#PAGE_PARAMS# #ADS_HEAD_SCRIPTS# #MICRODATA#

RNA Polymerase II CTD phosphatase Rtr1 fine-tunes transcription termination


Authors: Jose F. Victorino aff001;  Melanie J. Fox aff001;  Whitney R. Smith-Kinnaman aff001;  Sarah A. Peck Justice aff001;  Katlyn H. Burriss aff001;  Asha K. Boyd aff001;  Megan A. Zimmerly aff001;  Rachel R. Chan aff001;  Gerald O. Hunter aff001;  Yunlong Liu aff002;  Amber L. Mosley aff001
Authors place of work: Department of Biochemistry and Molecular Biology, Indiana University School of Medicine, Indianapolis, Indiana, United States of America aff001;  Department of Medical and Molecular Genetics, Indiana University School of Medicine, Indianapolis, Indiana, United States of America aff002;  Center for Computational Biology and Bioinformatics, Indiana University School of Medicine, Indianapolis, Indiana, United States of America aff003
Published in the journal: RNA Polymerase II CTD phosphatase Rtr1 fine-tunes transcription termination. PLoS Genet 16(3): e32767. doi:10.1371/journal.pgen.1008317
Category: Research Article
doi: https://doi.org/10.1371/journal.pgen.1008317

Summary

RNA Polymerase II (RNAPII) transcription termination is regulated by the phosphorylation status of the C-terminal domain (CTD). The phosphatase Rtr1 has been shown to regulate serine 5 phosphorylation on the CTD; however, its role in the regulation of RNAPII termination has not been explored. As a consequence of RTR1 deletion, interactions within the termination machinery and between the termination machinery and RNAPII were altered as quantified by Disruption-Compensation (DisCo) network analysis. Of note, interactions between RNAPII and the cleavage factor IA (CF1A) subunit Pcf11 were reduced in rtr1Δ, whereas interactions with the CTD and RNA-binding termination factor Nrd1 were increased. Globally, rtr1Δ leads to decreases in numerous noncoding RNAs that are linked to the Nrd1, Nab3 and Sen1 (NNS) -dependent RNAPII termination pathway. Genome-wide analysis of RNAPII and Nrd1 occupancy suggests that loss of RTR1 leads to increased termination at noncoding genes. Additionally, premature RNAPII termination increases globally at protein-coding genes with a decrease in RNAPII occupancy occurring just after the peak of Nrd1 recruitment during early elongation. The effects of rtr1Δ on RNA expression levels were lost following deletion of the exosome subunit Rrp6, which works with the NNS complex to rapidly degrade a number of noncoding RNAs following termination. Overall, these data suggest that Rtr1 restricts the NNS-dependent termination pathway in WT cells to prevent premature termination of mRNAs and ncRNAs. Rtr1 facilitates low-level elongation of noncoding transcripts that impact RNAPII interference thereby shaping the transcriptome.

Keywords:

Yeast – Phosphatases – Messenger RNA – RNA sequencing – Non-coding RNA sequences – Transcriptional control – Transcriptional termination – ChIP exo

Introduction

The termination of transcription by eukaryotic RNA Polymerase II (RNAPII) is tightly coupled with RNA processing, including small RNA processing, splicing, and mRNA cleavage and polyadenylation at the 3’-end of protein-coding genes [1, 2]. Recent studies have reaffirmed that transcription termination in eukaryotes is a highly dynamic process that can lead to different gene expression outputs through mechanisms, such as alternative polyadenylation site usage and premature transcription termination [310]. Transcription termination in yeast has been shown to be regulated through numerous termination factors as well as the phosphorylation status of the C-terminal domain (CTD) of RNAPII, which has the repetitive sequence (Tyr1-Ser2-Pro3-Thr4-Ser5-Pro6-Ser7)n [11, 12]. However, the exact mechanisms that underlie the role of CTD dephosphorylation in the regulation of elongation, termination, and the attenuation of these processes remain unclear. At least four phosphatases are components of the yeast transcription elongation/termination machinery: Rtr1, Ssu72, Glc7, and Fcp1 [1322]. There appears to be extensive interplay between the protein phosphatases and their control of the phosphorylation status of the RNAPII CTD. For instance, serine 5 (Ser5) dephosphorylation has been shown to be carried out by both Rtr1 and Ssu72 in both in vivo and in vitro studies [13, 2326]. Additionally, Ssu72 dephosphorylation of Ser5 serves as a prerequisite for Ser2 dephosphorylation by Fcp1 [14, 16]. However, it remains unclear how temporal dephosphorylation impacts the formation and/or recruitment of RNA processing complexes during transcription and the determination of the termination pathway that will be used.

One pathway that is heavily influenced by CTD phosphorylation is the Nrd1, Nab3, and Sen1 (NNS) polyadenylation independent transcription termination pathway [2732]. Nrd1 and Nab3 form a stable heterodimer, which cooperatively bind RNA with affinities in the low nM range [33]. It has also been shown that Nrd1-Nab3 can bind to RNA as multiple heterodimers in a cooperative manner, which could be mediated by low-complexity domains in Nab3 [3438]. Nrd1-Nab3 have been shown to have some preference for the consensus sequences UGUA and UCUU for Nrd1 and Nab3 respectively, as determined by UV-crosslinking based studies [31, 39]. However, a number of different RNA sequences are tolerated for binding by Nrd1 or Nab3 [39]. For instance, Nrd1 binding to its consensus motif has been described as semi-specific with a number of G-rich or AU-rich RNAs binding to Nrd1 with affinities in the low micromolar range [40]. In addition to RNA binding, Nrd1 contains domains important for protein-protein interactions, such as an RNAPII CTD interaction domain (CID) that preferentially interacts with a Ser5-P CTD [27]. In addition, a number of other proteins, such as the cap binding complex, the nuclear Rrp6-containing exosome and its partner protein Mpp6, and the polyA polymerase Trf4, have been shown to bind to Nrd1 [30, 41, 42]. Sen1 association with Nrd1 has been shown to be mediated through as many as three high-affinity interaction domains referred to as Nrd1 interaction motifs (NIMs) [43]. NIM association with the Nrd1 CID has been shown to occur with a higher binding affinity than observed for two canonical CTD Ser5-P repeats by fluorescence anisotropy [43]. Similarly, the interaction between the Trf4 NIM and the Nrd1 CID has also been shown to be 100-fold stronger than the interaction between the Nrd1 CID and two canonical CTD Ser5-P repeats [41]. However, these biophysical experiments may not fully capture potential in vivo dynamics which could include higher order interactions between the full length RNAPII CTD and multiple Nrd1-Nab3 heterodimers, which have been shown to facilitate cooperative binding to RNA [34].

The NNS complex regulates transcription termination of short non-coding transcripts and transcription elongation of a selection of protein coding genes [29, 4447]. Pcf11, a member of the cleavage factor Ia complex (CFIa), also contains a CID that has increased affinity for Ser2-P over Ser5-P modified RNAPII CTD [27]. Pcf11 has been shown to be required for both polyadenylation dependent termination and NNS termination [27, 48, 49]. While the cleavage and polyadenylation factor (CPF) complex does not contain any known CID containing proteins, both Ssu72 and Glc7 are integral subunits of CPF. Rtt103, a CID-containing protein with specificity to Ser2-P CTD, has been proposed to form a higher order complex with CFIa and CPF to possibly bridge the Rat1 exoribonuclease to the RNAPII CTD to trigger degradation of the cleaved 3’ end of the RNA transcription product and hence transcription termination [27, 50]. Rat1, Rtt103, and the decapping nuclease Rai1 are sufficient to terminate elongating RNAPII in vitro and have been shown to be required for RNAPII termination in vivo [5154]. However, numerous subunits of CFIa and CPF are required for fully efficient RNAPII termination in vivo suggesting that higher-order interactions between the transcription termination machinery, RNAPII, and the RNA are likely required in eukaryotes [55, 56]. Additional factors such as Rrp6, a subunit of the nuclear exosome, may also play a role in transcription termination through targeting of certain molecular states of RNAPII, such as the backtracked enzyme (previously described as the reverse torpedo model) [55, 5759]. We propose that the extensive control of RNAPII CTD dephosphorylation in eukaryotes serves as a critical regulator of co-transcriptional RNA processing and transcription termination with changes in timing of dephosphorylation by the four CTD phosphatases leading to the production of distinct transcriptional readouts.

We have previously shown that deletion of the Ser5-P CTD phosphatase Rtr1 results in an increase in global Ser5 RNAPII phosphorylation and disruption of termination at specific protein-coding genes [13, 26]. Although Rtr1 may have additional cellular targets other than the RNAPII CTD, this has not been fully explored to date. However, it has been previously shown that RNAPII is the major interaction partner of Rtr1 in yeast, suggesting that its major role may likely involve the regulation of RNAPII transcription at least in standard growth conditions [60]. Rtr1 has previously been implicated in the mediation of stress response to heat shock, however its role in this process has not been fully characterized [61]. In this work, data shows that RTR1 deletion leads to alterations in CID-containing protein interactions with RNAPII likely mediated by their interactions with the CTD. Additionally, the interactions between CFIa and CPF are decreased in the absence of Rtr1, suggesting that the timing of CTD dephosphorylation may regulate the formation of stable interactions between the termination machinery, the nascent RNA, and RNAPII. Transcriptome analysis reveals that rtr1Δ cells have decreased levels of a variety of noncoding transcripts, suggesting that their production is Rtr1-dependent. Analysis of Nrd1 occupancy shows that both Nrd1 and RNAPII accumulate at known Nrd1 binding sites in WT cells but that RTR1 deletion causes a decrease in RNAPII and Nrd1 levels at both noncoding and coding genes, which could suggest increased elongating RNAPII turnover through termination. This study shows that Rtr1 plays a role in fine-tuning the NNS termination pathway such that RTR1 deletion causes decreased steady-state levels of numerous noncoding RNAs and increases in premature termination of RNAPII at protein-coding genes by increasing the efficiency of termination through the NNS-dependent termination pathway. Since the NNS proteins are known to recruit the RNA exosome to carry out termination-coupled RNA processing and/or decay, the impact of rtr1Δ rrp6Δ was also explored, revealing that the decreases in noncoding RNA levels in rtr1Δ require Rrp6 activity. These findings clearly show that Rtr1 attenuates termination through the NNS pathway. Overall, our findings suggest that precise control of CTD dephosphorylation is required to maintain the balance between elongation and termination at a wide variety of target genes whose transcripts are produced by RNAPII and co-transcriptionally processed.

Results

Disruption-compensation analysis reveals changes in termination factor interactions in RTR1 deletion cells

The phosphorylation status of the RNAPII CTD plays a major role in the regulation of the mechanisms through which transcription termination occurs [6265]. We have recently shown that deletion of RTR1 causes global increases in CTD Ser5-P [26] and it has previously been shown that loss of RTR1 results in 3’-end processing defects at the polyA-dependent gene NRD1 [13]. To determine the role of Rtr1 in the regulation of RNAPII interactions with termination factors, we performed Disruption-Compensation (DisCo) network analysis. It has been postulated that genetic perturbations can cause edge-specific changes in protein-protein interaction (PPI) networks such as loss or gain of an edge or a change in the strength of the PPI [66, 67]. DisCo combines genetic perturbation and in-depth affinity purification-mass spectrometry (AP-MS) studies to obtain unique biological insights into the mechanisms that cause phenotypic changes in gene expression networks. For these studies, we generated dynamic protein-protein interaction networks using Significance Analysis of INTeractome (SAINT) probability scores in the presence or absence of Rtr1 [68, 69]. Four biological replicate affinity purifications were performed for Nrd1-TAP, Pcf11-TAP, and Ssu72-FLAG to represent the Nrd1-Nab3, cleavage factor Ia (CFIa), and cleavage and polyadenylation factor (CPF) complexes, respectively, from either WT or rtr1Δ cells. The resulting data matrix consists of 24 x 3,960 protein-level measurements in 668,969 peptide-spectrum matches (PSMs, S1 Table). We focused our analysis on high-confidence interactions between the protein components of the termination factor complexes along with the two largest subunits of RNAPII, Rpb1 and Rpb2, although a lower stringency network prepared using STRING v11 is also included (S1 Fig, [70]).

