Main

To systematically analyse how Ded1p influences translation initiation in cells, we first examined how a mutation in the enzyme altered the spectrum of ribosome footprints in cells7. We used the ded1-95 mutation (Ded1pT408I), which reduces the affinity of Ded1p for RNA, diminishes RNA unwinding and confers a temperature-sensitive growth defect to the budding yeast Saccharomyces cerevisiae8. The mutation does not affect pre-mRNA splicing or ribosome biogenesis9. We performed ribosome profiling on wild-type and ded1-95 strains before and after a temperature shift from 30 °C to 37 °C for 5 min (Extended Data Fig. 1a–h). The short time was chosen to minimize any secondary effects on ribosome footprints arising from broader translation defects.

At 30 °C, wild-type and ded1-95 strains showed virtually indistinguishable RNA expression and translation profiles (Extended Data Fig. 1i, j). After the temperature shift, translation broadly decreased in ded1-95, compared to the wild type (Extended Data Fig. 1k–n). These observations indicate that Ded1p promotes translation initiation for most mRNAs, consistent with previous findings10,11. However, translation of a subset of mRNAs coding for proteins involved in gluconeogenesis, cell wall synthesis and transcripts encoding histones were less affected by Ded1p than other mRNAs (Extended Data Fig. 2a, b).

The fraction of ribosomes on 5′ untranslated regions (UTRs) markedly increased upon temperature shift in ded1-95, compared to the wild type (Fig. 1a, Extended Data Fig. 3a). The majority of mRNAs showed higher ribosome occupancy of the 5′ UTR in the ded1-95 strain, which correlated with lower translation efficiency of the main open reading frame (ORF) (Fig. 1b, Extended Data Fig. 3b). To examine the link between increased ribosome occupancy in the 5′ UTR and diminished translation of the main ORF, we performed polysome fractionation with northern blot analysis of individual mRNAs. The PSA1 mRNA—the translation efficiency of which is markedly affected by Ded1p—showed a distinct shift from polysomes to monosomes in ded1-95 compared to the wild type, upon temperature shift but not at 30 °C (Fig. 1c, d, Extended Data Fig. 3c). TDH2 mRNA, which is largely unaffected by Ded1p, did not show a comparable shift (Fig. 1c, d). Collectively, these observations suggest that increased ribosome occupancy on 5′ UTRs correlates with binding of the mRNA to only a single ribosome. This notion is consistent with previous reports12. Ribosome profiling on only the 80S monosome fraction upon temperature shift also showed more footprints on 5′ UTRs in the ded1-95 strain, compared to the wild type (Extended Data Fig. 3d), indicating that ribosome occupancy on 5′ UTRs broadly correlates with the binding of mRNAs to single ribosomes.

Fig. 1: Defects in Ded1p activate alternative translation initiation sites.
figure 1

a, Ribosome profiling tracks of the 5′ UTR of PSA1 mRNA for wild-type DED1 and ded1-95, before and 5 min after a temperature shift. Bars mark the ribosome P-site, the star an alternative translation initiation site. Similar results were obtained in two independent experiments for each dataset. b, Correlation between change in translational efficiency (∆TE) and change in the centre of ribosome density of wild-type DED1, compared to ded1-95 (n = 2,837, 5 min, 37 °C). R, Pearson’s correlation coefficient. c, Representative RNA blots of PSA1 (log2∆TEPSA1 = −2.1) and TDH2 (log2∆TETDH2 = 0.7) after polysome fractionation for wild-type DED1 and ded1-95, 5 min after temperature shift. M, 80S monosomes; LP, light polysomes; HP, heavy polysomes. Similar results were obtained in three independent experiments. d, Quantification of PSA1 and TDH2 RNA blots. Bars indicate the fraction of the mRNA in monosomes (M), light polysomes (LP), and heavy polysomes (HP). P values from a two-tailed t-test. e, Representative ribosome profiling track for a segment in the 5′ UTR of PSA1 (indicated by the star in a in ded1-95 (5 min, 37 °C)). The near-cognate initiation codon is highlighted. Similar results were obtained in two independent experiments f, Fraction of near-cognate and cognate initiation codons at sites with marked ribosome accumulation (red bars, ATISs, n = 396), and at randomly chosen control positions (grey) in 5′ UTRs in ded1-95. P values from a two-tailed t-test. k, Mean ribosome occupancy 10 nt 3′ and 5′ of high-confidence ATIS on 5′ UTRs (moving average of ±1 nt) for ded1-95 (red) and wild-type DED1 (5 min, 37 °C).

Source data

A large number of sites on 5′ UTRs with increased ribosome footprints in the ded1-95 strain were enriched with near-cognate initiation codons (Fig. 1e–g), which differ from the canonical 5′-AUG-3′ initiation codon by a single nucleotide and can create alternative translation initiation sites (ATISs)13. Increased ribosome occupancy on AUG codons in 5′ UTRs was also seen in the ded1-95 strain (Fig. 1f), but only a few of these sites exist in the yeast transcriptome, compared to nine different near-cognate initiation codons, which constitute roughly 14% of all codons14. Ribosomes can translate from the ded1-95-activated ATISs, as demonstrated by ribosome profiles on small ORFs that start at these ATISs and finish at the respective termination codons, by the lack of ribosome accumulation at ATISs when translation was not arrested and by the periodicity of ribosome footprints starting from ATISs (Extended Data Fig. 3e–l). Collectively, the data indicate that defective Ded1p leads to ATIS activation in 5′ UTRs, which decreases polysome formation on the main ORFs and thereby overall protein production. We conclude that Ded1p function suppresses the use of ATISs.

Although ATIS activation in ded1-95 was extensive, only a subset of all near-cognate initiation codons was used. We detected no preferred length or register of the corresponding small ORFs relative to the main ORFs. However, in the ATISs, near-cognate codons from which translation initiation is most efficient were over-represented, whereas near-cognate codons from which translation initiation is least efficient were underrepresented15,16 (Extended Data Fig. 4a). These observations show that ATIS activation is influenced by inherent codon preferences of the pre-initiation complex (PIC), although these preferences do not fully explain the ATIS activation pattern (Extended Data Fig. 4b–e).

