Introduction

tRNA is the adapter molecule that physically links genetic information transcribed in mRNA to proteins1,2. tRNAs are post-transcriptionally decorated with a variety of chemical modifications that are incorporated by specific modifying enzymes3. These modifications maintain tRNA structural integrity, biochemical stability, and/or appropriate codon–anticodon interactions3. Human tRNAs collectively contain more than 40 modifications at specific positions4. The importance of these modifications is emphasized by the existence of more than 50 human tRNA modification enzymes with pathogenic mutations3,5,6. Deficiency of these enzymes frequently manifests as brain dysfunction, diabetes, short stature, and mitochondrial diseases. On the other hand, high expression of tRNA modification enzymes is frequently associated with cancer progression and is presumed to support cancer growth because knockout (KO) of these enzymes slows the growth of cancer cells in culture dishes and xenografts5,6,7.

Dihydrouridine (D) is a tRNA modification (Fig. 1a, b) conserved in all three domains of life8. It is found at tRNA positions 16, 17, 20, 20a, and 20b in the eponym D-loop and at position 47 in the variable loop8. D is formed by addition of two hydrogens at positions 5 and 6 of the uracil base, breaking the C=C double bond between positions 5 and 6 (Fig. 1a). The hydrogen additions remove the aromaticity and planar structure of the uracil base, destabilizing the C3′-endo ribose conformation and allowing less stacking of the base9,10,11. Thus, the D modification provides more conformational flexibility to the D-loop. Consistently, more D is found in psychrophilic bacteria and less D is observed in thermophilic archaea, likely reflecting the need to increase and decrease tRNA flexibility in order to adapt to environmental temperatures, respectively12,13.

Fig. 1: D modifications at positions 16 and 17 of human cytoplasmic tRNAs.
figure 1

a Chemical structure of D. b Secondary structure of human cytoplasmic tRNATyr(GUA) with modified nucleosides: 2-methylguanosine (m2G), 3-(3-amino-3-carboxypropyl)uridine (acp3U), N2,N2-dimethylguanosine (m22G), galactosyl-queuosine (galQ), pseudouridine (Ψ), N1-methylguanosine (m1G), 2’-O-methylpseudouridine (Ψm), 7-methylguanosine (m7G), 5-methylcytidine (m5C), 5-methyluridine (m5U), and N1-methyladenosine (m1A). Nucleoside positions are numbered following conventional guidelines61. Modified nucleosides are depicted according to a previous study60. c Neighbor-joining tree of DUS homologs from E. coli, S. cerevisiae, and H. sapiens generated by alignment of protein sequences and depiction of a cladogram using Clustal omega62 and its associated applications. The DUS homologs and their RefSeq numbers are as follows: E. coli DusA (NP_418473.3), DusB (NP_417726.1), and DusC (NP_416645.1); S. cerevisiae Dus1p (NP_013631.1), Dus2p (NP_014412.1), Dus3p (NP_013505.4), and Dus4p (NP_013509.1); and H. sapiens DUS1L (NP_071439.3), DUS2L (NP_001258691.1), DUS3L (NP_064560.2), and DUS4L (NP_853559.1).

The dihydrouridine synthase (DUS) family was first discovered in Escherichia coli and Saccharomyces cerevisiae14,15,16. In eubacteria, three DUS proteins have been identified, each of which is responsible for modifications at specific positions: DusA for D20 and D20a17,18,19, DusB for D1717, and DusC for D1620. In S. cerevisiae, four DUS proteins have been identified: Dus1p for D16 and D17, Dus2p for D20, Dus3p for D47, and Dus4p for D20a and D20b15. Biochemical and structural analyses revealed that DUS is a flavin mononucleotide-dependent enzyme that utilizes NADPH for hydrogen addition to the uracil base16,18,19,21. Yeast Dus2p requires other tRNA modifications for its activity, suggesting that Dus2p-mediated D formation is a later step in tRNA maturation22. Bacterial DusA, DusB, and DusC and the human DUS2L introduced below, commonly recognize the L-shaped structure of tRNA around the “elbow” region, but differ in recognition of tRNAs based on their different orientations, rotation of the target uracil, or use of specific tRNA recognition domains17,20,23,24,25,26.

Corresponding to yeast Dus1–4p, human has four DUS homologs, which were tentatively named DUS1-like (DUS1L) to DUS4-like (DUS4L) in databases. Similar to yeast Dus2p, human DUS2L is responsible for the tRNA D20 modification24. DUS2L is highly expressed in non-small cell lung carcinoma (NSCLC) tissues, and high DUS2L expression in NSCLC is associated with poor prognosis. DUS2L knockdown in a NSCLC cell line reduces proliferation rates. These data suggest that human DUS2L supports pulmonary carcinogenesis27. Human DUS1L, DUS3L, and DUS4L are necessary for D modification at tRNA positions 16, 47, and 20a, respectively28,29. However, these studies did not investigate the sufficiency of D modification enzymes. In addition, the D modification enzymes responsible for modifications at positions 17 and 20b remain unclear.

In this study, by performing DUS1L KO and in vitro reconstitution of D, we show that human DUS1L is responsible for D modifications at tRNA positions 16 and 17. DUS1L supported growth of a glioblastoma (WHO grade IV) cell line, and higher DUS1L expression was associated with poorer prognosis in glioma patients. Moreover, vector-mediated DUS1L overexpression inhibited 5′ and 3′ processing of the tRNATyr(GUA) precursor and reduced the levels of mature tRNATyr and corresponding UAN codon translation. DUS1L overexpression also increased the steady-state levels of several DUS1L-target tRNAs and the overall cellular translation level.

Results

DUS1L is required for D16 and D17 formation in human cytoplasmic tRNAs

Human DUS1L is an ortholog of S. cerevisiae Dus1p (Fig. 1c). To assess the capacity of DUS1L to mediate formation of D16 and potentially also D17 in human cytoplasmic tRNAs, DUS1L was knocked out in two glioblastoma cell lines (LNZ308 and U87) using the CRISPR/Cas9 system. Two single guide RNAs (sgRNAs) were designed against DUS1L exon 3, upstream of a conserved cysteine residue whose counterpart in E. coli DusA is required for D formation19. After transduction of LNZ308 and U87 cells with lentiviruses encoding CRISPR/Cas9, single cells were cloned. We confirmed nucleotide deletions or insertions in the DUS1L gene, which caused frameshifts and generation of premature termination codons (Fig. 2a, Supplementary Fig. 1). Control cells were also generated by the same procedures using control sgRNAs that did not target the human genome. The DUS1L mRNA level was reduced in DUS1L KO cells (Fig. 2b), which suggests that DUS1L mRNAs bared premature termination codons and were thus degraded by the nonsense-mediated mRNA decay pathway30.