Prey-prey correlation analysis was performed for all purifications from WT or RTR1 deletion genotypes. In brief, a high correlation value between two proteins suggests that they have a similar distribution of PSMs across the same set of purifications independent of the bait protein used for purification. Proteins that function within the same protein complex typically have the highest correlation values, as shown in Fig 1A. In addition, there is evidence of cross-association of cleavage factor Ia (CF1a), the cleavage and polyadenylation factor (CPF) and RNAPII (Pol II) in WT cells through a positive association between CFIa and the CPF subunits Fip1 and Pap1, which are both components of the recently described poly(A) polymerase module of CPF (Fig 1A, [71]). Analysis through SAINT probability calculation revealed association of additional CPF subunits with CFIa (Fig 1C). These data support a previous model, which suggested that the CTD-interaction domain (CID) of Pcf11 facilitates formation of a CPF-CF1a-RNAPII complex for stable 3’-end complex formation and mRNA polyadenylation [27]. However, in cells lacking the CTD phosphatase Rtr1, the cross-correlation between the CFIa complex and CPF complex subunits Fip1 and Pap1 is markedly reduced (Fig 1B). In addition, the correlation between CFIa and RNAPII is also reduced, suggesting that deletion of RTR1 leads to reduced interactions between Pcf11, the bait protein for CFIa, and RNAPII.

DisCo network analysis of complexes involved in RNAPII transcription termination in WT and <i>rtr1Δ</i> yeast.
Fig. 1. DisCo network analysis of complexes involved in RNAPII transcription termination in WT and rtr1Δ yeast.
A) Prey-prey correlation analysis of yeast termination complex affinity purification-mass spectrometry data from WT cells of n = 4 biological replicate purifications for each bait protein (Nrd1, Pcf11, and Ssu72, total n = 12). B) Prey-prey correlation analysis of yeast termination complex affinity purification-mass spectrometry data from rtr1Δ cells of ≥ 4 biological replicate purifications for each bait protein (Nrd1, Pcf11, and Ssu72, total n = 12). C) Ssu72-FLAG, Pcf11-TAP and Pcf11-FLAG, and Nrd1-TAP protein-protein interaction networks from WT and rtr1Δ cells depicting SAINT analysis of n = 4 biological replicate purifications for each genotype. The nodes represent a protein of interest whereas the edges represent the SAINT interaction probability as indicated in the legend. Each protein-protein interaction network represents a subset of the full protein-protein interaction network which is shown in the supplement with a fold-change cutoff value of 5 for each WT purification (S1 Fig).

SAINT analysis provides additional insights into the interactions between the termination machinery through interaction probability calculation. Ssu72 is a known member of the CPF complex [14], and all CPF subunits had SAINT probabilities of ≥ 0.95 (Fig 1C, S2 Table). When isolating protein complexes through affinity purification-mass spectrometry, we have found that protein complex subunits typically have SAINT probabilities in the 0.95–1 range, as observed with Ssu72-FLAG and the subunits of CPF. However, proteins that interact with the protein complex of interest can display a wide-range of SAINT probabilities that likely reflect the dynamic nature of some protein-protein interactions. Of note, dynamic interaction partners of RNAPII that regulate different stages of transcription were previously assigned SAINT probability values that ranged from 0.23 (Spt16) to 1 (Spt5, Pta1, Tfg1, Tfg2) in replicate Rpb3-TAP purifications [72]. In the Ssu72-FLAG purifications, the two largest subunits of RNAPII, Rpb1 and Rpb2, were detected as significant Ssu72 interacting proteins in both WT and rtr1Δ cells (Fig 1C). Rpb1 and Rpb2 have the highest number of detectable peptides amongst the twelve RNAPII subunits and therefore the highest probability of detection in an affinity purification-mass spectrometry approach [73]. Altogether, these data suggest that loss of Rtr1 function does not alter the interactions between the Ssu72-copurifying CPF complex and RNAPII, in agreement with the prey-prey correlation analysis (Fig 1). Higher correlation values were observed between the subunits of the CPF and RNAPII in rtr1Δ cells (Fig 1B). The SAINT probability values suggest that the increase in prey-prey correlation values is likely a consequence of the lack of detection of interactions between Pcf11/CFIa and RNAPII rather than an increased level of interaction between RNAPII and CPF.

CFIa has been characterized as a four-subunit protein complex containing Rna14, Rna15, Clp1, and the CID-containing protein Pcf11. SAINT analysis of biological replicates of Pcf11-TAP and Pcf11-FLAG purifications from WT and rtr1Δ cells (n = 4) identified these four CFIa subunits with probabilities ≥ 0.95, supporting their designation as a protein complex (Fig 1C, S2 Table). Previous studies have clearly shown that the purified Pcf11 CID domain has lower affinity for Ser5-P and Ser2,5-P than Ser2-P modified CTD [27]. In RTR1 deletion cells, we have shown that Ser2,5-P CTD repeats are present further downstream than in WT cells, as supported by increased histone H3K36me3 levels in rtr1Δ cells [26]. As illustrated in Fig 1C, we find that the RNAPII subunits Rpb1 and Rpb2 interact with Pcf11 in WT cells with a SAINT probability of 0.75 (S2 Table). However, no statistically significant interaction was detected between Pcf11 and Rpb1/Rpb2 when isolated from cells lacking Rtr1, as also observed from prey-prey correlation analysis (Fig 1B & 1C, S2 Table). It is possible that the interaction between Pcf11 and Rpb1 still occurs in RTR1 deletion cells, but that it was below the limit of detection for our affinity purification-mass spectrometry studies. Of interest, a number of CPF subunits were found in single-affinity step Pcf11-FLAG purifications at low levels, which are more apparent with SAINT analysis than through correlation analysis. Interestingly, the SAINT probabilities for CPF subunit interactions with CFIa were similar in WT and RTR1 deletion cells although fewer subunits of CPF were recovered (Fig 1C). These findings suggest that the interactions between the CFIa and CPF complexes occur in the absence of stable Pcf11-CTD interactions that were not detected in RTR1 deletion cells. However, it is likely that CFIa-CPF interactions are strengthened through association with the nascent RNA and are further stabilized through interactions with the RNAPII CTD.

Nrd1 has been proposed to function within a protein complex containing Nab3, Sen1, Cbc2, and Sto1 [30]. However, in our quantitative proteomics analysis of Nrd1 affinity purifications, only Nab3 was found to interact with Nrd1 with a SAINT probability of ≥ 0.95, and this interaction was also the only high correlation prey-prey value found (Fig 1). Sen1, Sto1, and Cbc2 were identified as Nrd1 interacting proteins, although their SAINT probability values indicate they are dynamic interacting partners of the bait, Nrd1 (Fig 1C). Additionally, subunits of the TRAMP complex and the nuclear exosome were also identified as dynamic interacting partners of Nrd1 consistent with previous findings (Fig 1C, [30]). The Nrd1 CID has been shown to have the highest affinity for Ser5-P and Ser2,5-P CTD repeats [63], whose abundance is increased in RTR1 deletion mutants [13, 26]. As illustrated in Fig 1C, an interaction between Nrd1 and Rpb1 was detected with a SAINT probability of 0.49 in RTR1 deletion cells (S2 Table). These findings suggest that loss of RTR1 increases the interaction between the Nrd1-Nab3 complex and RNAPII in vivo, likely due to increases in the number of Ser5-P modified CTD repeats. Even in rtr1Δ cells the interaction probability between Nrd1 and Rpb1 was lower than what was measured for Pcf11 from WT cells. This may suggest that the Nrd1-Rpb1 interaction occurs at a lower frequency than the Pcf11-Rpb1 interaction, which is consistent with previously measured binding affinities for each CID.

Rtr1 impacts global RNA expression

To determine how global RNAPII transcription was altered upon loss of the RNAPII CTD phosphatase Rtr1, we performed strand-specific RNA-Seq analysis of total RNA from four biological replicate RNA purifications. A spike-in control was included to detect the presence of global transcription defects [74]. Following alignment, differentially expressed transcripts were identified using edgeR analysis (Fig 2, S3 Table, [75]). Previously defined transcript annotations were used to distinguish multiple types of RNAPII transcripts including small nuclear/nucleolar RNAs (sn/snoRNAs), open reading frame transcripts (ORF-Ts), cryptic unstable transcripts (CUTs), stable unannotated transcripts (SUTs), and Nrd1-unterminated transcripts (NUTs) [76, 77]. Transcripts that are antisense to ORF-Ts were annotated as antisense transcripts (ASTs). In total, there was a reduction in 1,481 out of 11,151 transcripts in RTR1 deletion cells, including a significant number of ASTs and other ncRNA transcripts (Fig 2A & 2B). Two-hundred and seventy-six transcripts showed upregulation of more than 1.5-fold, many of which were ORF regions (Fig 2C). The most significantly reduced transcript was IMD2, whose expression is regulated by GTP levels as well as an NNS terminator (Fig 2A, labeled in green, [46, 47, 7880]). We also confirmed that Imd2 protein levels are significantly decreased in rtr1Δ cells using global proteomics analysis (S2 Fig, S4 Table). These data are also consistent with previous findings that RTR1 deletion cells show growth sensitivity to the IMD2 inhibitor mycophenolic acid [61, 81]. Several of the most strongly decreased transcripts in rtr1Δ cells are from subtelomeric genes, particularly the PAU genes (S3 Table). Of note, these subtelomeric genes have previously been shown to be silenced through a mechanism which requires Sen1 perhaps through its regulation of ASTs [80]. Overall, these data suggest that loss of Rtr1 activity results in the downregulation of a number of different classes of RNAs, although many of the transcripts for which large changes in RNA levels were observed have previously been implicated as dependent on the NNS termination pathway [2832, 77, 82]. These data suggest that the increased Ser5-P RNAPII CTD levels in rtr1Δ cells may facilitate elevated activity of the NNS-dependent termination pathway.

Loss of Rtr1 activity causes wide-spread changes in RNA expression in yeast.
Fig. 2. Loss of Rtr1 activity causes wide-spread changes in RNA expression in yeast.
A) Volcano plot for normalized RNA-Seq data for RNA transcripts in rtr1Δ vs. WT cells (n = 4, edgeR analysis). Data is shown for all annotated ORF-Ts cryptic unstable transcripts (CUTs), Nrd1 unterminated transcripts (NUTs), SNR extended transcripts, Ssu72 restricted transcripts (SRTs), and antisense transcripts (ASTs). False discovery rates (FDR) were calculated by edgeR and a cutoff of ≤ 0.05 is indicated with a dashed red line. Density plots are shown relative to each axis for comparison. B&C) Distribution of transcripts differentially expressed in rtr1Δ cells compared to WT across annotation categories. Downregulated transcripts are represented in the red diagram while upregulated transcripts are represented in the green diagram. In addition to the FDR cutoff of ≤ 0.05, transcripts were considered differentially expressed with a fold-change value of at least +/-1.5-fold.

RNAPII and Nrd1 occupancy was measured genome-wide through chromatin immunoprecipitation followed by exonuclease digestion and genome-wide sequencing (ChIP-exo) as described previously [83]. Considering that Nrd1 does not bind DNA directly, rather it binds nascent RNA and RNAPII CTD repeats at Ser5-P, we predicted that ChIP-exo of Nrd1-TAP would detect regions of DNA bound by RNAPII in complex with Nrd1 and possibly Nab3 and Sen1 (Fig 3A). To confirm that the binding patterns observed were specific to Nrd1, we compared the Nrd1-TAP ChIP-exo normalized read counts from WT cells to those of Rpb3-FLAG (Fig 3B). URA8 is known to be regulated by alternative start site selection that is dependent on nucleotide availability and is a known target for NNS-dependent early termination [31, 77]. The SOD1 locus is convergent with URA8 and lacks RNA binding sites for Nrd1. Fig 3B illustrates the differences seen in the binding patterns of total RNAPII (Rpb3-FLAG) and Nrd1-bound RNAPII (Nrd1-TAP) when comparing transcripts with high (URA8) and low (SOD1) levels of consensus Nrd1-Nab3 RNA binding sites. The consensus Nrd1 binding site of TTTGTAAAGTT is located 40 nt upstream of the URA8 ATG. The alternative start site is terminated by the NNS pathway in nutrient-rich conditions, such as growth in YPD as used in this study. Our ChIP-exo analysis of Rpb3-FLAG shows that RNAPII is localized at the 5’-end of the URA8 gene and throughout the SOD1 coding region (Fig 3B). The 5’-end localization of RNAPII at URA8 corresponds with the peak of Nrd1 binding in the same area, supporting previous work that found that the majority of URA8 transcript is terminated in early elongation by the NNS pathway, resulting in low-level transcription of full-length URA8 transcript [84]. The levels of Nrd1 association at the SOD1 gene are much lower than at URA8 even though total Rpb3-FLAG occupancy is relatively higher at SOD1 than at URA8 confirming that we are able to obtain selective enrichment of Nrd1 on chromatin using the ChIP-exo approach.

