1,886
Views
12
CrossRef citations to date
0
Altmetric
Extra View

The timing of pre-mRNA splicing visualized in real-time

&
Pages 11-14 | Received 05 Jan 2014, Accepted 30 Jan 2014, Published online: 30 Jan 2014

Abstract

Since it became clear that intervening sequences or introns are spliced out from precursor pre-mRNA molecules in the nucleus before mature mRNAs are exported to the cytoplasm, questions were raised about the timing of splicing. Does splicing start while RNA polymerase II is still transcribing? Is splicing a slow or a fast process? Is timing important to control the splicing reaction? Although our understanding on the mechanism and function of splicing is largely based on data obtained using biochemical and large-scale “omic” approaches, microscopy has been instrumental to address questions related to timing. Experiments done with the electron microscope paved the way to the discovery of splicing and provided unequivocal evidence that splicing can occur co-transcriptionally. More recently, live-cell microscopy introduced a technical breakthrough that allows real-time visualization of splicing dynamics. We discuss here some of the microscopy advances that provided the basis for the current conceptual view of the splicing process and we outline a most recent development that permits direct measurement, in living cells, of the time it takes to synthesize and excise an intron from individual pre-mRNA molecules.

This article refers to:

In the summer of 1977, a set of electron micrographs shown at the annual Cold Spring Harbor Symposium revolutionized the way molecular biologists perceived the structure of genes. Following the development of methodologies to observe isolated RNA and DNA molecules by electron microscopy, it was found that isolated RNA molecules can hybridize to double-stranded DNA by displacing one of the DNA strands.Citation1 Using this approach, the Roberts and Sharp teams independently realized that adenoviral polysomal RNAs do not hybridize in a continuous manner with the DNA from which they are transcribed. Rather, the mRNA sequence was complementary to multiple noncontiguous regions of the DNA.Citation2,Citation3 They concluded that adenoviral genes were split into several pieces, and because adenovirus is a mammalian virus, it seemed likely that at least some mammalian genes could have a similar structure. This was soon found to be the case. Although different mechanisms could result in joining together certain segments of genomic sequence into mature mRNA, evidence quickly accumulated supporting a model for mRNA synthesis that involves splicing of a larger precursor molecule or pre-mRNA.

Understanding the splicing mechanism required the development of cell-free systems where conditions could be controlled and the components purified and reconstituted. In vitro splicing assays based on extracts obtained by moderate salt solubilization of lysed cells or crude nuclei were able to accurately splice exogenously added pre-mRNA, indicating that splicing and transcription are independent processes (for a review see ref. Citation4). Yet, visualizing dispersed chromatin by electron microscopy showed that splicing was coupled to transcription in vivo. Back in 1969, Miller and Beatty reported that hypotonic lysis of cells released chromatin from the nucleus.Citation5 The released chromatin was then centrifuged onto a grid for observation in the electron microscope. This procedure resulted in a loosened two-dimensional array of chromatin fibers that maintained the nucleosomal structure. Nascent transcripts remained attached to the chromatin and appeared as ribonucleoprotein fibers extending from the chromatin backbone. The length of nascent transcripts, which could be precisely measured, increased with increasing distance from the transcription start site. Using this approach, Beyer and Osheim showed that the spliceosome formed shortly after synthesis of the 3′ splice site and that splicing of pre-mRNA often occurred on the nascent transcript.Citation6 This visual demonstration that introns could be removed prior to transcript release from the template and thus presumably prior to polyadenylation challenged the prevailing textbook view that splicing took place after polyadenylation. Beyer and Osheim assumed an elongation rate of 1500 nucleotides per minute to then deduce that splicing of introns removed cotranscriptionally occurred within 3 min after synthesis of the 3′ splice site. Similar conclusions were obtained when studying actively transcribed genes in Chironomus tentans salivary gland polytene chromosomes, which can be isolated by microdissection for direct observation in the electron microscope.Citation7 In vivo splicing rates in the range of 5 min or less were also measured biochemically for adenovirus and β-globin transcripts.Citation8-Citation10 Such rates differed substantially from the results obtained in cell-free assays, where a lag period of 15 to 45 min was observed at the start of the reaction before the onset of splicing, following which spliced RNA accumulated for 2–3 h.Citation4 Thus, splicing in vivo is at least 10-fold more rapid than in vitro. This discrepancy highlights the importance of nuclear organization: most likely, spliceosome components in the nucleus are organized in a way that favors efficient recognition of splice site sequences as soon as they emerge from RNA polymerase.