Fig. 2: Loss of D16 and D17 modifications in DUS1L KO human cells.
figure 2

a DUS1L alleles in DUS1L KO LNZ308 and U87 glioblastoma cell lines. Red lines indicate deletions and red letters indicate insertions. Aberrant amino acids generated by frameshifting are indicated in blue. DUS1L sgRNA 1 was used to generate LNZ308 and U87 KO1 cells. DUS1L sgRNA 2 was used to generate LNZ308 and U87 KO2 cells. Control cells were generated using sgRNAs that did not target the human genome. b RT-qPCR analysis of DUS1L mRNA levels normalized by GAPDH mRNA levels. Means ± s.e.m. from n = 4 samples each. c LC-MS analysis of total tRNA nucleosides generated by nuclease P1 digestion of total tRNA that was gel-excised after electrophoresis of total RNA. The values are the relative levels to the mean D levels of control 1 and control 2 cells, normalized by the m1G peak area of the same samples. Means ± s.e.m. from n = 4 samples each. d RT-qPCR analysis of DUS1L mRNA levels normalized by GAPDH mRNA levels in control 1 or DUS1L KO2 LNZ308 cells transduced with a lentivirus for stable expression of PAM-mutated synonymous DUS1L or with an empty control lentivirus. Means ± s.e.m. from n = 4 samples each. e Western blot analysis of lysates from the indicated cells using antibodies against DUS1L and ACTB. Endogenous DUS1L was not detected in control cells by western blotting, presumably due to its low expression, similar to many other mammalian tRNA modification enzymes56. f LC-MS analysis of total RNA nucleosides generated by nuclease P1 digestion of total RNA in indicated cells. Means ± s.e.m. from n = 8 samples each. g LC-MS analysis of tRNAPhe nucleosides generated by purification of tRNAPhe from total RNA of the indicated cells and nuclease P1 digestion. Means ± s.e.m. from n = 8 (Control + Empty) or n = 4 (Control + DUS1L and DUS1L KO + Empty) samples. h LC-MS analysis of RNase T1-digested fragments of tRNATyr(GUA) purified from total RNA of the indicated cells. Extracted-ion chromatograms (XICs) of unmodified, single D-modified, and doubly D-modified 3-mer fragments (positions 16–18) with the indicated m/z are shown. NL, normalization level, n.d., not detectable. bd, f, g ****P < 0.0001, ***P < 0.001, **P < 0.01, *P < 0.05, n.s., not significant by Brown–Forsythe and Welch ANOVA tests. Exact P-values are included in the Supplementary Data 1.

There was an approximately 30% decrease of D in total tRNA of DUS1L KO cells according to mass spectrometry of nucleosides obtained by degradation of gel-excised total cellular tRNAs (Fig. 2c). Given there are four human DUS paralogs, a 30% decrease of D upon DUS1L KO is reasonable. The decrease of D in DUS1L KO cells was confirmed to be due to loss of DUS1L by stably re-expressing sgRNA-resistant DUS1L in DUS1L KO cells (Fig. 2d, e, Supplementary Fig. 2), which recovered the D level in total RNA (Fig. 2f). DUS1L was also stably overexpressed in control cells (Fig. 2d–f, Supplementary Fig. 2), which we used to investigate the effect of DUS1L overexpression in later experiments. In addition, we attempted to overexpress catalytically dead mutant DUS1L as a negative control by mutating a cysteine residue whose counterpart in E. coli DusA is required for D formation19. However, upon expression of the DUS1L C107A mutant (equivalent to the E. coli DusA C114A mutant19), cell growth completely ceased (Supplementary Fig. 3), implying that this single point substitution has a dominant negative impact on cell growth, thus underscoring the importance of cysteine 107.

Various human cytoplasmic tRNAs, including tRNAPhe, tRNATyr(GUA), tRNAAla(AGC), tRNAAsn, tRNAHis, and tRNAVal, bear the D16 and/or D17 modifications4. To investigate whether or not DUS1L installs D16/D17 modifications into these tRNAs, cytoplasmic tRNAPhe, which has D at positions 16, 17, and 474, was purified using an oligo DNA probe and subsequently subjected to mass spectrometric analysis to monitor the D level. The level of D was reduced in tRNAPhe in DUS1L KO cells (Fig. 2g, red bar). Considering that D47 is modified by DUS3L28, this result strongly suggests that DUS1L is required for D16/D17 modification of tRNAPhe. To provide definitive evidence that DUS1L is required for D formation at tRNA positions 16 and 17, cytoplasmic tRNATyr(GUA) was purified using a DNA probe, digested with G-specific RNase T1, and subjected to mass spectrometric analysis. tRNATyr(GUA) fragments derived from control cells contained a 3-nt DDG fragment with both D16 and D17 modifications and lacked both the detectable unmodified UUG triplet as well as the single D-modified fragments (Fig. 2h). Thus, tRNATyr(GUA) in control cells was fully D-modified at positions 16 and 17. On the contrary, DUS1L KO cells completely lacked both D16 and D17 modifications in tRNATyr(GUA), and only the unmodified UUG fragment was detected. The loss of D16 and D17 modifications in DUS1L KO cells was rescued by stable re-expression of DUS1L (Fig. 2h). Therefore, DUS1L is required for both D16 and D17 modifications in cytoplasmic tRNA of human cells.

In vitro reconstitution of D16 and D17 with recombinant DUS1L

To investigate whether DUS1L is the enzyme that catalyzes tRNA D16 and D17 modifications, we performed in vitro reconstitution of D modifications using recombinant DUS1L protein. For the substrate RNA, rather than using an in vitro synthesized, unmodified tRNA transcript, we used total RNA of DUS1L KO cells. This is because confocal microscopic observation of C-terminally Myc-tagged human DUS1L (DUS1L-Myc) showed a predominantly cytoplasmic localization with a minor population localized to the nucleus (Fig. 3a), implying that DUS1L functions in the late steps of tRNA maturation. In addition, human DUS2L similarly localizes predominantly in the cytoplasm27, and yeast Dus2p requires prior installation of other tRNA modifications for its DUS activity22. We expressed hexahistidine-tagged DUS1L (His-DUS1L) in E. coli and purified His-DUS1L using Co2+-chelating beads (Fig. 3b). After the in vitro reconstitution reaction, the D level in total RNA increased only in the presence of both recombinant His-DUS1L and NADPH (Fig. 3c). Following the in vitro D reconstitution, tRNATyr(GUA) was further purified from total RNA, digested with RNase T1, and subjected to mass spectrometric analysis. In the presence of both NADPH and His-DUS1L, uridines at positions 16 and 17 were fully D-modified in tRNATyr(GUA) (Fig. 3d). Therefore, the human DUS1L enzyme catalyzes tRNA D16 and D17 modifications in vitro, without the need for additional partner proteins.

