Edinburgh Research Explorer...Abeer Fadda,1† Mark Ryten,2 Dorothea Droll,1 Federico Rojas,3...

21
Edinburgh Research Explorer Transcriptome-wide analysis of trypanosome mRNA decay reveals complex degradation kinetics and suggests a role for co- transcriptional degradation in determining mRNA levels Citation for published version: Fadda, A, Ryten, M, Droll, D, Rojas, F, Färber, V, Haanstra, JR, Merce, C, Bakker, BM, Matthews, K & Clayton, C 2014, 'Transcriptome-wide analysis of trypanosome mRNA decay reveals complex degradation kinetics and suggests a role for co-transcriptional degradation in determining mRNA levels', Molecular Microbiology, vol. 94, no. 2, pp. 307-326. https://doi.org/10.1111/mmi.12764 Digital Object Identifier (DOI): 10.1111/mmi.12764 Link: Link to publication record in Edinburgh Research Explorer Document Version: Publisher's PDF, also known as Version of record Published In: Molecular Microbiology Publisher Rights Statement: © 2014 The Authors. Molecular Microbiology published by John Wiley & Sons Ltd. This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction in any medium, provided the original work is properly cited General rights Copyright for the publications made accessible via the Edinburgh Research Explorer is retained by the author(s) and / or other copyright owners and it is a condition of accessing these publications that users recognise and abide by the legal requirements associated with these rights. Take down policy The University of Edinburgh has made every reasonable effort to ensure that Edinburgh Research Explorer content complies with UK legislation. If you believe that the public display of this file breaches copyright please contact [email protected] providing details, and we will remove access to the work immediately and investigate your claim. Download date: 22. Jun. 2021

Transcript of Edinburgh Research Explorer...Abeer Fadda,1† Mark Ryten,2 Dorothea Droll,1 Federico Rojas,3...

  • Edinburgh Research Explorer

    Transcriptome-wide analysis of trypanosome mRNA decayreveals complex degradation kinetics and suggests a role for co-transcriptional degradation in determining mRNA levels

    Citation for published version:Fadda, A, Ryten, M, Droll, D, Rojas, F, Färber, V, Haanstra, JR, Merce, C, Bakker, BM, Matthews, K &Clayton, C 2014, 'Transcriptome-wide analysis of trypanosome mRNA decay reveals complex degradationkinetics and suggests a role for co-transcriptional degradation in determining mRNA levels', MolecularMicrobiology, vol. 94, no. 2, pp. 307-326. https://doi.org/10.1111/mmi.12764

    Digital Object Identifier (DOI):10.1111/mmi.12764

    Link:Link to publication record in Edinburgh Research Explorer

    Document Version:Publisher's PDF, also known as Version of record

    Published In:Molecular Microbiology

    Publisher Rights Statement:© 2014 The Authors. Molecular Microbiology published by John Wiley & Sons Ltd.This is an open access article under the terms of the Creative Commons Attribution License, which permits use,distribution andreproduction in any medium, provided the original work is properly cited

    General rightsCopyright for the publications made accessible via the Edinburgh Research Explorer is retained by the author(s)and / or other copyright owners and it is a condition of accessing these publications that users recognise andabide by the legal requirements associated with these rights.

    Take down policyThe University of Edinburgh has made every reasonable effort to ensure that Edinburgh Research Explorercontent complies with UK legislation. If you believe that the public display of this file breaches copyright pleasecontact [email protected] providing details, and we will remove access to the work immediately andinvestigate your claim.

    Download date: 22. Jun. 2021

    https://doi.org/10.1111/mmi.12764https://doi.org/10.1111/mmi.12764https://www.research.ed.ac.uk/en/publications/5207f6fa-5679-45ab-b2c4-6b73a3168ad6

  • Transcriptome-wide analysis of trypanosome mRNA decayreveals complex degradation kinetics and suggests a role forco-transcriptional degradation in determining mRNA levels

    Abeer Fadda,1† Mark Ryten,2 Dorothea Droll,1

    Federico Rojas,3 Valentin Färber,1‡

    Jurgen R. Haanstra,4,5 Clemetine Merce,1

    Barbara M. Bakker,4,5 Keith Matthews3 andChristine Clayton1*1Zentrum für Molekulare Biologie der UniversitätHeidelberg (ZMBH), DKFZ-ZMBH Alliance, ImNeuenheimer Feld 282, 69120 Heidelberg, Germany.2Flat 1,172 Gloucester Terrace, London, UK.3Centre for Immunity, Infection and Evolution, Institutefor Immunology and infection Research, School ofBiological Sciences, Ashworth Laboratories, Universityof Edinburgh, West Mains Road, Edinburgh EH9 3JT,UK.4Department of Pediatrics and Systems Biology, Centrefor Energy Metabolism and Ageing, University MedicalCenter Groningen, University of Groningen, Groningen,The Netherlands.5Department of Molecular Cell Physiology, Faculty ofEarth and Life Sciences, VU University Amsterdam,Amsterdam, The Netherlands.

    Summary

    African trypanosomes are an excellent system forquantitative modelling of post-transcriptional mRNAcontrol. Transcription is constitutive and polycis-tronic; individual mRNAs are excised by trans splic-ing and polyadenylation. We here measure mRNAdecay kinetics in two life cycle stages, bloodstreamand procyclic forms, by transcription inhibition andRNASeq. Messenger RNAs with short half-lives tendto show initial fast degradation, followed by a slowerphase; they are often stabilized by depletion of the5′–3′ exoribonuclease XRNA. Many longer-livedmRNAs show initial slow degradation followed byrapid destruction: we suggest that the slow phasereflects gradual deadenylation. Developmentally

    regulated mRNAs often show regulated decay, andswitch their decay pattern. Rates of mRNA decay aregood predictors of steady state levels for shortmRNAs, but mRNAs longer than 3 kb show unexpect-edly low abundances. Modelling shows that varia-tions in splicing and polyadenylation rates cancontribute to steady-state mRNA levels, but this iscompletely dependent on competition between pro-cessing and co-transcriptional mRNA precursordestruction.

    Introduction

    The rates of transcription, RNA splicing, localization anddecay all affect the levels of cellular mRNAs, while proteinlevels are additionally affected by the rates of translationand protein turnover. In mammalian cells, mRNA abun-dance makes a contribution of up to 40% to final proteinlevels (Schwanhäusser et al., 2011). Results from highthroughput techniques have revealed that decay rates arecritical in overall mRNA homeostasis; genes encodingproteins with similar functions share similar transcriptdecay rates, and the decay rates respond to externalstimuli (Shalem et al., 2008; Munchel et al., 2011; Rabaniet al., 2011; Schwanhäusser et al., 2011).

    Traditionally, decay rates have been measured byextracting mRNA at various times after transcription shut-off. More recently, a method involving pulse-chase with4-thiouracil has enabled less invasive measurements, butthis method is limited to cell types permissive to the labeluptake and incorporation into newly synthesized RNA(e.g. Munchel et al., 2011; Rabani et al., 2011). Indepen-dently of the method used, transcriptome-wide decaystudies can provide insight into the specificity (Grigullet al., 2004; Mnaimneh et al., 2004; Houalla et al., 2006;Chekanova et al., 2007; Kiss and Andrulis, 2010) anddynamics (Gudipati et al., 2012; Harigaya and Parker,2012; Schneider et al., 2012; Deneke et al., 2013) of thedecay machinery. Some of these studies revealed thepresence of feedback loops between synthesis and decay(Shalem et al., 2008; Trcek et al., 2011; Sun et al., 2012;Haimovich et al., 2013). In many cases, decay kineticsare not exponential; this can be explained by models in

    Accepted 19 August, 2014. *For correspondence. E-mail [email protected]; Tel. (+49) 6221 546876; Fax (+49) 6221546891. Present affiliations: †Sidra Medical and Research Center,P.O. Box 2699, Doha, Qatar; ‡Fresenius Kabi Deutschland GmbH,Bad Homburg, Germany.

    Molecular Microbiology (2014) 94(2), 307–326 ■ doi:10.1111/mmi.12764First published online 15 September 2014

    © 2014 The Authors. Molecular Microbiology published by John Wiley & Sons Ltd.This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution andreproduction in any medium, provided the original work is properly cited.

    mailto:[email protected]:[email protected]://creativecommons.org/licenses/by/3.0/

  • which the ‘age’ of an mRNA influences its decay rate(Deneke et al., 2013).

    African trypanosomes are unicellular parasites with amode of gene expression that is uniquely suited to thestudy of post-transcriptional mRNA control mechanisms.Transcription is polycistronic, with no apparent regulationof RNA polymerase II initiation (Palenchar and Bellofatto,2006; Daniels et al., 2010). Mature mRNAs are transspliced, with addition of a capped spliced leadersequence at the 5′ end (Michaeli, 2011); polyadenylationat the 3′ end is determined by, and coupled to, the transsplicing of the next transcript downstream (Michaeli,2011). Measurements that focused on trans splicing of thetubulin mRNA (in an in vitro system) (Ullu et al., 1993) andthe PGKB mRNA (in vivo) both found precursor half-livesof about 2 min (Haanstra et al., 2008), which suggeststhat the mRNA may be spliced once the RNA polymeraseis 1–2 kb further downstream (Haanstra et al., 2008).Results for the HSP70 tandem repeat were consistentwith these estimates (Huang and van der Ploeg, 1991).However, it is not known whether this trans splicing rate isgenerally applicable.

    Two trypanosome life cycle stages are routinely cul-tured in the laboratory: the bloodstream form, which isequivalent to the form in mammalian blood, and the pro-cyclic form, equivalent to the form that grows in the midgutof the Tsetse vector. Most open reading frames have morethan one possible spliced leader addition site, and often,several polyadenylation sites are used; moreover, thechoice of both can vary between life cycle stages (Kolevet al., 2010; Nilsson et al., 2010; Siegel et al., 2010; 2011;Zhang et al., 2010). From RNASeq experiments, approxi-mately 400 genes were judged to be developmentallyregulated, with approximately equal numbers being pref-erentially expressed in each form (Siegel et al., 2010).Numerous studies have shown that developmental regu-lation of individual mRNAs is associated with differencesin their decay rates (Fernández-Moya and Estévez, 2010;Kramer and Carrington, 2011; Clayton, 2013). Oneexample, that of PGKB, has been subjected to quantita-tive modelling. The PGKB mRNA is very unstable inbloodstream forms, and stable in the procyclic form.Measurement and modelling of its synthesis and degra-dation kinetics indicated that the level of the PGKB mRNAwas predominantly controlled by the rates of transcriptionand degradation, and life cycle regulation was achievedby changes in the degradation rate (Haanstra et al.,2008). The extent to which differences in mRNA degrada-tion were responsible for developmental regulation on atranscriptome-wide scale was however unknown.

