Abstract
Dual reporters encoding two distinct proteins within the same mRNA have played a crucial role in identifying and characterizing new instances of unconventional eukaryotic translation mechanisms. These mechanisms include initiation by internal ribosomal entry sites (IRESs), ribosomal frameshifting, stop codon readthrough, and reinitiation. This design allows one reporter’s expression to be influenced by the specific mechanism under investigation, while the other reporter serves as an internal control. However, challenges arise when intervening test sequences are placed between these two reporters. Such sequences can inadvertently impact the expression or function of either reporter, independently of translation-related changes. These effects may occur because of cryptic elements inducing or affecting transcription initiation, splicing, polyadenylation, and antisense transcription, as well as unpredictable effects of the translated test sequences on the stability and activity of the reporters. Unfortunately, these unintended effects may lead to incorrect conclusions being published in the scientific literature. To address this issue and assist the scientific community in accurately interpreting dual reporter experiments, we have developed comprehensive guidelines. These guidelines cover experiment design, interpretation, and the minimal requirements for reporting results. They are designed to aid researchers conducting these experiments, as well as reviewers, editors, and other investigators who seek to evaluate published data.
Background
Fusions of genes expressing reporters have been used for characterizing mRNA translation mechanisms since the 1990s1–5. The popularity of this approach increased with the invention of the dual luciferase reporter in 19986. Since then, dual reporters have been used to study ribosomal frameshifting, stop codon readthrough, reinitiation, and internal initiation6–12. Dual reporter systems have proven instrumental, in combination with site-directed or random mutagenesis, in characterizing mRNA features involved in specific translation mechanisms11,13–23 and for the identification of underlying cellular factors impacting these mechanisms24–32. Dual reporters have also found application in the identification and study of drugs targeting specific mechanisms of translation25,33–39 as well as the mechanisms of disease associated polymorphisms40. Increasingly, dual reporters are being used for discovery of novel translational mechanisms and processes41,42.
The principle is based on encoding two distinguishable reporters within the same mRNA so that expression of one reporter is dependent on the studied mechanism while the other reporter is used for normalization, i.e. as an internal control (Figure 1). As reporter proteins are synthesized from the same mRNA, it is assumed that any differences in the expression of the reporter used to characterize the mechanism under study is independent of confounding variables, e.g. transfection efficiency or RNA stability. The assay can be performed in a single tube or a well with the readout being the ratio of upsteam to downstream reporter activity. However, differences in reporter expression can occur for several reasons other than differential translation, potentially leading to false interpretation of reporter readout (Figure 1). Furthermore, since the readout of dual reporter assays is usually the ratio of downstream to upstream reporter, then changes in the upstream internal control reporter without changes to the downstream reporter will result in misinterpretations without careful consideration of absolute reporter values.
Figure 1.
Principles of a dual reporter strategy and the most common artefacts (shaded background) arising from unexamined assumptions. It is expected that a single RNA species is transcribed from the DNA sequence shown at the top of the illustration. Even if a single RNA species is transcribed, it is assumed that the fusion of the insert sequences to either reporter does not alter reporter protein activity or stability. However, several shorter possible RNA species are shown that could lead to artefactual reporter expression.
There are various possible sources of dual reporter artefacts. Firstly, DNA-encoded dual reporters may contain cryptic promoters, cryptic splice sites, or cryptic polyadenylation signals that could generate unexpected mRNA transcripts encoding only one of the two reporters11,43–54. Plasmid DNA can also produce unexpected antisense transcripts that may affect reporter expression55. Secondly, the protein extensions encoded by the test sequence may alter the stability or activity of one or both reporters, if it is synthesized as a part of the same polypeptide chain12,56. Thirdly, in certain applications, the downstream reporter is placed under a known (control) or a putative (test) IRES. However, IRES activity may be influenced by surrounding sequences20,21,57,58; it has also been shown that the presence of an IRES may influence mRNA stability59 and translation60,61 of the upstream reporters in a sequence-dependent manner.
Although various artefacts generated by experiments involving dual reporters have been discussed extensively11,12,46–50,52,57,62–70, the misinterpretation of these assays continues to result in inaccurate conclusions53,54,56.
Therefore, we developed guidelines for the design, interpretation, and reporting of experiments involving single mRNA-encoded dual reporters. In the first part, Principles and Recommendations (PR), we cover the design and interpretation of dual reporter experiments. We provide guidance on the selection of appropriate vectors to reduce risks of false positive findings and design of critical controls and additional experiments that may help to identify and avoid erroneous conclusions for a wide range of mechanisms. The second part, Minimal INformation on Dual expression Reporters (MINDR), outlines proposed reporting requirements, which establish the essential data that should accompany any publication describing dual reporter experiments. MINDR is designed to facilitate reproducibility and equip editors, reviewers, and readers with the information necessary for evaluating the reliability of the employed dual reporter strategies and assessment of the likelihood of false positive findings.
PR: Principles and Recommendations
In the following subsections we consider reporter design, possible controls, and data interpretation. These are intended to equip researchers with the means to investigate and avoid potential artefacts. We first describe general aspects regarding most dual reporter applications followed by considerations relevant to specific types of assay or phenomena which are also summarsied in Table 1.
Table 1.
Recommended controls and orthogonal evidence
Phenomenon | Negative control | Positive control | Orthogonal evidence |
---|---|---|---|
Frameshifting | Mutate putative frameshift site; Introduce stop codon downstream of frameshift site in frame with downstream reporter. | Introduce an indel to place reporters in same reading frame and make synonymous changes to the slip site. | Immunoblotting if reporters are fused; ribosomal profiling: when frameshifting produces longer proteoform, ribosome footprint density is expected dowsntream of zero frame stop codon with triplet periodicity phase matching new frame. When frameshifting is efficient and is predicted to produce a shorter proteoform, drop of ribosome footprint density is expected downstream of stop codon in the new frame. |
Stop codon readthrough | Introduce double stop codon (e.g. two TAA codons instead of a readthrough stop codon candidate. | Stop to sense codon substitution. | Immunoblotting if reporters are fused. Ribosome profiling: ribosome footprint density is expected downstream of the stop signal with triplet periodicity matching the zero frame. |
Reinitiation | Overlap between upstream and downstream reporter ORFs; introduce a stop codon or indel in the intercistronic region closer to the 2nd cistron. | For reinitiation after short uORFs, delete AUG of an upstream uORF; for long uORFs, reporters could be fused in-frame and separated with StopGo | Immunoblotting to show that reporters are NOT synthetized as a single fusion protein. Use several different reporters as a 2nd cistron. |
Internal initiation | In DNA construct, delete the promoter directing the bicistronic transcript. Compare with bicistronic constructs harboring long non-specific intergenic sequences. | Compare with known effective viral IRES (e.g. EMCV) and with the corresponding monocistronic constructs. | mRNA transfection instead of DNA reporters; compare bicistronic construct with the corresponding monocistronic ones (m7G-capped, A-capped); insert 5’-terminal stem-loop and/or long overlapping uORF; siRNA targeting upstream reporter; polysome analysis; using appropriate cell-free systems.. |
General sources of artefacts
Unexpected mRNA species
Assumptions and potential problems –
It is often assumed that all measured reporter activity is derived from translation of a single RNA species generated in strict accordance with the researcher’s design. However, unexpected mRNA species may be produced through either cryptic splicing, cryptic transcription, or internal polyadenylation (Figure 1). These aberrant monocistronic transcripts could be translated much more efficiently than their bicistronic counterparts and, even if they represent only a tiny fraction, could substantially distort the overall interpretation of the results.
Recommendations -
The most straightforward way to control for these possibilities is to introduce in vitro transcribed RNA encoding the reporter proteins instead of using DNA-based constructs. Any discrepancies between DNA-encoded and RNA-encoded reporter protein ratios may indicate the existence of unexpected cryptic transcripts, although there may be other factors explaining these differences as discussed below. There may also be situations when RNA transfections are impractical as described in the RNA vs. DNA Reporters section below.
For DNA transfections using reporters that are expected to express fusion proteins (e.g. by readthrough or frameshifting), western blotting is an effective and sensitive approach to ensure that downstream reporter protein expression is not derived from aberrant mRNA species producing shorter variants of fusion proteins.
Where products of dual reporters are not expected to be fused (e.g. in StopGo vectors and reinitiation or IRES studies), it may be desirable to identify possible aberrant mRNA species directly. Because even a minor mRNA species could produce a major signal if translated very efficiently, highly sensitive techniques for detecting low abundance mRNA species are required. To this end, emerging long read sequencing technologies (such as nanopore or PacBio) or RT-qPCR may be sufficiently sensitive approaches to detect low abundance RNA species whereas northern blotting is unlikely to be sensitive enough. RT-qPCR can be used for assessing the levels of different RNA segments46 with specific primers as shown in Figure 2. Careful design and validation of RT-qPCR amplicons with in vitro transcript standards and melting curves is necessary to determine whether the amounts of each amplicon are equal71. As the same PCR amplicons can be derived from alternative transcripts and because RT-qPCR optimization is not always straightforward, it is advisable that this method is not the sole approach used to analyze the occurrence of transcript isoforms68. An in vitro transcribed full-length control RNA can be used to assess the signal expected from a single RNA species containing both reporters (Figure 2).