Fig. 3: In vitro reconstitution of tRNA D16 and D17 modifications.
figure 3

a Subcellular localization of human DUS1L-Myc. U87 cells were transfected with a plasmid expressing DUS1L-Myc. Confocal microscopic images were obtained after immunostaining of Myc-tagged proteins (green) and DAPI staining (blue). Scale bar, 10 µm. b Purified His-DUS1L resolved by SDS-PAGE and stained with Coomassie Brilliant Blue. Protein size markers are indicated. c LC-MS analysis of total RNA nucleosides after incubating the reaction mixture with (+) or without (−) His-DUS1L or NADPH. Total RNA from DUS1L KO or control LNZ308 cells was used as the substrate RNA. The values are the relative levels to the mean D levels in total RNA of control cells, normalized by the uridine peak area of the same samples. Means ± s.e.m. from n = 5 samples each. d LC-MS analysis of RNase T1-digested fragments of tRNATyr(GUA) purified from total RNA after incubating the reaction mixture with or without His-DUS1L or NADPH, using total RNA from DUS1L KO LNZ308 cells as the substrate RNA. XICs of unmodified, single D-modified, and doubly D-modified 3-mer fragments (positions 16–18) with the indicated m/z are shown. e LC-MS analysis of total RNA nucleosides after incubating the reaction mixture in the same way as in (c), with (+) or without (−) NADPH or NADH. Means ± s.e.m. from n = 4 samples each. ****P < 0.0001, *P < 0.05, n.s., not significant by the Brown–Forsythe and Welch ANOVA tests. Exact P-values are included in the Supplementary Data 1.

Bacterial DusA can use not only NADPH, but also NADH for the D modification18. In the case of eukaryotic DUS, although one study showed that recombinant human DUS2L can oxidize NADH24, no study likely has examined whether eukaryotic DUS can use NADH instead of NADPH for the D modification of tRNAs. To investigate this, an in vitro reconstitution experiment was performed in the same way as in Fig. 3c, except that NADH was used (Fig. 3e). As a result, NADH has been shown to be as effective as NADPH for the installation of D modification, producing similar D modification levels between the two, at least in terms of the end point of the reactions.

DUS1L supports cell growth

To investigate if DUS1L is involved in cell proliferation, the proliferation rates of the cultured glioblastoma cell line LNZ308 were monitored. DUS1L KO reduced the growth rates to two-thirds of that of control cells, and this effect was fully reversed upon stable re-expression of CRISPR-resistant DUS1L (Fig. 4a). Moreover, DUS1L overexpression increased the growth rates of LNZ308 cells (Fig. 4b). Therefore, DUS1L supports growth of a glioblastoma cell line under a cultured condition.

Fig. 4: Effect of DUS1L expression on the growth speed of a glioblastoma cell line.
figure 4

Growth of DUS1L KO (a) and DUS1L OE (b) cells. The cell lines are the same as those used in Fig. 2. The cells were seeded at the same density at 0 h, and cell density was measured after 24, 48, 72, and 96 h using a Cell Counting Kit 8 and normalized by the cell density at 24 h. Means ± s.e.m. from n = 8 samples at each time point and cell type. ****P < 0.0001, n.s. not significant compared with control cells by a two-way ANOVA followed by a multiple comparison test. c Kaplan–Meier plot of glioma patients with higher or lower levels of tissue DUS1L mRNA expression and the P-value between the two groups (**P = 0.0024) acquired from the Human Protein Atlas database31. The cut-off FPKM value between high and low DUS1L expression was set to 7.47 in order to discern 76 patients with high DUS1L expression and 77 patients with low DUS1L expression.

Increased expression of tRNA modification enzymes is often associated with cancer progression and poor prognosis5,6,7. To investigate the possible association of DUS1L expression with cancer progression, the survival probabilities of glioma patients with higher and lower DUS1L mRNA expression levels were compared using data in the Human Protein Atlas database31. A Kaplan–Meier plot showed that increased expression of DUS1L was associated with poorer prognosis of glioma patients (Fig. 4c).

DUS1L overexpression differentially affects the steady-state levels of the selected DUS1L-target tRNAs and increases the total cellular translation

tRNA modifications are generally considered to support cell growth and functions by supporting protein synthesis3,5,6. To investigate the possible effect of DUS1L on tRNA levels and protein synthesis, we first monitored the steady-state levels of several D16- and/or D17-modified tRNAs4. Specifically, we performed northern blotting of D16/D17-bearing tRNATyr(GUA), tRNAPhe, and tRNAAla(AGC), as well as D16/D17-lacking initiator methionine tRNA (tRNAiMet) as a control and found no changes in the steady-state levels of these tRNAs upon DUS1L KO (Fig. 5a, Supplementary Figs. 4a, 5). The overall cellular translation, monitored by 35S-methionine pulse labeling of nascent proteins, also did not show a clear difference in the overall translation level between DUS1L KO and control cells (Supplementary Fig. 4b).

Fig. 5: Effects of DUS1L overexpression on mature tRNA levels and total cellular translation.
figure 5

Mature tRNA levels in DUS1L KO (a) and DUS1L OE (b) cells. Northern blot analysis of tRNATyr(GUA), tRNAPhe(GAA), tRNAAla(AGC), and initiator tRNAMet(CAU) was performed, followed by band quantification and normalization by 5.8S rRNA. The original images are shown in Supplementary Figs. 4 and 5. Means ± s.e.m. from n = 4 samples each. *P < 0.05 by the Mann–Whitney test. c Nascent cellular protein synthesis observed by 35S-methionine pulse labeling. A radiation image of the electrophoresed gels is shown on the left. Coomassie brilliant blue (CBB) staining of the same gel is shown on the right as a loading control. n = 4 samples each. d Quantification of the nascent protein levels in (c). Means ± s.e.m. from n = 4 samples each. *P < 0.05 by the Mann–Whitney test. Exact P-values are included in the Supplementary Data 1.

By contrast, using the same stable DUS1L-overexpressing (OE) cells as in the experiments presented in Figs. 2 and 4, we found that DUS1L OE reduced tRNATyr(GUA) by approximately half compared to the control cells (Fig. 5b). On the other hand, DUS1L overexpression slightly increased the steady-state level of D16/D17-modified tRNAPhe (Fig. 5b, Supplementary Figs. 4c, 5), although it did not change the D level in tRNAPhe (Fig. 2g). In addition, DUS1L overexpression also slightly increased the steady-state level of D17-modified4 tRNAAla(AGC), whereas the control D16/D17-lacking tRNAiMet was unchanged (Fig. 5b, Supplementary Figs. 4c, 5). Interestingly, a 35S-methionine pulse-labeling experiment showed that DUS1L overexpression also increased overall cellular translation (Fig. 5c, d). We conclude that DUS1L overexpression differentially impacts the steady-state levels of several D16- and/or D17-modified tRNAs and increases the overall cellular translation. Since the levels of some tRNAs rise up, while those of tRNATyr(GUA) drop down dramatically, the latter effect cannot be explained by a general increase in the amount of translational components, and its elucidation awaits further investigation.