    Degradation of trypanosome mRNAs follows the classiceukaryotic paradigm: the CAF1-NOT complex dead-enylates the mRNA (Erben et al., 2014), and this is fol-lowed by 5′–3′ degradation by XRNA (Li et al., 2006) or

    3′–5′ degradation by the exosome (Clayton and Estevez,2010). The PAN2/PAN3 deadenylase also has a role indegradation of some mRNAs, and there is, in addition, adeadenylation-independent pathway that is initiated by 5′degradation (Schwede et al., 2009; Fadda et al., 2013).Studies of mRNA degradation currently rely on transcrip-tion inhibition. Unfortunately, pulse-labelling with 4-thio-uracil is not feasible: not only is there hardly any uptake,but there is also a bypass pathway for uridine synthesis(Ali et al., 2013) and a very strong flux of the nucleotideinto glycosylation of surface proteins and other pathways(Ali et al., 2012). We previously obtained approximateestimates of bloodstream-form trypanosome mRNA half-lives by inhibiting transcription for 30 min. We demon-strated that for most mRNAs, deadenylation precedesdegradation (Manful et al., 2011; Fadda et al., 2013) andalso found that the 5′–3′ exoribonuclease XRNA wasimportant mainly for degradation of very unstable mRNAs(Manful et al., 2011). This confirmed previous observa-tions made with individual mRNAs (Schwede et al., 2008;2009). The analyses were however limited by the use of asingle time-point for transcription inhibition, meaning thathalf-life estimates were very approximate and complexkinetics could not be detected. Also, procyclic trypano-somes were not studied.

    The work described here comprises part of a study thataims to model trypanosome metabolic pathways fromgenome to metabolome. To allow quantitative modelling,as many parameters as possible must be measured intrypanosomes cultured under tightly defined conditions(Bakker et al., 2010). In this paper we describe a detailedanalysis of mRNA decay in both bloodstream form andprocyclic trypanosomes, including complex decay kinet-ics. We find extensive variations in mRNA decay, bothbetween mRNAs and between life cycle stages, but theseare usually insufficient to account for steady-state mRNAabundance. By modelling, we uncover a potentially impor-tant role for competition between mRNA processing andco-transcriptional RNA destruction in determining steady-state mRNA levels.

    Results

    Developmental regulation of steady-state mRNA levels

    In this work we aimed to measure the half-lives of alltrypanosome mRNAs in both bloodstream and procyclicforms, and to obtain an overview of trans splicing kinetics.To measure mRNA decay we inhibited trans splicing andtranscription for various times, using Sinefungin toprevent splicing and Actinomycin D to stop transcription;to measure splicing we used Actinomycin D alone. Weprepared RNA, depleted rRNA by subtraction and thenmeasured relative RNA amounts with RNAseq. All

    308 A. Fadda et al. ■

    © 2014 The Authors. Molecular Microbiology published by John Wiley & Sons Ltd., Molecular Microbiology, 94, 307–326

  • samples are listed in Supplementary Table S1, whichincludes the raw data. Supplementary Table S2 summa-rizes the main results. Details are in subsequent Supple-mentary Tables. Most analyses concentrate on a non-redundant list of genes in which repeated copies havebeen removed (Siegel et al., 2010).

    We first analysed steady state mRNA levels, comparingresults for rRNA-depleted mRNA from four bloodstreamform samples and four procyclic samples. Statistically sig-nificant differential expression was seen for 693 genes, ofwhich 430 transcripts were more abundant in bloodstreamforms (Supplementary Table S3, sheets 1 & 2). We com-pared the these results with previously published data forpoly(A) + mRNA (Siegel et al., 2010; Singh et al., 2014;Vasquez et al., 2014) (Supplementary Figs S1 and S2,and Supplementary Table S3, sheet 3). There were cleardifferences between the datasets, which could have beencaused by differences in the trypanosomes used, theculture conditions, the methods used for RNA preparationand sequencing, and the tools used for data extractionand analysis. This illustrates the fact that for quantitativework, it is vital to control all conditions and methods astightly as possible.

    The comparisons enabled us to compile a list of the 450most reproducibly developmentally regulated mRNAs (atleast twofold regulated in at least 3 of the 4 availableRNA-Seq datasets, Supplementary Table S3, sheet 4).We grouped these according to manually annotated func-tional categories. Rather surprisingly, almost twice asmany mRNAs were reproducibly upregulated in blood-stream forms than were upregulated in procyclics; andoddly, genes with no known function accounted for all ofthe difference. Figure 1 shows just the upregulated genesthat could be functionally assigned. The regulation ofenergy metabolism was as expected – mitochondrialpathways present in procyclic forms, and reliance on gly-colysis in bloodstream forms (Bringaud et al., 2006).Bloodstream forms are also known to have a high require-ment for vesicular transport and GPI anchor biosynthesis(Overath and Engstler, 2004). Bloodstream-form-specificexpression of genes encoding cytoskeletal proteins,protein kinases and phosphatases, and RNA-binding pro-teins was less expected. Levels of RNA, however, do notalways correlate with levels of protein (Gunasekera et al.,2012; Urbaniak et al., 2013). Translation (Vasquez et al.,2014) and protein degradation are also important.

    Our procyclic forms were grown with proline as theprincipal carbon source, with glucose only from the dilutedserum. In contrast, the procyclic forms used by others(Siegel et al., 2010; Vasquez et al., 2014) were grown inmedium with 5 mM glucose. Thirteen mRNAs reproduc-ibly showed at least a threefold greater increase in ourprocyclics [ribo- and poly(A) + datasets] than was seen inthe RNASeq data for glucose-grown cells (Supplementary

    Table S3, sheet 3, columns O–R; names in bold in columnC). The increased expression of proline dehydrogenase(Tb927.7.210) has been reported previously (Lamouret al., 2005), and glutamine synthetase (Tb927.7.4970) isnext on the same pathway (Bringaud et al., 2006). ThemRNAs encoding glycosomal malate dehydrogenase(Tb927.10.15410), ubiquinol-cytochrome c reductase(Tb927.4.4990) and an aquaporin (Tb927.6.1520) arealso upregulated in proline-cultured cells. Specificallyincreased mRNAs encoding proteins of unknown functionwere Tb927.4.3500, Tb927.9.13200, Tb927.11.1690 andTb927.10.2360. This shows again that carbon sourceavailability affects expression levels in trypanosomes.

    The heterogeneity of trans splicing and polyadenylationsites precludes transcriptome-wide rate measurement ofprocessing kinetics

    In order to model gene expression, we needed an esti-mate of the variability in pre-mRNA splicing rates. In anattempt to obtain this, we treated cells with Actinomycin Dalone for the shortest feasible time, 5 min, and subjectedrRNA-depleted RNA to RNASeq. Since the three previ-ously measured splice times were in the range of 1–2 min,time intervals in that range would have been preferable,but were not technically feasible. For the splicing meas-urements we used two replicates for bloodstream formsand one for procyclic forms, together with one untreatedsample for each. Reads aligning to intergenic regionswithin a transcription unit come from un-spliced mRNA.Trans splicing involves trans esterification followed bydestruction of the intergenic RNA by exoribonucleases.Most trypanosome mRNAs, however, have multiple pos-sible splice sites (Kolev et al., 2010; Nilsson et al., 2010;Siegel et al., 2010; 2011; Zhang et al., 2010), whichmeans that many RNA sequences can be either inter-genic or within an mRNA. We collected all documentedsplice acceptor sites for both procyclic and bloodstreamforms from TriTrypDB, and merged them with our ownidentified sites. For each site, the read counts in the20mer upstream region were extracted, normalized to theread counts for the corresponding gene, and the ratiobetween untreated and Actinomycin D-treated sampleswas calculated. We then calculated the pre-mRNA half-lifefor each site, assuming exponential kinetics. If an openreading frame had more than one splice site, we chosethe splice site with the shortest half-life to infer splicingkinetics for that open reading frame. 3029 measurementsof 5′ splicing for bloodstream forms and 1554 for procyclicforms are shown in Supplementary Table S2 and dis-played in Supplementary Fig. S3. The measurementswere limited by the numbers of reads, as well as by theuse of only a single time interval. The median 5′ pre-mRNA half-life measured in both life stages was a little

    Transcriptome-wide mRNA decay in trypanosomes 309

    © 2014 The Authors. Molecular Microbiology published by John Wiley & Sons Ltd., Molecular Microbiology, 94, 307–326

  • Mitochondrialpathway

    Mitochondrial electrontransport

    Oxidative phosphorylation

    Citric acid cycle

    Unknown function: 27Mitochondrial

    transport& membrane

    Mitochondrialbiogenesis

    Mitochondrialtranslation

    Mitochondrial RNA Glucose & glycerol

    Pentose phosphatepathway

    Amino acidmetabolism

    Lipids &fatty acids

    Sterols

    Prenylation Nucleotides

    Redox pathways

    Enzyme & Folate

    Ribosome Translation, tRNA

    RNA processing Membrane protein

    Protease or peptidase Chaperone

    Ion channelor pump

    PROCYCLIC FORM

    BLOODSTREAM FORM

    Cell cycle Protein phosphatase

    Vesicular transport

    GPI anchors& metabolism

    Lysosome Glycan synthesis& metabolism

    Chaperone

    Cytoskeleton

    Protein kinase,phosphatase,

    signalling

    RNA binding,zinc finger

    RNA synthesis,degradation,

    helicase

    DNA synthesis,modification,repair, chromatin

    Glucose &glycerol

    Lipids & fatty acids

    Nucleotides

    Redox pathways Enzyme & folate

    Phosphatase Protease, peptidase Ubiquitin & proteasome

    Transporter Membrane

    protein

    GRESAG,LRRP,PAG

    Unknown function: 160

    129 genes

    134 genes

    Fig. 1. Developmentally regulated mRNAs.For each stage, genes that were upregulated in at least three of the four available RNASeq datasets were selected (Supplementary Table S3,sheet 4) and grouped according to manually assigned functional categories. Genes encoding proteins of no known function (27 for procyclicforms, 160 for bloodstream forms) are not shown. ‘Upregulated’ means at least twofold more abundant than in the other stage; for our currentdataset it also means significantly regulated.

    310 A. Fadda et al. ■

    © 2014 The Authors. Molecular Microbiology published by John Wiley & Sons Ltd., Molecular Microbiology, 94, 307–326

  • over 3 min, and many of the measured half-times werebetween 1–2 min. Since these times are similar to thosemeasured for tubulin, HSP70 and PGK mRNAs, they maybe correct. However, there was no correlation in 5′ splic-ing times between the two life stages even for genessharing the same major acceptor site in the two stages(not shown). Also, many precursor half-lives were over4–5 min. Although this could reflect slow splicing, it couldalso indicate persistence of bicistronic RNAs, or of mono-cistronic mature mRNAs that had upstream trans splicesites. The un-spliced PGK precursor is degraded with ahalf-life of 8 min (Haanstra et al., 2008) so half-livesabove this are unlikely to reflect splicing, but may insteadreflect precursor degradation. It was thus apparent that formost mRNAs, the complexity of trypanosome mRNA splic-ing patterns precluded reliable kinetic analysis using thehigh-throughput method that we had adopted. Our resultstherefore could not be used to assess the kinetics ofsplicing. Polyadenylation sites are even more heteroge-neous so we made no attempt to analyse 3′-processingkinetics.