Imaging the spatial organization of genes, nascent RNAs, and spliceosome components in the intact cell nucleus was made possible through successive optimizations of another microscopic technique termed in situ hybridization. The pioneer method relied on synthesis of radioactive RNA complementary to the DNA of interest. Cells were made permeable and incubated with the radioactive probe, resulting in formation of RNA-DNA hybrids. The location of these hybrids was then detected at the cytological level by autoradiography.Citation11,Citation12 Over the next 30 y, this technology progressed from a laborious, time-consuming method limited to the detection of highly abundant nucleic acid species with low cytological resolution, to a relatively fast, highly precise, and sensitive imaging of single genes and RNA molecules in the cell. An important advance was the optimization of the in situ hybridization procedure for fluorescent detection of nuclear RNA using biotin-labeled DNA probes and fluorophore-conjugated avidin.Citation13 Next, the method was further optimized for the simultaneous detection of DNA, unspliced precursors, and spliced mRNA. Using these tools, several laboratories showed that spliced mRNA localized in close proximity to the DNA from which it was transcribed, consistent with the view that many introns are excised while the pre-mRNA is still tethered to the gene locus via RNA polymerase.Citation14-Citation16 There are, however, situations where splicing is uncoupled from transcription: in the case of regulated alternative splicing events, intron removal can be delayed until after release of the pre-mRNA from the site of transcription.Citation17 In situ hybridization procedures were also optimized to localize the small nuclear RNAs that form the building blocks of the spliceosome.Citation18 By combining in situ hybridization and immuno-labeling for fluorescent and electron microscopic visualization, it became evident that spliceosomal small nuclear ribonucleoproteins and splicing protein factors accumulated in dedicated sub-nuclear compartments termed Cajal bodies and nuclear speckles (for a review see ref. Citation19). Unlike other RNA-based macromolecular machines in the cell such as the ribosome, spliceosomes form anew on each nascent intron and disassemble after each splicing event.Citation20 As most cells contain a surplus of long-lived spliceosomal components in the nucleus, these molecules are dynamically oscillating between two states: being part of a spliceosome or waiting for the opportunity to assemble a new spliceosome. Since actively spliced genes are intimately associated with the periphery of nuclear speckles, one possibility is that clustering of spliceosomal components in dedicated compartments increases their local concentration in the neighborhood of nascent transcripts, thus enhancing spliceosome assembly on newly synthesized introns.Citation21 This could be a reason why splicing is so much faster in vivo than in vitro.

An imaging breakthrough was introduced by the development of genetically encoded fluorescent tags that combined with fluorescence-based microscopic approaches of increasingly higher spatial and temporal resolution made it possible to analyze protein movement in living cells. Time-series recordings of cells expressing the green fluorescent protein (GFP) fused to an essential splicing protein provided direct evidence that nuclear speckles supply spliceosomal components to nearby activated sites of transcription.Citation22 Subsequent developments of methods such as FRAP (fluorescence recovery after photobleaching) unravelled the kinetic properties of splicing proteins in the nucleus of live cells.Citation23 FRAP revealed that spliceosomal proteins are continuously moving in the nucleus, shuttling in and out of nuclear speckles within seconds. Surprisingly, trafficking kinetics was independent of ongoing transcription and splicing, challenging the view that spliceosomal components are stored in nuclear speckles until a signal triggers their recruitment to nascent introns; rather, it is more likely that the building blocks of the spliceosome are constantly roaming the nuclear space until they collide and transiently interact with either a nascent pre-mRNA, to form a spliceosome, or with other splicing proteins localized in nuclear speckles.Citation24 FRAP analysis further revealed that core spliceosomal snRNP proteins have a residence time of 15–30 s in the nucleoplasm, where spliceosomal snRNPs are thought to interact predominantly with pre-mRNA.Citation25 Based on these results it was suggested that splicing can be accomplished within 30 s, which is significantly more rapid than previously reported.Citation25