DUS1L overexpression suppresses 5′ and 3′ processing of the tRNATyr(GUA) precursor and reduces the mature tRNATyr(GUA) level

Northern blotting of tRNATyr(GUA) (Fig. 5b) revealed that DUS1L OE cells accumulated a tRNATyr band that was approximately 20–30 nt longer than wild type tRNATyr (Fig. 6a, Supplementary Fig. 5). We hypothesized that this band might be a tRNATyr precursor. To investigate this possibility, we first compared the D modification level of mature tRNATyr(GUA) versus the elongated band. The tRNATyr-containing RNAs were enriched from total RNA of DUS1L OE cells using a biotinylated oligo DNA probe, and both forms were gel excised and subjected to nucleoside digestion followed by mass spectrometric analysis of the D level normalized to the cytidine level. We found that the elongated band contained about half the amount of D compared with mature tRNATyr(GUA) (Fig. 6b). This result supported our precursor hypothesis and further suggested that DUS1L installs D16 and D17 in a step-wise manner, while stably interacting with this tRNATyr(GUA). To further characterize the elongated tRNATyr precursor, tRNATyr-containing RNAs were enriched from total RNA of DUS1L OE cells using a biotinylated oligo DNA probe, followed by gel excision of the elongated tRNATyr, 3′-adapter ligation, reverse transcription, PCR amplification, cloning, and Sanger sequencing (Fig. 6c). As expected, the elongated tRNATyr was identified as the tRNATyr(GUA) precursor retaining the 5′ leader and 3′ trailer sequences up to the oligo(U) transcriptional termination site but with the intron already spliced out (Fig. 6d, e). Accumulation of this tRNATyr(GUA) precursor in DUS1L OE cells was confirmed by northern blotting using a probe against the trailer sequence of tRNATyr(GUA) (Fig. 6f). Thus, DUS1L overexpression inhibits removal of the 5′ leader and 3′ trailer sequences of tRNATyr(GUA) and reduces the mature tRNATyr(GUA) level.

Fig. 6: Effects of DUS1L overexpression on tRNATyr(GUA) maturation.
figure 6

a Northern blot analysis of tRNATyr(GUA) in DUS1L OE and control LNZ308 cells. Each lane was derived from different cell dishes. The lower band corresponds to mature tRNATyr(GUA) and the upper band is elongated tRNATyr(GUA). The positions of single-stranded DNA size markers are shown on the left. b Comparison of D levels in mature tRNATyr(GUA) and the elongated band, which were enriched using an oligo DNA probe against tRNATyr(GUA) and gel excision, followed by nuclease digestion and mass spectrometry analysis, normalized by cytidine levels. Means ± s.e.m. from n = 4 samples each. *P = 0.0286 by the Mann–Whitney test. c Outline of the method to identify the sequence of the elongated tRNATyr(GUA). d Schematic of the gene structure, transcription, and post-transcriptional processing of human tRNATyr(GUA). Pol III transcription terminates after producing the oligo(U) stretch63. The tRNATyr(GUA) precursor contains the 5′ leader, intron, and 3′ trailer sequences, which are removed by RNase P, the tRNA splicing complex, and RNase Z, respectively, followed by the 3′ terminal CCA addition64. e Sanger sequencing results of the elongated tRNATyr(GUA). The upper four clones were analyzed by PCR using a forward primer designed to target the 5′ leader (underlined sequences) and a reverse primer designed against the 3′ adapter. The lower four clones were analyzed by PCR using a forward primer designed to target the 5′ end inside the tRNATyr(GUA) sequence (underlined sequences) and a reverse primer designed against the 3′ adapter. f Northern blot analysis using a probe designed against the tRNATyr(GUA) trailer sequence. Each lane was derived from different cell dishes.

DUS1L overexpression reduces translation of tyrosine codons and readthrough of UAA and UAG stop codons

To investigate if reduction of the tRNATyr(GUA) levels upon DUS1L overexpression affects efficiency of translation of the tyrosine UAC and UAU codons, we performed codon translational reporter experiments using constructs with five consecutive codons of interest between the firefly luciferase and renilla luciferase genes. Translation of UAC and UAU codons was reduced by 23% and 21% in DUS1L OE cells, respectively (Fig. 7a). Thus, the DUS1L overexpression-mediated reduction of tRNATyr(GUA) expectedly decreases translation efficiency of the UAC and UAU codons.

Fig. 7: Effects of DUS1L overexpression on translation of tyrosine codons and translational readthrough of near-cognate stop codons.
figure 7

a Effects of DUS1L overexpression on translation of tyrosine codons. Cells were transfected with a codon translation reporter plasmid with a 5′ firefly luciferase sequence and a 3′ renilla luciferase sequence connected via five consecutive TAC, TAT, or random codons (control). Plasmid expression was normalized by dividing renilla luciferase luminescence by firefly luciferase luminescence. Possible renilla-to-firefly luciferase translational bias between control and DUS1L OE cells due to codons in luciferase genes was normalized using the random codon linker control. b Effects of DUS1L overexpression on translational readthrough of UAA and UAG stop codons. Cells were transfected with a stop codon readthrough reporter plasmid with a 5′ renilla luciferase sequence and a 3′ firefly luciferase sequence connected via a TAA codon (or its control TTA codon) or TAG codon (or its control TTG codon). Plasmid expression was normalized by dividing firefly luciferase luminescence by renilla luciferase luminescence. Possible firefly-to-renilla luciferase translational bias due to codons in luciferase genes was normalized using the corresponding TTA/TTG linker control. Means ± s.e.m. from n = 6 samples each. **P < 0.01, *P < 0.05 by the Mann–Whitney test. Exact P-values are included in the Supplementary Data 1.

tRNATyr(GUA), which decodes UAC and UAU codons, can sometimes mistranslate near-cognate UAA and UAG stop codons, by assigning a tyrosine to them, allowing so-called non-programmed stop codon readthrough32. To investigate whether the reduction of tRNATyr(GUA) upon DUS1L overexpression affects stop codon readthrough, we performed stop codon readthrough reporter luciferase experiments. UAA and UAG readthrough was decreased in DUS1L OE cells (Fig. 7b). Therefore, the DUS1L overexpression-mediated reduction of tRNATyr(GUA) decreases mistranslation of UAA and UAG codons.