Figure 2.
Assessment of dual reporter transcriptional integrity. Red arrows show the design of forward and reverse primers for the dual reporter construct and double red lines depict the resulting cDNA products. Using a synthetic RNA construct (bottom) it is possible to estimate the ratio between cDNA products for the transcript containing all three regions. Horizontal blue lines represent siRNAs for knockdown experiments expected to downregulate expression of both reporters if they are produced from the full-length RNA template as indicated with vertical yellow and green arrows.
Furthermore, an elegant approach employing RNA interference can be utilized to regulate the authenticity of bicistronic RNA46. An siRNA probe designed to knock down expression of the upstream reporter directs specific degradation of the bicistronic reporter mRNA, thereby decreasing the activity of both the upstream and downstream reporters to a similar extent, unless they are expressed from different mRNA species (Figure 2). Transcript isoforms could also be explored by methods like Cap Analysis of Gene Expression (5’-CAGE)72 or with direct RNA sequencing using nanopore technologies, bearing in mind that low abundance alternative products can produce most of the reporter activity.
Potential existence of aberrant transcripts may also be explored with study-specific controls, the design of which is specific to the studied phenomena.
Altered activity or stability of fused reporter proteins
Assumptions and potential problems –
When testing relative activities from fused reporter proteins (e.g. for readthrough or frameshifting), it is often assumed that the reporter protein activities accurately reflect reporter mRNA translation. Frameshifting and stop codon readthrough efficiencies are generally estimated by determing the ratio of upstream to downstream reporter protein of a test construct with the reporter protein ratio from a positive control (100%), where both reporters are in the same, uninterrupted reading frame. However, the part of the protein fusion encoded by the test sequence may influence the stability or activity of one or both reporter proteins. This can be especially problematic if fusions to the upstream reporter only influence the in-frame control reporter protein and not the test reporter, or vice versa (Figure 1)12,56.
Recommendations –
Place StopGo/2A sequences on both sides of the test sequence12 (Figure 3). StopGo is an atypical translation mechanism where the elongating ribosome fails to form a peptide bond and then efficiently continues translation – the outcome of this phenomenon is the production of two peptide chains (from the same ORF) separated by the StopGo motif 73.
Figure 3.
Fused versus unfused dual reporters. Both systems have their merits and drawbacks. The advantage of fused reporters is the ability to estimate recoding efficiencies by western blotting or in vitro translation reactions while controlling for unintended D-Reporter translation by internal initiation or cryptic splicing. One potential downside of fused reporters is erroneous estimations of recoding efficiencies based solely on reporter protein activities because of possible changes in D-Reporter activity resulting from its fusion to the test and U-Reporter proteins. This can be mitigated with unfused reporter systems where the test peptide is released from the reporter proteins by flanking StopGo (SG) motifs. While western blotting or in vitro translations can still be used to estimate recoding efficiencies with unfused reporter systems, due to the similar sized D-Reporter protein produced, spurious D-Reporter protein expression by internal initiation or cryptic splicing may not be as obvious.
A peptide motif enabling StopGo was initially discovered during translation of the 2A region of foot-and-mouth disease virus74, and since then a number of 2A peptides with varying efficiencies have been described. While in theory StopGo use should produce reporter proteins with identical amino acid sequences irrespective of the inserted test sequence, this may not always be the case. StopGo is not 100% efficient (generally 80–90%75), therefore a certain amount of fusion between either of the reporters and the test sequence-encoded fragment will still be produced. Nonetheless, for example, a comparison of absolute Renilla luciferase activities (when used as an upstream reporter) from a series of fused and unfused test constructs indicated much more variability with fused reporters compared to StopGo-containing plasmid 12. Additionally, the potential for ribosome drop-off during StopGo that may influence the ratio between two reporters can be easily tested with an appropriate in-frame control construct. When using a new StopGo-containing plasmid, it is important to monitor the efficiency of StopGo by western blotting in the intended biological system. Although variability in StopGo activity has been reported76, it can be mitigated by using longer StopGo motifs (>30 aa), which are less sensitive to the effects of proximal sequences, including test inserts, while having higher activities and negligible ribosome drop-off75. It should be noted that, although StopGo is active in all eukaryotes tested so far, different 2A peptide variants have different activities in various organisms, and none have been found to function in bacteria77. Furthermore, the kinetics of StopGo is still poorly understood and it is likely to cause some ribosome pausing, which may interfere with some of the studied phenomena.
Alterations in reporter ratios due to artefactual changes in reporters’ absolute readouts
Assumptions and potential problems –
It seems reasonable to assume that when transfecting the same reporter construct into two different cell lines or under different conditions, the ratiometric readout should be similar. However, this is not always the case. Not surprisingly, absolute reporter values may be quite different between cell lines/conditions, sometimes spanning orders of magnitude. There may be several reasons for this, including differences in the reporter delivery efficiencies between the cell lines. Reporter RNAs may trigger the innate immune response resulting in global suppression of translation, e.g. through activation of PKR, cGAS, TLRs, NLRs, or RIG-I-like receptors78,79. Reporter expression in particular cell lines can be compromised at either the transcriptional (e.g. weak promoter activity in DNA reporters) or translational (e.g. low global translation rates) level. When dual reporters are used to compare the efficiency of the studied mechanism across cell lines or under specific conditions (including stress versus control, overexpression or depletion of specific factors, treatment with small molecule inhibitors, and NMD activation) it is important to understand that differences in global translation may affect the ratio of the measured reporter activities without affecting their real relative activities. This is because any measured activity is a combination of genuine reporter protein activity with background levels due to biological noise and technical limitations. We can represent the ratio of measured activities as , where , are bona fide activities of the two reporter proteins and is the constant background signal. It is clear then to see that if and are changed proportionally (e.g. reduced twice by general translation inhibiton) and ratio stays the same, the measured ratio will change.
Perhaps more problematically, there may be cell specific differential reporter activity due to the presence of a cell-specific repressor/activator that acts on one of the reporters. Another factor to consider is the reporter overexpression driven by a strong promoter that could further exacerbate cell-specific effects by e.g. titrating essential translation components.
Recommendations –
Here it is critically important to report absolute readouts of reporter and background activities to reveal potential misinterpretations of observed changes in relative activities of the reporters, as e.g. in Khan et al53.
Assay-specific principles and recommendations
RNA vs DNA reporters
The use of mRNA transfection is a powerful strategy to avoid the generation of unexpected RNA species by cryptic transcription, polyadenylation, and/or splicing events. However, its advantages are not limited to this. For several applications, RNA transfection may be preferred irrespective of transcriptional artefacts. Analysis of the immediate stress response is particularly difficult when using plasmid DNA reporters. There is little point in applying stress stimuli immediately after DNA transfection as substantial time is required for the accumulation of sufficient mRNA. In contrast, for RNA transfection, the stress stimuli can be applied immediately or shortly before/after the transfection (1–2 hours), here, newly synthesized protein products are responsible for most of the reporter activity. It is also difficult to synchronize the expression of DNA reporters in the entire cell population without synchronizing their cell cycle, as the plasmid only enters the nucleus during mitosis. Furthermore, transfection of non-dividing cells (e.g. matured neurons or cardiac myocytes) with plasmid DNA is highly inefficient.
Nonetheless RNA transfection is not a panacea and should be used appropriately. Firstly, activity values are generally much lower from mRNA transfections as compared to DNA transfections. For this reason, RNA transfections should use reporters with a high signal-to-background ratio such as luciferase reporters. Secondly, RNA is relatively unstable, so it is better to analyze the activity shortly after transfection, ideally in 1–4 hours post transfection. At later time-points, e.g. 8–10 hours post transfection, accumulated reporter products start to exceed newly synthesized ones – even more so than in the case of DNA transfections. It is also important to use a transfection protocol that minimally stresses the cells and to avoid stress stimuli like cell re-plating, electroporation, or plate/dish cooling immediately before or during transfection85. Thirdly and most importantly, one should keep in mind that transfected artificial mRNA has not passed through the nucleus and may lack specific and potentially critical features including epitranscriptomic marks or associated mRNA-binding proteins of nuclear origin. Furthermore, in primary cells, mRNA transfection often triggers innate immune responses, so additional efforts are needed to reduce its activation, like the use of transcripts with 5’-cap and modified nucleotides (e.g. m1Ψ in place of U)78,79. It is also important to note that lipocationic transfection impedes the analysis of reporter mRNA stability, as only a tiny fraction of transfected mRNA is released from endosomes into the cytosol86. Accordingly, total RNA extraction from the transfected cells yields mRNA that is not predominantly derived from the cytosolic fraction but from endosomal and endolysosomal compartments86,87. Finally, transfected RNA may still be processed by cytoplasmic RNA cleaving enzymes, such as IRE188,89 or RNase L90, or during alternative polyadenylation that generates translated RNA fragments downstream of it91. Therefore, one cannot exclude the possibility that even in the case of mRNA transfection, aberrant RNA species may also be present. A more detailed comparison of DNA and RNA transfections can be found in a dedicated review92.
