1 embryos

In addition to myotube elongation and sarcomere assembly, F-actin plays a key role in myoblast fusion (Schejter and Baylies, 2010). Our Act57B reporter gene experiments showed hoip1 embryos that expressed Tm2-GFP#3 contained more β-galactosidase (β-gal)-positive nuclei than hoip1 embryos (Fig. 5N,O). This result suggested that Tm2 promotes myoblast fusion in hoip1 embryos. The reporter gene RP298.nlacZ is expressed in muscle founders, and subsequently in nascent myotubes, and can be used to measure myoblast fusion. hoip1 embryos that expressed Tm2-GFP#3 showed more β-gal-positive myonuclei than hoip1 embryos (Fig. 6J). Importantly, myonuclei number in the DO2 muscle was reduced in hoip1 embryos compared with control embryos, even though the DO2 muscle often elongates (Fig. 6K,L; WT=9.5 DO2 nuclei, hoip1=2.5, n≥28). hoip1 embryos that expressed Tm2-GFP#3 showed an increased number of DO2 myonuclei compared with hoip1 embryos (Fig. 6M; 4.6 nuclei/DO2, n=30). Tm2 therefore promotes myoblast fusion in hoip1 embryos.

Muscle size does not dictate sarcomeric gene expression

One mechanism that could explain Tm2-mediated gene expression is that Tm2 regulates muscle size, and muscle size in turn dictates sarcomeric gene expression. To test this possibility we normalized Mhc and Tropomyosin protein expression to somatic muscle size, which we refer to as the expression index. The expression index for Mhc and Tropomyosin was significantly reduced in hoip1 embryos compared with controls. hoip1 embryos that expressed Tm2-GFP#3 showed a significantly greater Mhc and Tropomyosin expression index than hoip1 embryos (Fig. 6J). Therefore, the difference in Mhc expression between hoip1 and hoip1 Tm2-GFP#3 rescued embryos is not simply the result of an increase in muscle size.

Tropomyosin regulates Mhc expression

Our rescue studies showed that Tm2 promotes myoblast fusion, myotube elongation, and sarcomeric gene expression in hoip1 embryos. To further characterize the function of Tm2 during myogenesis, we generated a Tm2 null mutation (Tm2Δ8-261). The Drosophila genome encodes only two Tropomyosin isoforms, Tm1 and Tm2, and both isoforms are deleted by Df(3R)BSC741. Using qPCR, we confirmed St17 Tm2Δ8-261 homozygous embryos and St17 Tm2Δ8-261/Df(3R)BSC741 transheterozygous embryos did not express Tm2 (supplementary material Fig. S6A). In addition, these embryos showed reduced Act57B expression compared with control embryos (supplementary material Fig. S6A; FC=0.29 and 0.31), and Act57B expression could be partially restored in Tm2Δ8-261 embryos that expressed Tm2 under the control of RP298.gal4 (supplementary material Fig. S6A; FC=0.65). The remaining sarcomeric RNAs were expressed at near WT levels in Tm2 mutant embryos. Surprisingly, myotube elongation was largely unaffected in Tm2Δ8-261/Df(3R)BSC741 embryos, but a subset of transheterozygous embryos showed defects in LT1-3 or LL1 morphogenesis in at least one segment (25%, n=12; Fig. 7A,B). Tm2Δ8-261/Df(3R)BSC741 embryos also showed a significant reduction in the Mhc expression index compared with controls but somatic muscle size was unaffected (Fig. 7I). Tm2Δ8-261 embryos that expressed Tm2 showed improved LT1-3 and LL1 morphogenesis (14% of embryos with elongation defects, n=14), and improved Mhc expression compared with transheterozygous embryos (Fig. 7C,I).

Fig. 7.

