+ All Categories
Home > Documents > Active ribosome profiling with RiboLaceActive ribosome profiling with RiboLace . Massimiliano...

Active ribosome profiling with RiboLaceActive ribosome profiling with RiboLace . Massimiliano...

Date post: 26-Jan-2021
Category:
Upload: others
View: 3 times
Download: 0 times
Share this document with a friend
21
Active ribosome profiling with RiboLace Massimiliano Clamer 1,2* , Toma Tebaldi 1 , Fabio Lauria 3 , Paola Bernabò 3 , Rodolfo F. Gómez-Biagi 4 , Elena Perenthaler 3 , Daniele Gubert 3,5 , Laura Pasquardini 4 , Graziano Guella 6 , Ewout J.N. Groen 7 , Thomas H. Gillingwater 7 , Alessandro Quattrone 1 , Gabriella Viero 3* 1 Centre for Integrative Biology, University of Trento, Via Sommarive, 9 Povo (Italy). 2 IMMAGINA Biotechnology srl, Via alla cascata 56/c, Povo (Italy). 3 Institute of Biophysics, CNR Unit at Trento, Via Sommarive, 18 Povo (Italy). 4 Fondazione Bruno Kessler-LaBSSAH, Via Sommarive 18, Povo, Trento, Italy. 5 Department of Information Engineering and Computer Science, University of Trento, Povo (Italy). 6 Department of Physics, University of Trento, Povo (Italy). 7 Edinburgh Medical School: Biomedical Sciences, University of Edinburgh, Edinburgh, UK *corresponding authors Ribosome profiling, or Ribo-Seq, is based around large-scale sequencing of RNA fragments protected from nuclease digestion by ribosomes. Thanks to its unique ability to provide positional information concerning ribosomes flowing along transcripts, this method can be used to shed light on mechanistic aspects of translation. However, current Ribo-Seq approaches lack the ability to distinguish between fragments protected by ribosomes in active translation or by inactive ribosomes. To overcome these significant limitation, we developed RiboLace: a novel method based on an original puromycin-containing molecule capable of isolating active ribosomes by means of an antibody-free and tag-free pull-down approach. RiboLace is fast, works reliably with low amounts of input material, and can be easily and rapidly applied both in vitro and in vivo, thereby generating a global snapshot of active ribosome footprints at single nucleotide resolution. The tightly regulated process of protein synthesis is a core regulator of numerous critical physiological pathways, ranging from cell growth 1 and development 2,3 through to immune responses 4 . Local protein synthesis in neurons 5 also plays fundamental roles in memory formation 68 and synaptic plasticity 9 . Hence, dysregulation of translation is a major driver of important pathologies, such as cancer 10,11 and neurodegenerative diseases 12 . Over the last few years, newly developed methodological approaches such as ribosome profiling (Ribo-Seq) 13 , have contributed to considerable new insights into the translation process. Ribo-Seq has been largely employed to identify novel translated RNAs (coding and non-coding), map novel upstream Open Reading Frames (ORFs), and estimate translation levels in different biological conditions. Indeed, Ribo-Seq has the potential to estimate translation efficiencies and the protein synthesis levels14,15 in a variety of organisms, from prokaryotes 15 , to yeast 13 , C. elegans 16 , zebrafish 17,18 , plants 19 , the mouse 20 , and human cell lines 2123 . Despite its undoubtful discrimination power and wide applicability, Ribo-Seq still faces a number of challenges and presents with numerous limitations. For example, translationally inactive mRNAs can be sequestered into ribonucleoprotein particles (mRNP) or monosomes (80S), whose translational status remains a controversial not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which was this version posted August 22, 2017. ; https://doi.org/10.1101/179671 doi: bioRxiv preprint
Transcript
  • Active ribosome profiling with RiboLace

    Massimiliano Clamer1,2*, Toma Tebaldi1, Fabio Lauria3, Paola Bernabò3, Rodolfo F.

    Gómez-Biagi4, Elena Perenthaler3, Daniele Gubert3,5, Laura Pasquardini4, Graziano

    Guella6, Ewout J.N. Groen7, Thomas H. Gillingwater7, Alessandro Quattrone1,

    Gabriella Viero3*

    1 Centre for Integrative Biology, University of Trento, Via Sommarive, 9 Povo (Italy). 2 IMMAGINA

    Biotechnology srl, Via alla cascata 56/c, Povo (Italy). 3 Institute of Biophysics, CNR Unit at Trento, Via

    Sommarive, 18 Povo (Italy). 4 Fondazione Bruno Kessler-LaBSSAH, Via Sommarive 18, Povo,

    Trento, Italy. 5 Department of Information Engineering and Computer Science, University of Trento,

    Povo (Italy). 6 Department of Physics, University of Trento, Povo (Italy). 7 Edinburgh Medical School:

    Biomedical Sciences, University of Edinburgh, Edinburgh, UK *corresponding authors

    Ribosome profiling, or Ribo-Seq, is based around large-scale sequencing of RNA fragments

    protected from nuclease digestion by ribosomes. Thanks to its unique ability to provide

    positional information concerning ribosomes flowing along transcripts, this method can be

    used to shed light on mechanistic aspects of translation. However, current Ribo-Seq

    approaches lack the ability to distinguish between fragments protected by ribosomes in

    active translation or by inactive ribosomes. To overcome these significant limitation, we

    developed RiboLace: a novel method based on an original puromycin-containing molecule

    capable of isolating active ribosomes by means of an antibody-free and tag-free pull-down

    approach. RiboLace is fast, works reliably with low amounts of input material, and can be

    easily and rapidly applied both in vitro and in vivo, thereby generating a global snapshot of

    active ribosome footprints at single nucleotide resolution.

    The tightly regulated process of protein synthesis is a core regulator of numerous

    critical physiological pathways, ranging from cell growth1 and development2,3 through

    to immune responses4. Local protein synthesis in neurons5 also plays fundamental

    roles in memory formation6–8 and synaptic plasticity9. Hence, dysregulation of

    translation is a major driver of important pathologies, such as cancer10,11 and

    neurodegenerative diseases12.

    Over the last few years, newly developed methodological approaches such as

    ribosome profiling (Ribo-Seq)13, have contributed to considerable new insights into

    the translation process. Ribo-Seq has been largely employed to identify novel

    translated RNAs (coding and non-coding), map novel upstream Open Reading

    Frames (ORFs), and estimate translation levels in different biological conditions.

    Indeed, Ribo-Seq has the potential to estimate translation efficiencies and the

    “protein synthesis levels”14,15 in a variety of organisms, from prokaryotes15, to

    yeast13, C. elegans16, zebrafish17,18, plants19, the mouse20, and human cell lines21–23.