The use of dual reporters for massively parallel assays
In recent years, dual reporter vectors have become popular in Massively Parallel Reporter Assays (MPRA) that simultaneously evaluate thousands of test sequences. This is a powerful approach that allows screening of a diverse pool of sequences for specific regulatory properties, e.g. driving internal initiation41,42, ribosomal frameshifting34 and other translation mechanisms. It can also be applied to screen a pool of all possible variants of a particular sequence to comprehensively characterize cis-acting regulatory elements.
While many of the general principles that should guide the use of dual reporters outlined above also apply here, the high-throughput nature entails several specific considerations associated with measuring reporter protein activity in a pooled manner. Certain guidelines such as reporting absolute measured expression levels of the reporter genes is typically harder to achieve than in the case of single reporter measurements. On the other hand, dual reporter-MPRAs also allow for the inclusion of a much larger number of controls that can be measured in the same experiment.
Particularly suitable for MPRA are dual fluorescent reporter constructs, as they allow fluorescence-activated cell sorting based on relative reporter expression levels. This is then followed by DNA sequencing-based identification and quantification of the underlying sequences.
An important consideration for MPRA is to ensure equal vector copy number and similar expression levels among the cells. Utilization of systems for genome integration93 enables the integration of a single copy of the test vector in the same genetic locus for all cells. However, this may not be strictly required. When this issue was carefully examined recently94, a high degree of correlation was found between the expression of two cistrons across a polyclonal population of cells, regardless of integration site and number of integrated copies of the bicistronic construct.
A significant challenge in screening large pools of diverse sequences is discerning artifacts amidst the positive hits. Reliance on validating only a subset of hits can lead to misinterpretation, as each individual sequence may possess unique properties, and even a single nucleotide change can alter cryptic promoter activity. Given that MPRA-based screens typically yield numerous hits, it is impractical to validate each one individually, making it difficult to estimate the rate of false positives41,42. Therefore, general conclusions should not be drawn from limited validated cases. Instead, conducting a meta-analysis of positive hit sequences may help identify shared features and specific types of artifacts. While rare artifacts are unlikely to significantly impact conclusions drawn from MPRAs aimed at dissecting the regulatory code of a gene regulatory mechanism, findings related to specific “positive” events should be limited to candidates that have been appropriately validated to minimize the potential for misinterpretation.
Cell-free translation systems and other in vitro assays
Another potential source of false positives from dual reporter assays is inappropriate use of in vitro translation systems. For example, the commercially available and widely used nuclease-treated Rabbit Reticulocyte Lysate (ntRRL), which is prepared from specialized cells with a limited range of RNA-binding proteins, has been repeatedly shown to inaccurately reproduce conditions found in normal cells62,69. Firstly, exogenously added mRNAs translated in ntRRL exhibit a relatively weak reliance on the 5’ cap51,95, although optimizing the buffer conditions can substantially increase cap-dependency and start site recognition96. Moreover, as some eIF4G molecules are sequestered by the capped 5’-terminal mRNA fragments remaining in the ntRRL after limited hydrolysis of endogenous reticulocyte transcripts, the addition of cap-dependent initiation inhibitors (such as m7GTP, 4E-BP, or proteases that cleave eIF4G) may release this factor and artificially stimulate translation of uncapped mRNAs. This system also does not recapitulate the cap/poly(A) synergy97. Taken together, all this is often misinterpreted as an indication of cap-independent translation of a particular studied mRNA. Secondly, ntRRL is prone to aberrant internal initiation at AUG codons located within extended unstructured regions98, causing artificial expression of the second cistron in bicistronic reporters even in the absence of bona fide IRESs. Finally, translation in RRL is highly sensitive to even moderately stable RNA secondary structures in the 5’ leader, which is not the case in living cells95.
It is likely that similar artefacts are present in cell-free systems derived from budding yeast and wheat germ, at least under some conditions67,99. In contrast, such effects are not typically observed in cytosolic extracts of cultured mammalian cells51,95. However, in any in vitro system, results can greatly depend on the specific preparation conditions and component concentrations. For example, varying polyamine concentrations in the yeast extract can modulate stop codon readthrough efficiency as much as 4-fold100. Therefore, caution is necessary when comparing findings from a specific cell-free system with cultured cells or in vivo observations.
Less commonly, bicistronic constructs are used in in vitro systems reconstituted from purified components101,102. Although such analysis is very informative, it also should be done with caution. Similar to RRL, in these systems, which are usually devoid of mRNA-binding proteins, the ribosomes are able to bind to internal AUG codons located within long unstructured regions103,104. This risks confusion of an authentic mechanism with an artificial one that should be excluded by validation in complete in vitro or in vivo systems.
Another potential source of artefacts specific for in vitro assays is related to partial hydrolysis of an in vitro synthesized reporter mRNA that can produce truncated versions of bicistronic constructs. This leads to inappropriate ribosome loading to the second cistron on transcripts that do not contain the first cistron. Thus, the extensive analysis of mRNA integrity should be carried out. In some cases (i.e. IRES studies), analysis of polysome-associated mRNA fractions can be helpful. In particular, total RNA isolated separately from monosome and polysome fractions of each sample tested would be subject to RT-qPCR using primers covering either each cistron alone or both cistrons together to determine the integrity of the full-length reporter mRNA.
Phenomenon-specific principles and recommendations
Specific considerations in the assessment of ribosomal frameshifting and stop codon readthrough
The assessment of ribosomal frameshifting and stop codon readthrough using dual reporters involves a calculation of the relative reporter activities (downstream to upstream ratio). The relative reporter activity of the test construct is then compared to that of a control construct. It is not advisable to calculate reporter activities relative to a negative control rather than a positive control, as this can result in the unintended consequence of an insignificant number having an outsized effect. An appropriate positive control has both reporters encoded within the same ORF (in-frame control). Often a single in-frame control is compared to several test sequences. We caution against this practice; when the sequence of the test constructs is considerably different from the sequence of a single in-frame control as some of the test sequences may contain cryptic splice sites or promoters. The ideal in-frame control should have no amino acid differences from the expected frameshift or readthrough product and minimal nucleotide differences.
For frameshifting, an in-frame control can be obtained with either insertion or deletion of a single nucleotide in the putative frameshifting site depending on the expected direction of frameshifting (−1 or +1). For stop codon readthrough, the stop codon should be replaced with a sense codon, ideally one encoding the same amino acid that is expected to be inserted in place of the stop codon. However, the identity may not be known in advance and there may be one of several possible amino acids inserted105,106. Reporter ratios in such positive control constructs is considered to correspond to 100% efficient frameshifting or stop codon readthrough. For studying frameshifting, it is advisable to introduce synonymous changes to disrupt the putative slippery sequence within the in-frame control to create a more accurate readout of 100% frameshifting. If the frameshifting site is not disrupted in an in-frame control and frameshifting occurs (say at 10% efficiency), only 90% of the ribosomes would synthesize the downstream reporter. Thus, ideally, each tested sequence should have its own positive in-frame control. Including a +1 or −1 frame termination codon (depending on the reading frame of the downstream reporter) 5’ of a putative frameshifting site is also worth considering to ensure that only frameshifting within the test sequence is reported.
When using fused reporters, orthogonal validation of frameshifting by western blotting is highly desirable when reporting novel instances. Most reporter proteins can be detected by commercially available antibodies that can detect frameshifting or stop codon readthrough efficiencies as low as 1%. As mentioned above, when using fused reporters, western blotting can also control for cryptic splicing and cryptic promoter activity.
Another important way to validate ribosomal frameshifting and stop codon readthrough is the use of a negative control. Ribosomal frameshifting normally occurs at specific frameshifting sites accompanied with stimulatory elements such as specific mRNA structures. While single point mutations in stimulatory signals rarely abolish frameshifting completely, disruptions of the frameshifting site are expected to eliminate frameshifting as tRNA re-pairing in the new frame is precluded. A useful negative control for stop codon readthrough is the insertion of tandem in-frame TAA stop codons that in most organisms represent the most efficient terminators.