To better understand this pattern, we examined whether remodelling of mRNA secondary structure by Ded1p is linked to ATIS activation. As an RNA helicase, Ded1p has been implicated in RNA structure remodelling11,17, but it is not known which mRNA structures Ded1p alters in cells. To delineate the cellular mRNA structures that are remodelled by Ded1p, we used dimethyl sulfate (DMS) probing in vivo18 and measured changes in mRNA structure in ded1-95 and wild-type strains upon temperature shift (Extended Data Fig. 5a, b). Unwinding of mRNA structure by Ded1p was most pronounced in 5′ UTRs, compared to other mRNA regions (Fig. 2a, Extended Data Fig. 5c). Notably, ded1-95 activated ATISs were generally located 5′ of unwound RNA regions (Fig. 2b, c). Even near-cognate codons for which translation initiation is least efficient were activated, if they were located 5′ of mRNA structure (Fig. 2b). Our observations link the inability of ded1-95 to resolve mRNA structure to ATIS activation, suggesting that Ded1p suppresses ATIS activation by unwinding mRNA structure.

Fig. 2: mRNA structure unwinding by Ded1p and ATIS activation.
figure 2

a, Metagene profile of mRNA unwinding by Ded1p on 5′ UTRs and the 5′ moiety of ORFs (moving average of ±1 nt). Similar results were obtained in two independent experiments. TSS, transcription start site; AUG, translation start site. b, Representative differential DMS mutational profiling with sequencing (DMS-MaPseq) track for the 5′ UTR of the PSA1 mRNA (upper track, mRNA regions unwound in wild-type DED1 (WT) marked by red bars, the more negative the value, the stronger the unwinding). Similar results were obtained in two independent experiments. For comparison, ribosome profiling traces of the 5′ UTR of PSA1 mRNA for wild-type DED1 and ded1-95 are shown. Near-cognate codons are colour-coded according to their initiation efficiency, as indicated. Green lines mark activated ATISs. c, Localization of unwound mRNA structure 3′ of activated ATISs. Enrichment of differential DMS-MaPseq counts. Negative values indicate unwound mRNA regions in wild-type DED1 within 20 nt of high-stringency ATISs (n = 274), compared to all other near-cognate codons (n = 60,666; excluding ATIS). The shaded area marks a significant difference in secondary RNA structure between the regions downstream of an ATIS and downstream of any near-cognate codon (P = 0.00004, two-tailed t-test). The dashed line marks the 5′ nucleotide of the ATISs or the near-cognate codon.

Source data

To investigate how Ded1p physically accomplishes this function, we determined which cellular RNAs bound to wild-type Ded1p using an high-throughput cross-linking-based approach (cross-linking-aided RNA affinity precipitation with sequencing (XL-RAP–seq)) and the individual-nucleotide resolution cross-linking and immunoprecipitation (iCLIP) technique to map Ded1p-binding sites on these RNAs19 (Extended Data Fig. 6a–c). Ded1p cross-linked predominantly to mRNAs and ribosomal RNA (Extended Data Fig. 6b), especially to the 40S ribosomal subunit (Fig. 3a), which is part of the PIC that scans 5′ UTRs20. The most frequently cross-linked position maps to helix 16, located at the mRNA entry channel (Fig. 3a). Notable cross-linking was also observed at helix 26, which is located at the mRNA exit site, and in extension segment 6 (around nucleotide 720), which is located in the vicinity of the other cross-link sites on the solvent side of the 40S subunit (Fig. 3a). Ded1p binding to helices 16 and 26 is consistent with reported interactions between Ded1p and eIF3c and the eIF3b–eIF3g–eIF3i sub-complex, that binds near these sites21,22 (Extended Data Fig. 7a, b). Human DDX3X also binds to helix 163.

Fig. 3: Ded1p cross-linking to the 40S ribosomal subunit and to mRNAs.
figure 3

a, Left, fraction of iCLIP reverse transcription stops on 18S rRNA. Moving average of ±2 nt, values represent the average from two independent experiments. Numbers denote predominant cross-linking sites. Right, the position of the three predominant Ded1p cross-linking sites (red) in the crystal structure of the 40S ribosomal subunit30. RNA, grey; ribosomal proteins, cyan; Ded1p cross-link sites, red. ES6, extension segment 6. b, Metagene profile of Ded1p association to mRNAs, calculated from two independent iCLIP experiments (moving average of ±1 nt). Stop, translation stop site; PAS, polyadenylation site. c, Ded1p cross-linking to the 5′ UTR of the PSA1 mRNA. Top, fraction of reverse transcription stops per nucleotide, normalized to transcript length. For comparison, differential DMS-MapSeq (middle) and ribosome profiling (bottom) tracks of the 5′ UTR of PSA1 mRNA for wild-type DED1 and ded1-95 are shown (5 min, 37 °C). Similar results were obtained in two independent experiments. d, Enrichment of Ded1p cross-linking within 20 nt of ATISs (n = 274) normalized to the background distribution of Ded1p binding (moving average of ±1 nt, reverse transcription stops normalized for each mRNA). The dashed line marks the ATIS position. The shaded area marks a significant difference in Ded1p binding between the regions in the vicinity of an ATIS and in the vicinity of a random position within the same 5′ UTR (P = 0.013, two-tailed t-test). e, Enrichment of differential DMS-MaPseq counts (Fig. 2c) within 40 nt of Ded1p binding sites (n = 178, high-stringency ATIS) on 5′ UTRs. The shaded area marks a significant difference in RNA structure between the regions downstream of a Ded1p binding site and downstream of a random position within the same 5′ UTR (P = 0.008, two-tailed t-test).

Source data

Ded1p further cross-linked to virtually all expressed mRNAs, predominantly in 5′ UTRs (Fig. 3b, c). This cross-linking pattern is consistent with the physical contact of Ded1p to the PIC. Aside from a modest preference for A and U, no sequence motifs could be identified in the mRNA cross-linking sites (Extended Data Fig. 7c). However, peaks of Ded1p cross-linking on 5′ UTRs were frequently proximal to ded1-95-activated ATISs (Fig. 3c, d), and unwound mRNA structure was located 3′ of Ded1p cross-linking sites (Fig. 3c, f).

Collectively, the data link Ded1p binding to mRNA, unwound mRNA structure, ATIS location and binding of Ded1p to the PIC. This link is illustrated by a segment of the PSA1 mRNA as an example (Fig. 4a). Ded1p binding is most pronounced 5′ of unwound RNA structure, indicating that Ded1p does not exclusively contact mRNA structure, but also regions that are 5′ of the structure. This finding is consistent with the notion that Ded1p functions in the context of the scanning PIC. The scanning process is slowed by RNA structure20,23, and a slowed PIC conceivably permits Ded1p to survey the mRNA for structured regions that it then unwinds. Biochemical data show higher functional affinity of Ded1p for unstructured RNA, compared to structured RNA8, rationalizing the contacts of Ded1p to unpaired mRNA that is 5′ of unwound mRNA structure, as the helicase travels in a 5′ to 3′ direction with the PIC.