    Despite the limitations in the results, they did suggestthat many trypanosome mRNAs might be spliced within afew minutes of synthesis. Based on this, we adopted aconservative splicing half-time of 2 ± 1.5 min in the mod-elling (below).

    Trypanosome mRNAs show complex decay kinetics

    We next analysed the data for parasites treated with bothSinefungin and Actinomycin D, using multiple time inter-vals and at least two replicates for each time point. Tonormalize the data we hybridized Northern blots with aprobe complementary to the spliced leader (see Methodsand Supplementary Fig. S4). The spliced leader provides agood measure of the number of mRNAs present, becauseit is present at the 5′-end of every mRNA. Unlike the poly(A)tail, it does not vary in length; and it will be removed byXRNA very fast after the mRNA has been decapped.

    To calculate half-lives, we considered three possiblescenarios: simple exponential decay, or one of two non-exponential patterns (Deneke et al., 2013). The non-exponential models represent a multi-step degradationprocess in which different mRNA molecules are in differ-ent stages of degradation; hence the ‘age’ of the mRNApopulation (i.e. the length of time after transcription hasstopped) plays a role in the shape of overall decay curve(Deneke et al., 2013). In the first scenario, the mRNApopulation initially seems relatively stable but becomesincreasingly unstable over time: we call this the ‘slow-fast’scenario. For example, the actin mRNA shows little deg-radation over the first 10 minutes, but then the rateincreases (Fig. 2A). In the second scenario, part of theRNA is degraded very rapidly, but a fraction appears to be

    much more stable (fast-slow scenario). An example of thisis shown in Fig. 2B. The model that fit the data best(smallest least-squares error) was used to calculate asingle transcript lifetime and half-life for each mRNA andthe resulting half-lives (Supplementary Table S4) wereused in all further analyses. The results are summarizedin Table 1. Sometimes the fit was always poor, especiallywhen there was extensive scatter at early time points(Supplementary Fig. S5A and B), while in other cases,moving from an exponential model to one of the othermodels clearly gave a better fit (Supplementary Fig. S5Cand D). As noted previously (Deneke et al., 2013), tran-scripts with short half-lives tended to fit the fast-slowmodel, while those with long half-lives tended to fit aslow-fast model (Fig. 2C and D).

    Some mRNAs showed no significant degradation overthe incubation period, and for some, the abundance wastoo low for a half-life measurement. However, the use ofmultiple time points and complex kinetic models enabledus to estimate half-lives for most mRNAs. Bloodstreamtrypanosome mRNAs had a modal half-life of 9 min,median 12 min, and arithmetic mean 16 min (Fig. 3),similar to the numbers that we previously reported (Manfulet al., 2011). Procyclic trypanosomes are larger thanbloodstream forms, grow at 27°C, have about twice asmany mRNAs per cell and divide more slowly (Haanstraet al., 2008); correspondingly, their mRNAs had a modalhalf-life of 12 min, median of 20 min, and arithmetic meanof 34 min (Fig. 3, Supplementary Table S3 sheet 5).

    RNASeq half-lives are similar to those found byother methods

    We previously had estimated half-lives for bloodstreamforms by measuring the amount of mRNA remaining after30 min of transcription inhibition (Manful et al., 2011). Thecorrelation between our new and old datasets was bestfor mRNAs with longer half-lives (Supplementary Fig. S6).One likely reason for this is that the reliability of RNASeq

    Table 1. Summary of half-life results.

    Category Bloodstream Procyclic

    Slow-fast 1903 2632Fast-slow 3074 1622Exponential 1495 1704Stable 92 149Not measured 480 993

    Results given are for 6951 unique open reading frames. mRNAs thathad half-lives over 120 min (bloodstream forms) or 240 min (procyclicforms) were mostly fitted using the exponential model, but wereclassified as stable since the half-life could not be measured wellusing the time points that we had. ‘Not measured’ usually means thatthe numbers of counts were insufficient to calculate a half-life and/orthat the data could not be fitted with any of the three models.

    Transcriptome-wide mRNA decay in trypanosomes 311

    © 2014 The Authors. Molecular Microbiology published by John Wiley & Sons Ltd., Molecular Microbiology, 94, 307–326

  • decreases at low read densities, and unstable RNAs givevery few reads after 30 min of degradation.

    The half-lives that we measured by RNA-Seq were inquite good agreement with previously reported half-livesmeasured by real-time PCR or Northern blotting, withcorrelation coefficients (for log-transformed data) of over0.8 (Fig. 2A and B; Fig. 4, Table 2). When we comparedthe RNASeq results with our previous measurementsmade using the same protocols for RNA synthesis inhibi-tion, differences in the kinetics were rather minor (Sup-plementary Fig. S7). RNA-Seq might be somewhat moresensitive at the later time-points, when RNA levelsdropped even below the levels that could be measuredreliably by q-PCR. Disagreements with literature values

    (Table 2) could be due to differences in the cell cultureconditions as well as in methodology.

    Changes in mRNA decay partially account fordevelopmental regulation

    We next asked whether mRNAs encoding proteins withparticular functions showed similar half-lives. We groupedmRNAs into functional classes (Supplementary Table S2)and plotted half-lives for each class (Fig. 5). A one-wayANOVA showed significant life-time differences betweenthe different classes (P-value

  • stages: large amounts of mRNA are produced from only1–3 open reading frames (Manful et al., 2011). Although allRNA classes were overall more stable in the procyclicstage (T-test results in Supplementary Table S5), morepronounced shifts were apparent for functions related toprocyclic energy metabolism, such as amino acids, citricacid cycle, mitochondrial biogenesis and oxidative phos-phorylation. The small number of mRNAs involved in folateand sterol metabolism also showed greater increases inhalf-life than average. These results were consistent withthe idea that developmental regulation is, at least in part,mediated through changes in half-life.

    To analyse the relationship further we plotted half-liferatios (bloodstream half-life/procyclic half-life) against theabundance ratios (bloodstream abundance/procyclicabundance). There was almost no correlation, except forvery strongly regulated mRNAs (Supplementary Fig. S8).Moreover, not all mRNAs with regulated turnover wereregulated in abundance. Restriction of the dataset to themRNAs that were rapidly spliced in both forms, or usingonly those mRNAs that had the same splice site in bothforms, did not increase the correlation (not shown).However, this analysis almost certainly under-estimatedthe contribution of mRNA turnover to developmental

    0

    500

    1000

    1500

    2000

    0-5 5-10

    10-15

    15-20

    20-25

    25-30

    30-35

    35-40

    40-45

    45-50

    50-55

    55-60

    60-12

    0>1

    20

    No.

    of g

    enes

    mRNA half-life (min)

    Procyclic form

    Bloodstream form

    Fig. 3. Most trypanosome mRNAs havehalf-lives of less than 30 min.All mRNAs with measurable half-lives (stableones are excluded) were placed into classesaccording to half-life range. The numbers ofmRNAs in each class are plotted. All data arefor ribo-minus RNA preparations.

    2

    3

    4

    5

    6

    7

    8

    3 4 5 6 7 8

    y = -1.4 + 1.2x R= 0.82

    PROCYCLICBLOODSTREAM

    RNA-Seq

    Nor

    ther

    n / q

    PC

    R

    RNA-Seq

    Nor

    ther

    n / q

    PC

    R

    2

    2.5

    3

    3.5

    4

    4.5

    5

    5.5

    6

    2.5 3 3.5 4 4.5 5 5.5 6 6.5y = 0.2 + 0.9x

    R= 0.81

    Fig. 4. Comparison between RNASeq and other methods.All available RNA half-lives measured by qRT-PCR or Northern blotting were compared with the half-lives measured by RNA-Seq. Both axesshow the log (base 2) of the half-life in min. References are in Table 2. If more than one Northern measurement was available the arithmeticmean was used. Only measurements with sufficient time points were included: for example, the numbers designated as ‘more than’ or ‘lessthan’ a certain half-life in Table 2 were excluded. The formulae and correlation coefficients were calculated with log2-transformed numbers.

    Transcriptome-wide mRNA decay in trypanosomes 313

    © 2014 The Authors. Molecular Microbiology published by John Wiley & Sons Ltd., Molecular Microbiology, 94, 307–326

  • Tab

    le2.

    Com

    paris

    onof

    RN

    AS

    eqha

    lf-liv

    esw

    ithN

    orth

    ern

    blot

    resu

    lts.

    Man

    fulB

    SB

    Sne

    wB

    SN

    orth

    ern/

    qPC

    RP

    Cne

    wP

    CN

    orth

    ern

    Ref

    Not

    regu

    late

    dT

    b927

    .5.4

    170

    His

    tone

    H4

    6845

    #30

    –45

    stab

    le>

    60m

    inP

    C(n

    p);

    BS

    (Sch

    wed

    eet

    al.,

    2009

    )(F

    ärbe

    ret

    al.,

    2013

    )T

    b927

    .9.8

    850

    Act

    inA

    2822

    #15

    –25

    33#

    ∼90

    ,52

    #,70

    #B

    Se.

    g.(n

    p),

    (Hot

    zet

    al.,

    1997

    );P

    C(n

    p),

    (Est

    évez

    ,20

    08;

    Kra

    mer

    etal

    .,20

    08)

    Tb9

    27.1

    .233

    0b-

    TU

    B70

    66#

    45–6

    0#13

    5#∼

    120

    (Man

    fule

    tal.,

    2011

    )#

    (Kra

    mer

    etal

    .,20

    08;

    Sch

    wed

    eet

    al.,

    2008

    )T

    b927

    .7.3

    780

    PA

    P-li

    ke9

    98

    15#

    na(M

    anfu

    leta

    l.,20

    11)

    Tb9

    27.8

    .281

    0X

    RN

    C11

    66

    9na

    (Man

    fule

    tal.,

    2011

    )T

    b927

    .4.4

    10C

    AF

    4026

    23#

    3046

    #na

    (Man

    fule

    tal.,

    2011

    )T

    b927

    .11.

    1660

    0E

    AP

    29

    128

    5na

    (Man

    fule

    tal.,

    2011

    )T

    b927

    .11.

    1631

    0M

    yosi

    n16

    3114

    11na

    (Man

    fule

    tal.,

    2011

    )T

    b927

    .11.

    4710

    Hyp

    910

    813

    na(M

    anfu

    leta

    l.,20

    11)

    Tb9

    27.1

    1.79

    0H

    yp11

    66

    18na

    (Man

    fule

    tal.,

    2011

    )T

    b927

    .11.