Tm2 regulates Mhc expression and directs myotube elongation. (A-C) St16 embryos labeled for Tropomyosin (Tm, green) and Mhc (red). (A,B) Tm2Δ8-261/Df(3R)BSC741 embryos expressed significant Tropomyosin protein, but showed reduced Mhc protein expression compared with control embryos. Tm2Δ8-261/Df(3R)BSC741 embryos also showed elongation defects at a low frequency (LT muscles from two segments are outlined). (C) Tm2.GFP restored Mhc protein expression in Tm2Δ8-261 embryos. (D-F) Live St17 embryos that expressed Tm2-cDNA.GFP or Tm2E54K-cDNA.GFP under the control of RP298.gal4. LL1 and LT somatic muscles that expressed Tm2E54K showed elongation defects. Tm2E54K also disrupted DO5 muscle morphology (notice the bifurcated DO5 in F). (G,H) DO2 muscles from St16 RP298.nlacZ embryos that co-expressed Tm2-cDNA.GFP or Tm2E54K-cDNA.GFP. The number of DO2 nuclei is reduced in embryos that expressed Tm2E54K-cDNA.GFP. (I) DO2 muscle size and Mhc/Tm expression indexes. (J) Quantification of Tm2E54K-cDNA.GFP expression, myotube elongation, and DO2 nuclei relative to Tm2-cDNA.GFP. *P<0.05, **P<0.01, ***P<0.001; error bars represent s.e.m. Scale bars: 10 µm. See supplementary material Fig. S1 for a diagram of somatic muscles.

Fig. 7.

Tm2 regulates Mhc expression and directs myotube elongation. (A-C) St16 embryos labeled for Tropomyosin (Tm, green) and Mhc (red). (A,B) Tm2Δ8-261/Df(3R)BSC741 embryos expressed significant Tropomyosin protein, but showed reduced Mhc protein expression compared with control embryos. Tm2Δ8-261/Df(3R)BSC741 embryos also showed elongation defects at a low frequency (LT muscles from two segments are outlined). (C) Tm2.GFP restored Mhc protein expression in Tm2Δ8-261 embryos. (D-F) Live St17 embryos that expressed Tm2-cDNA.GFP or Tm2E54K-cDNA.GFP under the control of RP298.gal4. LL1 and LT somatic muscles that expressed Tm2E54K showed elongation defects. Tm2E54K also disrupted DO5 muscle morphology (notice the bifurcated DO5 in F). (G,H) DO2 muscles from St16 RP298.nlacZ embryos that co-expressed Tm2-cDNA.GFP or Tm2E54K-cDNA.GFP. The number of DO2 nuclei is reduced in embryos that expressed Tm2E54K-cDNA.GFP. (I) DO2 muscle size and Mhc/Tm expression indexes. (J) Quantification of Tm2E54K-cDNA.GFP expression, myotube elongation, and DO2 nuclei relative to Tm2-cDNA.GFP. *P<0.05, **P<0.01, ***P<0.001; error bars represent s.e.m. Scale bars: 10 µm. See supplementary material Fig. S1 for a diagram of somatic muscles.

A dominant Tm2 allele disrupts myotube elongation

One explanation for the lack of a strong elongation phenotype in Tm2 embryos is that Tm2 is maternally contributed. Using qPCR and immunohistochemistry, we confirmed there is a maternal contribution of both Tm1 and Tm2 mRNAs (supplementary material Fig. S6B-F). However, females with Tm2Δ8-261 or Df(3R)BSC741 homozygous mutant germ lines did not lay eggs. Previous studies showed Tm1 is required for border cell migration in the ovary (Kim et al., 2011), which suggested Tropomyosin performs essential functions in the germ line. Our results further argue that Tm1 and Tm2 are essential for oogenesis.

To further investigate a role for Tm2 during myotube elongation, we generated a dominant Tm2 allele. The human genome encodes four Tropomyosin isoforms (TPM1-4), and congenital myopathies have been associated with dominant Tropomyosin mutations (Marttila et al., 2014; Olson et al., 2001). Tropomyosin is a coiled-coil protein comprised of seven pseudo-repeat domains, and each domain contains an α-zone that interacts with F-actin when muscle is relaxed (Marttila et al., 2014). Several dominant Tropomyosin mutations occur in α-zone residues that are conserved between humans and Drosophila, including E54K (supplementary material Fig. S2). We hypothesized that the E54K mutation would disrupt Tropomyosin function in a dominant fashion, and expressed Tm2E54K-GFP during myotube elongation with RP298.gal4. Otherwise, WT embryos that expressed Tm2E54K-GFP in developing somatic muscles showed a significant reduction in the percent of completely elongated myotubes (Fig. 7D-F,J; line#1: LL1=75.0%, LT1-3=80.6%; line#2: LL1=76.4%, LT1-3=81.3%) and in the number of DO2 nuclei (Fig. 7G,H) compared with embryos that expressed WT Tm2.GFP. These studies further demonstrate that Tm2 is required for myotube elongation and myoblast fusion.