Fig. 4: Ded1p function on 5′ UTRs.
figure 4

a, DMS-MapSeq constrained secondary structure model of a fragment of the PSA1 mRNA 5′ UTR. The ATIS is marked by a line. Shading indicates Ded1p cross-linking (iCLIP). Triangles indicate unwinding (DMS-MapSeq) for each nucleotide (log2 ratio of normalized DMS-MapSeq counts of wild type/ded1-95 in two categories). Yellow triangles, 0.35–0.7 (moderately unwound) and red triangles, >0.7 (strongly unwound). b, Representative RNA blots, after sucrose gradient centrifugation for wild-type DED1 and ded1-95, (5 min, 37 °C) for constructs expressing PSA1FLAG mRNA with wild-type 5′ UTR (ATIS and secondary structure (Δ2°)) or with a mutated ATIS (∆ATIS). Similar results were obtained in three independent experiments. c, Quantification of RNA blot experiments shown in b, indicating the fold change in the fraction of PSA1 mRNA in monosomes in ded1-95, compared to the wild type upon temperature shift. Data are from three independent biological replicates, lines mark the mean. P value from a two-tailed t-test. d, Schematic of Ded1p function on 5′ UTRs. mRNA is depicted as a black line, mRNA structure as a hairpin, the PIC as a grey shape, Ded1p as a red oval, and the near-cognate codon as a green rectangle. e, Ribosome occupancy tracks (5 min, 37 °C) of ded1-95 and wild-type DED1 (vegetative control and anaphase II) for GLY1 mRNA. ATISs are marked by dashed lines. Similar results were obtained in two (vegetative control) and four (anaphase II) independent experiments. mORF, main ORF. f, Mean ribosome occupancy 10 nt 3′ and 5′ of a high-confidence ATIS on 5′ UTRs (moving average ± 1nt) for wild-type DED1 (vegetative control and anaphase II). g, Representative western blot of Ded1p and Hxk1p (loading control) in vegetative cells and cells in anaphase II. Numbers indicate the relative expression level of Ded1p from four independent experiments with s.d.

Source data

Our data collectively indicate that failure of Ded1p to resolve mRNA structure leads to ATIS activation. To directly probe the link between mRNA unwinding and ATIS activation, we generated a PSA1 mRNA with a mutation in an activated ATIS 5′ of unwound RNA structure. (Fig. 4b). The mutation markedly diminished the sensitivity to Ded1p-deficiency seen with the native PSA1 mRNA (Fig. 4b, c). Alterations in the RNA structure 3′ of the ATIS also decreased sensitivity to Ded1p (Extended Data Fig. 8a, b). Identical observations were made for mutations in an ATIS and the corresponding RNA structure in the ATP5 mRNA (Extended Data Fig. 8c–f). These results show that the effect of Ded1p on translation initiation depends not only on RNA unwinding, but also on proximal ATISs. Without a proximal ATIS, failure of Ded1p to unwind 5′ UTR structures does not abrogate scanning of the PIC and subsequent translation of the main ORF (Extended Data Fig. 9). This finding challenges the notion that cellular 5′ UTR structures alone are insurmountable hindrances for the scanning PIC.

Together, our results suggest the following function for Ded1p on 5′ UTRs (Fig. 4d). The enzyme associates with the PIC in the vicinity of the mRNA entry site of the small ribosomal subunit (Fig. 3a). This site is in close proximity to eIF4G and eIF4A (Extended Data Fig. 7b), both of which bind Ded1p with high affinity and might therefore be important for recruitment and function of Ded1p on the PIC8,24,25. The density of Ded1p cross-linking sites on 5′ UTRs increases with distance from the 5′ cap (Fig. 3b), suggesting gradual recruitment of Ded1p to the mRNA entry site during the scanning process. This notion is consistent with the reported increase of Ded1p function with greater distance from the 5′ cap and with 5′ UTR length11. The mRNA binding pattern of Ded1p further suggests that Ded1p is targeted to its sites of action through association with the scanning PIC. This is an effective way to deploy the enzyme exactly at sites at which it is needed, even though these sites lack common sequence or defined structure signatures. If Ded1p is missing or defective, mRNA structure persists, the PIC stalls and either dissociates from the mRNA, continues slowed scanning through the structure, or undergoes subunit joining and translation initiation if a near-cognate codon is present (Fig. 4d). Ribosomes initiating on an ATIS block subsequent scanning ribosomes from reaching the canonical initiation site, thereby decreasing translation efficiency for the main ORF (Fig. 1b). Unless an ATIS marks an N-terminal extension of the main ORF, PICs initiating at an ATIS are likely to be deterred from translating the main ORF. PICs encountering 5′ UTR structures without a proximal ATIS also interfere with scanning, but the kinetic pause introduced by PIC stalling, slowed scanning through the structure or a combination thereof is shorter than on an activated ATIS. Slowed PICs will eventually reach the main ORF (Extended Data Fig. 9), and therefore 5′ UTR structure alone affects main ORF translation less in isolation than it does in combination with proximal ATISs. Our model for Ded1p function does not preclude additional roles of the enzyme before the PIC scanning process25. However, the Ded1p function outlined above largely accounts for the observed Ded1p interactions with mRNA, and therefore, additional roles of Ded1p are probably restricted to transient Ded1p–mRNA interactions.

Finally, our data reveal a straightforward mechanism for activation of upstream ORFs. The mRNA structures in the 5′ UTRs represent a large set of riboswitches that are sensitive to Ded1p. Active Ded1p turns the switches off, suppresses ATIS activation and allows efficient translation of the main ORF. Inactivation of the helicase by post-translational modifications1, by metabolites such as AMP26, by decreased Ded1p levels or by sequestration of Ded1p in RNP granules25,27 turns the switches on, activating the ATISs and thereby inhibiting translation from the corresponding main ORFs. Certain peptides that are translated from activated ATISs might also have direct biological functions28, but the regulation described here appears to be independent of functional peptides.