    8710

    Hyp

    109

    69

    na(M

    anfu

    leta

    l.,20

    11)

    Tb9

    27.6

    .281

    0A

    BC

    119

    516

    na(M

    anfu

    leta

    l.,20

    11)

    Tb9

    27.7

    .698

    0S

    F3b

    1514

    1728

    na(M

    anfu

    leta

    l.,20

    11)

    Tb9

    27.6

    .434

    0LS

    m5

    1013

    #9

    15na

    (Man

    fule

    tal.,

    2011

    )T

    b927

    .11.

    1133

    0H

    SP

    7021

    38#

    3910

    2na

    (Man

    fule

    tal.,

    2011

    )T

    b927

    .6.3

    740

    PD

    I263

    55na

    156

    ∼60

    (Kel

    lyet

    al.,

    2012

    )T

    b927

    .11.

    5050

    FR

    m23

    36na

    85#

    37(E

    stév

    ez,

    2008

    )T

    b927

    .11.

    1092

    0H

    yp11

    12na

    31∼

    10(S

    tern

    etal

    .,20

    09)

    Tb9

    27.1

    0.12

    700

    PY

    DH

    E53

    36na

    stab

    le45

    (Ste

    rnet

    al.,

    2009

    )B

    lood

    stre

    am-s

    peci

    ficT

    b927

    .10.

    7090

    AO

    Xst

    able

    stab

    lest

    able

    na3.

    2h

    (Cha

    udhu

    riet

    al.,

    2002

    )T

    b927

    .10.

    5620

    ALD

    stab

    lest

    able

    90#

    182

    ∼4

    h#(M

    ayho

    etal

    .,20

    06)

    Tb9

    27.2

    .600

    0G

    PI-

    PLC

    stab

    le53

    #30

    #13

    5(W

    ebb

    etal

    .,20

    05)

    Pro

    cycl

    ic-s

    peci

    ficT

    b927

    .10.

    1025

    0E

    Ppr

    ocyc

    linna

    na5–

    1098

    200

    (Hot

    zet

    al.,

    1997

    )an

    dot

    hers

    Tb9

    27.1

    .410

    0C

    OX

    IV14

    21#

    <30

    216

    >4

    h#(M

    ayho

    etal

    .,20

    06)

    Tb9

    27.9

    .317

    0C

    OX

    V15

    9#<

    3020

    3∼

    4h

    (May

    hoet

    al.,

    2006

    )T

    b927

    .10.

    280

    CO

    XV

    I15

    11#

    <30

    198

    ∼3

    h(M

    ayho

    etal

    .,20

    06)

    Tb9

    27.3

    .141

    0C

    OX

    VII

    2612

    <30

    stab

    le∼

    6h#

    (May

    hoet

    al.,

    2006

    )T

    b927

    .4.4

    620

    CO

    XV

    III17

    5<

    30st

    able

    ∼6

    h#(M

    ayho

    etal

    .,20

    06)

    Tb9

    27.1

    0.83

    20C

    OX

    IX26

    14#

    <30

    stab

    le∼

    8h#

    (May

    hoet

    al.,

    2006

    )T

    b927

    .7.7

    420

    AT

    Psy

    nth

    a17

    22#

    <30

    199

    >60

    #(B

    row

    net

    al.,

    2001

    )T

    b927

    .3.1

    380

    AT

    Psy

    nth

    b32

    27#

    <30

    152

    >90

    #(B

    row

    net

    al.,

    2001

    )T

    b927

    .5.1

    710

    AT

    Psy

    nth

    930

    22#

    <30

    stab

    le>

    240

    (Bro

    wn

    etal

    .,20

    01)

    Tb9

    27.6

    .217

    0G

    rpE

    3830

    #45

    130#

    na(M

    anfu

    leta

    l.,20

    11)

    Tb9

    27.1

    1.62

    80P

    PD

    K36

    417*

    *st

    able

    na(H

    aile

    etal

    .,20

    03)

    Tb9

    27.4

    .473

    0A

    AT

    P11

    1828

    #na

    104#

    82#,

    40#

    #(E

    stév

    ez,

    2008

    ),(S

    tern

    etal

    .,20

    09)

    Tb9

    27.7

    .497

    0G

    S12

    10na

    75#

    65#

    #(E

    stév

    ez,

    2008

    )T

    b927

    .10.

    1077

    0G

    CS

    131

    29na

    60#

    60(E

    stév

    ez,

    2008

    )T

    b927

    .3.4

    500

    FR

    c16

    38na

    stab

    le11

    6(E

    stév

    ez,

    2008

    )T

    b927

    .6.1

    520

    AQ

    P3

    8141

    #na

    stab

    le∼

    80(S

    tern

    etal

    .,20

    09)

    Man

    fulB

    S–

    Blo

    odst

    ream

    -for

    mR

    NA

    Seq

    resu

    ltsfr

    om(M

    anfu

    leta

    l.,20

    11).

    BS

    new

    ,PC

    new

    :res

    ults

    from

    the

    curr

    enta

    naly

    sis;

    BS

    =bl

    oods

    trea

    mfo

    rms.

    PC

    =pr

    ocyc

    licfo

    rms.

    BS

    orP

    CN

    orth

    ern:

    Est

    imat

    esfr

    omN

    orth

    ern

    blot

    s,us

    ing

    data

    from

    the

    refe

    renc

    esci

    ted

    onth

    erig

    htha

    ndco

    lum

    n.N

    ote

    that

    for

    man

    yof

    thes

    e,es

    timat

    esof

    accu

    racy

    are

    not

    avai

    labl

    ean

    dth

    em

    etho

    dsfo

    rin

    hibi

    ting

    tran

    scrip

    tion

    varie

    d.A

    llha

    lf-liv

    esar

    ein

    min

    utes

    .N

    on-s

    tand

    ard

    gene

    abbr

    evia

    tions

    are:

    b-T

    UB

    –be

    tatu

    bulin

    ;M

    yosi

    nH

    C–

    myo

    sin

    heav

    ych

    ain;

    Hyp

    –hy

    poth

    etic

    open

    read

    ing

    fram

    een

    codi

    ngpr

    otei

    nof

    unkn

    own

    func

    tion

    AB

    C–

    AB

    Ctr

    ansp

    orte

    r;P

    AP

    –po

    ly(A

    )po

    lym

    eras

    e-lik

    e;A

    TP

    synt

    h–

    AT

    Psy

    ntha

    se(s

    ubun

    itsa-

    alph

    a,b-

    beta

    ,an

    d9)

    .V

    alue

    sw

    ithas

    teris

    ksar

    efo

    rre

    port

    erm

    RN

    As

    with

    the

    3′-U

    TR

    ofth

    ege

    neco

    ncer

    ned,

    synt

    hesi

    zed

    byR

    NA

    poly

    mer

    ase

    I(*

    )or

    T7

    poly

    mer

    ase

    (**)

    .Tr

    ansc

    riptio

    nby

    T7

    poly

    mer

    ase

    mig

    htaf

    fect

    the

    half-

    life

    (Col

    asan

    teet

    al.,

    2007

    ).(n

    p)–

    our

    data

    ,no

    tpu

    blis

    hed.

    #Slo

    w-f

    ast

    kine

    tics.

    314 A. Fadda et al. ■

    © 2014 The Authors. Molecular Microbiology published by John Wiley & Sons Ltd., Molecular Microbiology, 94, 307–326

  • regulation. We had to exclude not only all stable mRNAs,but also those for which abundance in one stage was toolow for the half-life to be measured. The latter included29 procyclic-form-specific and 172 bloodstream-form-specific mRNAs (Supplementary Table S3, sheet 3).Excluding these, we found 179 mRNAs in which mRNAdecay played a clear role in regulation (SupplementaryTable S3, sheet 6). We saw regulated mRNA decay notonly for components of the cytochrome oxidase (Mayhoet al., 2006) and ATP synthase (Brown et al., 2001) com-plexes (Table 2), but also for 67 other mitochondrialproteins, the AATP11 amino acid transporter, and 9 glyco-somal or cytosolic enzymes involved in procyclic energymetabolism. For bloodstream form-specific expression, wedetected regulation for the GPI-phospholipase C (Webbet al., 2005); the mRNAs encoding fructose bisphosphatealdolase and the alternative oxidase were relatively stablein procyclic forms but showed no decay at all during ourassay in bloodstream forms (Table 2).

    Overall, more mRNAs fit a slow-fast model in procyclicthan in bloodstream forms (Table 1), and no fewer than1110 transcripts had different decay models in the two lifestages. 976 mRNAs switched from fast-slow in blood-

    stream forms to slow-fast in procyclics. These were par-ticularly enriched for mitochondrial processes that aredevelopmentally upregulated in procyclics: mitochondrialtranslation, RNA processing, electron transport, biogen-esis and other pathways. Other classes that showed thisswitch were prenylation, RNA processing and the catch-all ‘enzyme’ group. The 134 that switched the other waywere mostly upregulated in bloodstream forms and wereenriched for GPI anchors and metabolism, vesiculartransport, calcium regulation and DNA synthesis, modifi-cation and repair.

    Many mRNAs are less abundant than expected fromtheir half-lives

    Since trypanosome mRNAs are constitutively transcribed,gene copy number and mRNA half-life are expected tomake a significant contribution to steady-state mRNAabundance. We therefore normalized mRNA copynumbers per cell to the gene copy number, then com-pared these values with the half-life. Results suggestedthat half-life contributes to, but cannot completely accountfor, the mRNA level (Fig. 6). High abundance was

    0

    50

    100

    150

    200

    Hal

    f-life

    (min

    )

    Aci

    doca

    lcis

    ome

    Am

    ino

    acid

    met

    abol

    ism

    Am

    ino

    acid

    tran

    spor

    ter

    Arg

    inin

    e m

    ethy

    latio

    nC

    alci

    umC

    ell c

    ycle

    Cha

    pero

    neC

    hrom

    atin

    Citr

    ic a

    cid

    cycl

    eC

    ytos

    kele

    ton

    DN

    A s

    ynth

    esis

    , mod

    ifica

    tion

    & re

    pair

    Enz

    yme

    ES

    AG

    Fola

    teG

    luco

    se &

    gly

    cero

    lG

    lyca

    n sy

    nthe

    sis

    & m

    etab

    olis

    mG

    lyco

    som

    e bi

    ogen

    esis

    GP

    I anc

    hors

    & m

    etab

    olis

    mG

    RE

    SA

    GIo

    n ch

    anne

    l or p

    ump

    Lipi

    ds &

    fatty

    aci

    dsLR

    RP

    Lyso

    som

    eM

    embr

    ane

    prot

    ein

    Mito

    chon

    dria

    l bio

    gene

    sis

    Mito

    chon

    dria

    l DN

    AM

    itoch

    ondr

    ial e

    lect

    ron

    trans

    port

    Mito

    chon

    dria

    l mem

    bran

    eM

    itoch

    ondr

    ial p

    athw

    ayM

    itoch

    ondr

    ial R

    NA

    Mito

    chon

    dria

    l tra

    nsla

    tion

    Mito

    chon

    dria

    l tra

    nspo

    rtN

    ucle

    o-cy

    topl

    asm

    ic tr

    ansp

    ort

    Nuc

    leot

    ides

    Oxi

    dativ

    e ph

    osph

    oryl

    atio

    nP

    ento

    se p

    hosp

    hate

    pat

    hway

    Pho

    spha

    tase

    Pre

    nyla

    tion

    Pro

    teas

    e or

    pep

    tidas

    eP

    rote

    in k

    inas

    eP

    rote

    in p

    hosp

    hata

    seR

    edox

    pat

    hway

    sR

    ibos

    ome

    RN

    A b

    indi

    ngR

    NA

    deg

    rada

    tion

    RN

    A h

    elic

    ase

    RN

    A p

    roce

    ssin

    gR

    NA

    syn

    thes

    isrR

    NA

    tRN

    A p

    roce

    ssin

    gS

    igna

    lling

    Ste

    rols

    Tran

    slat

    ion

    Tran

    slat

    ion

    tRN

    ATr

    ansp

    orte

    rU

    biqu

    itin

    & p

    rote

    asom

    eV

    esic

    ular

    tran

    spor

    tZi

    nc fi

    nger

    BSPC

    Fig. 5. Half-lives of mRNAs in particular functional classes.Unique coding regions were placed into functional classes based on their genome annotation, user comments and publications. The classesused are indicated in Supplementary Table S5. In this box-and-whisker plot the bottom and top of the box are the first and third quartiles, theband inside the box is the median, the whiskers represent the lowest datum still within 1.5 of the lower quartile, and the highest datum stillwithin 1.5 of the upper quartile and the circles represent outliers.