    Despite its undoubtful discrimination power and wide applicability, Ribo-Seq still

    faces a number of challenges and presents with numerous limitations. For example,

    translationally inactive mRNAs can be sequestered into ribonucleoprotein particles

    (mRNP) or monosomes (80S), whose translational status remains a controversial

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

  • issue24–26. Importantly, mRNAs can be trapped within stalled or paused polysomes,

    as has been shown especially in neurons12,27–33. As such, Ribo-Seq does not

    necessarily discriminate “true” protected footprints of translating polysomes from

    RNA fragments protected by the 80S ribosome or stalled ribosomes in polysomes,

    leading to possible misinterpretations of translation occupancy profiles. Therefore,

    Ribo-Seq still requires further optimization and refinements, from the bench to the

    data analysis34, in order to generate maximal insights into the translation process.

    Here, we present RiboLace, a novel methodological approach based on a newly

    developed reagent, a puromycin analog molecule. RiboLace greatly improves the

    study of translation both in vitro and in vivo, by selectively portraying mRNA

    fragments protected by bona fide active ribosomes at single nucleotide resolution

    and with unprecedented simplicity, requiring 30 times less biological material than

    current protocols.

    RESULTS

    Design and synthesis of a new analog of puromycin

    Puromycin is an aminonucleoside antibiotic able to bind the catalytic center of

    the ribosome and the nascent peptide chain, causing ribosome disassembly and

    disruption of protein synthesis35–38. Over the years, it has been used extensively to

    quantify global of protein synthesis, taking advantage of radioactive39 and

    biotinylated molecules40 or anti-puromycin antibodies41. Leveraging its ability to keep

    contact with the ribosome42–45, puromycin has also been employed to covalently link

    an mRNA to the corresponding protein during its synthesis46,47. In addition,

    puromycin can be modified to create cell-permeable analogues suitable for direct

    and in situ imaging of newly synthesized proteins48,49. All these methods require the

    irreversible reaction of the α-amino group of puromycin with the carbon on its

    carbonyl group, acylating the 3ʼ hydroxyl group of the peptidyl-tRNA buried in the P-

    site of the ribosome.

    Motivated by evidence that molecules containing α-amino group modified puromycin

    can bind the large subunit of the active ribosome45, we covalently coupled puromycin

    to a biotin moiety through two 2,2ʼ-ethylenedioxy-bis-ethylamine units, to obtain a

    new compound still able to bind ribosomes by mimicking the tRNA entrance in the

    acceptor site (A-site). We synthesized the new molecule at purity higher than 90%,

    characterized it by NMR (Fig. 1a and Supplementary Fig. 1-6) and called it 3P.

    Then, we verified the activity of the biotin moiety, taking advantage of its absorbance

    spectrum and tested the binding on polystyrene or agarose beads. We observed that

    the biotin group allows the binding of the 3P to commercially available streptavidin-

    beads (Fig. 1b).

    To demonstrate that 3P molecule maintains an inhibitory effect on translation,

    we compared its effects to that exerted by puromycin using an eukaryotic in vitro

    cell-free transcription-translation system (IVTT, rabbit reticulocyte lysate) and the

    firefly luciferase as a reporter gene. We monitored total protein production by SDS-

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

  • PAGE (Fig.1c, Supplementary Fig.7) and luminescence assay (Supplementary

    Fig.7) in the presence of puromycin and 3P at different concentrations. As

    expected, puromycin induced conspicuous decay of protein production at nanomolar

    concentrations. In the case of 3P, we observed a decreased level of translation,

    which reached ̴70% of inhibition at concentrations higher than 1 µM (Fig. 1c). We

    concluded that, even if with slightly lower efficiency than puromycin, 3P can inhibit

    eukaryotic translation in vitro, interfering with ribosome function, and can be used to

    produce functionalized beads.

    3P-functionalized magnetic beads captures mRNAs in active translation

    in vitro.

    Our finding that 3P is able to interact with the translation process, likely

    through its puromycin moiety, prompted us to investigate whether 3P can capture

    mRNAs under active translation.

    First, we monitored the ability of 3P-functionalized magnetic beads (3P-beads) to

    purify transcripts of reporter genes with different levels of protein expression in in

    vitro translation systems. To purify mRNAs with 3P-beads, we developed the

    following protocol (Fig. 2a): (i) 3P-beads and control beads functionalized with a

    biotin-glycol conjugate (mP-beads, see Fig. 2a legend and methods for details) were

    added to the in vitro translation system and the suspension was incubated for one

    hour at 4°C in an appropriate buffer containing cycloheximide, to antagonize the

    dissociation of ribosomal subunits exerted by the puromycin component of the

    molecule; (ii) beads were pulled down using a magnet and washed two times; (iii)

    protein and/or RNA were extracted for downstream analyses. In parallel, the

    production of protein was followed by immunoblotting of whole protein extracts.

    As expected, the EGFP reporter gene showed differential efficiency in protein

    production depending on the expression vector used (pBluescript II KS+50, low

    performance protein production and IPR- IBA251, high performance protein

    production) (Fig. 2b, upper panel and Supplementary Fig. 8). Complete inhibition

    of protein production was observed after addition of the translation inhibitor

    harringtonine, as a control. RNA was purified and the relative abundance of the

    EGFP mRNA, in both low and high performant conditions, was monitored by qRT-

    PCR. We observed on 3P functionalized beads a 7 to 10-fold enrichment of the

    reporter transcript in conditions of active translation, with respect to samples treated

    with harringtonine (Fig. 2b lower panel, left), and in the absence of transcriptional

    changes (Fig. 2b lower panel, right). The decreased enrichment in highly

    performant conditions and at 120 min of incubation can be related to the fact that the

    in vitro system is a closed one, causing the reaction to stop in the absence of

    continuous supply of reactants during later stages of incubation.

    Then, to demonstrate that this result was not dependent on the reporter used,

    we applied RiboLace to a luciferase reporter system. In this case we observed a >

    1.6-fold enrichment of luciferase transcript with respect to negative controls (mP-

    beads) and an enrichment with respect to samples where translation had been

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

  • inhibited (Fig. 2c), as previously observed for EGFP (Supplementary Fig. 9).

    Finally, to understand if the observed enrichments were dependent on the puromycin

    moiety of 3P, we pre-saturated the system with puromycin and induced ribosome

    drop-off. Under these conditions, we found no evidence for mRNA enrichment.

    Overall, these findings support the claim that 3P-beads can be used to capture

    transcripts undergoing translation in eukaryotic in-vitro systems. We named the

    method RiboLace.