Our study has revealed a novel function for Tm2 during myogenesis. Functional rescue experiments showed Tm2 is epistatic to hoip and that Hoip regulates Tm2 protein expression via the coding region. Although Tropomyosins are known regulators of actin dynamics in migratory cells, we found that Tm2 regulates F-actin levels during myotube elongation and sarcomere assembly. In addition, Tm2 promoted myoblast fusion in hoip1 embryos and regulated Mhc protein expression. To our knowledge, this is the first study to show that Tropomyosins regulate muscle development prior to sarcomere assembly. Lastly, myotubes that expressed the gain-of-function allele Tm2E54K showed both fusion and elongation defects. This finding might have important implications in understanding Tropomyosin-related myopathies.

Post-transcriptional regulation of sarcomeric mRNAs

Tm2 protein expression is Hoip dependent, and sequences in the 5′ end of the Tm2 coding region are required for Tm2 expression in somatic muscles (Fig. 3E-J). Hoip orthologs direct spliceosome assembly (Schultz et al., 2006), and we had previously shown that an Mhc cDNA restored Mhc protein expression in hoip embryos. These findings suggested that Hoip regulates splicing, so we were surprised to discover that Hoip is not required for Tm2 pre-mRNA splicing or Tm2 mRNA nuclear export. However, our experiments with Mhc and Tm2 cDNAs did produce similar results. In the case of Mhc, we detected Mhc protein expression in hoip embryos homozygous for the Mhc transgene. In the case of Tm2, a high-level expressing insertion or two copies of a low-level insertion produced near-WT levels of Tm2. GFP in hoip embryos (Fig. 2P; supplementary material Fig. S4). These data argue that Hoip promotes robust protein expression from endogenous mRNAs, but this requirement can be overcome in hoip embryos when the transcripts are overexpressed.

Despite these similarities, Tm2 performs a distinct functional role to enhance sarcomeric protein expression. Tm2 rescued Mhc protein expression in hoip embryos (Fig. 6C), but Mhc did not rescue Tropomyosin expression (Johnson et al., 2013). In fact, several sarcomeric mRNAs were enriched in hoip embryos that expressed high levels of Tm2.GFP (Fig. 5Q). Thus, negative regulation of sarcomeric mRNAs is not simply offset by overexpressing any sarcomeric mRNA. Tm2 mutant embryos also showed reduced Mhc protein expression (Fig. 7B), but Mhc mRNA levels were unaffected (supplementary material Fig. S6A). Our results demonstrate that Tm2 performs a specific regulatory function, which is distinct at least from Mhc, to promote sarcomeric protein expression. Although Tm2.GFP ameliorated multiple myogenic defects in hoip embryos, Tm2.GFP was not sufficient to completely restore myogenesis. This incomplete rescue could be due to a number of factors including the spatial and temporal onset of Tm2.GFP expression or that Hoip has at least one additional target mRNA required for myogenesis.

Importantly, Tm2 regulates Mhc expression independent of muscle size (Figs 6J and 7I). Muscle size is thought to be dependent in part on myoblast fusion (Schejter and Baylies, 2010), and we do see a correlation between myonuclei number and muscle size in St16 embryos (Fig. 6J). Tm2.GFP promoted myoblast fusion in hoip embryos (Fig. 6K-L), and Tm2E54K reduced myoblast fusion when expressed in nascent myotubes (Fig. 7G,H). It is possible that Tm2 regulates sarcomeric protein expression by promoting myoblast fusion, which would increase the number of sarcomeric loci, and in turn transcription of sarcomeric genes. However, the Mhc expression index is significantly different between embryos with reduced muscle size and myonuclei number (Fig. 6J). Although we cannot absolutely rule out the possibility that sarcomeric gene transcription contributes to the phenotypes we have reported, it is clear that muscle size, myonuclei number, and gene transcription alone do not account for Tm2-mediated regulation of sarcomeric protein expression.