Analysis of RNAPII and Nrd1 occupancy throughout the genome in WT cells.
Fig. 3. Analysis of RNAPII and Nrd1 occupancy throughout the genome in WT cells.
A) Simplified schematic of immunoprecipitation of Nrd1-TAP by IgG-sepharose beads. Nrd1 binds to sequences in the RNA (red rectangle) via its RNA Recognition Motif (RRM domain). B) Graphical representation of Rpb3-FLAG (black) and Nrd1-TAP (green) occupancy in the WT strain as determined by ChIP-exo sequencing reads mapped to URA8-SOD1. Data is shown from a representative biological replicate. Points of interest are noted by arrows. C) Occupancy of RNAPII and Nrd1 at the NRD1 gene by ChIP-exo. Points of interest are noted by arrows. D) Occupancy of RNAPII and Nrd1 at the SNR13 gene by ChIP-exo. Points of interest are noted by arrows. E) Average gene analysis of Nrd1 occupancy by ChIP-exo at the top 100 most enriched Nrd1 RNA binding locations throughout the yeast genome. F) Average gene analysis of RNAPII (Rpb3) occupancy by ChIP-exo at the top 100 most enriched Nrd1 RNA binding locations throughout the yeast genome.

Upon analysis of the Nrd1 ChIP-exo dataset, a pronounced peak of RNAPII was observed upstream of the location of well-positioned Nrd1-RNAPII complexes (Fig 3C, arrow 1 vs. arrow 2). These peaks were observed at both protein-coding gene NNS targets such as NRD1 (Fig 3C) and noncoding genes such as SNR13 (Fig 3D, compare arrow 1 for upstream peak to arrow 2 for Nrd1-RNAPII peak). To further explore these findings, we used previously published Nrd1 PAR-CLIP datasets to annotate the top 100 most intense sites and then averaged the Nrd1 and RNAPII intensities surrounding the Nrd1 consensus RNA binding sites [31, 32]. In Fig 3E, we observed a narrow peak of Nrd1-RNAPII complexes located just downstream of the genomic location of the Nrd1 consensus motif (marked with a red line). A similar peak is observed from the Rpb3 ChIP-exo, although the peak is somewhat 5’ shifted, perhaps as an average of the Nrd1 bound and unbound RNAPII populations (Fig 3F). Globally, Nrd1 was also found to localize to the 5’-end of most mRNA encoding genes and, in agreement with previous studies using ChIP-microarray analysis, the mRNA peak of Nrd1 occupancy occurs 93 +/- 3 nucleotides downstream of the annotated mRNA transcription start sites (TSS, S3 Fig).

RNA Polymerase II and Nrd1 occupancy are reduced at SNR genes in rtr1Δ

Total RNA-Seq analysis revealed changes in a number of ncRNA classes, including SNR transcripts (Fig 2) and SNR transcript 3’-ends that we have manually annotated as extended transcripts (ETs), which are the regions downstream of SNR transcripts that are within the zone of termination [85]. Full SNR transcripts are subsequently subjected to 3’-end processing through the NNS-termination pathway in coordination with the TRAMP complex and the Rrp6-containing RNA exosome [30, 82, 8589]. Average gene analysis was performed using the ChIP-exo datasets for Rpb3 and Nrd1 for the SNR genes aligned to the TSS with 500 bp of data upstream and 1kb downstream [90]. The overall decrease in RNAPII occupancy is around 9% at snRNA genes (p-value = 6.97e-7), but this has an impact on a number of mature SNR transcripts with many showing significant decreases in abundance in the absence of RTR1 (Fig 4A, S3 Table). Nrd1 occupancy is decreased at 17% of SNR genes in rtr1Δ deletion cells (p-value = 9.63e-07, Fig 4A). This decrease in Nrd1 occupancy could be a consequence of more efficient NNS termination leading to lower steady-state levels of Nrd1 at SNR genes and/or a consequence of lower Nrd1 protein levels given that NRD1 mRNA levels are decreased (Fig 2A). Global proteomics analysis of RTR1 deletion cells found a slight reduction (~12%) in Nrd1 protein levels (S2 Fig, S4 Table). Regardless, the decrease in RNAPII levels at SNR genes suggests a higher degree of RNAPII turnover, which could occur through increased NNS-dependent termination. However it is also possible that there are decreases in RNAPII initiation. Deletion of the exosome subunit Rrp6 leads to accumulation of many SNR gene extended transcripts through defective 3’ end processing of sn/snoRNA transcripts as well as altered RNAPII termination [57]. Direct comparison of the changes in rrp6Δ cell sn/snoRNA transcript levels to those seen in rtr1Δ deletion cells is shown in Fig 4B. Significantly altered sn/snoRNA transcripts in rtr1Δ deletion cells have been highlighted, and are referred to as either “Rtr1-dependent (Rtr1-D) sn/snoRNA transcripts” for the mature yeast genome annotations for each sn/snoRNA (highlighted in blue) or “Rtr1-D sn/snoRNA extended transcripts” (ETs,highlighted in red) for RNAs that extend past their yeast genome annotated 3’ end boundaries (Fig 4B). A majority of the changes observed in Rtr1-D sn/snoRNA full length or extended transcripts are decreases in abundance (~40%). A total of 33 significantly decreased mature sn/snoRNA transcripts and 27 significantly decreased sn/snoRNA ETs were measured in rtr1Δ cells with many of the same transcripts increasing in abundance in rrp6Δ deletion cells (Fig 4B). The fold-increase in sn/snoRNA ETs in rrp6Δ is much larger than the decreased seen in rtr1Δ cells, however, this may be expected considering that sn/snoRNA ETs are produced at very low levels in WT cells. The quantitation for a subset of the ETs is shown in Fig 4C shows the relative heterogeneity seen in this transcript class, likely as a consequence of their termination in zones rather than distinct sites [85].

Loss of Rtr1 activity leads to decreased RNAPII occupancy at <i>SNR</i> genes and shortened <i>SNR</i> transcripts.
Fig. 4. Loss of Rtr1 activity leads to decreased RNAPII occupancy at SNR genes and shortened SNR transcripts.
A) Average gene analysis of RNAPII and Nrd1 occupancy at SNR genes in WT (left) or rtr1Δ (right) cells. The data is shown as average read count per million (RPM) values calculated using ngs.plot -500 and +1000 nucleotides relative to the SNR gene annotated TSSs in either WT or rtr1Δ cells. B) Comparison of SNR transcript fold-change values from rtr1Δ/WT and rrp6Δ/WT datasets. Data shown are color coded as indicated in the legend above the figure for all SNR transcripts and Rtr1-dependent (Rtr1-D) transcripts. C) RNA-Seq read count values for manually annotated extended transcript (ET) regions from a subset of SNR genes. ETs were manually annotated based on the extended RNAPII signal observed in Rpb3-3XFLAG ChIP exo datasets for RRP6 deletion mutants [57]. Data is shown as average read count values with standard deviations shown and significance where appropriate based on paired student’s t test.

Global levels of RNA Polymerase II and Nrd1 occupancy are altered in RTR1 deletion cells

Although Nrd1 recruitment is highest at RNAPII target genes containing RNA binding sites for Nrd1-Nab3 (such as URA8), average gene analysis in this study and others has shown that Nrd1 is recruited just downstream of the peak of RNAPII Ser5-P CTD phosphorylation at protein-coding genes [31, 49]. At the model protein-coding gene PMA1, RNAPII occupancy is relatively constant across the entire length of the gene (Fig 5A, upper panel). Nrd1 binding, in contrast, peaks in the 5’ end of the gene, ~270–321 nucleotides past the annotated transcription start site of PMA1 (Fig 5A, lower panel). To measure molecular changes that occur at the site of transcription, occupancy of Nrd1-TAP and Rpb3-FLAG in cells +/- RTR1 were compared at non-overlapping protein-coding genes 1000 nucleotides downstream of the TSS (Fig 5B & 5D). MNase-Seq analysis of histone occupancy data from WT cells is also included for reference and shows that RNAPII occupancy by ChIP-exo nicely tracks with the average nucleosome occupancy across the gene (Fig 5B & 5D). The overall RNAPII and Nrd1 occupancy observed relative to the annotated transcript end site (TES) +/- 200 nucleotides is also shown. Average Rpb3 localization is slightly higher at the 5’ end of protein-coding genes in RTR1 deletion cells relative to WT. However, Rpb3 occupancy decreases in rtr1Δ more than in WT cells as RNAPII progresses towards the 3’-end of these genes and at the TES (Fig 5B). The overall change in RNAPII levels suggests a higher degree of premature transcription termination, considering that Rpb3-FLAG occupancy is higher in rtr1Δ cells than in WT cells, which does not indicate a transcription initiation defect (Fig 5B). To explore the overall change in more detail, the ratio of RNAPII occupancy at the TSS relative to the TES for each protein coding gene is shown (Fig 5C). The data was separated into quartiles based on their RNAPII occupancy levels in WT cells. An increase in the ratio between the TTS and TES suggests premature RNAPII termination. A significant change in the TSS/TES ratio is observed for all quartiles in RTR1 deletion cells relative to WT (Fig 5C). Nrd1 levels show a small but consistent decrease in rtr1Δ samples relative to WT across the entire average gene and at the TES (Fig 5D). Interestingly, the transition to lower levels of RNAPII occupancy in rtr1Δ relative to WT occurs just following the peak of Nrd1 recruitment (overlaid data shown in S4 Fig).

Deletion of <i>RTR1</i> increases the ratio of RNAPII at the transcription start site relative to the transcription termination site.
Fig. 5. Deletion of RTR1 increases the ratio of RNAPII at the transcription start site relative to the transcription termination site.
A) Normalized read count data for RNAPII (Rpb3-3XFLAG) and Nrd1-TAP at the PMA1 protein-coding gene. B) Average gene analysis of RNAPII (Rpb3) levels at 2,426 protein-coding genes -200 nucleotides upstream and +1000 downstream of the TSS and -200/+200 upstream/downstream of the annotated TES. The data is shown as average RPM values calculated using ngs.plot for each genotype designated by the colors in the legend. MNase-Seq analysis of total histones from WT cells is also shown as a reference for chromatin organization along the average gene. C) Average ratio of RNAPII occupancy at the TSS/TES for 2,426 protein coding genes. Statistical analysis of the ratios for WT and RTR1 deletion cells was determined by paired student’s t-test. D) Average gene analysis for Nrd1-TAP ChIP-exo as shown for panel B.

IMD2 is a well-described NNS target gene whose expression is controlled by RNAPII transcription start site selection with four of its previously mapped upstream start sites (USS) located upstream of a consensus Nab3 binding site important for an IMD2 cryptic unstable transcript (CUT) terminator (Fig 6A) [46, 47, 79]. Regulation of start site selection has been shown to control basal levels of IMD2 expression until low GTP levels allow for preferential use of the downstream TSS most proximal to the IMD2 ATG (Fig 6A) [46, 47, 79]. However, previous RNA-Seq studies using polyA purified RNA have shown that transcripts can be produced from the USS at chromosome VIII at 554174 resulting in its annotation as the 5’ end of the IMD2 transcript (abbreviated as 174, Fig 6A, [91]). In rich media conditions, such as those used for these experiments, an upstream IMD2 CUT is produced from different TSSs, as indicated, and is often terminated within the region indicated in Fig 6A by NNS-dependent termination. However, a portion of IMD2 transcripts are able to escape termination, resulting in a basal level of IMD2 mRNA expression as observed in this study via RNA-Seq in WT cells (S3 Table). The IMD2 CUT terminator has been shown to have a strong functional requirement for Nab3, however both Nab3 and Nrd1 individually interact with IMD2 RNA in this region as previously shown by PAR-CLIP studies [31, 32, 92]. By ChIP-exo, both RNAPII and Nrd1 can be mapped to the upstream CUT region in WT cells with low levels of RNAPII observed in the IMD2 coding region (Fig 6B & 6C). MACS (Model-based Analysis for ChIP-seq) identified two significant Nrd1-TAP ChIP-exo peaks suggesting that this is a significant binding event (data provided in S5 Table, apex of each peak labeled in Fig 6A–6C). However, both RNAPII and Nrd1 levels are significantly reduced in rtr1Δ cells (Fig 6B & 6C).

Rtr1 is required for basal <i>IMD2</i> expression.
Fig. 6. Rtr1 is required for basal IMD2 expression.
A) Schematic representation of the IMD2 start sites. Previously defined upstream start sites (USS) are indicated with black arrows with the downstream start site (TSS), preferred in low GTP conditions, indicated with a red arrow [79]. The orange box designates the annotation information for the IMD2 upstream CUT whereas the black box designates the annotation information for the IMD2 mRNA. The red box designates the location of the previously characterized IMD2 terminator region in both A and B [46, 79]. These have been drawn to scale based on each annotation and previously published observations. MACS analysis of Nrd1-RNAPII peaks from the ChIP-exo datasets is indicated with aqua lines. The translation start site is also indicated as a red line. Occupancy of RNAPII (B) and Nrd1 (C) at the IMD2 gene. Data derived from WT cells is in black, and those from the rtr1Δ strain are in orange. The location and direction of transcription for all analyzed annotations are diagrammed below the graphs, each to scale as in A. D) Expression of the IMD2 mRNA RNA-Seq (n = 4) with FDR values calculated using ANOVA with a Tukey’s multiple comparison test. The symbols designate FDR values ≤ 0.001 as follows: * significantly different from WT, # significantly different from rtr1Δ, † significantly different from rrp6Δ, and § significantly different from rtr1Δ rrp6Δ. E) Expression of IMD2 upstream CUT by RNA-Seq (n = 4) with statistics performed as indicated for D. F) Northern blot analysis of IMD2 expression in various genotype cells as indicated. The location of the probe used for this northern is indicated by an aqua line under panel C.