    RiboLace captures active ribosomes and associated mRNAs from whole

    cellular lysates

    Next, we wanted to establish whether RiboLace was capable of isolating

    ribosomes and mRNAs under active translation from more complex samples than in

    vitro mixtures. We used RiboLace on whole cellular lysates and under different

    translational states, and monitored the resulting proteins and mRNAs associated

    with the beads (Fig. 3a). We took advantage of established cellular stimuli that can

    induce cells into translationally active or inactive states. To shut-down translation we

    used cell starvation, oxidative, proteotoxic and heat stresses, known to globally

    suppress protein synthesis52 (Supplementary Fig. S10a). To specifically activate

    protein synthesis, we rescued cells from starvation by Epithelial Growth Factor

    (EGF) or by Fetal Bovine Serum (FBS) stimulation53.

    First, we monitored the enrichment on RiboLace of functional and structural

    markers of ribosomes in lysates of immortalized human cells (HEK-293T).

    Importantly, we used 2 x 105 cells, representing ̴1/30th of what classical polysomal

    profiling approaches require. We monitored the ability of RiboLace to purify proteins

    known to be associated with the translation machinery (eEF1α, calnexin, RPL26 and

    RPS6). The elongation factor eEF1α is responsible for the delivery of aminoacyl-

    tRNAs to the translation machinery and is associated to ribosomes in active

    translation54,55. Calnexin is a chaperone protein in the endoplasmic reticulum that

    associate with ribosomes, helping protein folding during translation56. Finally, RPL26

    and RPS6 are ribosomal proteins belonging to the large and small subunits of the

    ribosome, respectively. After immunoblot on the RiboLace eluted proteins, in lysates

    from untreated cells (nt, Fig 3b and Fig. 3b), we observed an enrichment of all four

    proteins (Fig. 3c and Fig.3d) with respect to serum starvation (st). When cells were

    stimulated with EGF after starvation (st. + EGF 1 µg/mL, 4 hours, after starvation),

    we observed a slight increase in the signal of the translational markers (Fig. 3c).

    Since the background signal in controls (mP-beads) was the same in all conditions,

    our results suggest that RiboLace can pull-down active ribosomes and therefore

    monitoring the translational state of cells.

    To further confirm this result, we compared the relative abundances on

    RiboLace of eEF1α and eEF2 between no-stress and stress conditions. It is known

    that eEF2-mediated translocation, as well as the switch of ribosome conformation

    from the non-rotated to the rotated state57, is inhibited by cycloheximide58. We

    observed that the two proteins differentially co-sedimented along the polysomal

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

  • profile, with eEF1α mainly detected in the heavy fractions (Supplementary Fig. 10b)

    indicating a preferential association of eEF1α to ribosomes in cycloheximide treated

    cells. Then, in agreement with the hypothesis that RiboLace captures active

    ribosomes, we found enrichment of eEF1α, but not of eEF2 (Fig. 3e and

    Supplementary Fig. 10c and Fig 11), suggesting the capture of the pre-

    translocation complex in the non-rotated conformation.

    Next, we tested RiboLace in a different cell line, the widely used human tumor

    cell line (MCF7), in control (nt) or starved (st) and control conditions , again

    establishing levels of translational markers associated to RiboLace. In addition to

    ribosomal proteins, we detected additional proteins known to be associated to

    polysomes (PABP, eIF4B), or a marker of active translation (me(K9)H3)59 (Fig 3f

    and Supplementary Fig. S10d). In agreement with results obtained for HEK-293T

    cells, the decrease in translational markers associated to the beads in starved cells

    suggests that RiboLace captures less ribosomes when global translation is

    downregulated in different cell lines. To further validate this finding, we applied other

    stresses known to elicit repression of global protein synthesis (i.e. proteotoxic stress,

    heat shock and sodium arsenite). In all cases, translation markers were reduced

    (Supplementary Fig. 10e). We then tested RiboLace on a mouse motor-neuron like

    cell line, NSC-34, in normal growth conditions. Also in this case we observed an ̴8-

    fold enrichment of RPL26 and ̴ 4-fold enrichment of RPS6 with respect to control

    beads (Supplementary Fig. 10f, right), demonstrating that RiboLace can efficiently

    capture ribosomes from both human and mouse cell lines. Lastly, we wanted to

    establish whether RiboLace could capture components of the eukaryotic surveillance

    mechanisms that target stalled elongation complexes. Among them, Pelota

    (mammalian orthologue of the yeast Dom34)60,61 is a protein known to promote the

    dissociation of stalled ribosomes. Strikingly, in control, starved or arsenite treated

    lysates of HEK-293T and MCF7 cells, Pelota was not enriched on RiboLace beads

    (Fig. 3g).

    These findings prompted us to investigate whether RiboLace can provide an

    improved estimation of translation efficiency, with respect to the use of total RNA or

    polysomal RNA in profiling experiments (Fig. 3h and Supplementary Fig. S12). We

    coupled RiboLace to Next Generation Sequencing (NGS) and identified sets of

    differentially expressed mRNAs associated to RiboLace before and after EGF

    stimulation (Supplementary Fig. 13 and Supplementary Fig. 14). We compared

    these data with transcripts identified by classical transcriptome (RNA-Seq) or

    translatome (POL-Seq, i.e. polysomal profiling62) analyses. Eight found differentially

    expressed genes were selected for validation by RT-qPCR (PALLD, PLK3, IL27RA,

    NCS1, VEGFA, DUSP5, PDCD4, PAPSS2), showing a good agreement with NGS

    (Pearsonʼs r = 0.89) (Supplementary Fig.15). Then, we compared the protein levels

    of four genes (PALLD, PLK3, hb-EGF and CYP27A1) to the relative RNA

    abundances. PALLD and PLK3 proteins, evaluated by immunoblotting and

    normalised for three different housekeeping proteins (ACTB, GAPDH and RPL26),

    did not change upon EGF treatment (Fig 3i). Importantly, only RiboLace (RL) did not

    show significant variation on both RT-qPCR and RNA-Seq measurements. The

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

  • same concordance between protein level and RiboLace was obtained with hb-EGF

    and CYP27A1, whose protein levels increased coherently with RiboLace

    (Supplementary Fig. S16). Overall, these results establish the important proof-of-

    concept that RiboLace can capture ribosomes under active translation and

    determine protein variations more precisely than total or polysomal profiling, at least

    in our case study.

    In vivo active-ribosome profiling using RiboLace

    Given the relative simplicity of the RiboLace protocol, and its ability to be used

    with low amounts of input material, we next wanted to combine our purification

    strategy with Ribo-Seq. This would establish whether RiboLace could be used to

    capture active ribosome dynamics along transcripts, improving in vivo ribosome

    profiling experiments. To facilitate this, we modified our original protocol by including

    an endonuclease digestion step and applied it to lysates from mouse tissues (Fig.

    4a).