Discussion

In the present study, we identified DUS1L as the human enzyme responsible for D modifications at tRNA positions 16 and 17. Many tRNA modifications support mRNA translation and cell growth3,5. However, upon DUS1L KO, global nascent protein synthesis was not detectably reduced (Supplementary Fig. 4b), but cell growth was decreased (Fig. 4a). To understand this apparent discrepancy, the tRNA and codon translation levels may need to be monitored globally. Many enzymes that were previously regarded as tRNA modification enzymes were recently recognized to also serve as mRNA modification enzymes33,34,35. In the case of D, hundreds of D sites were recently found in yeast and human mRNAs29,36. Moreover, yeast Dus1–4p are responsible for D modifications of yeast mRNAs29. Our study may serve as a foundation to investigate the roles of D modifications not only in human tRNAs but also in human mRNAs. To investigate the possible roles of D modifications in mRNAs, performance of a combination of analyses such as ribosome profiling and transcriptome-wide mapping of D sites would be useful.

In DUS1L OE cells, the tRNATyr(GUA) intron that extends from the anticodon loop of the tRNATyr(GUA) precursor was spliced out, while 5′ and 3′ processing of the tRNATyr(GUA) precursor was inhibited. RNase P is responsible for maturation of the 5′ end of precursor tRNAs. Human nuclear RNase P interacts with the T-loop, D-loop, and acceptor stem of precursor tRNAs37,38. By contrast, the tRNA splicing endonuclease interacts with intron-containing tRNAs by recognizing the anticodon stem and some regions of the D-arm and acceptor stem but without touching the tRNA “elbow”39,40. Thus, the small fraction of DUS1L that was observed in the nucleus (Fig. 3a) may have been bound to the tRNATyr(GUA) precursor immediately after its transcription around the tRNA “elbow”, similar to other DUS enzymes that bind to this region24,26. Suppression of tRNA processing by a tRNA modification enzyme likely has not been reported previously. Thus, although it is in an overexpression context, this study may expand our knowledge of the functions of tRNA modification enzymes. It should be stressed here that tRNA 5′ and 3′ processing was likely not inhibited in other monitored D16/D17-containing tRNAs (tRNAPhe and tRNAAla(AGC)) (Supplementary Figs. 4c, 5). Currently, the cause for these differences is unclear; for example, it might be due to different DUS1L modes of binding to or affinity for different tRNA species.

As could have been anticipated, suppression of 5′ and 3′ processing of tRNATyr(GUA) upon DUS1L OE was associated with reduction in the mature tRNATyr(GUA) levels. Correctly processed 5′ and 3′ ends in the nucleus are recognized by the Exportin-t, which only exports such processed tRNAs to the cytoplasm41,42. Thus, upon DUS1L overexpression, suppression of 5′ and 3′ processing of precursor tRNATyr(GUA) likely inhibited nuclear export of unprocessed tRNATyr(GUA), allowing unprocessed tRNATyr(GUA) to be targeted by the nuclear RNA degradation machinery.

We unexpectedly observed slight but significant increases of mature tRNAPhe and tRNAAla(AGC) levels upon DUS1L overexpression (Fig. 5b, Supplementary Figs. 4c, 5). The increased amounts of several tRNAs might partially explain the increases of both cellular translation (Fig. 5c, d) and growth rates (Fig. 4b). However, this explanation must be taken with caution, because: (1) tRNATyr(GUA) levels drop down dramatically and (2) there are other D16/D17-modified tRNAs4 with so far unknown impact of DUS1L overexpression on their steady-state levels. In yeast and Drosophila, overexpression of elongation factor 1 A, which binds to aminoacylated tRNAs, protects certain hypomodified tRNAs from degradation by the rapid tRNA decay pathway43,44. It is tempting to speculate that overexpressed DUS1L protein might similarly bind to and protect D16/D17-target tRNAs from degradation, but such a speculation demands future investigations.

Upon DUS1L overexpression, translation of UAC and UAU tyrosine codons and translational readthrough of near-cognate UAA and UAG stop codons were fittingly decreased (Fig. 7), which is likely attributable to a decreased level of mature tRNATyr(GUA) (Fig. 5b). In addition, higher DUS1L expression was associated with poorer prognosis of glioma patients and DUS1L affected the growth rates of the selected cancer cell line (Fig. 4). UAA and UAG stop codon readthrough and tyrosine codon translation may affect the quality, quantity, and constituents of the proteome, and thus the cell growth. Further investigations are required to determine whether increased DUS1L expression in cancer tissues, which is likely lower than the expression level achieved in vector-mediated DUS1L OE cells (Fig. 2d), changes the proteome of cancer cells. To investigate this possibility in glioma patients, it is necessary to first investigate if glioma tissues with higher DUS1L expression contain more tRNATyr(GUA) precursor, tRNAPhe, and tRNAAla(AGC) and less mature tRNATyr(GUA). In addition, ribosome profiling would be useful to examine possible changes in sense codon dwelling times and stop codon readthroughs on UAR stops on translatome-wide level with a single nucleotide-level resolution.

In summary, we identified DUS1L as the human tRNA D16 and D17 modification enzyme, providing the foundation to study the roles of D modifications in physiology and pathophysiology.

Methods

Cell culture

The glioblastoma cell lines LNZ308 and U87 (gifts from Dr. Takahiro Yamamoto, Kumamoto University) were grown in Dulbecco’s modified Eagle’s medium (Thermo Fisher Scientific, 11995) supplemented with 10% fetal bovine serum (FBS) at 37 °C in a humidified atmosphere containing 5% CO2. Mycoplasma tests have not been performed. For cell growth analysis, cells were trypsinized and 1000 cells per 100 µL for each well were seeded in four 96-well plates and cultured in a CO2 incubator for 24, 48, 72, or 96 h. As blank controls, 100 µL of medium were transferred to four 96-well plates and put in a CO2 incubator for 24, 48, 72, or 96 h. The relative cell density was measured using a Cell Counting Kit 8 (Dojindo, CK04) according to the manufacturer’s protocol. Briefly, 10 µL of Cell Counting Kit 8 reagent was added to the wells containing cells or control medium and pipetted several times, followed by incubation in a 37 °C CO2 incubator for 1 h. Subsequently, the cell-containing 96-well plate and blank control 96-well plate were subjected to absorbance measurement at 450 nm using a Sunrise plate reader (TECAN). The absorbance in the blank medium well was subtracted as a background. The cell density was normalized by that at 24 h.