Following the advent of genetically encoded fluorescent protein tags, Belmont and colleagues pioneered a method to image chromatin dynamics in vivo. They introduced bacterial lac operator repeats into the genome of yeast and mammalian cells that expressed a GFP-lac repressor fusion protein.Citation26 Binding of the fluorescent repressor to its target sequence makes that particular region of chromatin visible in the nucleus. A similar approach was later developed to visualize RNAs in living cells by genetically inserting the binding sites for the MS2 bacteriophage coat protein in the RNA of interest.Citation27 The resulting reporter gene was then integrated in the genome of cells that expressed the MS2 coat protein fused to GFP. Insertion of the MS2 binding sites in the terminal exon of reporter genes revealed kinetic properties of the entire mRNA life cycle, from transcription to transport in the nucleus and export to the cytoplasm.Citation28,Citation29 Based on these results, a prediction was that an intronic insertion of MS2 binding sites should suffice to track splicing in real-time. Bertrand and colleagues combined such an approach with FRAP to measure co-transcriptional splicing kinetics.Citation30 They analyzed an ensemble population of pre-mRNAs synthesized from a gene cluster comprising ~20 copies of a reporter gene that contains MS2 binding sites in an artificial short intron derived from the adenovirus genome. Upon bleaching the MS2-GFP fluorescence associated with introns at the transcription site, they measured a half-life of 105 s for fluorescence recovery and estimated a mean splicing time of 162 s. This value was in good agreement with the splicing rates previously deduced from electron microscopyCitation6,Citation7 and biochemical pulse-chase experiments.Citation31 However, inferring the splicing rate from FRAP curves required the use of numerical models because fluorescence recovery is influenced by many variables in addition to splicing. These include rate of transcription by RNA polymerase II, diffusion and binding of MS2-GFP, and diffusion or degradation of excised introns. Modeling of these results is further complicated by the simultaneous presence, in the region analyzed, of multiple asynchronous transcripts. To circumvent these significant limitations and the potential problems introduced into the interpretation of the data, we directly tracked in time the fluorescence emitted by single pre-mRNAs labeled with MS2-GFP in the intron.Citation32 We integrated a single reporter gene in the genome of human cells and used a spinning-disk confocal microscope. Spinning disk confocal microscopy combines high sensitivity with high speed optical sectioning with minimal photobleaching, making it one of the methods of choice to analyze intracellular dynamics of single molecules with high spatial and temporal resolution. The fluorescence associated with a single transcription site in the nucleus was detected as a diffraction-limited object and its intensity was determined as a function of time. Increments in fluorescence intensity were due to de novo transcription of binding sites for MS2-GFP, while reductions reflected intron excision followed by degradation or diffusion. These fluorescence intensity fluctuations were used to determine the intron lifetime, defined as the time it takes for an intron to be transcribed, spliced, and degraded (or released from the transcription site). The fluorescence intensities at any given time point varied from cell to cell, presumably reflecting variation in the rate of transcription initiation: the higher the transcription rate the more nascent pre-mRNAs are simultaneously present at