    To demonstrate that RiboLace can capture isolated ribosomes after

    endonuclease digestion, we first measured the enrichment of eEF1α, calnexin,

    RPL26 and RPS6 on RiboLace applied to control (nt), harringtonine treated (harr),

    serum starved (st) or heat-shocked (h.s.) HEK-293T and HeLa cells (Fig. 4b). Our

    results confirmed that RiboLace can selectively capture isolated ribosomes under

    conditions of active translation. Then, we confirmed that RiboLace was able to enrich

    ribosome protected fragments, by urea-gel electrophoresis and by the use of the

    Bioanalyzer (Supplementary Fig. S17). After that, we probed RiboLace on 15 µL ( ̴

    1/50thof the total lysate) samples of whole brain lysates from wild-type (WT) and

    mice affected by the “wasted” mutation, consisting in a deletion of the elongation

    factor eEF1α2 (Fig.4c), resulting in defective translation63. Our results, in agreement

    with those previously obtained from cells, showed that RiboLace could selectively

    enrich RPL26 only in the WT mouse tissue, extending its functionality from cell lines

    to tissues.

    Having established that RiboLace can capture isolated ribosomes from very

    low amounts of starting material, we next isolated ribosome protected fragments 25-

    35 nt long (RPFs) from ribosomes pulled down by RiboLace and, in parallel, from

    polysomes isolated from whole mouse brain, following nuclease digestion (polysomal

    Ribo-Seq, Fig. 4d and Supplementary Fig S18). After sequencing, we analyzed

    both RiboLace and polysomal Ribo-Seq RPFs using the dedicated pipeline

    RiboWaltz64 to obtain sub-codon information and identification of the trinucleotide

    periodicity. The distribution of the reads showed a main population of length at ̴ 28

    nt (Supplementary Fig. S19), in agreement with what has previously been observed

    for ribosomes trapped on the mRNA by cycloheximide65,66. As expected for ribosome

    footprints, we observed an enrichment of signal along the coding sequence in both

    RiboLace and polysomal Ribo-Seq data (Fig. 4e). Occupancy meta-profiles showed

    the typical trinucleotide periodicity of the ribosome P-site along coding sequences,

    which is suggestive of signal derived from translating ribosomes (Fig.4f and g). The

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

  • comparison between meta-profiles obtained with RiboLace and polysomal Ribo-Seq

    highlights an accumulation of ribosomes at the start codon and at the 5th codon (Fig.

    4h). Interestingly, the latter feature is associated with ribosome pauses necessary for

    a productive elongation phase of translation67.

    Taken together, these results confirm that RiboLace is capable of providing

    positional data with nucleotide resolution and of enriching samples with active

    ribosomes, thereby facilitating reliable descriptions of bona fide translational events

    in vitro and in vivo.

    DISCUSSION

    During their lifetime in the cellular cytoplasm, mRNAs are regularly stored, degraded,

    and transported, with only a fraction being actively translated to produce

    proteins68,69. All these stages of the mRNA lifecycle are governed by cis- and trans-

    factors that tightly regulate the uploading of mRNAs on polysomes, and subsequent

    production of proteins. In order to generate a better understanding of these

    sophisticated and dynamic processes, different methodological approaches have

    been developed to determine, at a genome-wide scale changes in RNA steady state

    levels (e.g. RNA-seq), the change in engagement with the translational machinery

    (e.g. Ribo-Seq, polysomal profiling)20,62, and the change in protein production (e.g.,

    SILAC, PUNCH-P)40,7071. Although Ribo-Seq remains a complex technology that

    requires relatively large tissue samples, it has been shown to be extremely powerful

    for identifying ORFs and translation initiation sites (i) from cell lysates or ribosome

    pellets; (ii) from purified polysomal fractions (polysomal Ribo-Seq)19 or, more

    recently, (iii) from tagged ribosomes72,73. Unfortunately, however, the use of cell

    lysates and ribosome pellets often introduces unwanted background signals,

    presumably mostly due to the presence of stalled ribosomes and fragments

    protected by RNPs, or by the 80S monosome.

    Here, we sought to significantly enhance Ribo-Seq approaches by developing a new

    molecule (3P) that facilitates the selective capture of ribosomes under active

    translation. We focused our attention on puromycin, the well-known structural

    analogue of the 3′ end of aminoacyl-tRNA, and we suppressed its irreversible activity

    by tethering the α-amino group to a biotinylated linker. Despite this modification on

    the primary amino group of puromycin, we observed that 3P could still interfere with

    eukaryotic translation in vitro. We used 3P- functionalized magnetic beads (a method

    we called RiboLace) to capture and enrich transcripts undergoing translation in

    eukaryotic in vitro and in vivo systems. We observed that the elongation factor

    eEF1α, a key protein involved in delivering tRNAs to the ribosome, was the most

    enriched protein on RiboLace in all our experiments. This may be explained with the

    binding of 3P to the A-site of the ribosome in the not-rotated state, when the

    acceptor site accommodates the aminoacyl-tRNA engaged by eEF1α. We then

    demonstrated that RiboLace is capable of providing positional data with nucleotide

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

  • resolution of translational events when used for ribosome profiling. Remarkably, we

    observed > 95% of RPFs on the coding region with the characteristic trinucleotide

    periodicity, suggestive of active ribosomes flowing along the transcripts. In our

    biological model, RiboLace Ribo-Seq showed almost no signal on the 5′- and 3′-

    untranslated regions of the mRNAs, and a peculiar pause of ribosomes at the 5th

    codon. Overall, our data suggest that RiboLace possess significant advantages, in

    terms of both sample yield and accuracy in RPFs detection, over classical Ribo-Seq

    and polysomal Ribo-Seq approaches.

    RiboLace protocols can be further adjusted to (i) isolate ribosomes from other

    organisms, (ii) isolate ribosomes from specific eukaryotic cellular compartments such

    as the Endoplasmic Reticulum or organelles like mitochondria, or (iii) to

    characterized active specialized-ribosomes2,72,74 induced by different cell conditions.

    We specifically designed RiboLace for ribosome profiling experiments, to facilitate

    improved understanding of ribosome dynamics along transcripts and to allow better

    estimates of translation levels based on ribosome footprints.

    In summary, we report a major advancement in the technology available to

    undertake ribosome profiling, providing the unique benefit of capturing ribosome in

    active translation. Given its ability to enrich actively translating ribosome protected

    fragments, RiboLace can be used in difficult samples with low input material, and

    empowers accurate conclusions to be drawn concerning the actual translational

    state of a biological system, paving the way for more detailed and accurate studies

    of translation in the future.

    METHODS

    Methods and Supplementary Figures are available in the file Supplementary

    Information.

    ACKNOWLEDGMENTS

    We thank Tocris Bioscience (a Biotechne brand) for the support in scaling up the 3P

    synthesis. We thank Divya Kandala and Luca Minati for the helpful discussions. We

    also thank Veronica Desanctis and Roberto Bertorelli of the CIBIO NGS for technical

    support and Daniele Arosio for the kindly gift of the IPR- IBA2 plasmid.