Although not always possible, it is desirable to avoid AUG codons within the test sequence that are in the same reading frame as the downstream reporter ORF, as they may serve as initiation codons on cryptic transcripts missing the upstream reporter, entirely or partly. However, a deliberate insertion of an AUG codon in-frame with the downstream reporter may be used as a control for the existence of cryptic transcripts. In the absence of such transcripts, introduction of an AUG codon in a good Kozak context should not substantially alter the activity of the downstream reporter unless there is reinitiation. Although reinitiation is extremely rare after translation of long ORFs, there are certain signals that could enable reinitiation even after translation of long ORFs107–110. Another consideration is potential initiation at near-cognate starts which is generally inefficient, but is known to be highly productive and even comparable to that of AUG in certain contexts, therefore it is advisable to examine the sequences for the presence of such near-cognate start codons in a good Kozak context. Additional evidence may be derived from orthogonal methodological approaches, such as ribosome profiling by accessing publically available data in dedicated browsers111,112. While a detailed discussion of this technique is beyond the scope of this paper, we recommend that interested readers consult specialized reviews113,114.
Specific considerations in the assessment of reinitiation
Reinitiation is a process in which a ribosome initiates translation downstream of the stop codon at which it terminates (reviewed in115,116). Usually, this process is not very efficient in eukaryotes, unless the translated upstream ORF is short or mRNA-specific mechanisms are used107–110,115,117. However, reinitiation can be greatly facilitated under some physiological stress conditions or when ribosome recycling factors are artificially depleted29,118.
The rate of translation reinitiation after long ORFs can be assessed with the dual reporter assay; other methods are more suitable to examine reinitiation after short uORFs (see115,116 and references therein). However, due to the inefficiency of this process under normal conditions, absolute values of reporter activity and appropriate background correction should be thoroughly considered. As in other cases, cryptic promoters in intercistronic spacers must be excluded and only appropriate cell-free systems should be used for in vitro studies to exclude false (or true) internal initiation. Moreover, stop codon readthrough or frameshifting can be erroneously attributed to reinitiation, so these possibilities should also be excluded. The most appropriate control is to extend the coding region of the upstream ORF into the downstream ORF by mutating the stop codon of the former in a way to avoid matching reading frames of both ORFs – such a construct should allow zero reinitiation unless the AUG of the first ORF is “leaky” for initiation, which would still generate some background activity from the downstream reporter. Additionally, readthrough can be specifically excluded by inserting an additional stop codon (or a tandem of stops) between the translated upstream ORF and the putative reinitiation site, which would reduce readthrough, but not reinitiation. Similarly, the possibility of ribosomal frameshifting can be reduced by the insertion or deletion of a nucleotide within the intercistronic region to disrupt the reading frame and eliminate the possibility of frameshifting without affecting reinitiation (unless the nucleotide indel disrupts a cis-acting signal responsible for reinitiation). Alternatively, western blotting could be used to rule out the possibility of either readthrough or frameshifting.
Specific considerations in the assessment of internal initiation
When testing for IRES-dependent translation, the mRNA sequence suspected to promote internal initiation is inserted between two reporters to enable the translation of the downstream reporter. Unlike scenarios involving frameshifting or stop codon readthrough, no fusion product is expected. Therefore, the potential for the translation of the first cistron to interfere with the proposed IRES must be minimized. This can be achieved by incorporating additional stop codons to prevent any readthrough and by positioning the IRES at a sufficient distance from the first cistron stop codon to ensure that terminating ribosomes do not compromise the structural integrity of the IRES. However, it is important to consider the possibility that integrating an RNA fragment into an unnatural context may affect any potential IRES activity.
Although distant elements affecting stop codon readthrough have been reported119,120 and may also exist for frameshifting, it is often the case that only short motifs or structures are examined in those instances. Conversely, since internal initiation relies on more elaborate structures, testing longer sequences is essential. Accordingly, all the aforementioned concerns regarding cryptic promoters and splice sites are particularly relevant in the context of IRES testing. It is important to consider that nearly any arbitrarily selected long fragment of a mammalian 5’ UTR could exhibit cryptic promoter activity, attributable to the abundance of transcription factor binding sites within these regions. Furthermore, a significant proportion of human genes contain alternative transcription start sites (TSS)121, and TSS switching can occur during acute stress122–124. Additionally, long insertions increase the risk of false positives when using in vitro systems like ntRRL, which are not strongly cap-dependent.
To reveal cryptic splicing (but not cryptic promoter) mediated events, the independence of first and second cistron expression can be verified by designing constructs containing a sufficiently long uORF in the 5’ UTR (preferentially overlapping with the main AUG of the first cistron and with the uAUG codon in strong Kozak context) or by inserting a stable hairpin at the very 5’ end of the transcript to diminish the translation efficiency of the first cistron62. However, this approach requires analysis of the transcript level, as the downstream reporter may be affected if such modifications alter the stability of the whole mRNA.
In IRES research, the use of mRNA reporters is distinctly preferable to DNA reporters62. Using in vitro transcribed reporters eliminates artefacts arising from cryptic promoter activity, unintended splicing, or premature transcription termination. When employing DNA transfection to evaluate putative IRES activity, promoterless and siRNA-mediated controls (see above) are essential.
Most importantly, however, the bicistronic assay used to examine IRES activity possesses a notable intrinsic limitation that was not initially apparent. Comparing the expression ratios of upstream and downstream cistrons between bicistronic reporters containing different IRESs is often uninformative. This is because any two long non-specific arbitrary sequences placed between reporters are highly unlikely to yield equal readouts, while different bona fide IRESs can also exhibit significantly different activities (Figure 4A). As a result, it becomes challenging to confidently design negative or positive controls for such assays.
Figure 4.
Peculiarities of dual reporter assays in the study of IRES activity. A. Simulated results from typical bicistronic assays and their potential interpretation. Unambiguous conclusions cannot be drawn solely from comparisons of the activities of different bicistronic mRNAs with each other, as control values (both negative and positive ones) may exhibit significant variability. B. A mechanism of ribosome recruitment employed by a specific mRNA fragment (for example, a 5’ UTR) can be elucidated through three related dual reporter assays using the mRNA constructs shown in the upper subpanel. Three distinct translation initiation mechanisms can be distinguished: canonical cap-dependent scanning, CITE-directed, or IRES-directed, as indicated below. To prepare m7G-capped transcripts, it is advisable to employ CleanCap, ARCA, or post-transcriptional enzymatic capping instead of m7GpppG, as these methods yield more natural, non-immunogenic, and efficiently translated mRNAs. Additionally, ApppG or ApppA can be utilized to produce transcripts that are stabilized by non-functional cap analogs at the 5’ ends. The bottom panel represents typical simulated results. Bkg, the background level of the reporter activity in mock-transfected cells.
Many putative IRES sequences are derived from natural 5’ UTRs, which, in the case of cellular and many viral mRNAs, are naturally capped and therefore can be bound and scanned by the canonical initiation machinery. Therefore, even if the sequence under investigation suggests internal initiation in a bicistronic context, it is not straightforward to evaluate the contribution of internal initiation to overall translation in its natural context. These complications can be addressed by comparing expression from bicistronic and monocistronic reporters, allowing for an assessment of the relative contributions of different initiation mechanisms (Figure 4B).
Finally, it is important to note that the term “IRES” specifically refers to internal ribosome entry rather than cap-independence. While the latter arises as a consequence of the former, some mRNAs that definitely lack an IRES and strictly require a free 5’ end can nevertheless be efficiently translated in an uncapped (or capped with the artificial non-functional A-cap analog) form. In such cases, specific elements known as cap-independent translation enhancers (CITE) facilitate cap-independent translation62. In contrast to IRESs, CITEs cannot direct translation of the second cistron in a bicistronic mRNA (Figure 4B). Furthermore, views regarding the widespread presence of IRESs in the 5’ UTRs of cellular mRNAs are largely based on the assumption of low processivity of the translation initiation complex, which is, however, significantly underestimated51,125,126. This has contributed to the perception that any long 5’ UTR must use a non-canonical translation initiation pathway, which does not seem to be the case. Given the considerable proportion of purported “cellular IRESs” that have been challenged to date11,44,46,48–52,54,62,65,67–70,116, one should bear in mind that the detected cap-independent activity of a sequence under investigation may easily be attributed to CITEs or even some of the experimental artefacts described above.
MINDR: Minimal INformation on Dual expression Reporters.
Based on the caveats described above, we propose the following three minimal reporting requirements that should mitigate many of these issues and should accompany any study with data obtained with dual reporters:
1. A list of positive and negative controls
As discussed above, the use of appropriate positive and negative controls is critical. However, it may be too impractical to design all the controls described in the previous sections. Nevertheless, the level of confidence in the reported results depends on the specific controls used in an experiment. Therefore, the authors should explicitly describe which positive and negative controls have been used for detecting potential artefacts, to help reviewers and readers assess the reliability of the study.