    Transcriptome-wide mRNA decay in trypanosomes 315

    © 2014 The Authors. Molecular Microbiology published by John Wiley & Sons Ltd., Molecular Microbiology, 94, 307–326

  • 316 A. Fadda et al. ■

    © 2014 The Authors. Molecular Microbiology published by John Wiley & Sons Ltd., Molecular Microbiology, 94, 307–326

  • possible only if the mRNA was relatively stable: forexample, nearly all mRNAs present at more than 16copies/cell/gene in bloodstream forms or 32 copies percell in procyclics had half-lives longer than 30 min.

    We previously developed a model for gene expressionin bloodstream trypanosomes, based on detailed meas-urements for the PGK and tubulin genes (Haanstra et al.,2008). We expanded this model in order to account for ournew results (Fig. 7A, Table 3). In the published model, thefirst step is transcription; we had set the overall rate (ini-tiation plus elongation) so that it was sufficient to producethe measured steady state level of mRNA (step 1 inFig. 7A). The next step (step 2 in Fig. 7A) is 5′ splicing,with a fixed, half-time of 1 min. Splicing competes withprecursor degradation rate (step 4 in Fig. 7A); for this, weused a half-life of ∼ 8.7 min, as determined for the PGKprecursor (Haanstra et al., 2008). Mature mRNA degra-dation (step 6 in Fig. 7A) includes degradation in both thenucleus and the cytoplasm. The model also includes dilu-tion of RNA species via parasite growth (specific growthrate, μ). To start the new analysis, we used this publishedmodel to calculate the expected amounts of mRNA atsteady state for various half-lives (grey squares, Fig. 6A).We found that they were towards the upper edge of thecloud of data points. This indicates that the publishedmodel predicts almost the maximum amount of mRNA

    that can be obtained per gene in vivo. Notably, however,the levels of many mRNAs fell well below the estimatesfrom the published model. The first, and most obvious,explanation was that we had not allowed for the possibilitythat splicing rates might vary. We therefore went back toour measurements, and plotted the half-lives and abun-dances of mRNAs with relatively short measured 5′ pre-cursor half-lives (Fig. 6C and D): although the correlationwas marginally better than for the bulk population, thiscould be an artefact of the lower sample size. Returning tothe published model, we plotted the RNA abundances fora 5′ trans splicing reaction half-time of 10 min (greycircles, Fig. 6A). This still failed to account for the unex-pectedly low abundances of many mRNAs. Obviously,something was missing from the published model. OurmRNA decay rates (Fig. 7A, step 6) already included bothintra-nuclear and cytoplasmic degradation of the maturemRNA, so differences in nuclear export of mRNA couldnot explain the discrepancies between the model andreality. We identified transcription, polyadenylation andprecursor decay as possible variables.

    Modelling predicts an important role forco-transcriptional mRNA decay

    The first issue that we needed to address was the assump-tion that the transcription rate is constant within, andbetween, transcription units. It has previously been shownthat genes that encode mRNAs that persist after heatshock are located towards the end of transcription units(Kelly et al., 2012). This presumably enables synthesis ofthe mRNAs to continue for some time after stress-inducedshut-off of transcription initiation (Kelly et al., 2012). Thedistance of a gene from the transcription start site also borea relationship with cell cycle control (Kelly et al., 2012). Wetherefore considered the possibility that the position of thegene within the transcription unit was having an effect onthe mRNA abundance at steady state. To assess this, weplotted the half-life : abundance ratios for each geneacross the chromosomes. The plot for chromosome 10 is in

    Fig. 6. Relationship between half-life and mRNA abundance.A. Relationship between the number of mRNAs per cell per gene and the half-life for 6367 non-redundant open reading frames in bloodstreamforms. All mRNAs with apparent half-lives of less than 2 min were excluded since (given the 5 min Sinefungin treatment) this is outside therange of the assay. The mRNAs that appeared stable have been arbitrarily assigned half-lives of 240 min. The Pearson correlation coefficient(R) was calculated from log-transformed data. The grey squares indicate the results of modelling steady-state mRNA abundance in thepublished model (Haanstra et al., 2008) assuming a splicing half-time of 1 min; the circles are for an assumed 5′-trans splicing half-time of10 min. The gene copy numbers were estimated from random shotgun sequences of genomic DNA.B. As for (A), but for 5831 non-redundant open reading frames in procyclic forms. The mRNAs that appeared stable have been arbitrarilyassigned half-lives of 360 min; their median number of transcripts per gene per cell is 9.C. As for (A) but including only mRNAs with precursor half-times of less than 3 min.D. As for (B) but including only mRNAs with precursor half-times of less than 4 min.E. As for (A), but including only the 600 longest (red) and shortest (blue) mRNAs. The length ranges and correlation coefficients are alsoshown for the two groups. The grey squares indicate the results of modelling steady-state mRNA abundance in the published model (Haanstraet al., 2008) assuming a splicing half-time of 1 min. Lengths were taken from TritrypDB; all mRNAs lacking annotated poly(A) sites wereexcluded.F. As for (B), but including only the 600 longest (red) and shortest (blue) mRNAs.

    Table 3. The reactions in the model and their rate equations.

    Rate For reaction Rate equation or value

    v1 Transcription intoprecursor

    0.24 min−1

    v2 5′ trans-splicing ofunspliced precursor

    k2 × [unsplicedprecursor]

    v3 3′ poly-adenylation k3 × [5′splicedprecursor]v4 Degradation of unspliced

    precursor(k4 + μ) × [unsplicedprecursor]

    v5 Degradation of 5′ trans-spliced precursor

    (k5 + μ) × [5′splicedprecursor]

    v6 Degradation of maturemRNA

    (k6 + μ) × [mRNA]

    Transcriptome-wide mRNA decay in trypanosomes 317

    © 2014 The Authors. Molecular Microbiology published by John Wiley & Sons Ltd., Molecular Microbiology, 94, 307–326

  • Supplementary Fig. S9. Although more detailed analysismight reveal some trend for particular mRNA classes, wefound no gradient within transcription units, and no differ-ences between transcription units. We therefore did notadd variations in transcription rate to the model.

    Next, we added a step that was obviously missingfrom the published model: polyadenylation (step 3 inFig. 7A). This was assigned the same kinetics as fortrans splicing (Ullu et al., 1993). We assumed that it, liketrans splicing, was competing with precursor degradation

    Fig. 7. The kinetic model for trypanosome gene expression.A. Cartoon showing the main aspects of the model. *Steps 3 and 5 are extensions compared to (Haanstra et al., 2008).B. Trypanosome mRNA behaviour was modelled for three different groups: 550 mRNAs of 0.4–1, 4408 of 1–3 and 551 of 3–20. Thisdistribution approximates to that seen in our dataset. For each set, half-lives were distributed as in Fig. 4 up to 120′. The amount of mRNA atsteady state was then modelled using the following parameters: Transcription rate 0.24 molecules cell−1 min−1; Splicing and polyadenylationhalf-lives were sampled from a normal distribution around 2 ± 1.5′. The constant for degradation of the 5′ spliced precursor k5 (0.08 min−1) wasmultiplied by the ratio (mRNA length in nt)/(600 nt). Shown is the mature mRNA half-life range of 5–60 min, since this includes mosttranscripts.C. Measured abundances of mRNAs of 0.4–1 kb with half-lives of 5–60 min were plotted over the results from the model. The regression lineformulae and correlation coefficients are for log2-transformed data. Median mRNAs/cell/gene are also indicated.D. As (C) but for 1–3 kb mRNAsE. As (D) but for mRNAs longer than 3 kb.

    318 A. Fadda et al. ■

    © 2014 The Authors. Molecular Microbiology published by John Wiley & Sons Ltd., Molecular Microbiology, 94, 307–326

  • – in this case, degradation of a primary transcript thathas been 5′ trans spliced, but not yet polyadenylated(step 5 in Fig. 7A).

    Meanwhile, we discovered that the correlation betweenhalf-life and abundance was much better for shortermRNAs than for longer ones (Fig. 6E and F andFig. S10A), and that the shortest mRNAs fitted the pub-lished model relatively well (Fig. 6E). Longer mRNAswere on average less abundant than shorter ones with thesame half-life (Supplementary Fig. S10B). We concludedthat at least one of the steps in the model must include anegative influence of mRNA length.

    We now set up a simulation that incorporated variationsin processing, splicing and polyadenylation rates as wellas in mRNA decay, and that included a negative influenceof mRNA length. Splicing and polyadenylation were variedsomewhat conservatively such that the half-time for eachstep was 2 ± 1.5 min. To mimic length-dependent co-transcriptional destruction we multiplied kinetic constantk5 (degradation of the trans spliced precursor) by a factorof (mRNA length in nt)/600. (The number 600 was chosenempirically.) Our rationale for this was that a longer tran-script has more chance to be degraded before it is pro-cessed, since it takes longer for the polymerase to makea precursor that contains the downstream trans splicingsignal. In addition, slow (inefficient) polyadenylation ofan mRNA should also increase the likelihood of co-transcriptional destruction.