    AUTHOR CONTRIBUTION

    MC and GV conceived the experiments. MC performed all the biological experiments

    and contributed to the 3P synthesis. TT, DG and FL analyzed the RNA-seq and Ribo-

    Seq data. PB and EP performed the Poly Ribo-Seq and helped with the RiboLace

    Ribo-Seq. LP performed control experiments with mP-beads. RG performed a

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

  • preliminary synthesis of 3P. GG designed the 3P synthesis, performed MS analysis,

    COSY and NMR analysis of the 3P. EJNG and THG generated and provided the

    mouse tissues. MC and GV drafted the manuscript. TT, FL and DG wrote the

    methods related to RNA-seq and Ribo-Seq data analysis. GG wrote the methods

    related to the chemical synthesis and characterization. GV, MC, TT, FL, GG, THG

    and AQ reviewed the manuscript. All authors read and approved the final manuscript.

    FUNDINGS

    This work was supported by IMMAGINA BioTechnology s.r.l. and by the Provincia

    Autonoma di Trento, Italy (AxonomiX research project)with additional grant funding

    from the Wellcome Trust (to EJNG & THG) and UK SMA Research Consortium

    (SMA Trust to THG).

    COMPETING FINANCIAL INTEREST

    MC is founder, director and a share-holder of IMMAGINA BioTechnology s.r.l., a

    company engaged in the development of new technologies for gene expression

    analysis at the ribosomal level. AQ and GG are shareholders and scientific

    advisors of IMMAGINA BioTechnology s.r.l. GV is scientific advisor of

    IMMAGINA BioTechnology s.r.l. All other authors declare no competing financial

    interests. RiboLace is an IMMAGINA s.r.l. patented technology (WO2017013547

    A1 - PCT/IB2016/054210).

    Data availability. Raw and analyzed data for Ribosome profiling have been

    deposited under GEO: GSE102354 for RiboLace Ribo-Seq, and GEO: GSE102318

    for Poly Ribo-Seq

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

  • REFERENCES

    1. Goyer, C. et al. TIF4631 and TIF4632: two yeast genes encoding the high-molecular-weight subunits of the cap-binding protein complex (eukaryotic initiation factor 4F) contain an RNA recognition motif-like sequence and carry out an essential function. Mol. Cell. Biol. 13, 4860–4874 (1993).

    2. Kondrashov, N. et al. Ribosome-Mediated Specificity in Hox mRNA Translation and Vertebrate Tissue Patterning. Cell 145, 383–397 (2011).

    3. Xue, S. et al. RNA regulons in Hox 5′ UTRs confer ribosome specificity to gene regulation. Nature 517, 33–38 (2014).

    4. Piccirillo, C. A., Bjur, E., Topisirovic, I., Sonenberg, N. & Larsson, O. Translational control of immune responses: from transcripts to translatomes. Nat. Immunol. 15, 503–511 (2014).

    5. Jung, H., Gkogkas, C. G., Sonenberg, N. & Holt, C. E. Remote control of gene function by local translation. Cell 157, 26–40 (2014).

    6. Martin, K. C. et al. MAP Kinase Translocates into the Nucleus of the Presynaptic Cell and Is Required for Long-Term Facilitation in Aplysia. Neuron 18, 899–912 (1997).

    7. Kandel, E. R., Dudai, Y. & Mayford, M. R. The Molecular and Systems Biology of Memory. Cell 157, 163–186 (2014).

    8. Fioriti, L. et al. The Persistence of Hippocampal-Based Memory Requires Protein Synthesis Mediated by the Prion-like Protein CPEB3. Neuron 86, 1433–1448 (2015).

    9. McCamphill, P. K., Farah, C. A., Anadolu, M. N., Hoque, S. & Sossin, W. S. Bidirectional Regulation of eEF2 Phosphorylation Controls Synaptic Plasticity by Decoding Neuronal Activity Patterns. J. Neurosci. 35, 4403–4417 (2015).

    10. Topisirovic, I. & Sonenberg, N. Translation and cancer. Biochim. Biophys. Acta - Gene Regul. Mech. 1849, 751–752 (2015).

    11. Bhat, M. et al. Targeting the translation machinery in cancer. Nat. Rev. Drug Discov. 14, 261–278 (2015).

    12. Darnell, J. C. et al. FMRP Stalls Ribosomal Translocation on mRNAs Linked to Synaptic Function and Autism. Cell 146, 247–261 (2011).

    13. Ingolia, N. T., Ghaemmaghami, S., Newman, J. R. S. & Weissman, J. S. Genome-wide analysis in vivo of translation with nucleotide resolution using ribosome profiling. Science 324, 218–23 (2009).

    14. Ingolia, N. T. et al. Ribosome Profiling Reveals Pervasive Translation Outside of Annotated Protein-Coding Genes. Cell Rep. 8, 1365–1379 (2014).

    15. Li, G.-W., Burkhardt, D., Gross, C. & Weissman, J. S. Quantifying Absolute Protein Synthesis Rates Reveals Principles Underlying Allocation of Cellular Resources. Cell 157, 624–635 (2014).

    16. Stadler, M., Artiles, K., Pak, J. & Fire, A. Contributions of mRNA abundance, ribosome loading, and post- or peri-translational effects to temporal repression of C. elegans heterochronic miRNA targets. Genome Res. 22, 2418–2426 (2012).

    17. Chew, G.-L. et al. Ribosome profiling reveals resemblance between long non-coding RNAs and 5′ leaders of coding RNAs. Development 140, 2828–2834 (2013).

    18. Bazzini, A. A. et al. Identification of small ORFs in vertebrates using ribosome footprinting and evolutionary conservation. EMBO J. 33, 981–993 (2014).

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

  • 19. Juntawong, P., Girke, T., Bazin, J. & Bailey-Serres, J. Translational dynamics revealed by genome-wide profiling of ribosome footprints in Arabidopsis. Proc. Natl. Acad. Sci. U. S. A. 111, E203-12 (2014).

    20. Ingolia, N. T., Lareau, L. F. & Weissman, J. S. Ribosome profiling of mouse embryonic stem cells reveals the complexity and dynamics of mammalian proteomes. Cell 147, 789–802 (2011).

    21. Liu, B., Han, Y. & Qian, S.-B. Cotranslational Response to Proteotoxic Stress by Elongation Pausing of Ribosomes. Mol. Cell 49, 453–463 (2013).

    22. Lee, S. et al. Global mapping of translation initiation sites in mammalian cells at single-nucleotide resolution. Proc. Natl. Acad. Sci. 109, E2424–E2432 (2012).