the transgene locus, resulting in higher fluorescence intensity. To determine the fluorescence intensity emitted by a single pre-mRNA molecule, we treated cells with spliceostatin A (SSA), a potent splicing inhibitor that causes release of unspliced pre-mRNAs to the nucleoplasm.Citation33 After release from the transcription site, pre-mRNAs diffused throughout the nucleus making it possible to resolve individual transcripts. Having determined the fluorescence intensity of a single pre-mRNA, we then searched for cells that synthesized one reporter transcript at the time, i.e., cells with fluorescence fluctuations that started at background level, increased to a value in the range corresponding to a single pre-mRNA, and then returned again to background. As the MS2 binding sites were inserted very close to the 5′ splice site, the duration of these fluorescence cycles provided, for the first time, a direct measure of the intron lifetime. We found that transcription and excision of short introns (1.3–1.4 kb) occurs in 20–30 s which implies a splicing rate much faster than previously reported for the adenovirus-derived short intron based on FRAP experiments.Citation30 We believe the reason for this discrepancy derives from the higher resolution of our single-molecule analysis: the rapid fluorescence fluctuations that we observed for a single pre-mRNA molecule are probably hidden in the bulk measurement of fluorescence recovery from a multitude of molecules at different stages of the splicing cycle. We then addressed the longstanding question of whether some introns are spliced faster than others. First, we compared the lifetimes of the first and second introns of the β-globin gene and detected a 1.5-fold difference: the most frequent lifetime value was 20 s for the first intron and 30 s for the second; to rule out the influence of transcription, the second intron was shortened by an internal deletion so that both introns have approximately the same length.Citation32 Next, we compared the lifetimes of two introns derived from the mouse immunoglobulin µ gene. The two introns had exactly the same length but differed in the polypyrimidine tract sequence. We found that the lifetime of an intron with a weak splice site sequence was 1.4-fold longer than the lifetime of an intron with a strong splice site sequence.Citation32 Finally, we analyzed the influence of transcription on intron lifetime by comparing two introns with the same splice site sequences but differing in length by ~1000 nucleotides inserted downstream of the MS2 binding sites. We found that the cycles of fluorescence gain and loss were ~18 s longer for the longer intron. We also estimated that the array of MS2 binding sites is transcribed at a rate of approximately 4 kb per minute. At this rate, transcription of the extra 1000 nucleotides present in the long intron is expected to take ~15 s, an interval very close to the difference in duration of the fluorescence cycles for the long and short introns (~18 s). This suggests that longer introns take longer to be spliced not because splicing is slower but rather because they take longer to be transcribed. We could also infer that splicing is a much more rapid process than transcription.Citation32 To conclude, we developed an experimental system that we hope will prove valuable to further reveal the kinetic properties of splicing, particularly in the context of alternative splicing decisions. This system could also be useful to study in real-time how delays in splicing may regulate the timing of expression of specific genes, as recently proposed for the sequential appearance of distinct mRNAs in response to TNF induction.Citation34