2. Full sequences of all vectors and inserts
Sequences responsible for potential artefacts, such as transcription enhancers or donor/acceptor splice sites, may be distant from the sequence encoding reporters52,55. Therefore, for the reproducibility and future interrogation of reported results, the exact sequence of all plasmids used should be provided.
3. Absolute readout values for each reporter and for the background
The use of dual reporters in general requires the analysis of their ratios rather than absolute values, which are subject to high variability due to technical reasons, such as transfection efficiencies.
Nonetheless, as discussed earlier, consistent differences in absolute values are often indicative of artefacts. Therefore, it is important that in addition to providing the ratios between the reporter activities, the absolute raw readouts should be made available for each replicate of each construct including background values from control experiments without reporter constructs.
ACKNOWLEDGMENTS
The authors wish to acknowledge many colleagues who informally supported and encouraged development of these guidelines. The following authors wish to acknowledge funding support.
MM by grants PID2020–115122GA-I00 from the Ministerio de Ciencia, Innovación y Universidades of Spain. MSS by the NIH (R01GM148702). PVB by the Wellcome Trust (210692/Z/18/]) and Science Foundation Ireland (20/FFP-A/8929). LSV by the Lead Agency (DFG & CSF) grant 23–08669L and the Praemium Academiae grant provided by the Czech Academy of Sciences and CZ.02.01.01/00/22_008/0004575 RNA for therapy by ERDF and MEYS. SED from the Russian Science Foundation (23–14-00218, the IRES part). CSF by the NIH (R01 GM092927). WVG by the NIH (R01GM132358). YC by the NIH (R01GM130838).
Footnotes
CONFLICTS OF INTERESTS
G.L. and P.V.B. are cofounders and shareholders of EIRNABio.
References
- 1.Pelletier J & Sonenberg N Internal initiation of translation of eukaryotic mRNA directed by a sequence derived from poliovirus RNA. Nature 334, 320–325 (1988). [DOI] [PubMed] [Google Scholar]
- 2.Jang SK, Kräusslich HG, Nicklin MJ, Duke GM, Palmenberg AC & Wimmer E A segment of the 5’ nontranslated region of encephalomyocarditis virus RNA directs internal entry of ribosomes during in vitro translation. J Virol 62, 2636–2643 (1988). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Reil H, Höxter M, Moosmayer D, Pauli G & Hauser H CD4 Expressing Human 293 Cells as a Tool for Studies in HIV-1 Replication: The Efficiency of Translational Frameshifting Is Not Altered by HIV-1 Infection. Virology 205, 371–375 (1994). [DOI] [PubMed] [Google Scholar]
- 4.Stahl G, Bidou L, Rousset J-P & Cassan M Versatile vectors to study recoding: conservation of rules between yeast and mammalian cells. Nucl Acids Res 23, 1557–1560 (1995). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Bidou L, Stahl G, Grima B, Liu H, Cassan M & Rousset JP In vivo HIV-1 frameshifting efficiency is directly related to the stability of the stem-loop stimulatory signal. RNA 3, 1153–1158 (1997). [PMC free article] [PubMed] [Google Scholar]
- 6.Grentzmann G, Ingram JA, Kelly PJ, Gesteland RF & Atkins JF A dual-luciferase reporter system for studying recoding signals. RNA 4, 479–486 (1998). [PMC free article] [PubMed] [Google Scholar]
- 7.Chiba S, Jamal A & Suzuki N First Evidence for Internal Ribosomal Entry Sites in Diverse Fungal Virus Genomes. mBio 9, e02350–17 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Rakauskaite R, Liao P-Y, Rhodin MHJ, Lee K & Dinman JD A rapid, inexpensive yeast-based dual-fluorescence assay of programmed--1 ribosomal frameshifting for high-throughput screening. Nucleic Acids Res 39, e97 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Cardno TS, Poole ES, Mathew SF, Graves R & Tate WP A homogeneous cell-based bicistronic fluorescence assay for high-throughput identification of drugs that perturb viral gene recoding and read-through of nonsense stop codons. RNA 15, 1614–1621 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Harger JW & Dinman JD An in vivo dual-luciferase assay system for studying translational recoding in the yeast Saccharomyces cerevisiae. RNA 9, 1019–1024 (2003). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Bert AG, Grépin R, Vadas MA & Goodall GJ Assessing IRES activity in the HIF-1alpha and other cellular 5’ UTRs. RNA 12, 1074–1083 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Loughran G, Howard MT, Firth AE & Atkins JF Avoidance of reporter assay distortions from fused dual reporters. RNA 23, 1285–1289 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Jacobs JL, Belew AT, Rakauskaite R & Dinman JD Identification of functional, endogenous programmed −1 ribosomal frameshift signals in the genome of Saccharomyces cerevisiae. Nucleic Acids Res 35, 165–174 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Baranov PV, Henderson CM, Anderson CB, Gesteland RF, Atkins JF & Howard MT Programmed ribosomal frameshifting in decoding the SARS-CoV genome. Virology 332, 498–510 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Charbonneau J, Gendron K, Ferbeyre G & Brakier-Gingras L The 5’ UTR of HIV-1 full-length mRNA and the Tat viral protein modulate the programmed −1 ribosomal frameshift that generates HIV-1 enzymes. RNA 18, 519–529 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Gendron K, Charbonneau J, Dulude D, Heveker N, Ferbeyre G & Brakier-Gingras L The presence of the TAR RNA structure alters the programmed −1 ribosomal frameshift efficiency of the human immunodeficiency virus type 1 (HIV-1) by modifying the rate of translation initiation. Nucleic Acids Res 36, 30–40 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Howard MT, Shirts BH, Petros LM, Flanigan KM, Gesteland RF & Atkins JF Sequence specificity of aminoglycoside-induced stop condon readthrough: potential implications for treatment of Duchenne muscular dystrophy. Ann Neurol 48, 164–169 (2000). [PubMed] [Google Scholar]
- 18.Harrell L, Melcher U & Atkins JF Predominance of six different hexanucleotide recoding signals 3’ of read-through stop codons. Nucleic Acids Res 30, 2011–2017 (2002). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Yordanova MM, Loughran G, Zhdanov AV, Mariotti M, Kiniry SJ, O’Connor PBF, Andreev DE, Tzani I, Saffert P, Michel AM, Gladyshev VN, Papkovsky DB, Atkins JF & Baranov PV AMD1 mRNA employs ribosome stalling as a mechanism for molecular memory formation. Nature 553, 356–360 (2018). [DOI] [PubMed] [Google Scholar]
- 20.Hennecke M, Kwissa M, Metzger K, Oumard A, Kröger A, Schirmbeck R, Reimann J & Hauser H Composition and arrangement of genes define the strength of IRES-driven translation in bicistronic mRNAs. Nucleic Acids Res 29, 3327–3334 (2001). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Bochkov YA & Palmenberg AC Translational efficiency of EMCV IRES in bicistronic vectors is dependent upon IRES sequence and gene location. Biotechniques 41, 283–284, 286, 288 passim (2006). [DOI] [PubMed] [Google Scholar]
- 22.Vallejos M, Ramdohr P, Valiente-Echeverría F, Tapia K, Rodriguez FE, Lowy F, Huidobro-Toro JP, Dangerfield JA & López-Lastra M The 5’-untranslated region of the mouse mammary tumor virus mRNA exhibits cap-independent translation initiation. Nucleic Acids Res 38, 618–632 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Baranov PV, Wills NM, Barriscale KA, Firth AE, Jud MC, Letsou A, Manning G & Atkins JF Programmed ribosomal frameshifting in the expression of the regulator of intestinal stem cell proliferation, adenomatous polyposis coli (APC). RNA Biol 8, 637–647 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Beznosková P, Wagner S, Jansen ME, von der Haar T & Valášek LS Translation initiation factor eIF3 promotes programmed stop codon readthrough. Nucleic Acids Res 43, 5099–5111 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Müller C, Schulte FW, Lange-Grünweller K, Obermann W, Madhugiri R, Pleschka S, Ziebuhr J, Hartmann RK & Grünweller A Broad-spectrum antiviral activity of the eIF4A inhibitor silvestrol against corona- and picornaviruses. Antiviral Res 150, 123–129 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Zinshteyn B, Sinha NK, Enam SU, Koleske B & Green R Translational repression of NMD targets by GIGYF2 and EIF4E2. PLoS Genet 17, e1009813 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Kobayashi Y, Zhuang J, Peltz S & Dougherty J Identification of a cellular factor that modulates HIV-1 programmed ribosomal frameshifting. J Biol Chem 285, 19776–19784 