In Fig 6B, RNAPII levels are similar near the TSS of IMD2 but the difference between Rpb3-FLAG occupancy in WT vs. rtr1Δ cells increases following Nrd1 recruitment, suggesting a higher efficiency RNAPII termination. In addition to RNA-Seq analysis of WT and rtr1Δ cells, we also performed RNA-Seq analysis of rrp6Δ and rtr1Δ rrp6Δ cells to determine the role of the Rrp6-containing exosome in the downregulation of ncRNA transcripts in rtr1Δ cells (S5A & S5B Fig, S6 Table). Deletion of RTR1 causes significant decreases in IMD2 mRNA levels as determined by ANOVA, although deletion of RRP6 did not significantly alter IMD2 mRNA levels (Fig 6D). However, the rtr1Δ rrp6Δ cells have intermediate levels of IMD2 mRNA, suggesting that Rrp6 does contribute to the decreased transcript levels in rtr1Δ. These findings, along with the well-described role of Rrp6 in degradation of NNS-terminated transcripts, suggests that premature IMD2 transcript termination occurs at elevated levels in cells lacking Rtr1 [30, 93, 94]. The incomplete recovery of IMD2 mRNA levels in rtr1Δ rrp6Δ could be a consequence of the overlapping roles of Rrp6 and Dis3 in RNA turnover in the nucleus, considering that prematurely terminated transcripts would not have proper 3’ end processing making them susceptible to 3’-5’ exonucleases [2, 95]. The IMD2 upstream CUT RNA is significantly decreased in rtr1Δ relative to rrp6Δ, rtr1Δ rrp6Δ, and WT cells as determined by ANOVA (Fig 6E). IMD2 upstream CUT levels are significantly increased in rrp6Δ cells relative to all other genotypes, as would be expected from previous studies [80, 96]. However, rtr1Δ rrp6Δ cells show similar levels of the IMD2 upstream CUT relative to WT, suggesting that Rrp6 is required for the IMD2 CUT decrease observed in rtr1Δ. Interestingly, IMD2 CUT levels do not reach the same levels observed in rrp6Δ in the rtr1Δ rrp6Δ suggesting that loss of Rtr1 causes an additional defect in IMD2 CUT production/stability that is not rescued by rrp6Δ. For instance, it is possible that Dis3 could contribute more to IMD2 CUT turnover in rtr1Δ than observed in rrp6Δ. Northern blot analysis of the IMD2 mRNA reflected similar ratios to RNA-Seq and confirmed that the terminator over-ride (TOV) mutant in the Ser5-P CTD phosphatase Ssu72 (ssu72-tov) also showed increased expression of the full length IMD2 transcript (Fig 6F, [45]). These data clearly show that while Ssu72 and Rrp6 are required for turnover of the IMD2 CUT RNA, Rtr1 is a positive regulator of IMD2 CUT and mRNA production such that knockout of RTR1 leads to decreased IMD2 transcript levels.

Rtr1 promotes elongation of antisense noncoding transcripts

The Rrp6-containing nuclear exosome and NNS machinery have previously been implicated in the modulation of sense transcripts regulated by antisense transcript (AST) elongation [97]. Considering the large number of ASTs whose expression is decreased in RTR1 deletion cells, we performed more in-depth analysis of RNA-Seq data for ASTs from the rrp6Δ and rtr1Δ rrp6Δ datasets relative to rtr1Δ (S4 Fig, S6 Table). A focus on the antisense transcripts (ASTs) that were significantly downregulated in rtr1Δ relative to WT (n = 104) revealed that the vast majority of the Rtr1-dependent ASTs depend on the Rrp6-containing exosome for their downregulation (Fig 7). These data further suggest that the downregulation of the Rtr1-dependent ASTs is likely related to increased premature ncRNA termination through the NNS pathway, which couples with the Rrp6-containing exosome for RNA degradation [42, 82, 87, 95]. Elongation of noncoding transcripts, including ASTs, has been shown to regulate protein coding gene expression through transcription interference [77, 97]. Upregulation of noncoding transcript elongation, promoted by Rtr1 in WT cells, may therefore alter overall transcriptome profiles. It has been shown that transcription of the protein-coding gene YKL151C is regulated by an NNS-terminated antisense transcript, which is labelled as YKL151C AS [31, 98]. Elongation of the YKL151C AS leads to transcription interference of YKL151C expression (Fig 8A). Northern blot analysis of YKL151C / YKL151C AS expression using single stranded RNA probes shows that YKL151C sense is up-regulated in rtr1Δ, which correlates with YKL151C AS transcript downregulation (Fig 8B). The opposite effect can be observed in the rrp6Δ, and rtr1Δ rrp6Δ strains, suggesting that decreased NNS-dependent termination leads to YKL151C sense elongation (Fig 8). In ssu72-tov cells, YKL151C sense transcripts are decreased as well with a coordinate increase in YKL151C antisense, although both changes occur to a lower extent than the changes seen in rrp6Δ (Fig 8B). Furthermore, strand specific RNA-Seq analysis confirmed that YKL151C is upregulated in rtr1Δ while the YKL151C AS is significantly downregulated (Fig 8C & 8D). Overall, these data suggest that Rtr1 activity promotes elongation of a number of antisense RNAs that may contribute to maintenance of proper sense transcript production through transcription interference as observed for PHO84 and YKL151C [31, 97, 98].

Abundance of antisense transcripts in <i>RTR1/RRP6</i> double deletion strain phenocopies <i>rrp6Δ</i> cells.
Fig. 7. Abundance of antisense transcripts in RTR1/RRP6 double deletion strain phenocopies rrp6Δ cells.
Heat map of normalized expression values for ASTs in rtr1Δ, rrp6Δ, and rtr1Δ rrp6Δ compared to WT cells, according to the scale at the right. The ASTs shown were selected for clustering analysis based on their significant decreased expression relative to WT in RTR1 deletion cells. Unsupervised hierarchical clustering analysis was performed using normalized read count values for each biological replicate and genotype (n = 16). Each genotype clustered together illustrating a high degree of reproducibility between biological replicates with WT values groups as Cluster IV, rtr1Δ values grouped as Cluster III, rtr1Δ rrp6Δ grouped as Cluster II, and rrp6Δ grouped as Cluster I. Genotypes are also indicated at the bottom of the figure for reference.
Analysis of YKL151C sense and antisense expression in WT and <i>rtr1Δ</i> cells.
Fig. 8. Analysis of YKL151C sense and antisense expression in WT and rtr1Δ cells.
A) Northern blot analysis of YKL151C sense and antisense transcripts using single-stranded RNA probes. Band intensity relative to WT is shown under each relevant panel with labels found to the right. B) Schematic representation of YKL151C sense and antisense expression. C&D) Summary of RNA-Seq expression analysis of YKL151C sense and antisense transcript expression, FDR values are displayed above/below bar as calculated by edgeR.

Discussion

Through analysis of DisCo networks, RNA-Seq, and ChIP-exo, this work shows that Nrd1/RNAPII co-localization and protein-protein interaction is increased and NNS-dependent termination is enhanced in the absence of the atypical phosphatase Rtr1. This would suggest that in WT cells, Rtr1 attenuates the NNS pathway whereas the Ssu72 CTD phosphatase stimulates termination through NNS. This represents a novel role for Rtr1-dependent CTD dephosphorylation in the regulation of RNAPII termination choice in eukaryotes. Although some of the changes observed in RTR1 deletion cells were subtle, it is important to note that the impact of NNS-dependent termination on the transcriptome is limited by at least two major mechanisms. The first is that the NRD1 gene is autoregulated by Nrd1 protein regulation of its own RNA production. As a consequence, any increase in Nrd1 termination efficiency would lead to a corresponding decrease in NRD1 mRNA and protein levels. Accordingly, NRD1 mRNA levels are significantly reduced in rtr1Δ (Fig 2) and a corresponding 10% reduction in Nrd1 protein levels was also observed (S2 Fig). Secondly, Nrd1/Nab3 binding sites have been shown to occur at a low frequency in protein coding mRNA transcripts yet occur at high frequency in ASTs [99]. The depletion of high affinity Nrd1/Nab3 binding sites from most protein coding genes likely limits premature termination of mRNAs. Despite this, the ratio of RNAPII occupancy at the TSS/TES in rtr1Δ cells suggests that premature termination occurs broadly at mRNA encoding genes as a consequence of reduced CTD dephosphorylation (Fig 5). Additionally, Fig 7 illustrates the broad impact of RTR1 deletion on ASTs and that their abundance change is dependent on Rrp6-containing exosome function. These data, in light of previous work from many groups, show that transcript classes that are enriched in Nrd1/Nab3 binding sites rely on Rtr1 for their production and stability such that in rtr1Δ ASTs are produced yet rapidly degraded as a consequence of increased NNS-termination. Overall, these findings suggest that the balance of premature termination vs. elongation of both coding and non-coding transcripts is maintained through the action of Rtr1.

DisCo network analysis was able to differentiate changes in protein complex interactions that occur between RNAPII, CFIa, CPF and NNS as a consequence of RTR1 deletion. The decreased interaction between CFIa and RNAPII was evident through both prey-prey correlation analysis and SAINT probability analysis which also revealed that interactions between CFIa and CPF were detected less frequently in rtr1Δ cells. Additionally, interactions between Nrd1 and RNAPII were increased in cells lacking Rtr1 activity supporting our transcriptome-level findings that show that RNA produced from NNS-dependent genes can terminate with higher efficiency (Figs 48). However, the protein-protein interactions studies did not find evidence of altered interactions between RNAPII and Ssu72 in cells lacking Rtr1. These data may suggest that Ssu72 cannot directly compensate for Rtr1. Although Ssu72 and Rtr1 are both CTD phosphatases that remove Ser5-P from the RNAPII CTD, this study provides strong evidence that they play different roles in transcription termination. However, we cannot rule out that they have additional roles in targeting specific termination factors for dephosphorylation considering that a number of the termination factors, including Nrd1, are phosphorylated [100]. Numerous studies on Ssu72 and NNS-dependent terminators have identified mutants in Ssu72 that cause defects in sn/snoRNA and some mRNA termination and/or processing [18, 29, 45, 101, 102]. Total loss of Ssu72 using degron mediated protein degradation leads to more extensive transcription readthrough than is observed with specific point mutants with RNAPII accumulating at the 3’ ends of genes as measured by ChIP, which could imply that mRNA 3’ end processing and/or RNAPII termination could be defective in the absence of Ssu72 [14, 16]. Surprisingly, deletion of RTR1 has many phenotypic impacts that appear to be opposite of those seen in mutants of Ssu72 mutants or degron strains. For instance, RNAPII levels at the 5’ end of protein coding genes in rtr1Δ are slightly increased with decreased levels at the 3’ end suggesting either more efficient or premature transcription termination (Fig 5). At the transcript level, sn/snoRNAs and their extended transcripts which in some cases may represent terminator readthrough events are decreased in rtr1Δ (Figs 2 & 4). The IMD2 terminator is readthrough in many Ssu72 mutants, including ssu72-TOV, leading to increased expression of IMD2 whereas basal expression of IMD2 mRNA in rtr1Δ is greatly diminished (Figs 2 & 6, [45, 101]). These seemingly paradoxical findings suggest that the role of the CTD phosphatases in the regulation of RNAPII elongation and termination remains enigmatic and that additional studies are needed to explore the crosstalk between Rtr1 and Ssu72.

Methods

Yeast strains

All yeast strains used are isogenic to BY4741. RTR1 was knocked out of wild type and the RRP6 deletion strain by homologous recombination with a kanamycin cassette to create the RTR1 deletion and RTR1/RRP6 double deletion strains. RRP6 deletion strain is from the yeast knockout collection (Open Biosystems) [103]. The Rpb3-3xFLAG (referred to as Rpb3-FLAG) WT strain has been previously described [57]. The Nrd1-TAP strain is from the yeast TAP-tag collection (Open Biosystems). The Rpb3-FLAG and Nrd1-TAP RTR1 deletion strains were made by amplification of the RTR1 knockout cassette from the RTR1 deletion strain and transformation into the wild-type Rpb3-FLAG and Nrd1-TAP strains respectively. All deletion strains were confirmed by PCR-based genotyping. To perform a single biological replicate for genomics or proteomics experiments, growths of all strains of interest were pre-cultured from a single colony obtained from a sequence-verified glycerol stock of the strain that had been plated on the appropriate selective medium and grown for 2 days. Liquid cultures of all genotypes for an individual biological experiment were grown up on the same day. Cells for subsequent biological replicates were grown on different days.