The role of F-actin during myogenesis

In migratory cells, adhesions with the extracellular matrix form and disassemble at the junction of the lamellipodia and the lamella (Ponti et al., 2004). Actomyosin contractions in the lamella use these adhesion sites to move the cell forward. The lamella is characterized by discrete foci of actin polymerization and depolymerization (Ponti et al., 2004). Migrating epithelial cells that overexpress TPM1 lacked a lamellipodia, but showed rapid migration. Mechanistically, TPM1 overexpression concentrated F-actin and myosin II to the lamella and produced more cellular adhesions (Gupton et al., 2005). A similar cellular mechanism appears to direct myotube elongation. Like TPM1 (Gupton et al., 2005), Tm2 localized to punctate foci near the leading edge (Fig. 1E-G). These foci might reflect sites of Tm2-mediated actomyosin contractions. As both TPM1 and Tm2 promote F-actin assembly at the leading edge membrane, the cellular mechanism of myotube elongation likely parallels that of migratory cells. The actomyosin network also directs myoblast fusion (Kim et al., 2015). Here, actomyosin tension in the founder cell promotes pore formation at the fusogenic synapse. It will be interesting to see if Tm2 directs the assembly or even the function of the actomyosin network to promote myoblast fusion.

Although we established that Hoip regulates Tm2 through a largely post-transcriptional mechanism, Act57B.-593/+2 reporter gene expression was reduced in hoip1 embryos (Fig. 5P), and restored in hoip1 embryos that expressed Tm2.GFP. Accordingly, leading edge F-actin and thin filament F-actin were downregulated in hoip1 embryos, and rescued by Tm2.GFP (Fig. 5P; supplementary material Fig. S5). The only known molecular function of Tropomyosins is to bind F-actin, so transcriptional regulation of Act57B is most likely indirect. F-actin polymerization initiates a feedback loop that drives actin transcription in some contexts (Mokalled et al., 2010), so the changes in Act57B transcription we observed are likely a downstream effect of reduced F-actin levels.

Redundancy, compensation, and maternal contribution of Tropomyosin isoforms

Both Tm1 and Tm2 are expressed in developing somatic muscles (BDGP insitu homepage – http://insitu.fruitfly.org/cgi-bin/ex/insitu.pl), and we confirmed these are the only Tropomyosin-encoding genes in the Drosophila genome by BLAST analysis. Tm1 mRNA was enriched in Tm2 mutant embryos at St11 and St12 (supplementary material Fig. S6B; FC=3.1, 2.1), but not at St17 (FC=0.75). These data argue that Tm1 compensates for Tm2 in Tm2 null embryos at the onset of zygotic transcription. Similarly, Tm1 mRNA was enriched in hoip embryos (Fig. 5Q; FC=5.0), and reduced in Tm2.GFP rescued hoip embryos (FC=2.5). However, elevated Tm1 levels alone were not sufficient to induce Tropomyosin protein expression in hoip embryos, whereas high-level Tm2.GFP mRNA produced significant levels of Tm2.GFP protein (Figs 2L-O and 5Q). Hoip is therefore required for the expression of both Tropomyosin protein isoforms, but different mechanisms appear to direct Tm1 and Tm2 protein expression in somatic muscle.

The modest Tm2 zygotic phenotype we characterized is consistent with other maternally contributed genes that regulate myogenesis. For example, rho and Rok single-mutant embryos develop normal somatic muscles whereas rho, Rok double-mutant embryos show myoblast fusion defects (Kim et al., 2015). Although the maternal contribution of Tm1 and Tm2 produced only a fraction of WT transcript levels (supplementary material Fig. S6B), Df(3R)BSC741 embryos showed robust Tropomyosin expression during myotube elongation and this expression persisted through St16 (supplementary material Fig. S6C-F). By contrast, a majority of somatic muscles failed to express Tropomyosin protein in hoip mutant embryos (Fig. 6B). Thus, maternally contributed Tm1/2 can be translated in the presence of Hoip to direct largely normal myogenesis.

Tropomyosin alleles are associated with myopathies

Dominant mutations in TPM1, TPM2, and TPM3 are associated with congenital myopathies (Marttila et al., 2014) and cardiomyopathies (Karibe et al., 2001; Olson et al., 2001). The TPM1 E54K allele was identified in patients with dilated cardiomyopathy and in vitro thin filament reconstitution experiments show the E54K protein overinhibits actomyosin interactions, which decreases force generation during systolic contraction (Bai et al., 2012). Our Tm2 E54K allele disrupted myogenesis (Fig. 7D-F), which argues that Tm2-mediated actomyosin contractions are indeed required for myotube elongation. A second TPM1 allele, E40K, also interrupts actomysosin contractions (Bai et al., 2012), and TPM2 E41K is associated with congenital myopathies (Marttila et al., 2014). This raises the possibility that myotube elongation is affected in patients with TPM2-associated congenital myopathies. It will be of particular interest to characterize additional Tropomyosin alleles in vivo to better understand the disease mechanisms that underlie these myopathies.