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Green L, Houck-Loomis B, Yueh A & Goff SP Large ribosomal protein 4 increases efficiency of viral recoding sequences. J Virol 86, 8949–8958 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Young DJ, Makeeva DS, Zhang F, Anisimova AS, Stolboushkina EA, Ghobakhlou F, Shatsky IN, Dmitriev SE, Hinnebusch AG & Guydosh NR Tma64/eIF2D, Tma20/MCT-1, and Tma22/DENR Recycle Post-termination 40S Subunits In Vivo. Mol Cell 71, 761–774.e5 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Schult P, Roth H, Adams RL, Mas C, Imbert L, Orlik C, Ruggieri A, Pyle AM & Lohmann V microRNA-122 amplifies hepatitis C virus translation by shaping the structure of the internal ribosomal entry site. Nat Commun 9, 2613 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.LaFontaine E, Miller CM, Permaul N, Martin ET & Fuchs G Ribosomal protein RACK1 enhances translation of poliovirus and other viral IRESs. Virology 545, 53–62 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Zhang H, Song L, Cong H & Tien P Nuclear Protein Sam68 Interacts with the Enterovirus 71 Internal Ribosome Entry Site and Positively Regulates Viral Protein Translation. J Virol 89, 10031–10043 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Bidou L, Hatin I, Perez N, Allamand V, Panthier J-J & Rousset J-P Premature stop codons involved in muscular dystrophies show a broad spectrum of readthrough efficiencies in response to gentamicin treatment. Gene Ther 11, 619–627 (2004). [DOI] [PubMed] [Google Scholar]
- 34.Mikl M, Pilpel Y & Segal E High-throughput interrogation of programmed ribosomal frameshifting in human cells. Nat Commun 11, 3061 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Cencic R, Robert F & Pelletier J Identifying small molecule inhibitors of eukaryotic translation initiation. Methods Enzymol 431, 269–302 (2007). [DOI] [PubMed] [Google Scholar]
- 36.Novac O, Guenier A-S & Pelletier J Inhibitors of protein synthesis identified by a high throughput multiplexed translation screen. Nucleic Acids Res 32, 902–915 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Ahn D-G, Lee W, Choi J-K, Kim S-J, Plant EP, Almazán F, Taylor DR, Enjuanes L & Oh J-W Interference of ribosomal frameshifting by antisense peptide nucleic acids suppresses SARS coronavirus replication. Antiviral Res 91, 1–10 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Sun Y, Abriola L, Niederer RO, Pedersen SF, Alfajaro MM, Silva Monteiro V, Wilen CB, Ho Y-C, Gilbert WV, Surovtseva YV, Lindenbach BD & Guo JU Restriction of SARS-CoV-2 replication by targeting programmed −1 ribosomal frameshifting. Proc Natl Acad Sci U S A 118, e2023051118 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Feng Y, Pinkerton AB, Hulea L, Zhang T, Davies MA, Grotegut S, Cheli Y, Yin H, Lau E, Kim H, De SK, Barile E, Pellecchia M, Bosenberg M, Li J-L, James B, Hassig CA, Brown KM, Topisirovic I & Ronai ZA SBI-0640756 Attenuates the Growth of Clinically Unresponsive Melanomas by Disrupting the eIF4F Translation Initiation Complex. Cancer Res 75, 5211–5218 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Hekman KE, Yu G-Y, Brown CD, Zhu H, Du X, Gervin K, Undlien DE, Peterson A, Stevanin G, Clark HB, Pulst SM, Bird TD, White KP & Gomez CM A conserved eEF2 coding variant in SCA26 leads to loss of translational fidelity and increased susceptibility to proteostatic insult. Hum Mol Genet 21, 5472–5483 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Chen C-K, Cheng R, Demeter J, Chen J, Weingarten-Gabbay S, Jiang L, Snyder MP, Weissman JS, Segal E, Jackson PK & Chang HY Structured elements drive extensive circular RNA translation. Mol Cell 81, 4300–4318.e13 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Weingarten-Gabbay S, Elias-Kirma S, Nir R, Gritsenko AA, Stern-Ginossar N, Yakhini Z, Weinberger A & Segal E Comparative genetics. Systematic discovery of cap-independent translation sequences in human and viral genomes. Science 351, aad4939 (2016). [DOI] [PubMed] [Google Scholar]
- 43.Lidsky PV, Dmitriev SE & Andino R Robust Expression of Transgenes in Drosophila Melanogaster. http://biorxiv.org/lookup/doi/10.1101/2022.10.30.514414 (2022) doi: 10.1101/2022.10.30.514414. [DOI] [Google Scholar]
- 44.Mäkeläinen KJ & Mäkinen K Testing of internal translation initiation via dicistronic constructs in yeast is complicated by production of extraneous transcripts. Gene 391, 275–284 (2007). [DOI] [PubMed] [Google Scholar]
- 45.Han B & Zhang J-T Regulation of gene expression by internal ribosome entry sites or cryptic promoters: the eIF4G story. Mol Cell Biol 22, 7372–7384 (2002). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Van Eden ME, Byrd MP, Sherrill KW & Lloyd RE Demonstrating internal ribosome entry sites in eukaryotic mRNAs using stringent RNA test procedures. RNA 10, 720–730 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Holcik M, Graber T, Lewis SM, Lefebvre CA, Lacasse E & Baird S Spurious splicing within the XIAP 5’ UTR occurs in the Rluc/Fluc but not the betagal/CAT bicistronic reporter system. RNA 11, 1605–1609 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Kozak M A second look at cellular mRNA sequences said to function as internal ribosome entry sites. Nucleic Acids Res 33, 6593–6602 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Baranick BT, Lemp NA, Nagashima J, Hiraoka K, Kasahara N & Logg CR Splicing mediates the activity of four putative cellular internal ribosome entry sites. Proc Natl Acad Sci U S A 105, 4733–4738 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Young RM, Wang S-J, Gordan JD, Ji X, Liebhaber SA & Simon MC Hypoxia-mediated selective mRNA translation by an internal ribosome entry site-independent mechanism. J Biol Chem 283, 16309–16319 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Andreev DE, Dmitriev SE, Terenin IM, Prassolov VS, Merrick WC & Shatsky IN Differential contribution of the m7G-cap to the 5’ end-dependent translation initiation of mammalian mRNAs. Nucleic Acids Res 37, 6135–6147 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Lemp NA, Hiraoka K, Kasahara N & Logg CR Cryptic transcripts from a ubiquitous plasmid origin of replication confound tests for cis-regulatory function. Nucleic Acids Res 40, 7280–7290 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Khan YA, Loughran G, Steckelberg A-L, Brown K, Kiniry SJ, Stewart H, Baranov PV, Kieft JS, Firth AE & Atkins JF Evaluating ribosomal frameshifting in CCR5 mRNA decoding. Nature 604, E16–E23 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Akirtava C, May GE & McManus CJ False-positive IRESes from Hoxa9 and other genes resulting from errors in mammalian 5’ UTR annotations. Proc Natl Acad Sci U S A 119, e2122170119 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Nejepinska J, Malik R, Moravec M & Svoboda P Deep sequencing reveals complex spurious transcription from transiently transfected plasmids. PLoS One 7, e43283 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Loughran G, Fedorova AD, Khan YA, Atkins JF & Baranov PV Lack of evidence for ribosomal frameshifting in ATP7B mRNA decoding. Mol Cell 82, 3745–3749.e2 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Terenin IM, Andreev DE, Dmitriev SE & Shatsky IN A novel mechanism of eukaryotic translation initiation that is neither m7G-cap-, nor IRES-dependent. Nucleic Acids Res 41, 1807–1816 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Payne AJ, Gerdes BC, Kaja S & Koulen P Insert sequence length determines transfection efficiency and gene expression levels in bicistronic mammalian expression vectors. Int J Biochem Mol Biol 4, 201–208 (2013). [PMC free article] [PubMed] [Google Scholar]
- 59.Shikama Y, Hu H, Ohno M, Matsuoka I, Shichishima T & Kimura J Transcripts expressed using a bicistronic vector pIREShyg2 are sensitized to nonsense-mediated mRNA decay. BMC Mol Biol 11, 42 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Jünemann C, Song Y, Bassili G, Goergen D, Henke J & Niepmann M Picornavirus internal ribosome entry site elements can stimulate translation of upstream genes. J Biol Chem 282, 132–141 (2007). [DOI] [PubMed] [Google Scholar]