This mechanism for activation of upstream ORFs is probably used in biological processes. This notion is supported by several lines of evidence. First, there is a marked increase in sequence conservation in the RNA regions around activated ATIS (Extended Data Fig. 10a). Second, the ded1-95 activated ATIS in the ALA1 transcript produces an N-terminal extension that targets Ala1p to the mitochondria29 (Extended Data Fig. 10b). Third, during meiosis, ATIS activation15 occurs in a pattern that is highly similar to the ATIS activation pattern seen with ded1-95 upon temperature shift (Fig. 4e, f, Extended Data Fig. 10c). Notably, we find reduced Ded1p levels during meiosis (Fig. 4g, Extended Data Fig. 10d). This link between Ded1p levels and the activation of ATISs proximal to 5′ UTR structures during meiosis suggests a role for the levels of Ded1p in this process. Collectively, our observations show that the regulatory program linking Ded1p to mRNA structure and ATIS activation is used in a physiological cellular process. The results indicate that intricate translation control and activation of upstream ORFs can be based on simple, ubiquitous elements: a helicase, mRNA structure and near-cognate initiation codons.

Methods

No statistical methods were used to predetermine sample size. The experiments were not randomized and the investigators were not blinded to allocation during experiments and outcome assessment.

Yeast strains, plasmids and oligonucleotides

Yeast strains used in this study are listed in Supplementary Table 1. Strains were grown at 30 °C unless stated otherwise. Primers, northern blot probes and other DNA oligonucleotides are listed in Supplementary Table 2, RNA oligonucleotides are listed in Supplementary Table 3, and DNA plasmids are listed in Supplementary Table 4.

Generation of a yeast strain expressing Ded1p–HTBH

Construction of plasmid pEJ21 containing the N-terminally haemagglutinin (HA)-tagged Ded1p has previously been described24. pEJ21was then used to generate the plasmid pEJ5. The HA tag was replaced by a sequence containing a HpaI and a SphI site (amplification with primers X1 and X2), generating pEJ1. The His6-TEV-Biotin (HTB) tag was amplified from pFA6-HTB-kanMX6 plasmid (gift from P. Kaiser) with primers X3 and X4.The resulting PCR product was cloned into pEJ1 via its HpaI and SphI sites yielding pEJ2. A second His6 tag was introduced by site-directed mutagenesis with primers X5 and X6 generating pEJ3. The C-terminal His6-TEV-Biotin-His6 (HTBH) tag was introduced into pEJ424 by amplification of pEJ3 with primers X8 and X9 and subcloning with PflMI and SpeI into pEJ4, yielding pEJ5. pEJ5 was linearized and used to transform BY4741 by standard lithium acetate transformation yielding yeast strain yEJ2.

Generation of a yeast strain expressing Ded1p–His6–FLAG3

Yeast strain yDPB740, containing a C-terminal His6–FLAG3 tag on the endogenous DED1 allele, was generated from BY4742 using standard methods. In brief, a homologous recombination template was designed comprising the 40 nucleotides upstream and downstream of the DED1 stop codon flanking the His6FLAG3 tag (with stop codon) and kanMX6 drug resistance cassette. This template was generated by amplifying from pFA6a–6×His–3×FLAG–kanMX6 plasmid with primers DW1 and DW2. PCR product was used to transform BY4742 by standard lithium acetate transformation yielding yeast strain yDPB740.

Generation of a yeast strain expressing wild-type, ∆ATIS and secondary structure mutants of PSA1 and ATP5 mRNAs

FLAG-tagged PSA1 and ATP5 strains were generated from the respective cDNAs using standard methods as described above (pEJ14, pEJ15 and pEJ18, pEJ19, respectively). The FLAG-tag was appended at the 3′ terminus of the PSA1 and ATP5 ORF, respectively. For PSA1, mutations in the ATIS in the 5′ UTR (PSA1-∆ATIS, pEJ16, Fig. 4b, c) contained the following changes: (c.-58A>C (c.-58 indicates the 58th nucleotide 3′ of the A of the AUG), c.-56A>C, c.-52T>A). Mutations in the 5′ UTR mRNA structure (PSA1-∆2°, pEJ17, Extended Data Fig. 8a, b), contained the following changes: (c.-39_- 38AG>TC, c.-36_-35TA>AT, c.-32A>T,c.-24_-22AAA>TCT, c.-19_-18AA>CT). For ATP5, mutations in the ATIS in the 5′ UTR (ATP5-∆ATIS, pEJ20, Extended Data Fig. 8e, f) contained the following changes: (c.-126A>C, c.-120_119TT>CC, c.-109A>C). Mutations in the 5′ UTR mRNA structure (ATP5-∆2°, pEJ22, Extended Data Fig. 8e, f), contained the following changes: (c.-102C>A, c.-99C>A, c.-96G>A, c.-83G>A, c.-81C>A, c.-76G>A, c.-74G>A).

Polysome analysis

Polysome analysis using 20U (A260) lysate was performed as described31, with a lower final concentration of cycloheximide (50 μg ml−1). In brief, after centrifugation through a 15–45% (w/v) sucrose gradient, sixteen fractions were collected at a pump speed of S = 0.9 ml min−1. RNA in each fraction was precipitated by adding two volumes of ice-cold ethanol and incubating overnight at −80 °C. RNA was extracted with phenol–chloroform according to standard protocols. Samples were applied to a 1.4% agarose gel containing 6% (v/v) formaldehyde, and electrophoresis was performed as described31. RNA was visualized with ethidium bromide. The amount of 18S rRNA in each fraction of the gradient was quantified with ImageQuant 5.2 software (Molecular Dynamics).

Northern blot analysis

For ribosome association of individual mRNAs, gel electrophoresis was performed after polysome analysis and fractionation as described above. RNA was subsequently transferred to nitrocellulose membranes (AmershamHybond–N, GE Healthcare) and further processed as described31. DNA oligonucleotides (Supplementary Table 2; X85, X96, X105, X114) were radiolabelled with PNK according to standard procedures. Probes were incubated with the membranes in hybridization buffer (6× saline sodium citrate (SSC), 0.1% SDS, 10× Denhardt’s reagent) overnight at 42 °C. Membranes were subsequently washed three times with wash buffer (6× SSC, 0.1% SDS). Probe signals were visualized using a Molecular Dynamics Phosphorimager (GE Healthcare) and quantified with ImageQuant 5.2 software (Molecular Dynamics). Normalized signal intensities were compiled for fractions corresponding to monosomes, light and heavy polysomes and averaged from at least three biological replicates.

Western blot analysis

Lysates from yeast strain A14201 at vegetative phase and stage 11 (ndt80 release time course) were prepared as described15. After loading equal amounts of protein on a 10% NEXT gel, denaturing gel electrophoresis and transfer to a PVDF membrane, western blotting was performed with anti-Ded1p (rabbit; 1:5,000) and anti-hexokinase (rabbit;1:10,000; US Biological) antibodies. Chemiluminescence was quantified by Imagequant software. Hexokinase served as a loading control for normalization.