    23. Fritsch, C. et al. Genome-wide search for novel human uORFs and N-terminal protein extensions using ribosomal footprinting. Genome Res. 22, 2208–2218 (2012).

    24. WARNER, J. R., KNOPF, P. M. & RICH, A. A multiple ribosomal structure in protein synthesis. Proc. Natl. Acad. Sci. {U.S.A.} 49, 122–129 (1963).

    25. Munro, A. J., Jackson, R. J. & Korner, A. Studies on the nature of polysomes. Biochem. J. 92, 289–99 (1964).

    26. Heyer, E. E. & Moore, M. J. Redefining the Translational Status of 80S Monosomes. Cell 164, 757–769 (2016).

    27. Chapman, R. E. & Walter, P. Translational attenuation mediated by an mRNA intron. Curr. Biol. 7, 850–9 (1997).

    28. Verrier, S. B. & Jean-Jean, O. Complementarity between the mRNA 5’ untranslated region and 18S ribosomal RNA can inhibit translation. RNA 6, 584–97 (2000).

    29. Sivan, G., Kedersha, N. & Elroy-Stein, O. Ribosomal slowdown mediates translational arrest during cellular division. Mol. Cell. Biol. 27, 6639–46 (2007).

    30. Doma, M. K. & Parker, R. Endonucleolytic cleavage of eukaryotic mRNAs with stalls in translation elongation. Nature 440, 561–564 (2006).

    31. Higashi, S. et al. TDP-43 associates with stalled ribosomes and contributes to cell survival during cellular stress. J. Neurochem. 126, 288–300 (2013).

    32. Ortiz, J. O. et al. Structure of hibernating ribosomes studied by cryoelectron tomography in vitro and in situ. J. Cell Biol. 190, 613–621 (2010).

    33. Graber, T. E. et al. Reactivation of stalled polyribosomes in synaptic plasticity. Proc. Natl. Acad. Sci. U. S. A. 110, 16205–10 (2013).

    34. Aeschimann, F., Xiong, J., Arnold, A., Dieterich, C. & Großhans, H. Transcriptome-wide measurement of ribosomal occupancy by ribosome profiling. Methods 85, 75–89 (2015).

    35. Yarmolinsky, M. B. & Haba, G. L. Inhibition By Puromycin of Amino Acid Incorporation Into Protein. Proc. Natl. Acad. Sci. U. S. A. 45, 1721–1729 (1959).

    36. Welch, M., Chastang, J. & Yarus, M. An Inhibitor of Ribosomal Peptidyl Transferase Using the Transition-State Analogy. Biochemistry 34, 385–390 (1995).

    37. Nissen, P., Hansen, J., Ban, N., Moore, P. B. & Steitz, T. A. The structural basis of ribosome activity in peptide bond synthesis. Science 289, 920–930 (2000).

    38. Wilson, D. N. Ribosome-targeting antibiotics and mechanisms of bacterial resistance. Nat. Rev. Microbiol. 12, 35–48 (2014).

    39. Gambetti, P., Hirt, L., Stieber, A. & Shafer, B. Distribution of puromycin

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

  • peptides in mouse entorhinal cortex. Exp. Neurol. 34, 223–228 (1972). 40. Aviner, R., Geiger, T. & Elroy-Stein, O. Genome-wide identification and

    quantification of protein synthesis in cultured cells and whole tissues by puromycin-associated nascent chain proteomics (PUNCH-P). Nat. Protoc. 9, 751–60 (2014).

    41. Schmidt, E. K., Clavarino, G., Ceppi, M. & Pierre, P. SUnSET, a nonradioactive method to monitor protein synthesis. Nat. Methods 6, 275–277 (2009).

    42. Pestka, S., Rosenfeld, H., Harris, R. & Hintikka, H. Studies on transfer ribonucleic acid-ribosome complexes. XXI. Effect of antibiotics on peptidyl-puromycin synthesis by mammalian polyribosomes. J. Biol. Chem. 247, 6895–6900 (1972).

    43. Odom, O. W., Picking, W. D. & Hardesty, B. Movement of tRNA but not the nascent peptide during peptide bond formation on ribosomes. Biochemistry 29, 10734–10744 (1990).

    44. Kukhanova, M. et al. The donor site of the peptidyltransferase center of ribosomes. FEBS Lett. 102, 198–203 (1979).

    45. Schmeing, T. M. et al. A pre-translocational intermediate in protein synthesis observed in crystals of enzymatically active 50S subunits. Nat Struct Biol 9, 225–230 (2002).

    46. Roberts, R. W. & Szostak, J. W. RNA-peptide fusions for the in vitro selection of peptides and proteins. Proc. Natl. Acad. Sci. U. S. A. 94, 12297–12302 (1997).

    47. Biyani, M., Husimi, Y. & Nemoto, N. Solid-phase translation and RNA-protein fusion: A novel approach for folding quality control and direct immobilization of proteins using anchored mRNA. Nucleic Acids Res. 34, (2006).

    48. Ge, J. et al. Puromycin Analogues Capable of Multiplexed Imaging and Profiling of Protein Synthesis and Dynamics in Live Cells and Neurons. Angew. Chemie Int. Ed. n/a-n/a (2016). doi:10.1002/anie.201511030

    49. Starck, S. R., Green, H. M., Alberola-Ila, J. & Roberts, R. W. A general approach to detect protein expression in vivo using fluorescent puromycin conjugates. Chem. Biol. 11, 999–1008 (2004).

    50. Higashi, K. et al. Enhancement of +1 frameshift by polyamines during translation of polypeptide release factor 2 in Escherichia coli. J. Biol. Chem. 281, 9527–9537 (2006).

    51. Arosio, D. et al. Spectroscopic and Structural Study of Proton and Halide Ion Cooperative Binding to GFP. Biophys. J. 93, 232–244 (2007).

    52. Liu, B. & Qian, S.-B. Characterizing inactive ribosomes in translational profiling. Translation 4, e1138018 (2016).

    53. Thomas, G. The effect of serum, EGF, PGF2α and insulin on S6 phosphorylation and the initiation of protein and DNA synthesis. Cell 30, 235–242 (1982).

    54. Andersen, G. R., Valente, L., Pedersen, L., Kinzy, T. G. & Nyborg, J. Crystal Structures of Nucleotide Exchange Intermediates in the eEF1A-eEF1Balpha Complex. Nat. Struct. Biol. 8, 531–534 (2001).

    55. Lamberti, a et al. The translation elongation factor 1A in tumorigenesis, signal transduction and apoptosis: review article. Amino Acids 26, 443–448 (2004).

    56. Lakkaraju, A. K. et al. Palmitoylated calnexin is a key component of the ribosome-translocon complex. EMBO J. 31, 1823–1835 (2012).