Drosophila genetics

The stocks used in this study were hoip1 (Johnson et al., 2013), P{Gal4-kirrerP298} and P{lacZ-kirrerP298} (Nose et al., 1998), P{PTT-GC}Tm2ZCL2456 and P{PTT-GA}Zasp66ZCL0663 – referred to as Tm2GFP and Zasp66GFP protein traps (Buszczak et al., 2007), P{Act57B-lacZ.-593+2} (Kelly et al., 2002), P{UAS-mCD8.ChRFP}, P{UAS-mCD8.eGFP}, P{UAS-τGFP}, Df(3R)BSC741, P{neoFRT}82B, and P{ovoD1-18}3R (Bloomington Stock Center). Cyo, P{Gal4-Twi}, P{2X-UAS.eGFP}; Cyo, P{wg.lacZ}; and TM3, P{Gal4-Twi}, P{2X-UAS.eGFP} balancers were used to identify homozygous embryos.

Tm2 genomic transgenes were constructed by cloning genomic DNA or cDNA (RE15528) PCR products into pEntr (Life Technologies), and then recombined into a destination vector (TWG and AWG) as described (https://emb.carnegiescience.edu/drosophila-gateway-vector-collection). Transgenic insertions were generated by standard methods (Rainbow Transgenic Flies). Tm2-ΔcDNA deletion constructs were generated by PCR sewing as described (Johnson et al., 2013). Entry clones were fully sequenced.

The Tm2Δ8-261 allele was generated as described (Gratz et al., 2013), and homozygous embryos were sequenced to confirm the Tm2Δ8-261 mutation. Germ line clones were made by standard methods.

Immunohistochemistry and in situ hybridization

Antibodies used include anti-MHC (Kiehart and Feghali, 1986), anti-Tropomyosin (Abcam, MAC141), anti-β-Galactosidase (Promega, Z378A) and anti-GFP (Torrey Pines Laboratories; TP401). HRP-conjugated secondary antibodies and the TSA system (Molecular Probes; T20922, T20913) were used to detect primary antibodies. Antibody staining and in situ hybridization were performed as described (Johnson et al., 2013). RE15528 and LP10264 were used as templates for the Tm2 and TpnC47D in situ probes, respectively. Mhc and Act57B probes were generated from the templates described (Johnson et al., 2013; Kelly et al., 2002). Texas Red-conjugated phalloidin (Molecular Probes) was used to detect F-actin in fixed, hand devitellinated embryos.

Imaging and fluorescence quantification

Images were generated with an LSM700 confocal microscope (Zeiss). Control and mutant embryos were prepared and imaged in parallel where possible. Confocal imaging parameters were maintained between genotypes throughout the study. Normalized expression was calculated as mean fluorescence intensity over an entire DO2 muscle relative to salivary gland fluorescence (Tm2-GFP) or visceral muscle fluorescence (anti-Tm, anti-Mhc). Myotube leading edge F-actin levels were determined by normalizing mean fluorescence at the leading edge relative to internal fluorescence. Time-lapse microscopy was used to identify the position of the nascent myotubes prior to F-actin measurements (supplementary material Fig. S7). Act57B.-593/+2.nlacZ fluorescence was measured in single nuclei and normalized to background fluorescence. Muscle size was determined by outlining individual DO2 muscles to obtain an area. All measurements were performed with Zen2011 software (Zeiss).

Cell culture and western blots

Schneider 2 (S2) cells were grown in Schneider's media supplemented with 10% FBS and pen/strep. Cells were passaged biweekly and split the day prior to transfection. On the day of the transfection, cells were seeded to a density of 1.0×106 cells/ml in Schneider's media with FBS. 0.8 ml of cells were transferred to one well of a 12-well plate. Transfections were performed with Effectene according to the manufacturer's instructions. 48 h post-transfection, whole cell extracts were made by pelleting cells and resuspending in lysis buffer (20 mM Tris pH 7.5, 150 mM NaCl, 1% Triton X-100, and protease inhibitors). Cells were lysed on ice for 10 min, followed by 10 min centrifugation. Western blots were performed as described (Mokalled et al., 2010) and imaged using the ChemiDoc XRS+ system (BioRad). A minimum of 3 blots was performed from independent transfections for each experiment shown.