Disclosure of Potential Conflicts of Interest

No potential conflict of interest was disclosed.

10.4161/nucl.28056

References

  • Thomas M, White RL, Davis RW. Hybridization of RNA to double-stranded DNA: formation of R-loops. Proc Natl Acad Sci U S A 1976; 73:2294 - 8; http://dx.doi.org/10.1073/pnas.73.7.2294; PMID: 781674
  • Chow LT, Gelinas RE, Broker TR, Roberts RJ. An amazing sequence arrangement at the 5′ ends of adenovirus 2 messenger RNA. Cell 1977; 12:1 - 8; http://dx.doi.org/10.1016/0092-8674(77)90180-5; PMID: 902310
  • Berget SM, Moore C, Sharp PA. Spliced segments at the 5′ terminus of adenovirus 2 late mRNA. Proc Natl Acad Sci U S A 1977; 74:3171 - 5; http://dx.doi.org/10.1073/pnas.74.8.3171; PMID: 269380
  • Padgett RA, Grabowski PJ, Konarska MM, Seiler S, Sharp PA. Splicing of messenger RNA precursors. Annu Rev Biochem 1986; 55:1119 - 50; http://dx.doi.org/10.1146/annurev.bi.55.070186.005351; PMID: 2943217
  • Miller OL Jr., Beatty BR. Visualization of nucleolar genes. Science 1969; 164:955 - 7; http://dx.doi.org/10.1126/science.164.3882.955; PMID: 5813982
  • Beyer AL, Osheim YN. Splice site selection, rate of splicing, and alternative splicing on nascent transcripts. Genes Dev 1988; 2:754 - 65; http://dx.doi.org/10.1101/gad.2.6.754; PMID: 3138163
  • Baurén G, Wieslander L. Splicing of Balbiani ring 1 gene pre-mRNA occurs simultaneously with transcription. Cell 1994; 76:183 - 92; http://dx.doi.org/10.1016/0092-8674(94)90182-1; PMID: 8287477
  • Curtis PJ, Mantei N, Weissmann C. Characterization and kinetics of synthesis of 15S β-globin RNA, a putative precursor of β-globin mRNA. Cold Spring Harb Symp Quant Biol 1978; 42:971 - 84; http://dx.doi.org/10.1101/SQB.1978.042.01.098; PMID: 277329
  • Gattoni R, Keohavong P, Stévenin J. Splicing of the E2A premessenger RNA of adenovirus serotype 2. Multiple pathways in spite of excision of the entire large intron. J Mol Biol 1986; 187:379 - 97; http://dx.doi.org/10.1016/0022-2836(86)90440-7; PMID: 3009832
  • Audibert A, Weil D, Dautry F. In vivo kinetics of mRNA splicing and transport in mammalian cells. Mol Cell Biol 2002; 22:6706 - 18; http://dx.doi.org/10.1128/MCB.22.19.6706-6718.2002; PMID: 12215528
  • John HA, Birnstiel ML, Jones KW. RNA-DNA hybrids at the cytological level. Nature 1969; 223:582 - 7; http://dx.doi.org/10.1038/223582a0; PMID: 5799530
  • Gall JG, Pardue ML. Formation and detection of RNA-DNA hybrid molecules in cytological preparations. Proc Natl Acad Sci U S A 1969; 63:378 - 83; http://dx.doi.org/10.1073/pnas.63.2.378; PMID: 4895535
  • Lawrence JB, Singer RH, Marselle LM. Highly localized tracks of specific transcripts within interphase nuclei visualized by in situ hybridization. Cell 1989; 57:493 - 502; http://dx.doi.org/10.1016/0092-8674(89)90924-0; PMID: 2541917
  • Xing Y, Johnson CV, Dobner PR, Lawrence JB. Higher level organization of individual gene transcription and RNA splicing. Science 1993; 259:1326 - 30; http://dx.doi.org/10.1126/science.8446901; PMID: 8446901
  • Zhang G, Taneja KL, Singer RH, Green MR. Localization of pre-mRNA splicing in mammalian nuclei. Nature 1994; 372:809 - 12; http://dx.doi.org/10.1038/372809a0; PMID: 7997273
  • Custódio N, Carmo-Fonseca M, Geraghty F, Pereira HS, Grosveld F, Antoniou M. Inefficient processing impairs release of RNA from the site of transcription. EMBO J 1999; 18:2855 - 66; http://dx.doi.org/10.1093/emboj/18.10.2855; PMID: 10329631
  • Vargas DY, Shah K, Batish M, Levandoski M, Sinha S, Marras SA, Schedl P, Tyagi S. Single-molecule imaging of transcriptionally coupled and uncoupled splicing. Cell 2011; 147:1054 - 65; http://dx.doi.org/10.1016/j.cell.2011.10.024; PMID: 22118462
  • Carmo-Fonseca M, Tollervey D, Pepperkok R, Barabino SM, Merdes A, Brunner C, Zamore PD, Green MR, Hurt E, Lamond AI. Mammalian nuclei contain foci which are highly enriched in components of the pre-mRNA splicing machinery. EMBO J 1991; 10:195 - 206; PMID: 1824936