    57. Lareau, L. F., Hite, D. H., Hogan, G. J. & Brown, P. O. Distinct stages of the

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

  • translation elongation cycle revealed by sequencing ribosome-protected mRNA fragments. Elife 2014, 1–16 (2014).

    58. Schneider-Poetsch, T. et al. Inhibition of eukaryotic translation elongation by cycloheximide and lactimidomycin. Nat. Chem. Biol. 6, 209–217 (2010).

    59. Rivera, C. et al. Methylation of histone H3 lysine 9 occurs during translation. Nucleic Acids Res. 43, 9097–9106 (2015).

    60. Pisareva, V. P., Skabkin, M. A., Hellen, C. U., Pestova, T. V & Pisarev, A. V. Dissociation by Pelota, Hbs1 and ABCE1 of mammalian vacant 80S ribosomes and stalled elongation complexes. EMBO J. 3093, 1804–1817 (2011).

    61. Guydosh, N. R. & Green, R. Dom34 rescues ribosomes in 3′ untranslated regions. Cell 156, 950–962 (2014).

    62. Tebaldi, T. et al. Widespread uncoupling between transcriptome and translatome variations after a stimulus in mammalian cells. BMC Genomics 13, 220 (2012).

    63. Chambers, D. M., Peters, J. & Abbott, C. M. The lethal mutation of the mouse wasted (wst) is a deletion that abolishes expression of a tissue-specific isoform of translation elongation factor 1alpha, encoded by the Eef1a2 gene. Proc. Natl. Acad. Sci. U. S. A. 95, 4463–8 (1998).

    64. Lauria, F. et al. riboWaltz: optimization of ribosome P-site positioning in ribosome profiling data. BioRXiv (2017). doi:10.1101/169862

    65. Archer, S. K., Shirokikh, N. E., Beilharz, T. H. & Preiss, T. Dynamics of ribosome scanning and recycling revealed by translation complex profiling. Nature 535, 570–4 (2016).

    66. Lareau, L. F., Hite, D. H., Hogan, G. J. & Brown, P. O. Distinct stages of the translation elongation cycle revealed by sequencing ribosome-protected mRNA fragments. Elife 2014, 1–16 (2014).

    67. Han, Y. et al. Ribosome profiling reveals sequence-independent post-initiation pausing as a signature of translation. Cell Res. 24, 842–851 (2014).

    68. Wu, B., Eliscovich, C., Yoon, Y. J. & Singer, R. H. Translation dynamics of single mRNAs in live cells and neurons. Science (80-. ). 352, 1430–1435 (2016).

    69. Morisaki, T. et al. Real-time quantification of single RNA translation dynamics in living cells. Science (80-. ). 899, aaf0899 (2016).

    70. Ong, S.-E. & Mann, M. A practical recipe for stable isotope labeling by amino acids in cell culture (SILAC). Nat. Protoc. 1, 2650–60 (2006).

    71. Liu, T.-Y. et al. Time-Resolved Proteomics Extends Ribosome Profiling-Based Measurements of Protein Synthesis Dynamics. Cell Syst. 4, 636–644.e9 (2017).

    72. Simsek, D. et al. The Mammalian Ribo-interactome Reveals Ribosome Functional Diversity and Heterogeneity. Cell 169, 1051–1065.e18 (2017).

    73. Jan, C. H., Williams, C. C. & Weissman, J. S. Principles of ER cotranslational translocation revealed by proximity-specific ribosome profiling. Science (80-. ). 346, 1257521–1257521 (2014).

    74. Shi, Z. et al. Heterogeneous Ribosomes Preferentially Translate Distinct Subpools of mRNAs Genome-wide. Mol. Cell 67, 71–83.e7 (2017).

    75. Hansen, J. L., Moore, P. B. & Steitz, T. a. Structures of five antibiotics bound at the peptidyl transferase center of the large ribosomal subunit. J. Mol. Biol. 330, 1061–1075 (2003).

    76. Anger, A. M. et al. Structures of the human and Drosophila 80S ribosome.

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

  • Nature 497, 80–5 (2013). 77. Ban, N., Nissen, P., Hansen, J., Moore, P. B. & Steitz, T. a. The complete

    atomic structure of the large ribosomal subunit at 2. 4 A Resolut. Sci. 289, 905±920 (2000).

    78. Afshar-Kharghan, V., Li, C. Q., Khoshnevis-Asl, M. & Lopez, J. A. Kozak sequence polymorphism of the glycoprotein (GP) Ibalpha gene is a major determinant of the plasma membrane levels of the platelet GP Ib-IX-V complex. Blood 94, 186–191 (1999).

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

  • FIGURES

    Figure 1. A new analog of puromycin inhibits translation and can be used for

    functionalization of agarose and polystyrene beads. (a) Chemical formula of the

    new puromycin analog: 3P. (b) 3P binding to streptavidin coated agarose and

    polystyrene beads. Absorbance of the supernatant at 275 nm is measured after

    addition of streptavidin coated magnetic beads to 100 pmol of 3P. Data represent the

    mean of triplicate experiments (n = 3). The gray line identified the quantity of beads

    (polystyrene, µg or agarose, µL of 10% slurry suspension) used in all experiments for

    each sample. (c) Expression of the firefly luciferase in the presence of puromycin

    (left panel) and 3P (right panel). ε-Labeled biotinylated lysine-tRNAs is used to

    monitor the protein production by SDS–PAGE (top). The histograms represent the

    relative quantification of the representative bands reported in the immunoblot with

    respect to the control. The histogram on the right shows inhibition of luciferase

    expression in the presence of different concentrations of 3P (0 µM, 0.01 µM, 0.1 µM,

    1 µM and 10 µM (right panel). Error bars represent s.d. calculated from triplicate

    experiments (n =3); (**) = t-test p-val < 0.05.

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

  • Figure 2. 3P-beads can capture mRNAs in active translation in vitro. (a)

    Experimental design: 3P-nbeads are used to pull-down transcripts in a cell-free in

    vitro transcription-translation system. Briefly, from step 1 to 5: (1), Plasmids are

    added to the in vitro transcription-translation reaction; (2), Beads are functionalized

    with 3P; (3), The IVTT mix is added to 3P-beads and incubated for 1 hour in orbital

    rotation at 2 rpm at 4°C; (4), Beads are washed without detaching them from the

    magnet to remove unspecific binding; (4/a), The RNA is extracted, precipitated with

    isopropanol, and digested with DNase I (4/b) to avoid possible DNA contaminations.