Immunofluorescence

Immunofluorescence experiments were performed essentially as previously described45,46. DUS1L cDNA was amplified and a C-terminal Myc tag was added by reverse transcription and RT-PCR of total RNA from U87 cells using the primers listed in Supplementary Table 1. The amplified DNA was purified by agarose gel excision and cloned using a pcDNA3.4 TOPO Cloning Kit (Thermo Fisher Scientific, A14697) to generate the DUS1L-Myc expression vector (DUS1L-Myc/pcDNA3.4). U87 cells were grown on a 35 mm glass-bottom dish (Iwaki, 3911-035). The next day, the cells were transfected with 2.5 µg of DUS1L-Myc/pcDNA3.4 plasmid DNA using Lipofectamine 3000 (Thermo Fisher Scientific, L3000015). After 2 days, the cells were briefly washed with phosphate-buffered saline (PBS), fixed with 4% paraformaldehyde diluted in PBS (Wako, 10010) at room temperature for 10 min, washed thrice with PBS, and permeabilized with 0.2% Triton X-100 diluted in PBS at room temperature for 15 min. The cells were washed thrice with PBS containing 0.1% Tween (PBST), blocked with 1% bovine serum albumin diluted in PBST at room temperature for 1 h, and incubated with an anti-Myc antibody (MBL, M047-3, 1:500 dilution in PBST) at room temperature for 1.5 h. The cells were then washed thrice with PBST and incubated with an anti-mouse secondary antibody conjugated with Alexa Fluor 488 (Thermo Fisher Scientific, A21202, 1:500 dilution in PBST) at room temperature for 1 h. The cells were washed twice with PBST, stained with DAPI, and washed again with PBST. Finally, images were acquired using a FLUOVIEW FV3000 confocal microscope (Olympus).

Generation of DUS1L KO and control cells

DUS1L KO cells were generated using the CRISPR/Cas9 system, as described previously47,48. Briefly, each set of sense and antisense oligonucleotides (two sets targeting DUS1L and two sets that did not target the human genome as negative controls, Supplementary Table 1) encoding a sgRNA were selected from the human GECKO v2 library49. Sense and antisense oligonucleotides were annealed and cloned into the BsmBI sites of the lentiCRISPR v2 blast plasmid (Addgene, 12260). Lentiviruses were generated by transfecting HEK293FT cells with the sgRNA sequence-containing lentiCRISPR v2 blast plasmid, psPAX2 (Addgene, 12260), and pMD2.G (Addgene, 12259). After 1 day, the lentivirus-containing medium was collected, filtered through a 0.45 µm Millex-HV filter (Millipore, SLGVR33RS), and added to LNZ308 or U87 cells for transduction of the generated lentiviruses. After blasticidin selection of transduced cells, single clones were acquired by limiting dilution.

RNA extraction and RT-qPCR

LNZ308 or U87 cells in 10-cm dishes were briefly washed with PBS and lysed in 1 mL of TRI Reagent (MRC, TR118), and then total RNA was extracted according to the manufacturer’s protocol. RT-qPCR was performed as previously described50 using PrimeScript RT Master Mix (TAKARA, RR036A), Rotor Gene 2 (Qiagen), TB Green Premix Ex Taq II (TAKARA, RR820), and the primers listed in Supplementary Table 1.

Stable re-expression of DUS1L in LNZ308 cells

DUS1L-Myc was PCR-amplified from the DUS1L-Myc/pcDNA3.4 plasmid using the primers listed in Supplementary Table 1 to add the CACC sequence at the 5′ terminus. To generate the DUS1L-Myc/pENTR plasmid, the PCR product was gel-excised and subcloned into the pENTR plasmid using a Directional TOPO Cloning Kit (Thermo Fisher Scientific, K240020SP). To re-express DUS1L-Myc in DUS1L KO2 cells (which stably expressed CRISPR/Cas9 with a sgRNA against the DUS1L coding sequence), the PAM-containing sequence CTGGCT (PAM underlined, encoding a leucine and an alanine) was synonymously mutated to CTAGCT in order to destroy the PAM sequence. The mutation was performing by PCR of the DUS1L-Myc/pENTR plasmid using the primers listed in Supplementary Table 1. Subsequently, the PCR product was transformed into E. coli to obtain the PAM-destroyed DUS1L-Myc/pENTR plasmid. To generate C107A mutant DUS1L-Myc/pENTR, WT DUS1L-Myc/pENTR was modified using the primers listed in Supplementary Table 1. WT or C107A mutant DUS1L-Myc was subcloned into pLX302 (Addgene, 25896) using LR Clonase II Enzyme Mix (Thermo Fisher Scientific, 11791020). HEK293FT cells were transfected with PAM-destroyed DUS1L-Myc/pLX302 (or empty pLX302 as the empty control), psPAX2 (Addgene, 12260), and pCMV-VSVG (Addgene, 8454) using Lipofectamine 3000 according to the manufacturer’s protocol. The medium was replaced with fresh medium approximately 12 h post-transfection. At 24 h after the medium change, medium containing DUS1L-Myc-encoding lentivirus particles (or empty control lentivirus particles) was collected, filtered, and added to Control 1 or DUS1L KO2 LNZ308 cells. After 4 days of culture in DMEM containing 10% FBS, transduced LNZ308 cells were selected by culture in medium containing 2 µg mL−1 puromycin for 3 days.

Western blotting

Western blotting was performed as previously described51,52. Briefly, cells were trypsinized, washed with PBS, pelleted, lysed with lysis buffer (150 mM NaCl, 100 mM Tris-HCl pH 8, 0.5% NP-40, and a protease inhibitor cocktail (Roche, 05056489001)), and sonicated for 10 s. The protein concentration was determined using a Pierce BCA Protein Assay Kit (Thermo Fisher Scientific, 23225). Samples were electrophoresed on a sodium dodecyl sulfate (SDS) polyacrylamide gel and transferred to an Immobilon-P membrane (Merck Millipore, IPVH00010). The membrane was blocked with 5% skim milk prepared in TBST buffer (150 mM NaCl, 25 mM Tris-HCl pH 7.4, 2.7 mM KCl, and 0.05% Tween-20) and probed using a primary antibody against DUS1L (Santa Cruz, sc-393586, 1:100) or actin (MBL, M177-3, 1:5000) diluted in 5% skim milk prepared in TBST buffer at 4 °C overnight. The membrane was washed in TBST, probed using a secondary antibody (Dako, 0477, 1:4000) at room temperature for 1 h, and washed in TBST. Signals were detected using ECL Prime Western Blotting Detection Reagent (Cytiva, RPN2236) and a ChemiDoc imager (Bio-Rad).

RNA nucleoside mass spectrometry

RNA nucleoside mass spectrometry was performed as previously described53,54. Briefly, 20 µL of solution containing total RNA (3 µg for LCMS-8050 or 1 µg for LCMS-8060), 20 mM Hepes-KOH (pH 7.6), 1.6 U of nuclease P1 (Fujifilm, 145-08221), and 0.2 U of bacterial alkaline phosphatase (Takara, 2120 A) was incubated at 37 °C for 3 h. Subsequently, 3 µL of the nucleoside solution was injected into a LCMS-8050 or LCMS-8060 system (Shimadzu). The nucleosides were first separated by an Inertsil ODS-3 column (GL Science, 5020-84655) using a mobile phase that continuously changed from 100% solution A (5 mM ammonium acetate in water, pH 5.3) to 100% solution B (60% acetonitrile in water) over 17 min at a flow rate of 0.4 mL min−1, followed by electrospray ionization and triple quadrupole mass spectrometry in multiple reaction monitoring mode.