Ribosome profiling

Yeast cultures (500 ml) were grown at 30 °C in rich medium to mid-log phase (OD600 nm of approximately 0.4) and divided into two equal volumes. Cycloheximide (final concentration, 50 μg ml−1) was added to one sample. Cells were rapidly collected by centrifugation at 4,000g for 2 min, and snap-frozen on dry ice. One volume of pre-warmed medium (44 °C) was added to the remaining sample, resulting in a temperature of 37 °C for the entire volume. The temperature of 37 °C was verified. The entire sample was immediately moved to a shaking incubator at 37 °C. Five minutes after temperature shift, the yeast culture was treated with cycloheximide and cells were collected as mentioned above. For run-off experiments, yeast cells were collected in the absence of cycloheximide (Extended Data Fig. 3e).

Cell lysis, RNase I treatment (Ambion) and sucrose gradient centrifugation was performed as described32 for 25 units A260 per sample. In addition, lysates (approximately 150 μl) were treated with 3.25 μl Turbo DNase I for 1 h at 25 °C. Purification and processing of ribosome-protected fragments were carried out as described32, except that rRNA depletion with the RiboZero kit (Illumina) was omitted. Depletion of rRNA was performed at the level of circularized cDNA, as described7 (1 μl of a 5 μM mix of biotinylated DNA oligonucleotides, Supplementary Table 2 (X66–X80) and MyOne Streptavidin C1 DynaBeads (Invitrogen)). PCR amplification and sequencing was performed as described7.

Monosome-protected fragments were isolated as described12. In brief, sucrose gradient centrifugation of lysates was performed and fractions corresponding to monosomes were pooled and treated with 1/10 U RNase I per unit A260 nm lysate and 0.4U Turbo DNase I per unit A260 nm lysate for 1 h at 25 °C. Reactions were stopped by phenol–chloroform extraction of RNA. Monosome-protected fragments were processed as described above for ribosome-protected fragments.

The fragmented mRNA control libraries were generated as described32. Sizing, concentration and quality of each DNA library was assessed with the High Sensitivity DNA kit on an Agilent2100 Bioanalyzer system. Up to eight DNA libraries were pooled before performing 50 bp single end read sequencing on an Illumina HiSeq2500 V2 in rapid run mode.

Processing of the ribosome profiling data was performed as described32. In brief, adaptor sequences and ribosomal reads were removed. Remaining reads were mapped to the sacCer3 genome with the TopHat software (parameters set as: --no-novel-juncs -N 2 --read-edit-dist 2 --max-insertion-length 3 --max-deletion-length 3 -g 2 (https://www.yeastgenome.org)). All other parameters were kept at default settings33. The abundance of mRNAs in ribosome or monosome-protected fragments as well as in the fragmented RNA control libraries were determined using Cufflinks software33. These values were used to calculate translational efficiencies as described14. For the calculation of log2∆TE values we also included a constant factor reflecting the change in the overall size of the mRNA pool, derived from the spike-in of RNA controls (Supplementary Table 3).

P-sites in ribosome-protected fragments (RPFs) were determined using the 13th position from the 5′ end of reads with 28 or 29 nt14. The fraction of ribosomes on 5′ UTRs was calculated for each mRNA by counting all RPFs on the 5′ UTR (excluding positions −3 to −1), divided by the number of all RPFs mapped to the entire mRNA. mRNAs with 5′ UTRs containing fewer than 10 nt were excluded from the analysis. The centre of ribosome density (CRD) was calculated as described34.The shift in the CRD (∆CRD) in ded1-95 compared to wild-type DED1 upon temperature shift was defined relative to the entire length of the mRNA according to:

$${\rm{\Delta }}CRD=\left(\left(CRD\,ded1-{95}\right)-\left(CRD\,\mathrm{wild}-\mathrm{type}\,DED1\right)\right)/\mathrm{mRNA}\,\mathrm{length}$$

A negative ∆CRD value marks increased ribosome accumulation in the 5′ UTR in ded1-95.

ATIS were identified according to a previously described algorithm15. In brief, a position is considered an ATIS, (i) if minimal ribosome count value (±1 nt of the nucleotide under consideration) is greater than nine (high-stringency ATIS) or four (medium-stringency ATIS) in all replicates; (ii) if the ratio of ribosome occupancy between two neighbouring nucleotides 5′ to 3′ (positionn − 1/positionn) is greater than or equal to three (high-stringency ATIS) in all replicates, or greater or equal of three in one and greater than or equal to 1.75 in the other replicate (medium-stringency ATIS); and (iii) if the normalized ribosome count in ded1-95 cells 5′ after temperature shift to 37 °C is 1.5-fold higher that in wild-type DED1 in all replicates. This algorithm identified 396 high-stringency ATIS and 2,126 medium-stringency ATIS. Near-cognate codons were identified in 259 high-stringency ATIS (65%) and 1,382 medium-stringency ATIS (65%) within a moving window of ±1 nt. Canonical AUG initiation codons were found in 4% high-stringency ATIS, and in 3% medium-stringency ATIS. As a control set, we collected all near-cognate codons on 5′ UTRs of mRNA genes with a 5′ UTR length between 20 and 500 nt. After removal of near-cognate codons in medium-stringency ATIS, we identified 60,666 near-cognate codons.

XL-RAP–seq

Yeast cells containing HTBH-tagged DED1 were grown in rich medium to an OD600 nm of 1.0–1.5, collected by brief centrifugation at 4,000g, re-suspended in ice-cold water or remaining YPD medium, transferred to a Petri dish, and subjected to UV-light in a Stratalinker (600 mJ cm−2, 254 nm) on ice. Cells were washed in ice-cold water, sedimented by centrifugation for 5 min at 5,250g, frozen on dry ice and stored at −80 °C.