Affinity purification of protein complexes

Cells were grown to OD600 ≅ 1.5 in YPD broth overnight and collected by centrifugation for 10 minutes at 4000 x g, then washed in H2O and resuspended in 25mL TAP lysis buffer per 2.5 grams of pellet (40mM Hepes-KOH, pH 7.5; 10% glycerol; 100mM NaCl; 0.1% Tween-20; fresh yeast protease inhibitors (Sigma; diluted to 1X)). The cells were slowly transferred to liquid nitrogen using a syringe. The frozen cells were pulverized with a mortar and pestle and lysed further in a Waring Blender with dry ice. The frozen lysate was transferred to a new container and allowed to thaw at room temperature. The resulting extract was treated with 100units DNase I and 10μL of 30mg/mL heparin for 10 minutes at room temperature and clarified by centrifugation as previously described [73]. Tandem Affinity Purification (TAP) was performed as previously described [73]. For FLAG tagged purifications, the lysate was incubated with anti-FLAG agarose resin (Sigma) at 4°C overnight. The resin and bound proteins were removed from the lysate by gravity flow through a 30mL Bio-Rad Econoprep column and washed on the column with 60 mL TAP lysis buffer. The resin was resuspended 300μL of 50mM Ammonium bicarbonate pH 8.0 and transfer to a microcentrifuge tube for on bead digestion with 5μL of Trypsin Gold (0.1μg/μL) overnight with shaking at 37°C. The supernatant containing the digested proteins was removed and treated with 20μL of 90% formic acid to inactivate the trypsin.

MudPIT-LC/MS/MS and proteomics data analysis

Each affinity purified sample was loaded onto a two-phase MudPIT column containing strong cation exchange resin (Phenomenex), which binds positively charged ions, and reverse phase C18 resin (Phenomenex), which will retain peptides based on their hydrophobicity [104]. The samples were eluted off the column by the MudPIT protocol of 10 steps of increasing salt concentrations (50-350mM ammonium acetate) followed by an organic gradient (20–80% acetonitrile). All chromatography solutions also contained 1% formic acid. Peptides were analyzed by a ThermoFisher LTQ Velos for MS/MS analysis. Raw spectrum data from the MS analysis were submitted for protein identification by Proteome Discoverer software (Thermo) version 2.1 using SEQUEST-HT as the database search algorithm. Database searches were performed against a FASTA database from the yeast Uniprot proteome. The FASTA database also included a number of common protein contaminants such as keratins and IgGs.

Disruption–Compensation (DisCo) network analysis

DisCo analysis using protein-protein interaction tools was used to analyze protein-protein interaction dynamics as a consequence of genetic perturbation, in this case deletion of the CTD phosphatase RTR1. Statistical analysis of interactome (SAINT) was performed as previously described on at least four biological replicate purifications from each genotype [60, 68, 69]. In brief, PSMs for each copurified protein were annotated per purification by bait protein, genotype (WT or ‘rtr1D’ for rtr1Δ), replicate in the list format used for analysis through crapome.org [105]. SAINTexpress was used for the probability score calculation [69]. The output file from SAINT analysis was used as the input for ProHits-viz which was employed for prey-prey correlation analysis with the following key options: Abundance column = Spec (i.e. PSM), Score column = Saint score, Abundance cutoff for prey correlation = 20, Add bait counts = yes [106].

RNA Isolation

RNA was extracted using the hot acid phenol method described previously [57]. An Ambion DNase-turbo kit was used to degrade any contaminating DNA if the RNA was to be used for subsequent sequencing or PCR. The quality of the total RNA samples was determined with an Agilent Bioanalyzer before preparation of the sequencing libraries.

Illumina HiSeq 4000 sequencing methods

Illumina TruSeq total RNA standard methods were used for yeast whole transcriptome sequencing. Total RNA was isolated and DNase treated (Ambion DNase). RNA was evaluated for quantity and quality for a minimum RIN score of 7 or higher using Agilent Bioanalyzer 2100. RNA samples were spiked with ERCC ExFold RNA spike-in mix (Life Technologies, 4456739) prior to library preparation. Samples were depleted of Ribosomal RNA using Ribo-Zero Magnetic Gold Kit (Illumina, MRZY1324). cDNA libraries were prepared using RNA fragmentation, cDNA synthesis, ligation of index adaptors, and amplification as specified in TruSeq sample preparation guide (Illumina, 15031048). Total RNA was sequenced with the Illumina HiSeq 4000. Ggplot was used for volcano plot visualization [107].

ChIP-exo and MNase-seq

Chromatin IP followed by exonuclease treatment was performed using the protocol described by Rhee and Pugh with the following specifics [83]. Rpb3-FLAG WT and rtr1Δ and Nrd1-TAP WT and rtr1Δ cells were grown to an OD600 = 0.8–1 prior to crosslinking with formaldehyde. Immunoprecipitation was performed with 50μL of anti-FLAG agarose or anti-TAP sepharose beads (Sigma). The volume of beads used for immunoprecipitation was optimized by affinity purification followed by mass spectrometry [72, 108]. Subsequent sample processing steps including exonuclease treatment and sequencing library preparation were performed as previously described [83].

Micrococcal nuclease (MNase) digest and sequencing was performed through adaptation of the protocol by Wal and Pugh [109]. Following optimization of the digestion conditions, 15U of MNase was added to a chromatin slurry and incubated with shaking at 37°C for 20 minutes. The digestion was quenched by addition of 50 mM EDTA and 0.2% SDS. The digested DNA was cleaned up through phenol/chorloform extraction followed by ethanol precipitation with 20ug of glycogen (Sigma) as a carrier.

ChIP-exo and MNase library construction, EZBead preparation, and Next-Gen sequencing were completed using standard methods based on the Life Technologies SOLiD5500xl system as previously described [57].

Genomics data analysis

SOLiD reads were mapped to Saccharomyces cerevisiae sacCer3 reference genome using mapping pipelines that utilize bfast-0.7.0a [110]. Read counts per nucleotide were calculated using bamutils from NGSUtils [111]. The average gene analysis plots for different RNAPII gene classes were generated using data from two biological replicate experiments per genotype per plot with the program ngs.plot using data from bam files and further edited in Adobe Illustrator [90]. The plots include the standard error of the mean for the total number of genes (defined in the text and figure legends) used for average gene analysis calculated by ngs.plot. All raw and processed files from the ChIP and MNase sequencing performed for this study have been deposited to Gene Expression Omnibus [GEO] under the accession number GSE87657 and the RNA-Seq dataset has been uploaded as GSE135056.

Illumina reads were mapped to sacCer3 reference genome using STAR RNA-seq aligner [112]. Extended transcripts (ETs) were manually annotated and added to sacCer3 based on the change in read counts in RRP6 deletion cell RNA-Seq relative to WT past the 3’ end annotations from sacCer3 as previously described [57]. ET annotations were ended prior to any downstream gene annotations regardless of increases in read counts in RRP6 suggesting transcription readthrough the adjacent gene. Read count distribution across the genome for each nucleotide was assessed using bamutils from the NGSUtils package [111]. Following data alignment, noncoding transcripts were manually inspected individually using the Integrative Genomics Viewer [113]. To identify ASTs with significant changes in differential expression, the strand was reversed for all sense annotations for the coding region of each ORF-Ts and the text “AS_” was added in front of the ORF-T name. The annotations for the 5’ and 3’ UTR were not included. These annotations were then used for edgeR analysis and the annotations for ASTs that showed significant changes in rrp6Δ were used for subsequent differential expression analysis to generate the final dataset in S3 Table. Differential gene expression was analyzed using edgeR, which has been shown to work well with low replicate numbers [75, 114]. Four biological replicates were used for each genotype in the RNA-Seq analysis. All raw and processed files from the RNA sequencing performed for this study have been deposited to Gene Expression Omnibus [GEO] under the accession number GSE135056.

Northern Blot analysis

Northern blot analysis was performed as previously described [57]. Thirty micrograms of total RNA were loaded per lane on a 1% agarose gel and separated by electrophoresis at 120 volts for 1 hour at 4°C. The RNA was transferred to Bio-Rad Zeta-Probe® blotting membranes by capillary overnight. Transfer efficiency was determined by Methylene Blue staining. Strand specific RNA probes were expressed from a linearized pET-DEST42 (Invitrogen) containing the region of interest in the sense or antisense orientation by T7 transcription (MAXIscript) using 32P labeled UTP. The radiolabeled probe was purified and then hybridized to the RNA blot at 68˚C overnight. The membranes were then washed with 1xSSC/.1%SDS twice at room temperature and twice with .1xSSC/.1%SDS for 15 minutes at 68°C. Blots were exposed to a phosphorscreen followed by scanning using a phosphorimager (GE Healthcare).

Global proteome abundance analysis

WT and rtr1Δ cells were lysed in 8M urea 100mM Tris pH 8.5 for optimal protein extraction. Samples were digested using Trypsin Gold (Promega) and labeled with Tandem Mass Tag (TMT) reagents according to the manufacturer’s protocols (Thermo Fisher). TMT labeled peptide samples were combined for multiplexing, then subjected to high pH reversed-phase fractionation (8 fractions). The fractions were analyzed on an Orbitrap Fusion Lumos instrument using an SPS MS3 method and the data searched on Proteome Discoverer 2.3 using a yeast proteome downloaded from Uniprot in October 2017. The protein abundances were normalized using total peptide amounts per multiplexed channel.

Supporting information

S1 Table [xlsx]
Peptide–spectrum match (PSM) table for the termination factor purification dataset.

S2 Table [xls]
Statistical analysis of INTeractome (SAINT) dataset for the termination factor purifications.

S3 Table [fdr]
knockout (KO) cell transcriptome data relative to WT.

S4 Table [wt]
Global proteomics abundance dataset for WT and deletion cells.

S5 Table [xlsx]
Nrd1 ChIP-exo peak analysis from MACS.

S6 Table [fdr]
knockout (KO) and knockout cell transcriptome data relative to WT.

S1 Fig [tif]
STRING network analysis of termination factor complex data using a fold-change cutoff of 5 or more [].

S2 Fig [tif]
Global proteomics analysis of protein abundance changes in deletion vs.

S3 Fig [tss]
Enrichment of Nrd1 occupancy at protein coding genes.

S4 Fig [tif]
Average ChIP-exo occupancy profiles for RNAPII (Rpb3) and Nrd1 from WT and RTR1 knockout cells.

S5 Fig [a]
Volcano plots representing significant changes in the and knockout transcriptomes relative to WT.


Zdroje

1. Herzel L, Straube K, Neugebauer KM. Long-read sequencing of nascent RNA reveals coupling among RNA processing events. Genome Res. 2018;28(7):1008–19. Epub 2018/06/16. doi: 10.1101/gr.232025.117 29903723; PubMed Central PMCID: PMC6028129.

2. Peck SA, Hughes KD, Victorino JF, Mosley AL. Writing a wrong: Coupled RNA polymerase II transcription and RNA quality control. Wiley Interdiscip Rev RNA. 2019;10(4):e1529. Epub 2019/03/09. doi: 10.1002/wrna.1529 30848101; PubMed Central PMCID: PMC6570551.

3. Kamieniarz-Gdula K, Proudfoot NJ. Transcriptional Control by Premature Termination: A Forgotten Mechanism. Trends Genet. 2019. Epub 2019/06/20. doi: 10.1016/j.tig.2019.05.005 31213387.

4. Kamieniarz-Gdula K, Gdula MR, Panser K, Nojima T, Monks J, Wisniewski JR, et al. Selective Roles of Vertebrate PCF11 in Premature and Full-Length Transcript Termination. Mol Cell. 2019;74(1):158–72 e9. Epub 2019/03/02. doi: 10.1016/j.molcel.2019.01.027 30819644; PubMed Central PMCID: PMC6458999.

5. Schlackow M, Nojima T, Gomes T, Dhir A, Carmo-Fonseca M, Proudfoot NJ. Distinctive Patterns of Transcription and RNA Processing for Human lincRNAs. Mol Cell. 2017;65(1):25–38. Epub 2016/12/27. doi: 10.1016/j.molcel.2016.11.029 28017589; PubMed Central PMCID: PMC5222723.