- 61.Yordanova MM, Loughran G, Atkins JF & Baranov PV Stop codon readthrough contexts influence reporter expression differentially depending on the presence of an IRES. Wellcome Open Res 5, 221 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Terenin IM, Smirnova VV, Andreev DE, Dmitriev SE & Shatsky IN A researcher’s guide to the galaxy of IRESs. Cell Mol Life Sci 74, 1431–1455 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Kozak M New ways of initiating translation in eukaryotes? Mol Cell Biol 21, 1899–1907 (2001). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Schneider R, Agol VI, Andino R, Bayard F, Cavener DR, Chappell SA, Chen JJ, Darlix JL, Dasgupta A, Donzé O, Duncan R, Elroy-Stein O, Farabaugh PJ, Filipowicz W, Gale M, Gehrke L, Goldman E, Groner Y, Harford JB, Hatzglou M, He B, Hellen CU, Hentze MW, Hershey J, Hershey P, Hohn T, Holcik M, Hunter CP, Igarashi K, Jackson R, Jagus R, Jefferson LS, Joshi B, Kaempfer R, Katze M, Kaufman RJ, Kiledjian M, Kimball SR, Kimchi A, Kirkegaard K, Koromilas AE, Krug RM, Kruys V, Lamphear BJ, Lemon S, Lloyd RE, Maquat LE, Martinez-Salas E, Mathews MB, Mauro VP, Miyamoto S, Mohr I, Morris DR, Moss EG, Nakashima N, Palmenberg A, Parkin NT, Pe’ery T, Pelletier J, Peltz S, Pestova TV, Pilipenko EV, Prats AC, Racaniello V, Read GS, Rhoads RE, Richter JD, Rivera-Pomar R, Rouault T, Sachs A, Sarnow P, Scheper GC, Schiff L, Schoenberg DR, Semler BL, Siddiqui A, Skern T, Sonenberg N, Sossin W, Standart N, Tahara SM, Thomas AA, Toulmé JJ, Wilusz J, Wimmer E, Witherell G & Wormington M New ways of initiating translation in eukaryotes. Mol Cell Biol 21, 8238–8246 (2001). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Kozak M Alternative ways to think about mRNA sequences and proteins that appear to promote internal initiation of translation. Gene 318, 1–23 (2003). [DOI] [PubMed] [Google Scholar]
- 66.Jacobs JL & Dinman JD Systematic analysis of bicistronic reporter assay data. Nucleic Acids Res 32, e160 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Mardanova ES, Zamchuk LA & Ravin NV Contribution of internal initiation to translation of cellular mRNAs containing IRESs. Biochem Soc Trans 36, 694–697 (2008). [DOI] [PubMed] [Google Scholar]
- 68.Thompson SR So you want to know if your message has an IRES? Wiley Interdiscip Rev RNA 3, 697–705 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Jackson RJ The current status of vertebrate cellular mRNA IRESs. Cold Spring Harb Perspect Biol 5, a011569 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Shatsky IN, Dmitriev SE, Andreev DE & Terenin IM Transcriptome-wide studies uncover the diversity of modes of mRNA recruitment to eukaryotic ribosomes. Crit Rev Biochem Mol Biol 49, 164–177 (2014). [DOI] [PubMed] [Google Scholar]
- 71.Bustin SA Absolute quantification of mRNA using real-time reverse transcription polymerase chain reaction assays. J Mol Endocrinol 25, 169–193 (2000). [DOI] [PubMed] [Google Scholar]
- 72.Shiraki T, Kondo S, Katayama S, Waki K, Kasukawa T, Kawaji H, Kodzius R, Watahiki A, Nakamura M, Arakawa T, Fukuda S, Sasaki D, Podhajska A, Harbers M, Kawai J, Carninci P & Hayashizaki Y Cap analysis gene expression for high-throughput analysis of transcriptional starting point and identification of promoter usage. Proc Natl Acad Sci U S A 100, 15776–15781 (2003). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Atkins JF, Wills NM, Loughran G, Wu C-Y, Parsawar K, Ryan MD, Wang C-H & Nelson CC A case for ‘StopGo’: reprogramming translation to augment codon meaning of GGN by promoting unconventional termination (Stop) after addition of glycine and then allowing continued translation (Go). RNA 13, 803–810 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Ryan MD & Drew J Foot-and-mouth disease virus 2A oligopeptide mediated cleavage of an artificial polyprotein. EMBO J 13, 928–933 (1994). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Minskaia E, Nicholson J & Ryan MD Optimisation of the foot-and-mouth disease virus 2A co-expression system for biomedical applications. BMC Biotechnol 13, 67 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Kim JH, Lee S-R, Li L-H, Park H-J, Park J-H, Lee KY, Kim M-K, Shin BA & Choi S-Y High cleavage efficiency of a 2A peptide derived from porcine teschovirus-1 in human cell lines, zebrafish and mice. PLoS One 6, e18556 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Donnelly MLL, Hughes LE, Luke G, Mendoza H, Ten Dam E, Gani D & Ryan MD The ‘cleavage’ activities of foot-and-mouth disease virus 2A site-directed mutants and naturally occurring ‘2A-like’ sequences. J Gen Virol 82, 1027–1041 (2001). [DOI] [PubMed] [Google Scholar]
- 78.Sahin U, Karikó K & Türeci Ö mRNA-based therapeutics--developing a new class of drugs. Nat Rev Drug Discov 13, 759–780 (2014). [DOI] [PubMed] [Google Scholar]
- 79.Pardi N, Hogan MJ, Porter FW & Weissman D mRNA vaccines — a new era in vaccinology. Nat Rev Drug Discov 17, 261–279 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Juszkiewicz S & Hegde RS Initiation of Quality Control during Poly(A) Translation Requires Site-Specific Ribosome Ubiquitination. Molecular Cell 65, 743–750.e4 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Sundaramoorthy E, Leonard M, Mak R, Liao J, Fulzele A & Bennett EJ ZNF598 and RACK1 Regulate Mammalian Ribosome-Associated Quality Control Function by Mediating Regulatory 40S Ribosomal Ubiquitylation. Molecular Cell 65, 751–760.e4 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Matsuo Y, Ikeuchi K, Saeki Y, Iwasaki S, Schmidt C, Udagawa T, Sato F, Tsuchiya H, Becker T, Tanaka K, Ingolia NT, Beckmann R & Inada T Ubiquitination of stalled ribosome triggers ribosome-associated quality control. Nat Commun 8, 159 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Ikeuchi K, Tesina P, Matsuo Y, Sugiyama T, Cheng J, Saeki Y, Tanaka K, Becker T, Beckmann R & Inada T Collided ribosomes form a unique structural interface to induce Hel2-driven quality control pathways. The EMBO Journal 38, e100276 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Kisly I, Kattel C, Remme J & Tamm T Luciferase-based reporter system for in vitro evaluation of elongation rate and processivity of ribosomes. Nucleic Acids Research 49, e59–e59 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Akulich KA, Andreev DE, Terenin IM, Smirnova VV, Anisimova AS, Makeeva DS, Arkhipova VI, Stolboushkina EA, Garber MB, Prokofjeva MM, Spirin PV, Prassolov VS, Shatsky IN & Dmitriev SE Four translation initiation pathways employed by the leaderless mRNA in eukaryotes. Sci Rep 6, 37905 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Barreau C, Dutertre S, Paillard L & Osborne HB Liposome-mediated RNA transfection should be used with caution. RNA 12, 1790–1793 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Paramasivam P, Franke C, Stöter M, Höijer A, Bartesaghi S, Sabirsh A, Lindfors L, Arteta MY, Dahlén A, Bak A, Andersson S, Kalaidzidis Y, Bickle M & Zerial M Endosomal escape of delivered mRNA from endosomal recycling tubules visualized at the nanoscale. J Cell Biol 221, e202110137 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Hollien J, Lin JH, Li H, Stevens N, Walter P & Weissman JS Regulated Ire1-dependent decay of messenger RNAs in mammalian cells. J Cell Biol 186, 323–331 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Han D, Lerner AG, Vande Walle L, Upton J-P, Xu W, Hagen A, Backes BJ, Oakes SA & Papa FR IRE1alpha kinase activation modes control alternate endoribonuclease outputs to determine divergent cell fates. Cell 138, 562–575 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Karasik A, Jones GD, DePass AV & Guydosh NR Activation of the antiviral factor RNase L triggers translation of non-coding mRNA sequences. Nucleic Acids Res 49, 6007–6026 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Malka Y, Alkan F, Ju S, Körner P-R, Pataskar A, Shulman E, Loayza-Puch F, Champagne J, Wenzel C, Faller WJ, Elkon R, Lee C & Agami R Alternative cleavage and polyadenylation generates downstream uncapped RNA isoforms with translation potential. Mol Cell 82, 3840–3855.e8 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Andreev DE, Terenin IM, Dmitriev SE & Shatsky IN Pros and cons of pDNA and mRNA transfection to study mRNA translation in mammalian cells. Gene 578, 1–6 (2016). [DOI] [PubMed] [Google Scholar]
- 93.Matreyek KA, Stephany JJ, Chiasson MA, Hasle N & Fowler DM An improved platform for functional assessment of large protein libraries in mammalian cells. Nucleic Acids Res 48, e1 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.She R, Luo J & Weissman JS Translational fidelity screens in mammalian cells reveal eIF3 and eIF4G2 as regulators of start codon selectivity. Nucleic Acids Res 51, 6355–6369 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Dmitriev SE, Andreev DE, Ad’ianova ZV, Terenin IM & Shatskiĭ IN [Efficient cap-dependent in vitro and in vivo translation of mammalian mRNAs with long and highly structured 5’-untranslated regions]. Mol Biol (Mosk) 43, 119–125 (2009). [PubMed] [Google Scholar]