Frozen cells were lysed in QIA-1M buffer (100 mM NaH2PO4 pH 8, 10 mM Tris, 1M NaCl, 8 M Urea, 10 mM imidazole, 0.5% (w/v) IGEPAL, 2.5 mM β -mercaptoethanol, 1 mM PMSF, protease inhibitor cocktail (Roche)) with glass beads six times for 30 s in a Beadbeater system (Biospec products). Glass beads were removed, and lysates were centrifuged at 5,250g for 30 min. Cleared lysates were incubated with Ni2+–Agarose (40 μl slurry per g dry pellet weight, pre- equilibrated in buffer QIA-1M; Qiagen) overnight at 4 °C. Ni2+-beads were washed in 25 ml of wash buffer 1 (0.3 M NaCl, 10 mM Tris, 100 mM NaH2PO4, 8 M Urea, 10 mM imidazole) and sample was eluted with 10 ml elution buffer1 (0.3 M NaCl, 100 mM Tris, 50 mM NaH2PO4, 8 M Urea, 500 mM imidazole, 10% (v/v) glycerol). Eluates were then incubated with12.5 μl equilibrated streptavidin-conjugated agarose resin (Pierce Technologies) per g dry pellet weight overnight at 4 °C. Streptavidin beads were washed with 12.5 ml wash buffer 2 (0.3 M NaCl, 100 mM Tris, 8 M Urea, 0.5 mM EDTA)—containing 2% SDS, with wash buffer 2 (12.5 ml) without SDS, and with 1× TEV buffer. Beads were next incubated with 50 U AcTEV (Invitrogen) for 2 h at 4 °C. The sample was eluted with 2 × 0.9 ml TEV elution buffer (300 mM NaOAc pH 6, 8 M Urea, 0.5% NP-40). Eluates were incubated with 175 μl Proteinase K (4 mg ml−1; Roche), and the reaction was stopped by standard phenol–chloroform extraction. Released RNA was precipitated by ethanol precipitation overnight. RNA was re-suspended in 1× Turbo DNase buffer and incubated with 0.4 μl TurboDNase (Ambion) in a final volume of 20 μl for 20 min at 37 °C. Turbo DNase was inactivated according to the manufacturer’s instruction and the RNA was precipitated with ethanol.

cDNA was generated from the RNA sample using the ‘template switch’ activity of M-MLV reverse transcriptase35. Reactions were performed in 20 μl according to the instructions of the manufacturer mix (Invitrogen) with 0.5 μl Superscript II and 0.05 μM final concentration of tailed random hexamer primer (X13, Supplementary Table 2). Reaction conditions were as follows: 20 °C, 10 min; 37 °C, 10 min; 42 °C, 45 min. Next, 0.25 μM of primers (X14–X17, Supplementary Table 2) containing a 7-nt index and three guanosine ribonucleotides at their 3′ end were added to the reaction mix and reactions were performed at 42 °C for 30 min. Four microlitres of the resultant cDNA were used as template for amplification with primers X18 and X19 (Supplementary Table 2) and Advantage 2 polymerase mix (Clontech) for 30 cycles, according to the manufacturer’s instructions. PCR products were precipitated and washed with 75% ethanol. DNA libraries were sequenced as 36-bp single-end reads on an Illumina Genome Analyzer. Reads were mapped to sacCer2 genome with Bowtie software with default settings.

Reads were excluded from further analysis if their location was outside of the boundaries of an mRNA or other transcribed regions of the genome as previously defined36. Genes were only considered to be bound by Ded1p, if more than 10 FPKM mapped to the respective mRNA gene.

iCLIP

iCLIP experiments were performed independently with two different approaches, (i) using Ded1p–HTBH and (ii) Ded1p–His6–FLAG3, both on the endogenous DED1 allele. For iCLIP with Ded1p–HTBH, cell growth, cross-linking and tandem affinity purification on Ni2+-agarose and streptavidin beads was performed as described above for the XL-RAP–seq procedure. Streptavidin beads were washed twice in 1× PNK buffer and split into two samples (80%, L; 20%, H). 200 ng RNase I was added per gram dry pellet weight to the L-sample and 0.1 ng RNase I to the H-sample. The samples were incubated for 5 min at 37 °C on a rotator. Resins were subsequently washed with ~1.5 ml wash buffer 2 (2% (w/v) SDS), and then with 1.5 ml wash buffer 2 without SDS, and finally with 1.5 ml 1× PNK buffer. The supernatant was removed and beads were re-suspended in 67 μl RNase-free water, 3 μl alkaline phosphatase (NEB) and 2 μl RNasin (Roche), and incubated for 20 min at 37 °C. Beads were washed twice with 1.5 ml 1× PNK buffer.

3′ ligation was performed by re-suspending the resin in 32 μl RNase-free water with 8 μl of 20 μM RL3 RNA Linker (Supplementary Table 3). Reactions were performed overnight at 4 °C in 40 μl, containing 22 μl RNase-free water, 8 μl 10× T4 RNA ligase buffer, 8 μl BSA (0.2 μg μl−1), 3 μl T4 RNA ligase (all NEB).

The resin was washed with 1.5 ml 1× TEV-salt buffer (50 mM Tris pH 7.5, 300 mM NaCl, 0.5 mM EDTA, 1 mM DTT) and twice with 1.5 ml 1× PNK buffer. Beads were re-suspended in 64 μl RNase-free water, 8 μl 10× PNK Buffer, 4 μl 32P-γ-ATP, 4 μl T4 PNK (all NEB). Reactions were performed for 50 min in a shaking thermoblock at 37 °C. Beads were washed twice with 1.5 ml TEV salt buffer and twice with 1.5 ml TEV-elution buffer. Beads were subsequently mixed with SDS loading dye and subjected to PAGE on a 10% NEXT gel (Amresco) according to the manufacturer’s conditions. Gels were subsequently blotted with nitrocellulose membranes (Amersham Protran, GE Healthcare) and exposed to X-ray film. RNA was then liberated by Proteinase K treatment as described37. The purified RNAs were re-suspended in 89 μl RNase- free water, 11 μl 10× DNase I Buffer, 5 μl RNasin, 5 μl RQ1 DNase (all Promega) and incubated for 20 min at 37 °C. The reaction was stopped by standard phenol–chloroform extraction and RNA was ethanol-precipitated overnight.

The DNase-treated RNA was re-suspended in 1.4 μl RNase-free water, 0.2 μl 10 mM dNTPs, 1 μl 20 nM reverse transcription primer X97 (Supplementary Table 2), and incubated for 5 min at 65 °C. Next, 0.8 μl 5× first-strand buffer, 0.2 μl 1M DTT, 0.2 μl RNase Inhibitor, 0.2 μl Superscript III (all Invitrogen) were added and incubated for 30 min at 50 °C. RNA was degraded by alkaline hydrolysis (after addition of 0.5 μl of 1N NaOH and incubation at 98 °C for 15 min). After addition of loading buffer, cDNA was applied to 10% denaturing PAGE and staining with SYBR Gold. Fragments of 100–125 nt were cut from the gel. The gel slices were crushed and cDNA was recovered by incubation in 500 μl diffusion buffer (20 mM Tris-HCl pH 7.5, 250 mM NaOAc, 1 mM EDTA, 0.25% (w/v) SDS) overnight at 4 °C with subsequent ethanol precipitation.