6. Dhir A, Dhir S, Proudfoot NJ, Jopling CL. Microprocessor mediates transcriptional termination of long noncoding RNA transcripts hosting microRNAs. Nat Struct Mol Biol. 2015;22(4):319–27. Epub 2015/03/03. doi: 10.1038/nsmb.2982 25730776; PubMed Central PMCID: PMC4492989.

7. Wagschal A, Rousset E, Basavarajaiah P, Contreras X, Harwig A, Laurent-Chabalier S, et al. Microprocessor, Setx, Xrn2, and Rrp6 co-operate to induce premature termination of transcription by RNAPII. Cell. 2012;150(6):1147–57. Epub 2012/09/18. doi: 10.1016/j.cell.2012.08.004 22980978; PubMed Central PMCID: PMC3595997.

8. Fusby B, Kim S, Erickson B, Kim H, Peterson ML, Bentley DL. Coordination of RNA Polymerase II Pausing and 3' End Processing Factor Recruitment with Alternative Polyadenylation. Molecular and cellular biology. 2016;36(2):295–303. Epub 2015/11/04. doi: 10.1128/MCB.00898-15 26527620; PubMed Central PMCID: PMC4719304.

9. Brody Y, Neufeld N, Bieberstein N, Causse SZ, Bohnlein EM, Neugebauer KM, et al. The in vivo kinetics of RNA polymerase II elongation during co-transcriptional splicing. PLoS Biol. 2011;9(1):e1000573. Epub 2011/01/26. doi: 10.1371/journal.pbio.1000573 21264352; PubMed Central PMCID: PMC3019111.

10. Liu X, Freitas J, Zheng D, Oliveira MS, Hoque M, Martins T, et al. Transcription elongation rate has a tissue-specific impact on alternative cleavage and polyadenylation in Drosophila melanogaster. Rna. 2017;23(12):1807–16. Epub 2017/08/31. doi: 10.1261/rna.062661.117 28851752; PubMed Central PMCID: PMC5689002.

11. Corden JL. Tails of RNA polymerase II. Trends Biochem Sci. 1990;15(10):383–7. Epub 1990/10/01. doi: 10.1016/0968-0004(90)90236-5 2251729.

12. Buratowski S. Progression through the RNA polymerase II CTD cycle. Mol Cell. 2009;36(4):541–6. Epub 2009/11/28. doi: 10.1016/j.molcel.2009.10.019 19941815; PubMed Central PMCID: PMC3232742.

13. Mosley AL, Pattenden SG, Carey M, Venkatesh S, Gilmore JM, Florens L, et al. Rtr1 is a CTD phosphatase that regulates RNA polymerase II during the transition from serine 5 to serine 2 phosphorylation. Mol Cell. 2009;34(2):168–78. doi: 10.1016/j.molcel.2009.02.025 19394294; PubMed Central PMCID: PMC2996052.

14. Zhang DW, Mosley AL, Ramisetty SR, Rodriguez-Molina JB, Washburn MP, Ansari AZ. Ssu72 phosphatase-dependent erasure of phospho-Ser7 marks on the RNA polymerase II C-terminal domain is essential for viability and transcription termination. J Biol Chem. 2012;287(11):8541–51. Epub 2012/01/12. doi: 10.1074/jbc.M111.335687 22235117; PubMed Central PMCID: PMC3318730.

15. Jeronimo C, Collin P, Robert F. The RNA Polymerase II CTD: The Increasing Complexity of a Low-Complexity Protein Domain. J Mol Biol. 2016;428(12):2607–22. Epub 2016/02/16. doi: 10.1016/j.jmb.2016.02.006 26876604.

16. Bataille AR, Jeronimo C, Jacques PE, Laramee L, Fortin ME, Forest A, et al. A universal RNA polymerase II CTD cycle is orchestrated by complex interplays between kinase, phosphatase, and isomerase enzymes along genes. Mol Cell. 2012;45(2):158–70. Epub 2012/01/31. doi: 10.1016/j.molcel.2011.11.024 22284676.

17. Schreieck A, Easter AD, Etzold S, Wiederhold K, Lidschreiber M, Cramer P, et al. RNA polymerase II termination involves C-terminal-domain tyrosine dephosphorylation by CPF subunit Glc7. Nat Struct Mol Biol. 2014;21(2):175–9. Epub 2014/01/15. doi: 10.1038/nsmb.2753 24413056; PubMed Central PMCID: PMC3917824.

18. Nedea E, He X, Kim M, Pootoolal J, Zhong G, Canadien V, et al. Organization and function of APT, a subcomplex of the yeast cleavage and polyadenylation factor involved in the formation of mRNA and small nucleolar RNA 3'-ends. J Biol Chem. 2003;278(35):33000–10. Epub 2003/06/24. doi: 10.1074/jbc.M304454200 12819204.

19. Kong SE, Kobor MS, Krogan NJ, Somesh BP, Sogaard TM, Greenblatt JF, et al. Interaction of Fcp1 phosphatase with elongating RNA polymerase II holoenzyme, enzymatic mechanism of action, and genetic interaction with elongator. J Biol Chem. 2005;280(6):4299–306. Epub 2004/11/26. doi: 10.1074/jbc.M411071200 15563457.

20. Kobor MS, Archambault J, Lester W, Holstege FC, Gileadi O, Jansma DB, et al. An unusual eukaryotic protein phosphatase required for transcription by RNA polymerase II and CTD dephosphorylation in S. cerevisiae. Mol Cell. 1999;4(1):55–62. Epub 1999/08/13. doi: 10.1016/s1097-2765(00)80187-2 10445027.

21. Schwer B, Ghosh A, Sanchez AM, Lima CD, Shuman S. Genetic and structural analysis of the essential fission yeast RNA polymerase II CTD phosphatase Fcp1. Rna. 2015;21(6):1135–46. Epub 2015/04/18. doi: 10.1261/rna.050286.115 25883047; PubMed Central PMCID: PMC4436666.

22. Ghosh A, Shuman S, Lima CD. The structure of Fcp1, an essential RNA polymerase II CTD phosphatase. Mol Cell. 2008;32(4):478–90. Epub 2008/11/26. doi: 10.1016/j.molcel.2008.09.021 19026779; PubMed Central PMCID: PMC2645342.

23. Hsu PL, Yang F, Smith-Kinnaman W, Yang W, Song JE, Mosley AL, et al. Rtr1 is a dual specificity phosphatase that dephosphorylates Tyr1 and Ser5 on the RNA polymerase II CTD. J Mol Biol. 2014;426(16):2970–81. Epub 2014/06/22. doi: 10.1016/j.jmb.2014.06.010 24951832; PubMed Central PMCID: PMC4119023.

24. Krishnamurthy S, He X, Reyes-Reyes M, Moore C, Hampsey M. Ssu72 Is an RNA polymerase II CTD phosphatase. Mol Cell. 2004;14(3):387–94. Epub 2004/05/06. doi: 10.1016/s1097-2765(04)00235-7 15125841.

25. Meinhart A, Silberzahn T, Cramer P. The mRNA transcription/processing factor Ssu72 is a potential tyrosine phosphatase. J Biol Chem. 2003;278(18):15917–21. Epub 2003/02/28. doi: 10.1074/jbc.M301643200 12606538.

26. Hunter GO, Fox MJ, Smith-Kinnaman WR, Gogol M, Fleharty B, Mosley AL. The phosphatase Rtr1 regulates global levels of serine 5 RNA Polymerase II C-terminal domain phosphorylation and cotranscriptional histone methylation. Molecular and cellular biology. 2016. doi: 10.1128/MCB.00870-15 27247267.

27. Lunde BM, Reichow SL, Kim M, Suh H, Leeper TC, Yang F, et al. Cooperative interaction of transcription termination factors with the RNA polymerase II C-terminal domain. Nat Struct Mol Biol. 2010;17(10):1195–201. Epub 2010/09/08. doi: 10.1038/nsmb.1893 20818393; PubMed Central PMCID: PMC2950884.

28. Vasiljeva L, Kim M, Mutschler H, Buratowski S, Meinhart A. The Nrd1-Nab3-Sen1 termination complex interacts with the Ser5-phosphorylated RNA polymerase II C-terminal domain. Nat Struct Mol Biol. 2008;15(8):795–804. Epub 2008/07/29. doi: 10.1038/nsmb.1468 18660819; PubMed Central PMCID: PMC2597375.

29. Kim M, Vasiljeva L, Rando OJ, Zhelkovsky A, Moore C, Buratowski S. Distinct pathways for snoRNA and mRNA termination. Mol Cell. 2006;24(5):723–34. Epub 2006/12/13. doi: 10.1016/j.molcel.2006.11.011 17157255.

30. Vasiljeva L, Buratowski S. Nrd1 interacts with the nuclear exosome for 3' processing of RNA polymerase II transcripts. Mol Cell. 2006;21(2):239–48. Epub 2006/01/24. doi: 10.1016/j.molcel.2005.11.028 16427013.

31. Creamer TJ, Darby MM, Jamonnak N, Schaughency P, Hao H, Wheelan SJ, et al. Transcriptome-wide binding sites for components of the Saccharomyces cerevisiae non-poly(A) termination pathway: Nrd1, Nab3, and Sen1. PLoS Genet. 2011;7(10):e1002329. Epub 2011/10/27. doi: 10.1371/journal.pgen.1002329 22028667; PubMed Central PMCID: PMC3197677.

32. Jamonnak N, Creamer TJ, Darby MM, Schaughency P, Wheelan SJ, Corden JL. Yeast Nrd1, Nab3, and Sen1 transcriptome-wide binding maps suggest multiple roles in post-transcriptional RNA processing. Rna. 2011;17(11):2011–25. Epub 2011/09/29. doi: 10.1261/rna.2840711 21954178; PubMed Central PMCID: PMC3198594.

33. Hobor F, Pergoli R, Kubicek K, Hrossova D, Bacikova V, Zimmermann M, et al. Recognition of transcription termination signal by the nuclear polyadenylated RNA-binding (NAB) 3 protein. J Biol Chem. 2011;286(5):3645–57. Epub 2010/11/19. doi: 10.1074/jbc.M110.158774 21084293; PubMed Central PMCID: PMC3030368.

34. Carroll KL, Ghirlando R, Ames JM, Corden JL. Interaction of yeast RNA-binding proteins Nrd1 and Nab3 with RNA polymerase II terminator elements. Rna. 2007;13(3):361–73. Epub 2007/01/24. doi: 10.1261/rna.338407 17237360; PubMed Central PMCID: PMC1800511.

35. Loya TJ, O'Rourke TW, Degtyareva N, Reines D. A network of interdependent molecular interactions describes a higher order Nrd1-Nab3 complex involved in yeast transcription termination. J Biol Chem. 2013. Epub 2013/10/09. doi: 10.1074/jbc.M113.516765 24100036.

36. Loya TJ, O'Rourke TW, Reines D. The hnRNP-like Nab3 termination factor can employ heterologous prion-like domains in place of its own essential low complexity domain. PLoS One. 2017;12(10):e0186187. Epub 2017/10/13. doi: 10.1371/journal.pone.0186187 29023495; PubMed Central PMCID: PMC5638401.

37. Loya TJ, O'Rourke TW, Reines D. Yeast Nab3 protein contains a self-assembly domain found in human heterogeneous nuclear ribonucleoprotein-C (hnRNP-C) that is necessary for transcription termination. J Biol Chem. 2013;288(4):2111–7. Epub 2012/11/30. doi: 10.1074/jbc.M112.430678 23192344; PubMed Central PMCID: PMC3554884.

38. O'Rourke TW, Reines D. Determinants of Amyloid Formation for the Yeast Termination Factor Nab3. PLoS One. 2016;11(3):e0150865. Epub 2016/03/10. doi: 10.1371/journal.pone.0150865 26954508; PubMed Central PMCID: PMC4783047.

39. Wlotzka W, Kudla G, Granneman S, Tollervey D. The nuclear RNA polymerase II surveillance system targets polymerase III transcripts. Embo J. 2011;30(9):1790–803. Epub 2011/04/05. doi: 10.1038/emboj.2011.97 21460797; PubMed Central PMCID: PMC3102002.

40. Bacikova V, Pasulka J, Kubicek K, Stefl R. Structure and semi-sequence-specific RNA binding of Nrd1. Nucleic Acids Res. 2014;42(12):8024–38. Epub 2014/05/27. doi: 10.1093/nar/gku446 24860164; PubMed Central PMCID: PMC4081072.

41. Tudek A, Porrua O, Kabzinski T, Lidschreiber M, Kubicek K, Fortova A, et al. Molecular basis for coordinating transcription termination with noncoding RNA degradation. Mol Cell. 2014;55(3):467–81. Epub 2014/07/30. doi: 10.1016/j.molcel.2014.05.031 25066235; PubMed Central PMCID: PMC4186968.