    We now used this extended model to simulate steady-state abundances for transcripts with varying lengths,half-lives, 5′ trans splicing and 3′ polyadenylation kinetics,using the known length distribution for the mRNAs and ahalf-life distribution based on our measurements. Theresults are shown in Fig. 7B. We then compared the simu-lated data for each length group with the real data(Fig. 7C–E). There was a good qualitative agreementbetween simulations and the data. In the simulations thegoodness-of-fit (R2) for the relation between mRNA half-life and abundance decreased with increasing transcriptlength, and in both the simulations and in the experimen-tal data, abundance was highest for the short transcriptsand lower for the longer ones. The real data, though, hadbigger variation (lower R or R2) than the simulations. Thismay hint at additional regulation, or a need to adjust theparameters further. For instance, compared with theextended model, in the real dataset the intermediatelength group (Fig. 7D) had more transcripts that werestable (half-life > 16′) and showed very low abundance.These might be mRNAs that are polyadenylated slowerthan the maximum half-time of 3.5 min that we allowed forin the model. (Unfortunately we could not model using thereal distribution of processing times because these couldnot be measured using the RNASeq approach.) Themodel also predicted a population of long, very unstable

    transcripts with lower abundance than was seen in thedata (less than one mRNA in every two parasites)(Fig. 7E). These would be very unstable, long transcriptsthat are processed slowly. It is possible that this particularcombination of properties (long and unstable) is rare.

    We also tested the effect of a theoretical length-dependent 3′ polyadenylation step – although it is difficultto imagine a mechanism for this (k3 decreasing withincreasing transcript length). This also resulted in thelinear relationship between mRNA half-life and abun-dance being poorer for longer transcripts but the effectwas not as pronounced as seen when we allowedfor co-transcriptional degradation (SupplementaryFig. S11A). Notably, if we also removed degradation of the5′ trans spliced precursor (step 5), we obtained a virtuallyperfect relation between half-life and abundance for alllengths (Supplementary Fig. S11B). This is a conse-quence of the steady-state assumption. Simulationsshowed that without degradation of the 5′ trans splicedprecursor, if transcription were induced from an ‘off’ state,slower 3′ polyadenylation lowered the mature mRNAabundance in the initial phase, but did not affect theabundance when the steady state was reached (notshown). If degradation of the un-spliced precursor wasremoved as well, a perfect correlation between abun-dance and half-life was obtained (SupplementaryFig. S11C). The modelling thus yielded the important con-clusion that length-dependent variations in processingcan only affect steady state mRNA abundance if process-ing is competing with degradation of the partly processedintermediates.

    Discussion

    At the start of this study it was already known that the rateof mRNA decay is an important determinant of trypano-some mRNA abundance, but detailed quantitative infor-mation, including time-courses for the degradation,was available only for a very small number of mRNAs(Table 2). We have now provided transcriptome-widedecay kinetics for both bloodstream and procyclic forms,and have developed a detailed model for gene expres-sion. We could confirm the important role of mRNA decayin determining steady state mRNA levels. We also foundthat a previously unrecognized step in processing has tobe present to explain the steady-state mRNA abundanceswe measured. This additional step is most likelyco-transcriptional degradation. Differences in mRNAmaturation rates can only affect mRNA abundance ifco-transcriptional degradation is also occurring.

    The kinetics of mRNA decay fell into at least threedifferent patterns. The ‘fast-slow’ group of mRNAsshowed initial fast degradation, followed by a slowerphase. This is also seen, in bloodstream forms, for

    Transcriptome-wide mRNA decay in trypanosomes 319

    © 2014 The Authors. Molecular Microbiology published by John Wiley & Sons Ltd., Molecular Microbiology, 94, 307–326

  • mRNAs bearing the 3′-UTR of the mRNA encoding themajor surface protein EP procyclin; in that case, the rapiddegradation phase occurs without deadenylation and isdependent on the 5′–3′ exoribonuclease XRNA (Li et al.,2006; Schwede et al., 2009). Those EP mRNAs that arenot attacked in this way are deadenylated more slowly,then degraded by XRNA and the exosome (Irmer andClayton, 2001; Haile et al., 2003; Li et al., 2006; Schwedeet al., 2009). Depletion of the exosome caused a slightdelay in the initiation of EP mRNA degradation (Haileet al., 2003), suggesting some cross-talk between the 5′and 3′ pathways. To see if this pattern was general weexamined previous RNASeq results for parasites defec-tive in different steps of mRNA decay. The fast-to-slowmethod of decay was indeed the predominant one in thecase of mRNAs whose half-lives were increased afterXRNA depletion (Manful et al., 2011) (Fisher’s testP-value

  • trans splicing, but may in fact be slightly slower, since it isthought to occur after splicing has been completed (Ulluet al., 1993). Second, in considering the effects of tran-script length on degradation of the spliced precursor, wehave opted for a simple length-dependence, which largelyleaves transcripts < 600 nt unaffected by intermediate deg-radation. We could have changed the boundary chosen, ormade this dependence more complex. Third, the modelwas evaluated at steady state, but we cannot rule out thatpre-steady-state kinetics influenced the experimentalresults. If there are variations in the overall transcriptioninitiation rate during the cell cycle, some slowly processedmRNAs may never reach steady state.

    The detailed mechanism of trypanosome co-transcriptional degradation is currently unknown. Co-transcriptional degradation certainly occurs in othersystems, although the details may not apply in trypano-somes. In Opisthokonts, the nuclear quality controlmachinery is triggered by failure of mRNA capping, and byfailure to splice or polyadenylate a terminated mRNA(Schmid and Jensen, 2010; 2013; Kilchert and Vasiljeva,2013). 5′->3′ degradation by Xrn2/Rat1 requires a free 5′end, and has mainly been studied in relation to transcrip-tion termination beyond polyadenylation sites (Schmid andJensen, 2010; 2013; Kilchert and Vasiljeva, 2013). Withtrypanosome polycistronic transcription, most mRNAshave no free 5′ end, since the mature 5′-end is created bytrans-esterification, and termination is relevant only for atiny proportion of genes. Polyadenylation might, however,fail after downstream trans splicing has been completed,and the resulting free 3′-ends should be attacked by thenuclear exosome (Cristodero and Clayton, 2007;Etheridge et al., 2009). There is, though, no reason to thinkthat this would be more efficient for long mRNAs than forshort ones and in fact, depletion of the exosome actuallycauses preferential loss of longer mRNAs (Fadda et al.,2013; Jha et al., 2014). The NOT deadenylation complex isassociated with RNA polymerase II in mammals, but it isthought to promote (not inhibit) elongation (Reese, 2013).Previous experiments that aimed at characterization ofRNA polymerase II subunits and transcription factors (e.g.Lee et al., 2009) were not designed for detection of pro-teins present in sub-stoichiometric amounts, but aftertandem affinity purification of the NOT complex, wedetected two peptides from a putative polymerase II elon-gation factor (Tb927.2.3580) (Färber et al., 2013).

    Our data are best explained by the existence of anendonuclease that intermittently cleaves the associatedmRNA precursor. This is highly speculative since there is,so far as we know, no real precedent for such a degrada-tion process (Schmid and Jensen, 2010; 2013; Kilchertand Vasiljeva, 2013). To recognize the precursor specifi-cally, we would expect the endonuclease to be associatedeither with chromatin, or with the elongating RNA poly-

    merase II complex. One consequence of such a mecha-nism would be that a decrease in the polymeraseelongation rate would increase the time that the precursoris accessible for digestion. Indeed, after depletion oftrypanosome CTR9, a probable transcription elongationfactor (Ouna et al., 2012), the abundances of longmRNAs were drastically decreased (Fadda et al., 2013;Jha et al., 2014). In animal cells and yeast, elongatingRNA polymerase II pauses, and even backtracks(Darzacq et al., 2007; Churchman and Weissman, 2011).We speculate that such pauses might, in trypanosomes,be a trigger for cleavage without detachment of the poly-merase, and could even be influenced by the DNAsequence. This would provide an additional means toregulate the formation of mRNA in the absence of anycontrol of transcription initiation.

    The simplicity of the trypanosome system has made itpossible to construct a mathematical model of mRNAcontrol which has revealed novel aspects and refined ourview of the likely contribution of mRNA processing anddegradation to mRNA abundance. When combined withribosome profiling to determine translation efficiency, ourmeasurements will enable us to mathematically modeltrypanosome gene expression from DNA to protein.

    Experimental procedures

    Sample preparation

    Bloodstream and procyclic form trypanosomes of the Lister427 strain, expressing the tet repressor, were used; the blood-stream forms were the 2T1 line described in (Alsford et al.,2005), while procyclic cells expressed the tet repressor fromplasmid pHD2060. Bloodstream forms were cultured in HMI-9medium containing 20% FCS, to a density of 1.0–1.2 × 106

    cells ml−1; procyclic forms were cultured in MEM-Pros to adensity of 2 × 106 (experiment 1) or 1.3 × 106 (experiment 2)cells ml−1. For the pre-mRNA decay assay, transcription wasinhibited briefly by incubating the cells with Actinomycin D(10 μg ml−1) and samples were harvested immediately(equivalent to a 5 min incubation); the bloodstream formexperiment was done twice; for procyclics a single experimentwas done. For the mRNA decay assay, cap methylation wasinhibited by adding 1 μg ml−1 Sinefungin for 5 min to preventsplicing (t0 samples), then Actinomycin D was added for up to120 min for bloodstream forms and 240 min for procyclics. Allsamples are listed in Supplementary Table S1. The parasiteswere harvested by centrifugation for 5 min (all times giveninclude the centrifugation time); RNA was extracted usingpeqGOLD-Trifast (peqLab) and depleted for rRNA usingRibominus Eukaryotic kit for RNA-seq (Invitrogen, no.10837-02 or -08). The RNA was then sequenced (Librarypreparation at the BioQuant facility, Heidelberg and HiSeq atEMBL). The first set of bloodstream-form decay samples wasmeasured using Illumina GAIIX and all remaining sampleswith Illumina HiSeq (Supplementary Table S1). All data areavailable at the NCBI Sequence read archive, with accessionnumber SRP042959, and a simplified version of the results

    Transcriptome-wide mRNA decay in trypanosomes 321

    © 2014 The Authors. Molecular Microbiology published by John Wiley & Sons Ltd., Molecular Microbiology, 94, 307–326

  • has been sent to TritrypDB. We cannot rule out the possibilitythat the Sinefungin orActinomycin D treatments influenced themRNA decay rates, especially after long incubation times, butat present alternative methods are not available.

    RNA decay analysis by Northern blotting

    To determine the overall mRNA decay rate, aliquots of thesequenced samples were separated by agarose gel electro-phoresis, blotted and probed with a spliced leader probe,using the rRNA signal for normalization, as previouslydescribed (Fadda et al., 2013). The measured signal wasused to fit an exponential decay curve with R (SupplementaryFig. S4) and the values corresponding to the curve were usedto normalize the RNAseq decay data.