Gateway technology was used to generate tagged Hoip constructs using the hoip cDNA clone (RE51843) as a PCR template.

Quantitative RT-PCR

Staged embryos were dechorionated and hand sorted to isolate homozygous mutants. RNA was extracted from embryos or S2 cells using TRizol, and cDNA was generated using Superscript III (Life Technologies). qPCR was performed with SYBR Select Master Mix using an ABI Prism 7000 (Life Technologies). Forward and reverse primers were designed to exons separated by at least one intron, except for splicing assay primers. Here, forward primers were designed to an exon and reverse primers were designed to the downstream intron (to detect unspliced transcripts) or to span the downstream splice donor/acceptor sites (to detect spliced transcripts). qPCR reactions were run in triplicate and normalized to RpL32 or GAPDH. See supplementary material Table S1 for primer sequences.

We thank Richard Cripps for reagents, Mayssa Mokalled for insights and discussions throughout this study and for critical reading of the manuscript, and Brenna Clay for assistance with embryology.

Author contributions

J.W., N.G.B., J.M.V. and A.N.J. designed and performed experiments. A.N.J. prepared the manuscript.

Funding

A.N.J. was supported by a Scientist Development Grant [12SDG12030160] from the American Heart Association, and by a Webb-Waring Biomedical Research Award from the Boettcher Foundation.