42. Kim K, Heo DH, Kim I, Suh JY, Kim M. Exosome Cofactors Connect Transcription Termination to RNA Processing by Guiding Terminated Transcripts to the Appropriate Exonuclease within the Nuclear Exosome. J Biol Chem. 2016;291(25):13229–42. Epub 2016/04/15. doi: 10.1074/jbc.M116.715771 27076633; PubMed Central PMCID: PMC4933236.

43. Zhang Y, Chun Y, Buratowski S, Tong L. Identification of Three Sequence Motifs in the Transcription Termination Factor Sen1 that Mediate Direct Interactions with Nrd1. Structure. 2019;27(7):1156–61 e4. Epub 2019/05/21. doi: 10.1016/j.str.2019.04.005 31104813; PubMed Central PMCID: PMC6610696.

44. Arigo JT, Carroll KL, Ames JM, Corden JL. Regulation of yeast NRD1 expression by premature transcription termination. Mol Cell. 2006;21(5):641–51. Epub 2006/03/02. doi: 10.1016/j.molcel.2006.02.005 16507362.

45. Loya TJ, O'Rourke TW, Reines D. A genetic screen for terminator function in yeast identifies a role for a new functional domain in termination factor Nab3. Nucleic Acids Res. 2012;40(15):7476–91. Epub 2012/05/09. doi: 10.1093/nar/gks377 22564898; PubMed Central PMCID: PMC3424548.

46. Jenks MH, O'Rourke TW, Reines D. Properties of an intergenic terminator and start site switch that regulate IMD2 transcription in yeast. Molecular and cellular biology. 2008;28(12):3883–93. Epub 2008/04/23. doi: 10.1128/MCB.00380-08 18426909; PubMed Central PMCID: PMC2423123.

47. Kopcewicz KA, O'Rourke TW, Reines D. Metabolic regulation of IMD2 transcription and an unusual DNA element that generates short transcripts. Molecular and cellular biology. 2007;27(8):2821–9. Epub 2007/02/14. doi: 10.1128/MCB.02159-06 17296737; PubMed Central PMCID: PMC1899919.

48. Grzechnik P, Gdula MR, Proudfoot NJ. Pcf11 orchestrates transcription termination pathways in yeast. Genes Dev. 2015;29(8):849–61. Epub 2015/04/17. doi: 10.1101/gad.251470.114 25877920; PubMed Central PMCID: PMC4403260.

49. Mayer A, Heidemann M, Lidschreiber M, Schreieck A, Sun M, Hintermair C, et al. CTD tyrosine phosphorylation impairs termination factor recruitment to RNA polymerase II. Science. 2012;336(6089):1723–5. Epub 2012/06/30. doi: 10.1126/science.1219651 22745433.

50. Kim M, Krogan NJ, Vasiljeva L, Rando OJ, Nedea E, Greenblatt JF, et al. The yeast Rat1 exonuclease promotes transcription termination by RNA polymerase II. Nature. 2004;432(7016):517–22. Epub 2004/11/27. doi: 10.1038/nature03041 15565157.

51. Fong N, Brannan K, Erickson B, Kim H, Cortazar MA, Sheridan RM, et al. Effects of Transcription Elongation Rate and Xrn2 Exonuclease Activity on RNA Polymerase II Termination Suggest Widespread Kinetic Competition. Mol Cell. 2015;60(2):256–67. Epub 2015/10/17. doi: 10.1016/j.molcel.2015.09.026 26474067; PubMed Central PMCID: PMC4654110.

52. Pearson EL, Moore CL. Dismantling promoter-driven RNA polymerase II transcription complexes in vitro by the termination factor Rat1. J Biol Chem. 2013;288(27):19750–9. Epub 2013/05/22. doi: 10.1074/jbc.M112.434985 23689372; PubMed Central PMCID: PMC3707679.

53. Baejen C, Andreani J, Torkler P, Battaglia S, Schwalb B, Lidschreiber M, et al. Genome-wide Analysis of RNA Polymerase II Termination at Protein-Coding Genes. Mol Cell. 2017;66(1):38–49 e6. Epub 2017/03/21. doi: 10.1016/j.molcel.2017.02.009 28318822.

54. Park J, Kang M, Kim M. Unraveling the mechanistic features of RNA polymerase II termination by the 5'-3' exoribonuclease Rat1. Nucleic Acids Res. 2015;43(5):2625–37. Epub 2015/02/28. doi: 10.1093/nar/gkv133 25722373; PubMed Central PMCID: PMC4357727.

55. Proudfoot NJ. Transcriptional termination in mammals: Stopping the RNA polymerase II juggernaut. Science. 2016;352(6291):aad9926. Epub 2016/06/11. doi: 10.1126/science.aad9926 27284201; PubMed Central PMCID: PMC5144996.

56. Eaton JD, West S. An end in sight? Xrn2 and transcriptional termination by RNA polymerase II. Transcription. 2018;9(5):321–6. Epub 2018/07/24. doi: 10.1080/21541264.2018.1498708 30035655; PubMed Central PMCID: PMC6150625.

57. Fox MJ, Gao H, Smith-Kinnaman WR, Liu Y, Mosley AL. The exosome component Rrp6 is required for RNA polymerase II termination at specific targets of the Nrd1-Nab3 pathway. PLoS Genet. 2015;11(2):e1004999. Epub 2015/02/14. doi: 10.1371/journal.pgen.1004999 25680078; PubMed Central PMCID: PMC4378619.

58. Lemay JF, Bachand F. Fail-safe transcription termination: Because one is never enough. RNA Biol. 2015;12(9):927–32. Epub 2015/08/15. doi: 10.1080/15476286.2015.1073433 26273910; PubMed Central PMCID: PMC4615224.

59. Lemay JF, Larochelle M, Marguerat S, Atkinson S, Bahler J, Bachand F. The RNA exosome promotes transcription termination of backtracked RNA polymerase II. Nat Struct Mol Biol. 2014;21(10):919–26. Epub 2014/09/23. doi: 10.1038/nsmb.2893 25240800.

60. Smith-Kinnaman WR, Berna MJ, Hunter GO, True JD, Hsu P, Cabello GI, et al. The interactome of the atypical phosphatase Rtr1 in Saccharomyces cerevisiae. Mol Biosyst. 2014;10(7):1730–41. Epub 2014/03/29. doi: 10.1039/c4mb00109e 24671508; PubMed Central PMCID: PMC4074173.

61. Gibney PA, Fries T, Bailer SM, Morano KA. Rtr1 is the Saccharomyces cerevisiae homolog of a novel family of RNA polymerase II-binding proteins. Eukaryot Cell. 2008;7(6):938–48. Epub 2008/04/15. doi: 10.1128/EC.00042-08 18408053; PubMed Central PMCID: PMC2446653.

62. Gudipati RK, Villa T, Boulay J, Libri D. Phosphorylation of the RNA polymerase II C-terminal domain dictates transcription termination choice. Nat Struct Mol Biol. 2008;15(8):786–94. Epub 2008/07/29. doi: 10.1038/nsmb.1460 18660821.

63. Vasiljeva L, Kim M, Mutschler H, Buratowski S, Meinhart A. The Nrd1-Nab3-Sen1 termination complex interacts with the Ser5-phosphorylated RNA polymerase II C-terminal domain. Nature structural & molecular biology. 2008;15(8):795–804. Epub 2008/07/29. doi: 10.1038/nsmb.1468 18660819; PubMed Central PMCID: PMC2597375.

64. Arndt KM, Reines D. Termination of Transcription of Short Noncoding RNAs by RNA Polymerase II. Annual review of biochemistry. 2015;84:381–404. doi: 10.1146/annurev-biochem-060614-034457 25747400.

65. Porrua O, Libri D. Transcription termination and the control of the transcriptome: why, where and how to stop. Nature reviews Molecular cell biology. 2015;16(3):190–202. doi: 10.1038/nrm3943 25650800.

66. Sahni N, Yi S, Zhong Q, Jailkhani N, Charloteaux B, Cusick ME, et al. Edgotype: a fundamental link between genotype and phenotype. Curr Opin Genet Dev. 2013;23(6):649–57. Epub 2013/11/30. doi: 10.1016/j.gde.2013.11.002 24287335; PubMed Central PMCID: PMC3902775.

67. Charloteaux B, Zhong Q, Dreze M, Cusick ME, Hill DE, Vidal M. Protein-protein interactions and networks: forward and reverse edgetics. Methods Mol Biol. 2011;759:197–213. Epub 2011/08/25. doi: 10.1007/978-1-61779-173-4_12 21863489.

68. Choi H, Larsen B, Lin ZY, Breitkreutz A, Mellacheruvu D, Fermin D, et al. SAINT: probabilistic scoring of affinity purification-mass spectrometry data. Nat Methods. 2011;8(1):70–3. Epub 2010/12/07. doi: 10.1038/nmeth.1541 nmeth.1541 [pii]. 21131968; PubMed Central PMCID: PMC3064265.

69. Teo G, Liu G, Zhang J, Nesvizhskii AI, Gingras AC, Choi H. SAINTexpress: improvements and additional features in Significance Analysis of INTeractome software. J Proteomics. 2014;100:37–43. Epub 2014/02/12. doi: 10.1016/j.jprot.2013.10.023 S1874-3919(13)00538-1 [pii]. 24513533; PubMed Central PMCID: PMC4102138.

70. Szklarczyk D, Gable AL, Lyon D, Junge A, Wyder S, Huerta-Cepas J, et al. STRING v11: protein-protein association networks with increased coverage, supporting functional discovery in genome-wide experimental datasets. Nucleic Acids Res. 2019;47(D1):D607–D13. Epub 2018/11/27. doi: 10.1093/nar/gky1131 30476243; PubMed Central PMCID: PMC6323986.

71. Casanal A, Kumar A, Hill CH, Easter AD, Emsley P, Degliesposti G, et al. Architecture of eukaryotic mRNA 3'-end processing machinery. Science. 2017;358(6366):1056–9. Epub 2017/10/28. doi: 10.1126/science.aao6535 29074584; PubMed Central PMCID: PMC5788269.

72. Bedard LG, Dronamraju R, Kerschner JL, Hunter GO, Axley ED, Boyd AK, et al. Quantitative Analysis of Dynamic Protein Interactions during Transcription Reveals a Role for Casein Kinase II in Polymerase-associated Factor (PAF) Complex Phosphorylation and Regulation of Histone H2B Monoubiquitylation. J Biol Chem. 2016;291(26):13410–20. Epub 2016/05/05. doi: 10.1074/jbc.M116.727735 27143358; PubMed Central PMCID: PMC4919428.

73. Mosley AL, Sardiu ME, Pattenden SG, Workman JL, Florens L, Washburn MP. Highly reproducible label free quantitative proteomic analysis of RNA polymerase complexes. Mol Cell Proteomics. 2011;10(2):M110 000687. Epub 2010/11/05. doi: 10.1074/mcp.M110.000687 M110.000687 [pii]. 21048197; PubMed Central PMCID: PMC3033667.

74. Jiang L, Schlesinger F, Davis CA, Zhang Y, Li R, Salit M, et al. Synthetic spike-in standards for RNA-seq experiments. Genome Res. 2011;21(9):1543–51. Epub 2011/08/06. doi: 10.1101/gr.121095.111 21816910; PubMed Central PMCID: PMC3166838.

75. Robinson MD, McCarthy DJ, Smyth GK. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics. 2010;26(1):139–40. Epub 2009/11/17. doi: 10.1093/bioinformatics/btp616 19910308; PubMed Central PMCID: PMC2796818.

76. Neil H, Malabat C, d'Aubenton-Carafa Y, Xu Z, Steinmetz LM, Jacquier A. Widespread bidirectional promoters are the major source of cryptic transcripts in yeast. Nature. 2009;457(7232):1038–42. Epub 2009/01/27. doi: 10.1038/nature07747 19169244.

77. Schulz D, Schwalb B, Kiesel A, Baejen C, Torkler P, Gagneur J, et al. Transcriptome surveillance by selective termination of noncoding RNA synthesis. Cell. 2013;155(5):1075–87. Epub 2013/11/12. doi: 10.1016/j.cell.2013.10.024 24210918.

78. Merran J, Corden JL. Yeast RNA-Binding Protein Nab3 Regulates Genes Involved in Nitrogen Metabolism. Molecular and cellular biology. 2017;37(18). Epub 2017/07/05. doi: 10.1128/MCB.00154-17 28674185; PubMed Central PMCID: PMC5574042.

79. Kuehner JN, Brow DA. Regulation of a eukaryotic gene by GTP-dependent start site selection and transcription attenuation. Mol Cell. 2008;31(2):201–11. Epub 2008/07/29. doi: 10.1016/j.molcel.2008.05.018 18657503.