The cDNA was suspended in 15 μl RNase-free water and circularized with CircLigase I (Epicentre) according to the manufacturer’s instructions. The circularized cDNA was used for amplification with Phusion polymerase (NEB) and primers X98 and X99 (Supplementary Information 2). PCR settings were: 30 s at 98 °C and then 24 PCR cycles (10 s at 98 °C, 30 s at 58 °C, 30 s at 72 °C). PCR products were applied to 10% non-denaturing PAGE and visualized by SYBR Gold. Products with 75–90bp were extracted from cut gel slices, and ethanol precipitated as described above. PCR products were amplified with Phusion polymerase and primers X100 and X101 for five cycles using the same PCR settings as above. PCR products were then separated on a 2% agarose gel, cut out and subjected to Illumina sequencing using primer X102.

For iCLIP with Ded1p–His6–FLAG3, cells were grown in SD–Trp medium to OD600 nm = 0.5–0.6. The culture was subsequently transferred to a 245 mm × 245 mm × 25 mm square Petri dish. UV cross-linking was performed in a Stratalinker 2400 (150 mJ cm−2, 254 nm) at room temperature. Cells were collected by centrifugation for 5 min at 2000g, washed twice in ice-cold PBS, frozen in liquid nitrogen, and stored at −80 °C.

Frozen cells were lysed in CLIP lysis buffer (50 mM Tris-HCl, pH 7.8, 300 mM NaCl, 1% Triton X-100, 1 mM PMSF, protease inhibitor cocktail (Roche)) with glass beads six times for 1 min in a Disruptor Genie system (Scientific Industries). Lysates were centrifuged after removal of the glass beads at 10,000g, twice for 5 min. Cleared lysates (~26.5 A260 units) were incubated with anti-FLAG M2 Magnetic Beads (20 μl slurry pre-equilibrated in CLIP Lysis Buffer; Sigma) in a total volume of 1 ml overnight at 4 °C. Beads were washed twice in 1 ml FLAG Wash Buffer (50 mM Tris-HCl pH 7.8, 1 M NaCl, 0.1% NP-40) and twice in 1 ml FLAG Elution Buffer (50 mM Tris-HCl pH7.8, 150 mM NaCl, 0.1% NP-40). Proteins were eluted twice in 95 μl FLAG Elution Buffer containing 150 ng μl 3× FLAG tag peptide (Sigma). Pooled eluates were incubated with 10 μl RNase I (Ambion), diluted 1:500,000 in FLAG elution buffer for 15 min at room temperature. Reactions were quenched with 960 μl 8 M guanidine–HCl, 90 μl Dilution Buffer (600 mM Tris-HCl pH 7.8, 3.93 M NaCl), 6.4 μl 2 M imidazole, 10.8 μl 10% NP-40and 12.8 μl 500 mM β-mercaptoethanol. RNase-treated eluates were incubated further with Ni-NTA magnetic agarose beads (50 μl slurry pre-equilibrated in Ni-NTA binding buffer (50mM Tris-HCl pH 7.8, 300 mM NaCl, 10 mM imidazole, 6 M guanidine-HCl, 0.1% NP-40, 5 mM β-mercaptoethanol); Qiagen) overnight at 4 °C.

Ni2+-beads were washed twice in 1 ml CLIP wash buffer I (50 mM Tris-HCl pH 7.8, 500 mM NaCl, 10 mM imidazole, 6 M guanidine-HCl, 0.1% NP-40, 5 mM β-mercaptoethanol) and three times in 1 ml 1× PNK Buffer (50 mM Tris-HCl pH 7.8, 10 mM MgCl2, 0.5% NP-40, 10 mM β-mercaptoethanol). Beads were subsequently incubated with 30 μl dephosphorylation mix (50 mM Tris-HCl pH 7.8, 10 mM MgCl2, 10 mM β-mercaptoethanol, 3 M BU TSAP (Promega), 30 U SUPERase-In (Ambion)) for 30 min at 37 °C in a thermomixer at 1000 rpm. Reactions were terminated by adding 1 ml CLIP wash buffer I, and beads were washed three times in 1 ml 1× PNK buffer, re-suspended in 30 μl ligation mix (50 mM Tris-HCl, pH 7.8, 10 mM MgCl2, 10 mM β-mercaptoethanol, 10% PEG8000 (NEB), 10% DMSO, 2 μM 3′ adenylated adaptor X103, 30 U T4 RNA Ligase 1 (NEB), 30 U SUPERase-In (Ambion)) and incubated for 3 h at 22 °C in a thermomixer at 1000 rpm. Reactions were terminated by adding 1 ml CLIP wash buffer I. Beads were washed three times in 1 ml 1× PNK Buffer, re-suspended in 30 μl Kinase Mix (50 mM Tris-HCl, pH 7.8, 10 mM MgCl2, 10 mM β-mercaptoethanol, 30 U PNK (NEB), 5 μCi γ- 32P-ATP, 30 U SUPERase-In (Ambion)) and incubated for 30 min at 37 °C in a thermomixer (1000 rpm). Reactions were terminated by adding 1 ml CLIP Wash Buffer I. Beads were washed three times in 1 ml CLIP Wash Buffer I, and three times in 1 ml CLIP wash buffer II (50 mM Tris-HCl, pH 7.8, 50 mM NaCl, 10 mM imidazole, 0.1% NP-40, 5 mM β-mercaptoethanol).

Proteins were eluted with 35 μl CLIP wash buffer II containing 200 mM imidazole three times for 5 min at 22 °C in a thermomixer (1000 rpm).