Baas
,
D.
,
Caussanel-Boude
,
S.
,
Guiraud
,
A.
,
Calhabeu
,
F.
,
Delaune
,
E.
,
Pilot
,
F.
,
Chopin
,
E.
,
Machuca-Gayet
,
I.
,
Vernay
,
A.
,
Bertrand
,
S.
, et al. 
(
2012
).
CKIP-1 regulates mammalian and zebrafish myoblast fusion
.
J. Cell Sci.
125
,
3790
-
3800
.
Bai
,
F.
,
Groth
,
H. L.
and
Kawai
,
M.
(
2012
).
DCM-related tropomyosin mutants E40K/E54K over-inhibit the actomyosin interaction and lead to a decrease in the number of cycling cross-bridges
.
PLoS ONE
7
,
e47471
.
Berger
,
S.
,
Schafer
,
G.
,
Kesper
,
D. A.
,
Holz
,
A.
,
Eriksson
,
T.
,
Palmer
,
R. H.
,
Beck
,
L.
,
Klambt
,
C.
,
Renkawitz-Pohl
,
R.
and
Onel
,
S.-F.
(
2008
).
WASP and SCAR have distinct roles in activating the Arp2/3 complex during myoblast fusion
.
J. Cell Sci.
121
,
1303
-
1313
.
Bongiovanni
,
A.
,
Romancino
,
D. P.
,
Campos
,
Y.
,
Paterniti
,
G.
,
Qiu
,
X.
,
Moshiach
,
S.
,
Di Felice
,
V.
,
Vergani
,
N.
,
Ustek
,
D.
and
d'Azzo
,
A.
(
2012
).
Alix protein is substrate of Ozz-E3 ligase and modulates actin remodeling in skeletal muscle
.
J. Biol. Chem.
287
,
12159
-
12171
.
Buckingham
,
M.
and
Rigby
,
P. W. J.
(
2014
).
Gene regulatory networks and transcriptional mechanisms that control myogenesis
.
Dev. Cell
28
,
225
-
238
.
Bugyi
,
B.
and
Carlier
,
M.-F.
(
2010
).
Control of actin filament treadmilling in cell motility
.
Annu. Rev. Biophys.
39
,
449
-
470
.
Buszczak
,
M.
,
Paterno
,
S.
,
Lighthouse
,
D.
,
Bachman
,
J.
,
Planck
,
J.
,
Owen
,
S.
,
Skora
,
A. D.
,
Nystul
,
T. G.
,
Ohlstein
,
B.
,
Allen
,
A.
, et al. 
(
2007
).
The carnegie protein trap library: a versatile tool for Drosophila developmental studies
.
Genetics
175
,
1505
-
1531
.
Chen
,
E. H.
and
Olson
,
E. N.
(
2004
).
Towards a molecular pathway for myoblast fusion in Drosophila
.
Trends Cell Biol.
14
,
452
-
460
.
Ciglar
,
L.
and
Furlong
,
E. E. M.
(
2009
).
Conservation and divergence in developmental networks: a view from Drosophila myogenesis
.
Curr. Opin. Cell Biol.
21
,
754
-
760
.
de Joussineau
,
C.
,
Bataillé
,
L.
,
Jagla
,
T.
and
Jagla
,
K.
(
2012
).
Diversification of muscle types in Drosophila: upstream and downstream of identity genes
.
Curr. Top. Dev. Biol.
98
,
277
-
301
.
Friedrich
,
B. M.
,
Fischer-Friedrich
,
E.
,
Gov
,
N. S.
and
Safran
,
S. A.
(
2012
).
Sarcomeric pattern formation by actin cluster coalescence
.
PLoS Comput. Biol.
8
,
e1002544
.
Gratz
,
S. J.
,
Cummings
,
A. M.
,
Nguyen
,
J. N.
,
Hamm
,
D. C.
,
Donohue
,
L. K.
,
Harrison
,
M. M.
,
Wildonger
,
J.
and
O'Connor-Giles
,
K. M.
(
2013
).
Genome engineering of Drosophila with the CRISPR RNA-guided Cas9 nuclease
.
Genetics
194
,
1029
-
1035
.
Gross
,
S. R.
(
2014
).
Actin binding proteins: their ups and downs in metastatic life
.
Cell Adh. Migr.
7
,
199
-
213
.
Gupton
,
S. L.
,
Anderson
,
K. L.
,
Kole
,
T. P.
,
Fischer
,
R. S.
,
Ponti
,
A.
,
Hitchcock-DeGregori
,
S. E.
,
Danuser
,
G.
,
Fowler
,
V. M.
,
Wirtz
,
D.
,
Hanein
,
D.
, et al. 
(
2005
).
Cell migration without a lamellipodium: translation of actin dynamics into cell movement mediated by tropomyosin
.
J. Cell Biol.
168
,
619
-
631
.
Johnson
,
A. N.
,
Mokalled
,
M. H.
,
Valera
,
J. M.
,
Poss
,
K. D.
and
Olson
,
E. N.
(
2013
).
Post-transcriptional regulation of myotube elongation and myogenesis by Hoi Polloi
.
Development
140
,
3645
-
3656
.
Karibe
,
A.
,
Tobacman
,
L. S.
,
Strand
,
J.
,
Butters
,
C.
,
Back
,
N.
,
Bachinski
,
L. L.
,
Arai
,
A. E.
,
Ortiz
,
A.
,
Roberts
,
R.
,
Homsher
,
E.
, et al. 
(
2001
).
Hypertrophic cardiomyopathy caused by a novel alpha-tropomyosin mutation (V95A) is associated with mild cardiac phenotype, abnormal calcium binding to troponin, abnormal myosin cycling, and poor prognosis
.
Circulation
103
,
65
-
71
.
Kawamura
,
K.
,
Takano
,
K.
,
Suetsugu
,
S.
,
Kurisu
,
S.
,
Yamazaki
,
D.
,
Miki
,
H.
,
Takenawa
,
T.
and
Endo
,
T.
(
2004
).