    Data processing

    Reads were aligned to the T. brucei 927 reference genome,and read counts were extracted using Bowtie, SamTools andcustom-made PERL scripts as previously described (Faddaet al., 2013). Read counts per ORF were first normalized forthe total number of counts (RPM, reads per million reads). Inorder to build the libraries, the same amount of input RNAwas used for each sample; and similar numbers of readcounts were obtained for each library. After addition of tran-scription inhibitors, the proportions of the unstable mRNAs godown (giving a decrease in RPM) while the proportions of thestable RNAs go up (giving a decrease in RPM). The stablemRNAs seem to increase because the overall reduction inmRNA levels is not reflected in the RNASeq data. This neces-sitates the use of an external control to detect – and correctfor – the decline in mRNA quantities in the cell with time. Wecould not normalize using endogenous stable RNAs,because they were excluded during the ribo-minus and sizeselection steps of library preparation. To measure globalmRNA decay we therefore used Northern blots to assay thedecline of the spliced leader (SL) signal, as previouslydescribed (Fadda et al., 2013). Then, to account for thedecreasing total amounts of RNA during the decay assay, theRPM for each time point were multiplied by the correspondingSL signal at that time. ORFs with an average of less than 50reads at time 0 were excluded from the half-life analysis.Lastly, the read counts were calculated as a ratio to time point0 (t0). This format was used as input for a C++ code tocalculate half-lives. Using the equations described byDeneke et al., (2013) we allowed for three alternative decaypatterns: simple exponential, initially fast, then slower (fast-slow, equation 9 of Deneke et al., 2013) and initially slow,then fast (slow-fast, equation 10 of Deneke et al., 2013). Forthe calculations each measurement was treated as an inde-pendent data point. Differential expression was analysed byDESeq (Anders and Huber, 2010).

    To measure splicing kinetics, read counts were extractedfor the 20mer upstream of all known splice acceptor sites foreach gene ( Kolev et al., 2010; Nilsson et al., 2010; Siegelet al., 2010; Zhang et al., 2010 and our data), as well as foreach coding region. If the read counts for the 20mer wereequal to, or more than, the length-normalized read counts incoding region, we assumed that the splice site was not being

    used in our sample, and the data for the site were removed;we also removed sites if no reads existed in the 20mer regionin the control (untreated) sample. Finally, zero values in the20mer region in Actinomycin D samples were converted to 1read. Read counts per 20mer were normalized to the readcounts per coding region of the corresponding gene in eachsample. The ratio of the normalized 20mer counts in thetreated/untreated samples was used to calculate a pre-mRNA half-life based on an exponential model. While inter-preting the results, we attribute this to splicing, althoughstrictly speaking pre-mRNA degradation also contributes.Probability density plots were done using a kernel densityfunction contained in the R package ‘ggplot’.

    To calculate gene copy numbers, we first did a control foruniformity of coverage. We mapped the reads from the refer-ence genome against itself using reads available from Dr A.Ivens (University of Edinburgh). Allowing each read to maponly once should yield a bell-shaped coverage plot. Next weallowed the reads to map up to 300 times and calculatedRPKM for each gene. We assumed that the modal RPKMvalue (M) corresponded to a gene copy number of 1, sincemost genes should be unique. Next, we aligned the reads fromsequenced, fragmented DNA from our lab strain, and mappedthem to the reference genome, allowing each read to align onlyonce. We again calculated the modal value (M′). If the numberof reads from a gene was M′, then that gene has the same copynumber as in the reference; if M′ has a value n × M, then thestrain has n times the number of copies in the reference.

    Calibrating decay models

    The code for decay calibration is written in C++ and is pro-vided as Supplementary file S1. We used the equationsdescribed by Deneke et al., (2013) to model yeast mRNAdecay; that paper explains the mathematical details.Complex, non-exponential mRNA degradation is modelled bya Markov chain with 5 states; each state represents a differ-ent biochemical process of mRNA degradation, where themolecule is likely to be degraded at any state. The rate oftransition between states is assumed to be the same for allstates and is denoted as λ. The degradation rate from anystate is denoted as μ. In one variant of the complex model theinitial rate of degradation μ1 is assumed bigger than all sub-sequent rates, denoted ν. This is described by the parametricdensity Equation 9 (Deneke et al., 2013) and referred to hereas ‘fast-slow’ model. The other model variant assumes alldegradation rates except the last to be equal, and the lastrate μn ≡ ν > 0. This is described by the parametric densityEquation 10 (Deneke et al., 2013) and here referred to as‘slow-fast’ model. We calibrated and compared three models:the exponential model, ‘fast-slow’ and ‘slow-fast’. The calibra-tion procedure worked in several stages. In the first stage wecalibrated each model. The fitting procedure using non-linearleast squares which we used (Dennis et al., 1981) isextremely fast but suffers numerically from multiple localminima. These local minima are obtained from different ‘initialparameter guesses’. For the three-parameter models wechose to calibrate the model using 125 different startingpoints, lying on a mesh between 0 and 3.0 with 5 points peraxis. For each calibration we stored the residual model errorand the calibrated parameters. In the next stage we ran a

    322 A. Fadda et al. ■

    © 2014 The Authors. Molecular Microbiology published by John Wiley & Sons Ltd., Molecular Microbiology, 94, 307–326

  • rejection/acceptance procedure for each of the parametricforms, and for each of a form’s 125 models. The rejectioncriteria were subjective, but were designed to reject modelsthat were completely implausible, or had collapsed to simplermodels. The rejection criteria for each model are found in thefunction includeTentativeSolution(). As an example, for a cali-brated model of Equation 9’s density we have parameters λ,μ, ν (Deneke et al., 2013). The model is rejected if λ or ν are0, or if any of the parameters have become infinite. The nextstage of calibration chooses the best fitting accepted modelfor each density type. We call the three best-fit models thethree ‘candidate’ models. The last stage of calibration selectsthe best-fit model from among the candidate models. The nullhypothesis is that the distribution is Exponential E. For acandidate model C ≠ E we look at the relative

    D C ER C R E

    R C R E,

    .( ) >

    ( ) − ( )⋅ ( ) + ( )0 5

    where R(M) is the residual sum of squares error from fittingwith model M. If D(C;E) > 5% we rejected the null hypothesisand accepted C. The choice of threshold is essentially arbi-trary. Deneke et al. (2013) used a boundary of 10%, we chose5% after visual examination of individual examples. Using thelower threshold leads to a lower proportion of curves beingdesignated as exponential. We did not favour either the Equa-tion 10 or Equation 9 density over each other and simply tookthe candidate from the two with the minimal residual. The ‘halflives’ for the mRNAs with non-exponential decay in all theTables are the time at which half of the mRNA is left (the valueof time T where 50% of the distribution lies above T, and 50%lies below T).

    Modelling mRNA processing and decay

    The PGK mRNA model developed in Haanstra et al. (2008)was initially used. Then we made a new, modified versionwhich applies to any transcript. The new model was extendedto include polyadenylation and length-dependent degradationof the trans spliced intermediate. Model simulations weredone, and extensions were made, in the open source soft-ware COPASI v.4.9.45 (Hoops et al., 2006). Figure 7A showsa graphical representation of the reactions in the model, andtheir rate equations are listed in Table 1.

    The constants k in the equations relate to half-lives (HL) ofthe corresponding species via

    k =( ) −ln min2 1

    HL

    For k4, k5 and k6 this relation has been explicitly included inthe model, which enables us to affect the k’s by changing thehalf-life through random sampling (see below).

    k2 and k3 each have a default value of 0.41 (half-life of 1.7′)(as in Haanstra et al., 2008). We assumed k4 and k5 to beequal to each other and took a value of 0.08 min−1 which isbased on the previously measured half-life of the PGK pre-cursor after inhibition of both transcription and splicing byActinomycin D and Sinefungin (Haanstra et al., 2008). μ isthe specific growth rate and is 0.0019 min−1 (a populationdoubling time of 6 h).

    The differential equations that describe the changes inspecies concentrations over time in the model are:

    ddt

    v kk

    unsplicedprecursorunsplicedprecursor

    [ ]= − × [ ]

    − +( ) ×1 2

    4 μ uunsplicedprecursor[ ]

    ddt

    k k[ ]′ = × [ ] −

    × ′

    5

    52 3

    splicedprecursorunsplicedprecursorspliceedprecursorsplicedprecursor

    [ ] − +( )× ′[ ]

    k55

    μ

    ddt

    k kmRNA

    splicedprecursor mRNA[ ]

    = × ′[ ] − +( ) × [ ]3 65 μ

    Concentrations are in molecules cell−1.To simulate effects that depend on the length of the mature

    mRNA, the relevant constant was multiplied or divided by(length in nt)/(600 nt): For length-dependent degradation ofthe 5′ spliced intermediate, k5 was multiplied by length/600 sothat longer transcripts would be affected faster. For length-dependent polyadenylation, k3 was multiplied by 600/lengthso that polyadenylation would be slowed down for longertranscripts.

    The model was used to simulate steady state abundancesfor individual transcripts (i.e. with their own values for thekinetic constants). To this end, 5510 transcripts were simu-lated by

    • Sampling the half-life of the un-spliced precursor in the 5′trans splicing reaction from a normal distribution with amean of 2′ and a standard deviation of 1.5′ (affects k2),

    • Sampling the half-life of the 5′ spliced precursor in the 3′poly-adenylation reaction from a normal distribution witha mean of 2′ and a standard deviation of 1.5′ (affects k3;only used when k3 was not multiplied by a length-dependent factor), and

    • Sampling the length of a transcript from a uniform dis-tribution for short (551 transcripts in the range 400–1000 nt), intermediate (4408 transcripts in the range1000–3000 nt) and long (551 transcripts in the range3000–20000 nt) transcripts

    For each length range, the half-life of the mature mRNA(affecting k6) was sampled from a uniform distribution withinthe half-life range in each bin in Fig. 3, except that for thefirst bin, the half-lives ranged between 1–5′. The number oftranscripts in the length range for each bin was taken pro-portionally to the distribution of lengths of the overalldataset (10% short, 80% intermediate and 10% longtranscripts).

    In some cases, the combination of parameters did notresult in a steady state and the final number of solutions wastherefore lower than the number of simulations.

    Acknowledgements

    We thank Janine Philipp for making the SL-hybridizedNorthern blots for bloodstream forms, and Ute Leibfried fortechnical assistance. We also thank Dr Carlus Deneke (MaxPlanck Institute of Colloids and Interfaces, Potsdam) foruseful discussions, and Alasdair Ivens (University of Edin-burgh) for the TREU927 shotgun reads.

    Transcriptome-wide mRNA decay in trypanosomes 323

    © 2014 The Authors. Molecular Microbiology published by John Wiley & Sons Ltd., Molecular Microbiology, 94, 307–326

  • Funding

    This work was supported by the collaborative Sysmo grant‘The silicon trypanosome’, with funds from the Bundesminis-terium für Bildung und Forschung, BBSRC (BB/I004602/1),and the Netherlands Organization for Scientific Research(NWO). Work in Edinburgh was also funded by the WellcomeTrust (088293; 095831). DD was supported by the DeutscheForschungsgemeinshaft, Grant no CL112/17-1, and BMB isfunded by a Rosalind Franklin Fellowship from the Universityof Groningen.