  • Handwerger KE, Gall JG. Subnuclear organelles: new insights into form and function. Trends Cell Biol 2006; 16:19 - 26; http://dx.doi.org/10.1016/j.tcb.2005.11.005; PMID: 16325406
  • Rino J, Carmo-Fonseca M. The spliceosome: a self-organized macromolecular machine in the nucleus?. Trends Cell Biol 2009; 19:375 - 84; http://dx.doi.org/10.1016/j.tcb.2009.05.004; PMID: 19616950
  • Misteli T. Cell biology of transcription and pre-mRNA splicing: nuclear architecture meets nuclear function. J Cell Sci 2000; 113:1841 - 9; PMID: 10806095
  • Misteli T, Cáceres JF, Spector DL. The dynamics of a pre-mRNA splicing factor in living cells. Nature 1997; 387:523 - 7; http://dx.doi.org/10.1038/387523a0; PMID: 9168118
  • Phair RD, Misteli T. High mobility of proteins in the mammalian cell nucleus. Nature 2000; 404:604 - 9; http://dx.doi.org/10.1038/35007077; PMID: 10766243
  • Rino J, Carvalho T, Braga J, Desterro JM, Lührmann R, Carmo-Fonseca M. A stochastic view of spliceosome assembly and recycling in the nucleus. PLoS Comput Biol 2007; 3:2019 - 31; http://dx.doi.org/10.1371/journal.pcbi.0030201; PMID: 17967051
  • Huranová M, Ivani I, Benda A, Poser I, Brody Y, Hof M, Shav-Tal Y, Neugebauer KM, Stanek D. The differential interaction of snRNPs with pre-mRNA reveals splicing kinetics in living cells. J Cell Biol 2010; 191:75 - 86; http://dx.doi.org/10.1083/jcb.201004030; PMID: 20921136
  • Robinett CC, Straight A, Li G, Willhelm C, Sudlow G, Murray A, Belmont AS. In vivo localization of DNA sequences and visualization of large-scale chromatin organization using lac operator/repressor recognition. J Cell Biol 1996; 135:1685 - 700; http://dx.doi.org/10.1083/jcb.135.6.1685; PMID: 8991083
  • Bertrand E, Chartrand P, Schaefer M, Shenoy SM, Singer RH, Long RM. Localization of ASH1 mRNA particles in living yeast. Mol Cell 1998; 2:437 - 45; http://dx.doi.org/10.1016/S1097-2765(00)80143-4; PMID: 9809065
  • Janicki SM, Tsukamoto T, Salghetti SE, Tansey WP, Sachidanandam R, Prasanth KV, Ried T, Shav-Tal Y, Bertrand E, Singer RH, et al. From silencing to gene expression: real-time analysis in single cells. Cell 2004; 116:683 - 98; http://dx.doi.org/10.1016/S0092-8674(04)00171-0; PMID: 15006351
  • Shav-Tal Y, Darzacq X, Shenoy SM, Fusco D, Janicki SM, Spector DL, Singer RH. Dynamics of single mRNPs in nuclei of living cells. Science 2004; 304:1797 - 800; http://dx.doi.org/10.1126/science.1099754; PMID: 15205532
  • Schmidt U, Basyuk E, Robert MC, Yoshida M, Villemin JP, Auboeuf D, Aitken S, Bertrand E. Real-time imaging of cotranscriptional splicing reveals a kinetic model that reduces noise: implications for alternative splicing regulation. J Cell Biol 2011; 193:819 - 29; http://dx.doi.org/10.1083/jcb.201009012; PMID: 21624952
  • Singh J, Padgett RA. Rates of in situ transcription and splicing in large human genes. Nat Struct Mol Biol 2009; 16:1128 - 33; http://dx.doi.org/10.1038/nsmb.1666; PMID: 19820712
  • Martin RM, Rino J, Carvalho C, Kirchhausen T, Carmo-Fonseca M. Live-cell visualization of pre-mRNA splicing with single-molecule sensitivity. Cell Rep 2013; 4:1144 - 55; http://dx.doi.org/10.1016/j.celrep.2013.08.013; PMID: 24035393
  • Martins SB, Rino J, Carvalho T, Carvalho C, Yoshida M, Klose JM, de Almeida SF, Carmo-Fonseca M. Spliceosome assembly is coupled to RNA polymerase II dynamics at the 3′ end of human genes. Nat Struct Mol Biol 2011; 18:1115 - 23; http://dx.doi.org/10.1038/nsmb.2124; PMID: 21892168
  • Hao S, Baltimore D. RNA splicing regulates the temporal order of TNF-induced gene expression. Proc Natl Acad Sci U S A 2013; 110:11934 - 9; http://dx.doi.org/10.1073/pnas.1309990110; PMID: 23812748

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.