Isolation of tRNAPhe and tRNATyr(GUA) from total RNA

tRNAPhe and tRNATyr(GUA) were isolated from total RNA as previously described47. Briefly, 2 nmol of the 3′ biotinylated DNA probe (Supplementary Table 1) was bound to streptavidin beads (GE Healthcare, 17511301) at room temperature in binding buffer (100 mM NaCl, 10 mM Hepes-KOH pH 7.6, and 5 mM EDTA) for 1 h, followed by washing with binding buffer and hybridization buffer (1200 mM NaCl, 30 mM Hepes-KOH pH 7.6, and 7.5 mM EDTA). The probe-bound beads were mixed with 100 µg of total RNA in hybridization buffer and incubated at 65 °C for 1 h with occasional agitation. The beads were washed ten times with wash buffer (600 mM NaCl, 30 mM Hepes-KOH pH 7.6, and 7.5 mM EDTA) at 65 °C and eluted with elution buffer (20 mM NaCl, 0.5 mM Hepes-KOH pH 7.6, and 0.25 mM EDTA) at 65 °C. Subsequently, the eluate was subjected to 7 M urea/Tris-borate-EDTA (TBE)/10% polyacrylamide gel electrophoresis (PAGE), followed by SYBR Gold staining and gel excision of the tRNA band.

RNase T1 fragment mass spectrometry

Mass spectrometry of tRNA fragments digested by RNase T1 was performed essentially as previously described55. The isolated tRNATyr(GUA) (approximately 50 ng) was digested in 10 µL of reaction mixture containing 20 mM NH4OAc (pH 5.3) and 2 U µL−1 RNase T1 at 37 °C for 30 min and then mixed with a 1/10 volume of 0.1 M triethylamine-acetate (pH 7.0). The digests were subjected to capillary liquid chromatography (LC)/nano-electron spray ionization mass spectrometry (ESI-MS) using a DiNa splitless nano HPLC system (Techno Alpha) connected to a Q Exactive Hybrid Quadrupole-Orbitrap mass spectrometer (Thermo Fisher Scientific). The digests were loaded on a trap column (L-column2 ODS, 5 μm particle size, 0.3 × 5 mm; Chemical Evaluation and Research Institute) and separated on a HiQ sil C18W-3 nanospray column (C18, 3 μm particle size, 0.1 × 100 mm; Techno Alpha) with a linear gradient of 2.5–50% methanol containing 0.4 M 1,1,1,3,3,3-hexafluoro-2-propanol (pH 7.0, adjusted with triethylamine) over 35 min at a flow rate of 300 nL min−1. The scan range was m/z 600–2000.

Expression and purification of recombinant His-DUS1L protein

The human DUS1L coding sequence with codons optimized for E. coli was synthesized by Azenta Life Sciences. The sequence was amplified using the primers listed in Supplementary Table 1 to add 5′ NheI and 3′ BamHI recognition sequences. The PCR amplicon was cloned into the pET28a (+) vector (Novagen) to obtain the expression vector pET-DUS1L, which produced N-terminal hexahistidine tag-fused DUS1L (His-DUS1L). One Shot BL21 Star DE3 cells (Thermo Fisher Scientific, C601003) were transformed with pET-DUS1L and cultured in LB media containing 50 µg mL−1 kanamycin. When the bacteria reached an OD of 0.5, protein expression was induced by addition of 100 µM isopropyl β-D-thiogalactopyranoside, and the cells were grown for 3 h at 37 °C. Cells were harvested, and the pellet (2 g) was suspended in 40 mL of lysis buffer (50 mM Hepes-KOH pH 7.6, 250 mM KCl, 1 mM 2-mercaptoethanol (2-ME), 1× protease inhibitor cocktail (Roche, 05056489001), 200 µg mL−1 lysozyme (Fujifilm, 129-06723), and 1/1000 volume DNaseI (Roche, 4716728001)). The lysate was sonicated on ice using a Branson 250 Sonifier (power 6, total of 3 min) and cleared by centrifugation at 12,000 × g for 20 min. Thereafter, 1 mL of 50% slurry TALON Metal Affinity Resin (TAKARA, 635502) was washed with equilibration buffer (lysis buffer without lysozyme and DNaseI), mixed with the E. coli lysate, and rotated at 1 h at 4 °C. The beads were cleared and washed with 10 mL of the following buffers: twice with equilibration buffer, four times with wash buffer 1 (20 mM Hepes-KOH, pH 7.6, 500 mM NaCl, 0.5% Triton X-100, and 1 mM 2-ME), four times with equilibration buffer, and once with wash buffer 2 (50 mM Hepes-KOH pH 7.6, 250 mM KCl, 1 mM 2-ME, and 10 mM imidazole pH 7.6). Ten elution fractions were collected each using 600 µL of elution buffer (50 mM Hepes-KOH pH 7.6, 250 mM KCl, 1 mM 2-ME, and 150 mM imidazole pH 7.6). Fractions 3–6 were collected, dialyzed overnight in dialysis buffer (50 mM Hepes-KOH pH 7.6, 150 mM KCl, and 1 mM dithiothreitol (DTT)), aliquoted, snap-frozen in liquid nitrogen, and stored at −80 °C until the in vitro reconstitution experiment.

In vitro reconstitution of D using recombinant His-DUS1L

In vitro D reconstitution was performed essentially as previously described for DUS2L24. The in vitro reconstitution reaction mixture (50 µL) contained 50 mM Tris-HCl (pH 7.5), 100 mM CH3COONa, 10 mM MgCl2, 2 mM DTT, 1 mM NADPH (or 1 mM NADH), 15% glycerol, 1 µM His-DUS1L, and 25 µg of total RNA from DUS1L KO LNZ308 cells. The reaction mixture was incubated at 30 °C for 2 h. The reaction was terminated by adding TRI Reagent, and RNA was recovered according to the manufacturer’s protocol.