    Finally, the cDNA is synthetized (4/c). (5), Samples are analyzed by RT-qPCR to

    detect the reporter gene. (b) Top panel, immunoblotting of total EGFP protein at

    different incubation time, without (-) or with harringtonine (+) at the reported

    concentration. Middle panel, immunoblotting showing the comparison between the

    total EGFP expression from the pPR-IBA2 plasmid and the EGFP expressed from

    the pBluescript II KS+ plasmid. Bottom panel, EGFP RNA enrichment on RiboLace

    (-, no harrigtonine) respect to the treated sample (harr, harringtonine treatment, 2

    µg/mL for 3 min). On the right, the the total RNA content (gray histograms) without (-)

    or with harringtonine (harr) for the two vectors as measured by RT-qPCR. (**) = t-test

    pval = 0.01 (25 min); pval = 0.03 (120 min); n = 3. (c) Top panel, Sketch of

    experimental protocol for in vitro transcription and translation of the firefly

    luciferase(luc). Harringtonine and puromycin (puro, 5 µg/mL) are added to the

    mixture. Then, (i) ribosomes in active translation are isolated with 3P-beads, by

    using mP- beads as control; (ii) RNA is extracted, treated with DNAse I and retro-

    transcribed to single stranded cDNA with random hexamers. (c) Bottom, Fold

    change values relative to the total amount of transcript captured by 3P-beads

    compared to the control beads (mP), henceforth referred as the ʼenrichmentʼ. (**) = t-

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

  • test pval = 0.03 (luv vs luc+harr), pval = 0.02 (luc vs luc+puro); n = 5. Error bars

    represent s.d.

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

  • Figure 3. RiboLace is able to capture ribosomes and associated mRNAs under

    active translation in cell cultures and tissues. (a) RiboLace protocol: magnetic

    beads coated with streptavidin are functionalized with the 3P molecule (step 1).

    RiboLace beads are then added to the cell lysate (step 2) (usually 5 - 20 µL,

    corresponding to ~ 1.2 - 5 x 105 cells) and washed (step 3). Finally, both proteins

    and RNA are recovered for further analysis (step 4). (b) Ponceau staining of a

    nitrocellulose membrane containing (from left to right) the total protein extract after

    applying the RiboLace protocol on HEK-293 and corresponding inputs in three

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

  • different conditions (nt, not treated; st, starvation (0.5% FBS); st + EGF, starvation +

    EGF stimulation). The lane intensity is reported as a percentage of the sum of the

    three lysates. (c) Immunoblotting of eEF1ɑ, calnexin, RPL26 and RPS6 isolated after

    applying the RiboLace protocol on HEK-293 (RiboLace, mP-beads, and in input). (d)

    Quantification of proteins isolated with RiboLace in different stress conditions applied

    to the cells. Immunoblots were scanned with a Chemidoc (Biorad) and quantified

    with ImageJ v1.45s, n = 3, * indicates pval < 0.05. (e) Comparison between the

    relative enrichment (no starvation vs starvation) of eEF2 and eEF1ɑ on RiboLace.

    (*) t-test p-val = 0.02, n = 4 (f) Immunoblotting of eIf4B, PABP, eEF1ɑ, RPL14,

    RPl26, me(K9)H3 and H3, detected on RiboLace in normal growing conditions (nt) or

    under serum starvation (st) with relative inputs in MCF7 cytoplasmic lysates. (g)

    Top, Immunoblotting of Pelota and eEF1α detected on RiboLace in HEK-293 not

    treated, under starvation or after EGF stimulation, with relative inputs. Bottom,

    Immunoblotting of Pelota and eEF1α detected on RiboLace in MCF7 treated (Ar) or

    not treated (nt) with Arsenite. (#) indicates the number of each biological replicate

    (h) Experimental design for identifying the global RNA repertoire of RNAs

    associated to RiboLace by RNA-seq. MCF7 cells treated with EGF or serum-starved

    in comparison to classical polysomal profiling (POL-Seq) and total RNA

    transcriptomics analysis. After proteinase K digestion, RNA is extracted from

    RiboLace and mP beads, and from both polyribosomal and total RNA from the same

    profile. RiboLace, RL; Polyribosomal, P; Total, T.. Libraries are prepared using the

    Illumina TruSeq library preparation kit and the sequencing performed with Illumina

    HiSeq 2000. (i) Top, histograms representing RNA-seq (white bars) and RT-qPCR

    (black bars) fold changes (FC) of four genes (NCS1, IL27RA, DUSP5, VEGFA) t-test

    p-val < 0.05 (*). Bottom, Comparison between protein fold change and RNA fold

    changes obtained with different methods (RL, P, T). The semi-quantitative analysis

    of the protein band intensity (n = 3) for PALLD and PLK3 is reported in the

    histogram. Black bars, RT-qPCR fold change; white bars, RNA-seq fold change; light

    gray bars, protein fold change. Housekeeping: GAPDH, beta actin; ribosomal protein

    L26. t-test (*) = p-val < 0.05.

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

  • Figure 4. RiboLace Ribo-Seq. (a) Schematic RiboLace protocol for the separation

    of single active ribosomes. Cell lysates (1) is treated with RNase I for 45 min (2) and

    the quenched with an RNase inhibitor (3). Then, RiboLace beads are incubated with

    the digested cell lysate (1h, 4°C) (4), washed (5) and then proteins or RNA extracted

    (6). See methods for details. (b) Immunoblotting of eEF1α, calnexin, RPL26 and

    RPS6 after applying the RiboLace protocol on HEK-293 and HeLa. Nt, not treated;

    st, starvation (0.5% FBS), st + EGF, starvation + EGF stimulation; harr,

    harringtonine; h.s., heat shock (c) Top, Schematic RiboLace protocol for the

    extraction of single active ribosomes from mouse brain. Bottom, immunoblotting on

    eEF1α, and RPL26 for RiboLace, mP-beads and relative inputs. (d) Scheme of the

    protocol for comparative active Ribo-Seq and poly Ribo-seq. (e) Left, percentage of

    P-sites mapping to the 5’ UTR, CDS and 3’ UTR of mRNAs from RiboLace and Poly

    Ribo-Seq data. Right, percentage of region lengths in mRNAs sequences. Both

    techniques show a clear enrichment in signal mapping to the CDS, consistent with

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

  • ribosome protected fragments. (f) Percentage of P-sites corresponding to the three

    possible reading frames along the 5’ UTR, CDS and 3’ UTR, stratified for read

    length, comparing RiboLace (top panel) and Poly Ribo-Seq (bottom panel). For each

    length and each region, the sum of the signal is normalized to 100%. The enrichment

    in frame 0 is CDS specific in both cases. (g) Meta-gene profiles showing the density

    P-sites around the translation initiation site (TIS) and translation termination site

    (TTS) for RiboLace (top panel) and Poly Ribo-Seq (bottom panel). The peak

    corresponding to the fifth codon is highlighted with an asterisk.

    not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted August 22, 2017. ; https://doi.org/10.1101/179671doi: bioRxiv preprint

    https://doi.org/10.1101/179671

Recommended