80. Steinmetz EJ, Warren CL, Kuehner JN, Panbehi B, Ansari AZ, Brow DA. Genome-wide distribution of yeast RNA polymerase II and its control by Sen1 helicase. Mol Cell. 2006;24(5):735–46. Epub 2006/12/13. doi: 10.1016/j.molcel.2006.10.023 17157256.

81. Hyle JW, Shaw RJ, Reines D. Functional distinctions between IMP dehydrogenase genes in providing mycophenolate resistance and guanine prototrophy to yeast. J Biol Chem. 2003;278(31):28470–8. Epub 2003/05/15. doi: 10.1074/jbc.M303736200 12746440; PubMed Central PMCID: PMC3367515.

82. Arigo JT, Eyler DE, Carroll KL, Corden JL. Termination of cryptic unstable transcripts is directed by yeast RNA-binding proteins Nrd1 and Nab3. Mol Cell. 2006;23(6):841–51. Epub 2006/09/16. doi: 10.1016/j.molcel.2006.07.024 16973436.

83. Rhee HS, Pugh BF. ChIP-exo method for identifying genomic location of DNA-binding proteins with near-single-nucleotide accuracy. Curr Protoc Mol Biol. 2012;Chapter 21:Unit 21 4. Epub 2012/10/03. doi: 10.1002/0471142727.mb2124s100 23026909; PubMed Central PMCID: PMC3813302.

84. Thiebaut M, Colin J, Neil H, Jacquier A, Seraphin B, Lacroute F, et al. Futile cycle of transcription initiation and termination modulates the response to nucleotide shortage in S. cerevisiae. Mol Cell. 2008;31(5):671–82. Epub 2008/09/09. doi: 10.1016/j.molcel.2008.08.010 18775327.

85. Hazelbaker DZ, Marquardt S, Wlotzka W, Buratowski S. Kinetic competition between RNA Polymerase II and Sen1-dependent transcription termination. Mol Cell. 2013;49(1):55–66. Epub 2012/11/28. doi: 10.1016/j.molcel.2012.10.014 23177741; PubMed Central PMCID: PMC3545030.

86. Vasiljeva L, Kim M, Terzi N, Soares LM, Buratowski S. Transcription termination and RNA degradation contribute to silencing of RNA polymerase II transcription within heterochromatin. Mol Cell. 2008;29(3):313–23. Epub 2008/02/19. doi: 10.1016/j.molcel.2008.01.011 18280237.

87. Fasken MB, Laribee RN, Corbett AH. Nab3 facilitates the function of the TRAMP complex in RNA processing via recruitment of Rrp6 independent of Nrd1. PLoS Genet. 2015;11(3):e1005044. Epub 2015/03/17. doi: 10.1371/journal.pgen.1005044 25775092; PubMed Central PMCID: PMC4361618.

88. Lemay JF, D'Amours A, Lemieux C, Lackner DH, St-Sauveur VG, Bahler J, et al. The nuclear poly(A)-binding protein interacts with the exosome to promote synthesis of noncoding small nucleolar RNAs. Mol Cell. 2010;37(1):34–45. Epub 2010/02/05. doi: 10.1016/j.molcel.2009.12.019 20129053.

89. LaCava J, Houseley J, Saveanu C, Petfalski E, Thompson E, Jacquier A, et al. RNA degradation by the exosome is promoted by a nuclear polyadenylation complex. Cell. 2005;121(5):713–24. Epub 2005/06/07. doi: 10.1016/j.cell.2005.04.029 15935758.

90. Shen L, Shao N, Liu X, Nestler E. ngs.plot: Quick mining and visualization of next-generation sequencing data by integrating genomic databases. BMC genomics. 2014;15:284. doi: 10.1186/1471-2164-15-284 24735413; PubMed Central PMCID: PMC4028082.

91. Nagalakshmi U, Wang Z, Waern K, Shou C, Raha D, Gerstein M, et al. The transcriptional landscape of the yeast genome defined by RNA sequencing. Science. 2008;320(5881):1344–9. Epub 2008/05/03. doi: 10.1126/science.1158441 18451266; PubMed Central PMCID: PMC2951732.

92. Webb S, Hector RD, Kudla G, Granneman S. PAR-CLIP data indicate that Nrd1-Nab3-dependent transcription termination regulates expression of hundreds of protein coding genes in yeast. Genome Biol. 2014;15(1):R8. Epub 2014/01/08. doi: 10.1186/gb-2014-15-1-r8 24393166; PubMed Central PMCID: PMC4053934.

93. Honorine R, Mosrin-Huaman C, Hervouet-Coste N, Libri D, Rahmouni AR. Nuclear mRNA quality control in yeast is mediated by Nrd1 co-transcriptional recruitment, as revealed by the targeting of Rho-induced aberrant transcripts. Nucleic Acids Res. 2011;39(7):2809–20. Epub 2010/11/30. doi: 10.1093/nar/gkq1192 21113025; PubMed Central PMCID: PMC3074134.

94. Marquardt S, Hazelbaker DZ, Buratowski S. Distinct RNA degradation pathways and 3' extensions of yeast non-coding RNA species. Transcription. 2011;2(3):145–54. Epub 2011/08/10. doi: 10.4161/trns.2.3.16298 21826286; PubMed Central PMCID: PMC3149692.

95. Gudipati RK, Xu Z, Lebreton A, Seraphin B, Steinmetz LM, Jacquier A, et al. Extensive degradation of RNA precursors by the exosome in wild-type cells. Mol Cell. 2012;48(3):409–21. Epub 2012/09/25. doi: 10.1016/j.molcel.2012.08.018 23000176; PubMed Central PMCID: PMC3496076.

96. Davis CA, Ares M Jr., Accumulation of unstable promoter-associated transcripts upon loss of the nuclear exosome subunit Rrp6p in Saccharomyces cerevisiae. Proc Natl Acad Sci U S A. 2006;103(9):3262–7. Epub 2006/02/18. doi: 10.1073/pnas.0507783103 16484372; PubMed Central PMCID: PMC1413877.

97. Castelnuovo M, Rahman S, Guffanti E, Infantino V, Stutz F, Zenklusen D. Bimodal expression of PHO84 is modulated by early termination of antisense transcription. Nat Struct Mol Biol. 2013;20(7):851–8. Epub 2013/06/19. doi: 10.1038/nsmb.2598 23770821.

98. Mostovoy Y, Thiemicke A, Hsu TY, Brem RB. The Role of Transcription Factors at Antisense-Expressing Gene Pairs in Yeast. Genome Biol Evol. 2016;8(6):1748–61. Epub 2016/05/18. doi: 10.1093/gbe/evw104 27190003; PubMed Central PMCID: PMC4943177.

99. Cakiroglu SA, Zaugg JB, Luscombe NM. Backmasking in the yeast genome: encoding overlapping information for protein-coding and RNA degradation. Nucleic Acids Res. 2016;44(17):8065–72. Epub 2016/08/06. doi: 10.1093/nar/gkw683 27492286; PubMed Central PMCID: PMC5041482.

100. Conrad NK, Wilson SM, Steinmetz EJ, Patturajan M, Brow DA, Swanson MS, et al. A yeast heterogeneous nuclear ribonucleoprotein complex associated with RNA polymerase II. Genetics. 2000;154(2):557–71. Epub 2000/02/03. 10655211; PubMed Central PMCID: PMC1460961.

101. Chen X, Poorey K, Carver MN, Muller U, Bekiranov S, Auble DT, et al. Transcriptomes of six mutants in the Sen1 pathway reveal combinatorial control of transcription termination across the Saccharomyces cerevisiae genome. PLoS Genet. 2017;13(6):e1006863. Epub 2017/07/01. doi: 10.1371/journal.pgen.1006863 28665995; PubMed Central PMCID: PMC5513554.

102. Ganem C, Devaux F, Torchet C, Jacq C, Quevillon-Cheruel S, Labesse G, et al. Ssu72 is a phosphatase essential for transcription termination of snoRNAs and specific mRNAs in yeast. Embo J. 2003;22(7):1588–98. Epub 2003/03/28. doi: 10.1093/emboj/cdg141 12660165; PubMed Central PMCID: PMC152886.

103. Winzeler EA, Shoemaker DD, Astromoff A, Liang H, Anderson K, Andre B, et al. Functional characterization of the S. cerevisiae genome by gene deletion and parallel analysis. Science. 1999;285(5429):901–6. Epub 1999/08/07. doi: 10.1126/science.285.5429.901 10436161.

104. Washburn MP, Wolters D, Yates JR 3rd. Large-scale analysis of the yeast proteome by multidimensional protein identification technology. Nat Biotechnol. 2001;19(3):242–7. Epub 2001/03/07. doi: 10.1038/85686 85686 [pii]. 11231557.

105. Mellacheruvu D, Wright Z, Couzens AL, Lambert JP, St-Denis NA, Li T, et al. The CRAPome: a contaminant repository for affinity purification-mass spectrometry data. Nat Methods. 2013;10(8):730–6. Epub 2013/08/08. doi: 10.1038/nmeth.2557 23921808; PubMed Central PMCID: PMC3773500.

106. Knight JDR, Choi H, Gupta GD, Pelletier L, Raught B, Nesvizhskii AI, et al. ProHits-viz: a suite of web tools for visualizing interaction proteomics data. Nat Methods. 2017;14(7):645–6. Epub 2017/07/01. doi: 10.1038/nmeth.4330 28661499; PubMed Central PMCID: PMC5831326.

107. Wickham H. ggplot2—Elegant Graphics for Data Analysis: Springer-Verlag, New York.; 2016. 260 p.

108. Burriss KH, Mosley AL. Methods review: Mass spectrometry analysis of RNAPII complexes. Methods. 2019;159–160:105–14. Epub 2019/03/25. doi: 10.1016/j.ymeth.2019.03.013 30902665.

109. Wal M, Pugh BF. Genome-wide mapping of nucleosome positions in yeast using high-resolution MNase ChIP-Seq. Methods Enzymol. 2012;513:233–50. Epub 2012/08/30. doi: 10.1016/B978-0-12-391938-0.00010-0 22929772; PubMed Central PMCID: PMC4871120.

110. Homer NM, Barry; Nelson, Stanley F. BFAST: an alignment tool for large scale genome resequencing. PLoS One. 2009;4(11):e7767. doi: 10.1371/journal.pone.0007767 19907642

111. Breese MR, Liu Y. NGSUtils: a software suite for analyzing and manipulating next-generation sequencing datasets. Bioinformatics. 2013;29(4):494–6. Epub 2013/01/15. doi: 10.1093/bioinformatics/bts731 23314324; PubMed Central PMCID: PMC3570212.

112. Dobin A, Davis CA, Schlesinger F, Drenkow J, Zaleski C, Jha S, et al. STAR: ultrafast universal RNA-seq aligner. Bioinformatics. 2013;29(1):15–21. Epub 2012/10/30. doi: 10.1093/bioinformatics/bts635 23104886; PubMed Central PMCID: PMC3530905.

113. Robinson JT, Thorvaldsdottir H, Winckler W, Guttman M, Lander ES, Getz G, et al. Integrative genomics viewer. Nat Biotechnol. 2011;29(1):24–6. Epub 2011/01/12. doi: 10.1038/nbt.1754 21221095; PubMed Central PMCID: PMC3346182.

114. Ching T, Huang S, Garmire LX. Power analysis and sample size estimation for RNA-Seq differential expression. Rna. 2014;20(11):1684–96. doi: 10.1261/rna.046011.114 25246651; PubMed Central PMCID: PMC4201821.


Článek vyšel v časopise

PLOS Genetics


2020 Číslo 3
Nejčtenější tento týden
Nejčtenější v tomto čísle
Kurzy

Zvyšte si kvalifikaci online z pohodlí domova

Svět praktické medicíny 1/2024 (znalostní test z časopisu)
nový kurz

Koncepce osteologické péče pro gynekology a praktické lékaře
Autoři: MUDr. František Šenk

Sekvenční léčba schizofrenie
Autoři: MUDr. Jana Hořínková

Hypertenze a hypercholesterolémie – synergický efekt léčby
Autoři: prof. MUDr. Hana Rosolová, DrSc.

Význam metforminu pro „udržitelnou“ terapii diabetu
Autoři: prof. MUDr. Milan Kvapil, CSc., MBA

Všechny kurzy
Kurzy Podcasty Doporučená témata Časopisy
Přihlášení
Zapomenuté heslo

Zadejte e-mailovou adresu, se kterou jste vytvářel(a) účet, budou Vám na ni zaslány informace k nastavení nového hesla.

Přihlášení

Nemáte účet?  Registrujte se

#ADS_BOTTOM_SCRIPTS#