N-WASP and WAVE2 acting downstream of phosphatidylinositol 3-kinase are required for myogenic cell migration induced by hepatocyte growth factor
.
J. Biol. Chem.
279
,
54862
-
54871
.
Kelly
,
K. K.
,
Meadows
,
S. M.
and
Cripps
,
R. M.
(
2002
).
Drosophila MEF2 is a direct regulator of Actin57B transcription in cardiac, skeletal, and visceral muscle lineages
.
Mech. Dev.
110
,
39
-
50
.
Kiehart
,
D. P.
and
Feghali
,
R.
(
1986
).
Cytoplasmic myosin from Drosophila melanogaster
.
J. Cell Biol.
103
,
1517
-
1525
.
Kim
,
J. H.
,
Cho
,
A.
,
Yin
,
H.
,
Schafer
,
D. A.
,
Mouneimne
,
G.
,
Simpson
,
K. J.
,
Nguyen
,
K.-V.
,
Brugge
,
J. S.
and
Montell
,
D. J.
(
2011
).
Psidin, a conserved protein that regulates protrusion dynamics and cell migration
.
Genes Dev.
25
,
730
-
741
.
Kim
,
J. H.
,
Ren
,
Y.
,
Ng
,
W. P.
,
Li
,
S.
,
Son
,
S.
,
Kee
,
Y.-S.
,
Zhang
,
S.
,
Zhang
,
G.
,
Fletcher
,
D. A.
,
Robinson
,
D. N.
, et al. 
(
2015
).
Mechanical tension drives cell membrane fusion
.
Dev. Cell
32
,
561
-
573
.
Kuhn
,
J. F.
,
Tran
,
E. J.
and
Maxwell
,
E. S.
(
2002
).
Archaeal ribosomal protein L7 is a functional homolog of the eukaryotic 15.5kD/Snu13p snoRNP core protein
.
Nucleic Acids Res.
30
,
931
-
941
.
Liu
,
S.
,
Li
,
P.
,
Dybkov
,
O.
,
Nottrott
,
S.
,
Hartmuth
,
K.
,
Luhrmann
,
R.
,
Carlomagno
,
T.
and
Wahl
,
M. C.
(
2007
).
Binding of the human Prp31 Nop domain to a composite RNA-protein platform in U4 snRNP
.
Science
316
,
115
-
120
.
Marttila
,
M.
,
Lehtokari
,
V.-L.
,
Marston
,
S.
,
Nyman
,
T. A.
,
Barnerias
,
C.
,
Beggs
,
A. H.
,
Bertini
,
E.
,
Ceyhan-Birsoy
,
Ö.
,
Cintas
,
P.
,
Gerard
,
M.
, et al. 
(
2014
).
Mutation update and genotype-phenotype correlations of novel and previously described mutations in TPM2 and TPM3 causing congenital myopathies
.
Hum. Mutat.
35
,
779
-
790
.
Mokalled
,
M. H.
,
Johnson
,
A.
,
Kim
,
Y.
,
Oh
,
J.
and
Olson
,
E. N.
(
2010
).
Myocardin-related transcription factors regulate the Cdk5/Pctaire1 kinase cascade to control neurite outgrowth, neuronal migration and brain development
.
Development
137
,
2365
-
2374
.
Nose
,
A.
,
Isshiki
,
T.
and
Takeichi
,
M.
(
1998
).
Regional specification of muscle progenitors in Drosophila: the role of the msh homeobox gene
.
Development
125
,
215
-
223
.
Olson
,
T. M.
,
Kishimoto
,
N. Y.
,
Whitby
,
F. G.
and
Michels
,
V. V.
(
2001
).
Mutations that alter the surface charge of alpha-tropomyosin are associated with dilated cardiomyopathy
.
J. Mol. Cell. Cardiol.
33
,
723
-
732
.
Ponti
,
A.
,
Machacek
,
M.
,
Gupton
,
S. L.
,
Waterman-Storer
,
C. M.
and
Danuser
,
G.
(
2004
).
Two distinct actin networks drive the protrusion of migrating cells
.
Science
305
,
1782
-
1786
.
Richardson
,
B. E.
,
Beckett
,
K.
,
Nowak
,
S. J.
and
Baylies
,
M. K.
(
2007
).
SCAR/WAVE and Arp2/3 are crucial for cytoskeletal remodeling at the site of myoblast fusion
.
Development
134
,
4357
-
4367
.
Ridley
,
A. J.
,
Schwartz
,
M. A.
,
Burridge
,
K.
,
Firtel
,
R. A.
,
Ginsberg
,
M. H.
,
Borisy
,
G.
,
Parsons
,
J. T.
and
Horwitz
,
A. R.
(
2003
).
Cell migration: integrating signals from front to back
.
Science
302
,
1704
-
1709
.
Rui
,
Y.
,
Bai
,
J.
and
Perrimon
,
N.
(
2010
).
Sarcomere formation occurs by the assembly of multiple latent protein complexes
.
PLoS Genet.
6
,
e1001208
.
Schejter
,
E. D.
and
Baylies
,
M. K.
(
2010
).
Born to run: creating the muscle fiber
.
Curr. Opin. Cell Biol.
22
,
566
-
574
.
Schnorrer
,
F.
and
Dickson
,
B. J.
(
2004
).
Muscle building: mechanisms of myotube guidance and attachment site selection
.
Dev. Cell
7
,
9
-
20
.
Schultz
,
A.
,
Nottrott
,
S.
,
Watkins
,
N. J.
and
Luhrmann
,
R.
(
2006
).
Protein-protein and protein-RNA contacts both contribute to the 15.5K-mediated assembly of the U4/U6 snRNP and the box C/D snoRNPs
.
Mol. Cell. Biol.
26
,
5146
-
5154
.

Competing interests

The authors declare no competing or financial interests.

Supplementary information