- 96.Kozak M Evaluation of the fidelity of initiation of translation in reticulocyte lysates from commercial sources. Nucleic Acids Res 18, 2828 (1990). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Soto Rifo R, Ricci EP, Décimo D, Moncorgé O & Ohlmann T Back to basics: the untreated rabbit reticulocyte lysate as a competitive system to recapitulate cap/poly(A) synergy and the selective advantage of IRES-driven translation. Nucleic Acids Res 35, e121 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Dmitriev SE, Bykova NV, Andreev DE & Terenin IM [Adequate system for investigation of translation initiation of the human retrotransposon L1 mRNA in vitro]. Mol Biol (Mosk) 40, 25–30 (2006). [DOI] [PubMed] [Google Scholar]
- 99.Iizuka N, Najita L, Franzusoff A & Sarnow P Cap-dependent and cap-independent translation by internal initiation of mRNAs in cell extracts prepared from Saccharomyces cerevisiae. Mol Cell Biol 14, 7322–7330 (1994). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Tuite MF & McLaughlin CS Polyamines enhance the efficiency of tRNA-mediated readthrough of amber and UGA termination codons in a yeast cell-free system. Curr Genet 7, 421–426 (1983). [DOI] [PubMed] [Google Scholar]
- 101.Skabkin MA, Skabkina OV, Hellen CUT & Pestova TV Reinitiation and other unconventional posttermination events during eukaryotic translation. Mol Cell 51, 249–264 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Zinoviev A, Hellen CUT & Pestova TV Multiple mechanisms of reinitiation on bicistronic calicivirus mRNAs. Mol Cell 57, 1059–1073 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Terenin IM, Dmitriev SE, Andreev DE, Royall E, Belsham GJ, Roberts LO & Shatsky IN A cross-kingdom internal ribosome entry site reveals a simplified mode of internal ribosome entry. Mol Cell Biol 25, 7879–7888 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Abaeva IS, Pestova TV & Hellen CUT Attachment of ribosomal complexes and retrograde scanning during initiation on the Halastavi árva virus IRES. Nucleic Acids Res 44, 2362–2377 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Beznosková P, Pavlíková Z, Zeman J, Echeverría Aitken C & Valášek LS Yeast applied readthrough inducing system (YARIS): an invivo assay for the comprehensive study of translational readthrough. Nucleic Acids Res 47, 6339–6350 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Beznosková P, Gunišová S & Valášek LS Rules of UGA-N decoding by near-cognate tRNAs and analysis of readthrough on short uORFs in yeast. RNA 22, 456–466 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Powell ML, Leigh KE, Pöyry TAA, Jackson RJ, Brown TDK & Brierley I Further characterisation of the translational termination-reinitiation signal of the influenza B virus segment 7 RNA. PLoS One 6, e16822 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Luttermann C & Meyers G A bipartite sequence motif induces translation reinitiation in feline calicivirus RNA. J Biol Chem 282, 7056–7065 (2007). [DOI] [PubMed] [Google Scholar]
- 109.Pöyry TAA, Kaminski A, Connell EJ, Fraser CS & Jackson RJ The mechanism of an exceptional case of reinitiation after translation of a long ORF reveals why such events do not generally occur in mammalian mRNA translation. Genes Dev 21, 3149–3162 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Gould PS, Dyer NP, Croft W, Ott S & Easton AJ Cellular mRNAs access second ORFs using a novel amino acid sequence-dependent coupled translation termination-reinitiation mechanism. RNA 20, 373–381 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Michel AM, Kiniry SJ, O’Connor PBF, Mullan JP & Baranov PV GWIPS-viz: 2018 update. Nucleic Acids Res 46, D823–D830 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Kiniry SJ, Judge CE, Michel AM & Baranov PV Trips-Viz: an environment for the analysis of public and user-generated ribosome profiling data. Nucleic Acids Res 49, W662–W670 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Andreev DE, O’Connor PBF, Loughran G, Dmitriev SE, Baranov PV & Shatsky IN Insights into the mechanisms of eukaryotic translation gained with ribosome profiling. Nucleic Acids Res 45, 513–526 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Ingolia NT, Hussmann JA & Weissman JS Ribosome Profiling: Global Views of Translation. Cold Spring Harb Perspect Biol 11, a032698 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Gunišová S, Hronová V, Mohammad MP, Hinnebusch AG & Valášek LS Please do not recycle! Translation reinitiation in microbes and higher eukaryotes. FEMS Microbiol Rev 42, 165–192 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Kozak M Constraints on reinitiation of translation in mammals. Nucleic Acids Res 29, 5226–5232 (2001). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Sorokin II, Vassilenko KS, Terenin IM, Kalinina NO, Agol VI & Dmitriev SE Non-Canonical Translation Initiation Mechanisms Employed by Eukaryotic Viral mRNAs. Biochemistry (Mosc) 86, 1060–1094 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Makeeva DS, Riggs CL, Burakov AV, Ivanov PA, Kushchenko AS, Bykov DA, Popenko VI, Prassolov VS, Ivanov PV & Dmitriev SE Relocalization of Translation Termination and Ribosome Recycling Factors to Stress Granules Coincides with Elevated Stop-Codon Readthrough and Reinitiation Rates upon Oxidative Stress. Cells 12, 259 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Brown CM, Dinesh-Kumar SP & Miller WA Local and distant sequences are required for efficient readthrough of the barley yellow dwarf virus PAV coat protein gene stop codon. J Virol 70, 5884–5892 (1996). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Xu Y, Ju H-J, DeBlasio S, Carino EJ, Johnson R, MacCoss MJ, Heck M, Miller WA & Gray SM A Stem-Loop Structure in Potato Leafroll Virus Open Reading Frame 5 (ORF5) Is Essential for Readthrough Translation of the Coat Protein ORF Stop Codon 700 Bases Upstream. J Virol 92, e01544–17 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Kimura K, Wakamatsu A, Suzuki Y, Ota T, Nishikawa T, Yamashita R, Yamamoto J, Sekine M, Tsuritani K, Wakaguri H, Ishii S, Sugiyama T, Saito K, Isono Y, Irie R, Kushida N, Yoneyama T, Otsuka R, Kanda K, Yokoi T, Kondo H, Wagatsuma M, Murakawa K, Ishida S, Ishibashi T, Takahashi-Fujii A, Tanase T, Nagai K, Kikuchi H, Nakai K, Isogai T & Sugano S Diversification of transcriptional modulation: large-scale identification and characterization of putative alternative promoters of human genes. Genome Res 16, 55–65 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Makhnovskii PA, Gusev OA, Bokov RO, Gazizova GR, Vepkhvadze TF, Lysenko EA, Vinogradova OL, Kolpakov FA & Popov DV Alternative transcription start sites contribute to acute-stress-induced transcriptome response in human skeletal muscle. Hum Genomics 16, 24 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Huang F, Gonçalves C, Bartish M, Rémy-Sarrazin J, Issa ME, Cordeiro B, Guo Q, Emond A, Attias M, Yang W, Plourde D, Su J, Gimeno MG, Zhan Y, Galán A, Rzymski T, Mazan M, Masiejczyk M, Faber J, Khoury E, Benoit A, Gagnon N, Dankort D, Journe F, Ghanem GE, Krawczyk CM, Saragovi HU, Piccirillo CA, Sonenberg N, Topisirovic I, Rudd CE, Miller WH & Del Rincón SV Inhibiting the MNK1/2-eIF4E axis impairs melanoma phenotype switching and potentiates antitumor immune responses. J Clin Invest 131, e140752, 140752 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Watt K, Dauber B, Szkop KJ, Lee L, Jovanovic P, Chen S, Masvidal L, Tandoc K, Palia R, Topisirovic I, Larsson O & Postovit L-M Epigenetic Coordination of Transcriptional and Translational Programs in Hypoxia. http://biorxiv.org/lookup/doi/10.1101/2023.09.16.558067 (2023) doi: 10.1101/2023.09.16.558067. [DOI] [Google Scholar]
- 125.Wang X-Q & Rothnagel JA 5’-untranslated regions with multiple upstream AUG codons can support low-level translation via leaky scanning and reinitiation. Nucleic Acids Res 32, 1382–1391 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Dmitriev SE, Andreev DE, Terenin IM, Olovnikov IA, Prassolov VS, Merrick WC & Shatsky IN Efficient translation initiation directed by the 900-nucleotide-long and GC-rich 5’ untranslated region of the human retrotransposon LINE-1 mRNA is strictly cap dependent rather than internal ribosome entry site mediated. Mol Cell Biol 27, 4685–4697 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]