Pooled eluates were digested with Proteinase K (Invitrogen) at 50 °C for 1 h. The reaction was stopped by standard phenol–chloroform extraction. Released RNA was precipitated by ethanol precipitation overnight in the presence of 1 μl GlycoBlue (Ambion) and 1 pmol reverse- transcription DNA primer X104. RNA was converted to first-strand cDNA in a 10 μl standard reaction mix with 0.5 μl Superscript III (Invitrogen) and the co-precipitated DNA primer. The reaction conditions were as follows: 25 °C for 5 min, 42 °C for 20 min, 50 °C for 40 min. RNA was degraded with 1.67 μl of 1 M NaOH for 10 min at 90 °C. The cDNA was ethanol precipitated with GlycoBlue for >2 h. cDNA fragments of 120–200 nt were gel purified on a urea-denaturing 6% acrylamide gel and ethanol precipitated with GlycoBlue overnight. cDNA was circularized in a 10 μl reaction mix using 0.5 μl CircLigase I (Epicentre) in the presence of 1 M betaine at 60 °C for 1h. The reaction was then supplemented with additional 0.5 μl CircLigase I and incubated at 60 °C for 1 h. The enzyme was inactivated at 80 °C for 10 min. PCR and formamide-gel purification of PCR products were performed as described38, using 20 cycles of PCR and isolating 120–200 nt ssDNA fragments.

Sequencing reads were processed as previously described37,39. In brief, after the trimming of adaptor sequences, reads were mapped to sacCer3 with Bowtie2 or TopHat software with similar settings used for the ribosome profiling data, outlined above. Identical reads were subsequently collapsed and duplications removed. The cross-link position of Ded1p to RNA was defined as 1 nt 5′ of the 5′ mapped nucleotide of a sequencing read37.

DMS-MaPseq

Yeast strains (wild-type DED1 and ded1-95) were grown in YPD at 30 °C. Overnight cultures were diluted to OD600 nm of ~ 0.09 and grown to an OD600 nm = 0.6. An equal volume of 44 °C YPD medium was added to achieve an immediate temperature shift to 37 °C, as outlined for the ribosome profiling experiments. Cultures were incubated in a 37 °C water bath for 3 min. At this time, DMS (Sigma) was added to a 5% (v/v) final concentration and incubation was continued with stirring for 3 min. DMS was quenched by adding 30 ml of ice-cold stop solution (30% β-mercaptoethanol, 50% (v/v) isoamyl alcohol). Cells were quickly transferred to ice, collected by centrifugation at 3,500g at 4 °C for 4 min, and washed with 10 ml 30% β- mercaptoethanol solution. Cells were re-suspended in 0.6 ml RNA lysis buffer (6 mM EDTA, 45 mM NaOAc, pH 5.5). Total RNA was purified with hot acid phenol (Ambion) and ethanol precipitation. Sequencing libraries were prepared as previously described17.

Raw fastq files were stripped of linker sequences and filtered for overall quality using the FASTX-Toolkit Clipper and Quality Filter functions (http://hannonlab.cshl.edu/fastx_toolkit/), respectively, requiring that 80% of sequenced bases have a quality score >25. Reads were aligned against the yeast genome (sacCer3) using Tophat v2.1.0 with Bowtie2 with the following settings for a 50 bp sequencing run: --no-novel-juncs -N 5 --read-gap-length 7 --read-edit-dist 7 --max-insertion-length 5 --max-deletion-length 5 -g 3. All non-uniquely aligned reads were then removed. Owing to empirically determined mutation enrichment from non-template addition, 2 nt was trimmed from the 5′ end of each read. Mismatches located within 3 nt of an indel were also discarded for future analysis.

Bioinformatic analyses

Yeast genomic sequence conservation scores were obtained from S. cerevisiae genome database (https://www.yeastgenome.org). Positional coordinates of mRNAs including transcription start sites and polyadenylation sites are based on sacCer3 and reported measurement40. Genome-wide datasets were visualized by IGV software41. Structural models of the small and large ribosomal subunits including initiation factors were generated with the Chimera software30,42. Analyses of Gene Ontology term enrichment were carried out with GOrilla software using a single ranked list of genes43.

RNA structure prediction was carried out with sequences 0–99 nt 3′ of the first nucleotide of an alternative start codon using the RNAfold web server44. Constraint settings were derived from DMS-MaPseq data as follows: a nucleotide was set as ‘unpaired’ if the DMS-MaPseq counts of a given nucleotide exceeded the value of 0.49 relative to the third highest count number in the range of 100 nt downstream of an ATIS.

Statistical significance of enrichment or depletion in certain regions (for example, Figs. 2c, 3d, e, Extended Data Fig. 10a) was determined by comparing weighted data vectors of the observed variable to the background value. To this end, we calculated t-values with the wtd.t.test function in R. The algorithm is based on the mean and 1/(standard errors)2 as an estimate of the means accuracy. The given P values correspond to a two-tailed t-test.

Further bioinformatic analyses and multiple linear regressions were performed with R45 with customized scripts using RStudio Software (https://www.rstudio.com/). Code is available upon request. Normalization of the datasets including ribosome protected fragments, monosome-protected fragments, Ded1p iCLIP–seq and DMS-MaPseq counts were performed relative to the total number of counts of the entire mRNA42.

To compute Ded1p binding density, DMS-MapSeq ratios, or sequence conservation values in the vicinity of ATIS, it was important to normalize for inherent positional trends within the exact region in the respective iCLIP, DMS-MapSeq and sequence conservation datasets. For example, values for DMS-MapSeq ratios (counts ded1-95/ counts wild type), and iCLIP reads show an upward trend with increasing distance from the 5′ cap in 5′ UTRs. To normalize for inherent positional trends, we calculated a background distribution for the vicinity of each ATIS. We randomly choose a position in the respective section of a given mRNA, and determined the signal distribution in the vicinity of this position (for example, position −5 relative to the 5′ nt of an ATIS). This process was repeated four times. The background value reflects the average of these five calculated values. Reported enrichment values represent the ratio of the measured signal over the background value at each indicated position. Values are given in all plots as log2(measured signal/background signal). Statistical significance of enrichment or depletion was determined by calculating the t-value of the observed variable on the basis of the mean and s.d. of the background value.

Metagene profiles were calculated by averaging normalized Ded1p iCLIP counts and DMS- MaPseq counts after binning transcript coordinates from 5′ UTRs, ORFs and 3′ UTRs in bins reflecting 2% of each section of mRNA. Ded1p binding sites and the midpoint of RNA secondary structures were determined by Piranha peak calling software (http://smithlabresearch.org).

Calling parameters were optimized on the basis of visual inspection. To call peak sites of RNA secondary structures, a genome-wide dataset of log2(counts of DMS-MaPseq wild type/ded1-95) was used as input file.

Reporting summary

Further information on experimental design is available in the Nature Research Reporting Summary linked to this paper.

Data availability

The data that support the findings of this study have been deposited in the Gene Expression Omnibus (GEO) repository with the accession code GSE93959. All other data are available from the corresponding author upon reasonable request.