Northern blotting

Northern blotting was performed as previously described56. Total RNA (1.5 µg) was first separated by 7 M urea/TBE/10% PAGE. The gel was stained with SYBR Gold to check RNA quality and then transferred to a nylon membrane (Merck Millipore, INYC00010) in 1× TBE using a wet transfer blotting system (Bio-Rad) on ice at 50 V for 80 min. Membranes were dried and crosslinked with ultraviolet light at 1200 × 100 µJ cm−2 using HL-2000 Hybrilinker (Funakoshi) and incubated in prehybridization buffer (6× saline sodium citrate (SSC), 0.1% SDS, and 1× Denhardt’s solution) at 42 °C for 1 h. The membranes were then hybridized with digoxigenin (DIG)-labeled probe DNA (Roche) in hybridization buffer (900 mM NaCl, 90 mM Tris-HCl pH 8, 6 mM EDTA, and 0.3% SDS) overnight at 50 °C. The probes for tRNAPhe, tRNAiMet, and 5.8S rRNA were the same as those used in a previous study57. The probes for tRNATyr(GUA) and tRNAAla(AGC) were designed for this study. The sequences of northern blot probes are listed in Supplementary Table 1. The membranes were washed with 1× SSC, blocked using a DIG Wash and Block Buffer Set (Roche, 11585762001), and probed with anti-DIG alkaline phosphatase Fab fragments (Roche, 11093274910) and CDP-Star (Roche, 12041677001). Images were acquired using a ChemiDoc imager (Bio-Rad). The band intensities were quantified using ImageJ software58.

35S-methionine labeling of nascent proteins

Pulse labeling of nascent proteins was performed essentially as previously described47,48. One day before radiolabeling, 6 × 105 cells were seeded into 6-cm dishes. The next day, the cells were briefly washed using 37 °C DMEM without Met, Cys, and Gln (Thermo Fisher Scientific, 21013024). A total of 3 mL of 37 °C preincubation medium (DMEM without Met, Cys, or Gln, supplemented with 2% FBS, 2 mM Gln, and 0.2 mM Cys) was added to the cells, and the cells were incubated in a CO2 incubator at 37 °C for 15 min. The medium was then exchanged with 1.25 mL of incubation medium (1.25 mL of preincubation medium containing 8.88 MBq of 35S-labeled methionine [PerkinElmer]), and the cells were incubated in a CO2 incubator at 37 °C for 30 min. Subsequently, the medium was removed and the cells were washed with PBS and collected using trypsin and DMEM containing 10% FBS. Cells were lysed and the protein concentration was measured using a BCA Protein Assay Kit (Thermo Fisher Scientific, 23225). Then, 15 µg of total protein were run on a Tricine PAGE gel (NOVEX, EC66252BOX). The gel was stained using Coomassie Brilliant Blue staining solution (Bio-Rad, 1610436), dried on a gel dryer, and photographed. A radiation image was acquired using an imaging plate and Typhoon FLA 7000 imager (GE Healthcare).

Identification of the tRNATyr-containing band that accumulated upon DUS1L overexpression

tRNATyr-containing RNAs were isolated from total RNA of DUS1L OE cells using a biotinylated complementary DNA probe as described above. The tRNATyr-containing RNA was subjected to urea-PAGE, and the gel was stained with SYBR Gold. The band situated approximately 20–30 nt above the tRNATyr band was gel-excised. The elongated tRNATyr was demethylated using E. coli-derived AlkB (Addgene, 79050) to enable reverse transcription59. In brief, the RNA was incubated in 45 mM Tris-HCl (pH 8), 0.9 mM α-ketoglutaric acid, 1.8 mM ascorbic acid, 67 µM (NH4)2Fe(SO4)2, and 2.5 µM AlkB at 37 °C for 2 h. Demethylation was confirmed by simultaneously performing yeast tRNA demethylation and RNA nucleoside mass spectrometry as a positive control. The RNA and Adenylated Linker A (BIOO) were incubated at 60 °C for 3 min for denaturation and then ligated using T4 RNA ligase 2 truncated (NEB, M0242), according to the manufacturer’s protocol. The 3′ adapter-ligated tRNA was urea gel-excised and reverse-transcribed using ReverTra Ace (Toyobo, TRT-101) according to the manufacturer’s protocol and a primer listed in Supplementary Table 1. cDNA was gel-excised and PCR-amplified using a forward primer corresponding to the 5′ leader or tRNA internal 5′ region and a reverse primer complementary to the 3′ adapter (Supplementary Table 1). The PCR products were cloned using a TA Cloning Kit (Thermo Fisher Scientific, A14697), followed by Sanger sequencing of the clones.

Luciferase reporter experiments

The translation reporter was constructed using a plasmid containing dual CMV promoters (PSF-CMV-CMV-SBF1-FLUC, Sigma-Aldrich, OGS608). cDNA encoding the Renilla luciferase gene was cloned into the XhoI and BamHI sites. Subsequently, oligonucleotides containing five consecutive TAC or TAT tyrosine codons (or five random codons for a control plasmid) were cloned at the 5′ end of the Renilla luciferase gene by insertion into NcoI and XhoI sites.

One day before plasmid transfection, 9 × 103 LNZ308 cells were transferred to a 96-well plate. The next day, the cells were transfected with 10 ng of the codon reporter or control plasmid using Lipofectamine 3000. At 24 h post-transfection, the cells were lysed and luminescence was measured using a Dual-luciferase Reporter Assay System (Promega, E1910), a specialized 96-well plate (Greiner, 675075), and a Centro XS3 LB960 luminometer (Bethold Technologies). Differences of luminescence due to the transfected plasmid level were normalized using the internal firefly luciferase. Possible biases of the renilla-to-firefly luminescence ratio due to codons in the two luciferases were normalized using the control plasmid with random codons between the renilla luciferase and firefly luciferase genes.

UAA and UAG readthrough reporter plasmids from previous studies32,60 were used. A UAA or UAG readthrough reporter plasmid expresses a transcript with a renilla luciferase coding sequence at the 5′ side connected via a UAA or UAG codon to a firefly luciferase coding sequence at the 3′ side. In addition, we generated corresponding control plasmids that expressed the renilla-firefly transcript with a UUA or UUG (leucine) codon linker instead of the UAA or UAG codon linker by performing PCR of the UAA or UAG codon reporter plasmids using the primers listed in Supplementary Table 1. The control plasmids were used to normalize possible biases of the renilla-to-firefly luminescence ratio due to codons in the two luciferases. Plasmid transfection and luciferase measurements were performed in the same way as described for codon translation measurement, except that 240 ng of plasmid was used for transfection.

Statistics and reproducibility

Each n represents a distinct sample derived from cells grown in separate dishes or separate in vitro reconstitution reaction. All numerical data were analyzed by GraphPad Prism 9 or 10 software. No data were excluded. Blinding was not performed due to time and personnel constraints. To assess differences between two groups, the Mann–Whitney test was used. A two-tailed P value less than 0.05 was considered significant. To assess differences between more than two groups, an analysis of variance (ANOVA) or its derivatives was used. Data are presented as means ± standard error of means (s.e.m.). Exact P-values are included in the Supplementary Data 1.

Reporting summary

Further information on research design is available in the Nature Portfolio Reporting Summary linked to this article.