    Author contributions

    Abeer Fadda did most of the data analysis. Mark Ryten wrotethe programme to fit the different decay models to the data andderive half-lives. Dorothea Droll, Federico Rojas and ValentinFärber were responsible for parasites, RNA samples andNorthern blots. Clementine Merce did the gene copy numbercalculations. Keith Matthews supervised the bloodstream-form work. Jurgen Haanstra was responsible for the model-ling, with supervision from Barbara Bakker. Christine Claytonwas responsible for overall supervision and coordinatingwriting of the paper.

    Conflict of interest

    The authors declare no conflict of interest.

    References

    Ali, J., Tagoe, D., Munday, J., Donachie, A., Morrison, L., andde Koning, H. (2013) Pyrimidine biosynthesis is not anessential function for Trypanosoma brucei bloodstreamforms. PLoS ONE 8: e58034.

    Ali, J.A., Creek, D.J., Burgess, K., Allison, H.C., Field, M.C.,Maser, P., and De Koning, H.P. (2012) Pyrimidine salvagein Trypanosoma brucei bloodstream forms and the trypano-cidal action of halogenated pyrimidines. Mol Pharmacol 83:439–453.

    Alsford, S., Kawahara, T., Glover, L., and Horn, D. (2005)Tagging a T. brucei RRNA locus improves stable transfec-tion efficiency and circumvents inducible expression posi-tion effects. Mol Biochem Parasitol 144: 142–148.

    Anders, S., and Huber, W. (2010) Differential expressionanalysis for sequence count data. Genome Biol 11: R106.

    Avliyakulov, N., Hines, J., and Ray, D. (2003) Sequenceelements in both the intergenic space and the 3′ untrans-lated region of the Crithidia fasciculata KAP3 gene arerequired for cell cycle regulation of KAP3 mRNA. EukaryotCell 2: 671–677.

    Bakker, B.M., Krauth-Siegel, R.L., Clayton, C., Matthews, K.,Girolami, M., Westerhoff, H.V., et al. (2010) The silicontrypanosome. Parasitology 137: 1333–1341.

    Bringaud, F., Riviere, L., and Coustou, V. (2006) Energymetabolism of trypanosomatids: adaptation to availablecarbon sources. Mol Biochem Parasitol 149: 1–9.

    Brooks, D.R., Denise, H., Westrop, G.D., and Coombs, G.H.(2001) The stage-regulated expression of Leishmaniamexicana CPB cysteine proteases is mediated by an

    intercistronic sequence element. J Biol Chem 276:47061–47069.

    Brown, B.S.V., Chi, T.B., and Williams, N. (2001) The Trypa-nosoma brucei mitochondrial ATP synthase is develop-mentally regulated at the level of transcript stability. MolBiochem Parasitol 115: 177–187.

    Chaudhuri, M., Sharan, R., and Hill, G. (2002) Trypanosomealternative oxidase is regulated post-transcriptionally at thelevel of RNA stability. J Eukaryot Microbiol 49: 263–269.

    Chekanova, J.A., Gregory, B.D., Reverdatto, S.V., Chen, H.,Kumar, R., Hooker, T., et al. (2007) Genome-wide high-resolution mapping of exosome substrates reveals hiddenfeatures in the Arabidopsis transcriptome. Cell 131: 1340–1353.

    Churchman, L.S., and Weissman, J.S. (2011) Nascent tran-script sequencing visualizes transcription at nucleotideresolution. Nature 469: 368–373.

    Clayton, C. (2013) The regulation of trypanosome geneexpression by RNA-binding proteins. PLoS Pathog 9:e1003680.

    Clayton, C., and Estevez, A. (2010) The exosomes of trypa-nosomes and other protists. In RNA Exosome. Jensen, T.(ed.). Austin, TX: Landes Bioscience, pp. 39–49.

    Colasante, C., Robles, A., Li, C.-H., Schwede, A., Benz, C.,Voncken, F., et al. (2007) Regulated expression of glyco-somal phosphoglycerate kinase in Trypanosoma brucei.Mol Biochem Parasitol 151: 193–204.

    Cristodero, M., and Clayton, C. (2007) Trypanosome MTR4in involved in ribosomal RNA processing. Nucleic AcidsRes 35: 7023–7030.

    Daniels, J., Gull, K., and Wickstead, B. (2010) Cell biology ofthe trypanosome genome. Microbiol Mol Biol Rev 74: 552–569.

    Darzacq, X., Shav-Tal, Y., de Turris, V., Brody, Y., Shenoy, S.,Phair, R., and Singer, R. (2007) In vivo dynamics of RNApolymerase II transcription. Nat Struct Mol Biol 14: 796–806.

    Deneke, C., Lipowsky, R., and Valleriani, A. (2013) Complexdegradation processes lead to non-exponential decay pat-terns and age-dependent decay rates of messenger RNA.PLoS ONE 8: e55442.

    Dennis, J., Gay, D., and Roy Welsch, R. (1981) Algorithm573: an adaptive nonlinear least-squares algorithm. ACMTrans Math Softw 7: 367–383.

    Erben, E., Chakraborty, C., and Clayton, C. (2014) TheCAF1-NOT complex of trypanosomes. Front Genet 4: 299.

    Estévez, A. (2008) The RNA-binding protein TbDRBD3 regu-lates the stability of a specific subset of mRNAs in trypa-nosomes. Nucleic Acids Res 36: 4573–4586.

    Etheridge, R.D., Clemens, D.M., Gershon, P.D., andAphasizhev, R. (2009) Identification and characterizationof nuclear non-canonical poly(A) polymerases from Trypa-nosoma brucei. Mol Biochem Parasitol 164: 66–73.

    Fadda, A., Färber, V., Droll, D., and Clayton, C. (2013)The roles of 3′-exoribonucleases and the exosome intrypanosome mRNA degradation. RNA 19: 937–947.

    Färber, V., Erben, E., Sharma, S., Stoecklin, G., and Clayton,C. (2013) Trypanosome CNOT10 is essential for the integ-rity of the NOT deadenylase complex and for degradationof many mRNAs. Nucleic Acids Res 41: 1211–1222.

    324 A. Fadda et al. ■

    © 2014 The Authors. Molecular Microbiology published by John Wiley & Sons Ltd., Molecular Microbiology, 94, 307–326

  • Fernández-Moya, S., and Estévez, A. (2010) Posttranscrip-tional control and the role of RNA-binding proteins in generegulation in trypanosomatid protozoan parasites. WileyInterdiscip Rev RNA 1: 34–46.

    Fluck, C., Salomone, J., Kurath, U., and Roditi, I. (2003)Cycloheximide-mediated accumulation of transcripts froma procyclin expression site depends on the intergenicregion. Mol Biochem Parasitol 127: 93–97.

    Grigull, J., Mnaimneh, S., Pootoolal, J., Robinson, M., andHughes, T. (2004) Genome-wide analysis of mRNA stabil-ity using transcription inhibitors and microarrays revealsposttranscriptional control of ribosome biogenesis factors.Mol Cell Biol 24: 5534–5547.

    Gudipati, R.K., Xu, Z., Lebreton, A., Seraphin, B., Steinmetz,L.M., Jacquier, A., and Libri, D. (2012) Extensive degrada-tion of RNA precursors by the exosome in wild-type cells.Mol Cell 48: 409–421.

    Gunasekera, K., Wuthrich, D., Braga-Lagache, S., Heller, M.,and Ochsenreiter, T. (2012) Proteome remodelling duringdevelopment from blood to insect-form Trypanosomabrucei quantified by SILAC and mass spectrometry. BMCGenomics 13: 556.

    Haanstra, J., Stewart, M., Luu, V.-D., van Tuijl, A.,Westerhoff, H., Clayton, C., and Bakker, B. (2008) Controland regulation of gene expression: quantitative analysis ofthe expression of phosphoglycerate kinase in bloodstreamform Trypanosoma brucei. J Biol Chem 283: 2495–2507.

    Haile, S., Estévez, A.M., and Clayton, C. (2003) A role for theexosome in the initiation of degradation of unstablemRNAs. RNA 9: 1491–1501.

    Haimovich, G., Medina, D.A., Causse, S.Z., Garber, M.,Millan-Zambrano, G., Barkai, O., et al. (2013) Geneexpression is circular: factors for mRNA degradation alsofoster mRNA synthesis. Cell 153: 1000–1011.

    Harigaya, Y., and Parker, R. (2012) Global analysis of mRNAdecay intermediates in Saccharomyces cerevisiae. ProcNatl Acad Sci USA 109: 11764–11769.

    Hoops, S., Sahle, S., Gauges, R., Lee, C., Pahle, J., Simus,N., et al. (2006) COPASI – a COmplex PAthway SImulator.Bioinformatics 22: 3067–3074.

    Hotz, H.-R., Hartmann, C., Huober, K., Hug, M., and Clayton,C.E. (1997) Mechanisms of developmental regulation inTrypanosoma brucei: a polypyrimidine tract in the3′-untranslated region of a trypanosome surface proteinmRNA affects RNA abundance and translation. NucleicAcids Res 25: 3017–3025.

    Houalla, R., Devaux, F., Fatica, A., Kufel, J., Barrass, D.,Torchet, C., and Tollervey, D. (2006) Microarray detectionof novel nuclear RNA substrates for the exosome. Yeast23: 439–454.

    Huang, J., and van der Ploeg, L. (1991) Maturation of polycis-tronic pre-mRNA in Trypanosoma brucei: analysis of transsplicing and poly(A) addition at nascent RNA transcriptsfrom the hsp70 locus. Mol Cell Biol 11: 3180–3190.

    Irmer, H., and Clayton, C.E. (2001) Degradation of the EP1mRNA in Trypanosoma brucei is initiated by destruction ofthe 3′-untranslated region. Nucleic Acids Res 29: 4707–4715.

    Jha, B., Fadda, A., Merce, C., Mugo, E., Droll, D., andClayton, C. (2014) Depletion of the trypanosome pumiliodomain protein PUF2 causes transcriptome changes

    related to coding region length. Eukaryot Cell 13:664–674.

    Kapotas, N., and Bellofatto, V. (1993) Differential response toRNA trans-splicing signals within the phosphoglyceratekinase gene cluster in Trypanosoma brucei. Nucleic AcidsRes 21: 4067–4072.

    Kelly, S., Kramer, S., Schwede, A., Maini, P., Gull, K., andCarrington, M. (2012) Genome organization is a majorcomponent of gene expression control in response tostress and during the cell division cycle in trypanosomes.Open Biol 2: 120033.

    Kilchert, C., and Vasiljeva, L. (2013) mRNA quality controlgoes transcriptional. Biochem Soc Trans 41: 1666–1672.

    Kiss, D.L., and Andrulis, E.D. (2010) Genome-wide analysisreveals distinct substrate specificities of Rrp6, D