Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Published: 26 June 2024

TnpB homologues exapted from transposons are RNA-guided transcription factors

  • Tanner Wiegand 1 ,
  • Florian T. Hoffmann   ORCID: orcid.org/0000-0003-0301-4259 1 ,
  • Matt W. G. Walker   ORCID: orcid.org/0000-0003-3345-8248 2 ,
  • Stephen Tang   ORCID: orcid.org/0000-0001-5492-9796 1 ,
  • Egill Richard 1 ,
  • Hoang C. Le 1 ,
  • Chance Meers   ORCID: orcid.org/0000-0003-2554-3987 1 &
  • Samuel H. Sternberg   ORCID: orcid.org/0000-0001-8240-9114 1  

Nature ( 2024 ) Cite this article

2932 Accesses

68 Altmetric

Metrics details

  • Bacteriophages
  • Gene regulation
  • Molecular evolution

Transposon-encoded tnpB and iscB genes encode RNA-guided DNA nucleases that promote their own selfish spread through targeted DNA cleavage and homologous recombination 1 , 2 , 3 , 4 . These widespread gene families were repeatedly domesticated over evolutionary timescales, leading to the emergence of diverse CRISPR-associated nucleases including Cas9 and Cas12 (refs. 5 , 6 ). We set out to test the hypothesis that TnpB nucleases may have also been repurposed for novel, unexpected functions other than CRISPR–Cas adaptive immunity. Here, using phylogenetics, structural predictions, comparative genomics and functional assays, we uncover multiple independent genesis events of programmable transcription factors, which we name TnpB-like nuclease-dead repressors (TldRs). These proteins use naturally occurring guide RNAs to specifically target conserved promoter regions of the genome, leading to potent gene repression in a mechanism akin to CRISPR interference technologies invented by humans 7 . Focusing on a TldR clade found broadly in Enterobacteriaceae , we discover that bacteriophages exploit the combined action of TldR and an adjacently encoded phage gene to alter the expression and composition of the host flagellar assembly, a transformation with the potential to impact motility 8 , phage susceptibility 9 , and host immunity 10 . Collectively, this work showcases the diverse molecular innovations that were enabled through repeated exaptation of transposon-encoded genes, and reveals the evolutionary trajectory of diverse RNA-guided transcription factors.

This is a preview of subscription content, access via your institution

Access options

Access Nature and 54 other Nature Portfolio journals

Get Nature+, our best-value online-access subscription

24,99 € / 30 days

cancel any time

Subscribe to this journal

Receive 51 print issues and online access

185,98 € per year

only 3,65 € per issue

Buy this article

  • Purchase on Springer Link
  • Instant access to full article PDF

Prices may be subject to local taxes which are calculated during checkout

transcription factor thesis

Similar content being viewed by others

transcription factor thesis

Bridge RNAs direct programmable recombination of target and donor DNA

transcription factor thesis

Structural mechanism of bridge RNA-guided recombination

transcription factor thesis

Identification of RNA structures and their roles in RNA functions

Data availability.

Next-generation sequencing data generated in this study were deposited in the NCBI SRA (BioProject accession PRJNA1029663 ) and GEO ( GSE245749 ). The published genome used for ChIP–seq analyses was obtained from NCBI (GenBank NC_000913.3 ). Publicly available RNA-seq data analysed for TldR–gRNA expression are in the NCBI SRA ( ERR6044061 ) and GEO ( GSE115009 ) databases. The published genomes used for bioinformatics analyses were obtained from NCBI (Supplementary Table 4 ). The ISfinder database can be accessed at https://www-is.biotoul.fr/index.php .

Code availability

Custom scripts used for bioinformatics, TAM library analyses, and ChIP–seq data analyses are available on request. The R script describing initial steps to discover TldRs is available at https://github.com/sternberglab/Wiegand_etal_2024 .

Altae-Tran, H. et al. The widespread IS200/IS605 transposon family encodes diverse programmable RNA-guided endonucleases. Science 374 , 57–65 (2021).

Article   ADS   CAS   PubMed   PubMed Central   Google Scholar  

Karvelis, T. et al. Transposon-associated TnpB is a programmable RNA-guided DNA endonuclease. Nature 599 , 692–696 (2021).

Meers, C. et al. Transposon-encoded nucleases use guide RNAs to promote their selfish spread. Nature 622 , 863–871 (2023).

Article   ADS   CAS   PubMed   Google Scholar  

Zedaveinyte, R. et al. Antagonistic conflict between transposon-encoded introns and guide RNAs. Preprint at bioRxiv https://doi.org/10.1101/2023.11.20.567912 (2023).

Kapitonov, V. V., Makarova, K. S. & Koonin, E. V. ISC, a novel group of bacterial and archaeal DNA transposons that encode Cas9 homologs. J. Bacteriol. 198 , 797–807 (2015).

Article   PubMed   Google Scholar  

Chylinski, K., Makarova, K. S., Charpentier, E. & Koonin, E. V. Classification and evolution of type II CRISPR–Cas systems. Nucleic Acids Res. 42 , 6091–6105 (2014).

Article   CAS   PubMed   PubMed Central   Google Scholar  

Larson, M. H. et al. CRISPR interference (CRISPRi) for sequence-specific control of gene expression. Nat. Protoc. 8 , 2180–2196 (2013).

Nakamura, S. & Minamino, T. Flagella-driven motility of bacteria. Biomolecules 9 , 279 (2019).

Samuel, A. D. et al. Flagellar determinants of bacterial sensitivity to χ-phage. Proc. Natl Acad. Sci. USA 96 , 9863–9866 (1999).

Wilson, D. R. & Beveridge, T. J. Bacterial flagellar filaments and their component flagellins. Can. J. Microbiol. 39 , 451–472 (1993).

Article   CAS   PubMed   Google Scholar  

Aziz, R. K., Breitbart, M. & Edwards, R. A. Transposases are the most abundant, most ubiquitous genes in nature. Nucleic Acids Res. 38 , 4207–4217 (2010).

Cosby, R. L., Chang, N. C. & Feschotte, C. Host–transposon interactions: conflict, cooperation, and cooption. Genes Dev. 33 , 1098–1116 (2019).

Jangam, D., Feschotte, C. & Betran, E. Transposable element domestication as an adaptation to evolutionary conflicts. Trends Genet. 33 , 817–831 (2017).

Gould, S. J. & Vrba, E. S. Exaptation — a missing term in the science of form. Paleobiology 8 , 4–15 (1982).

Article   Google Scholar  

Koonin, E. V. & Makarova, K. S. Mobile genetic elements and evolution of CRISPR–Cas systems: all the way there and back. Genome Biol. Evol. 9 , 2812–2825 (2017).

Koonin, E. V. & Makarova, K. S. Origins and evolution of CRISPR–Cas systems. Phil. Trans. R. Soc. B 374 , 20180087 (2019).

Nunez, J. K., Lee, A. S., Engelman, A. & Doudna, J. A. Integrase-mediated spacer acquisition during CRISPR–Cas adaptive immunity. Nature 519 , 193–198 (2015).

McGinn, J. & Marraffini, L. A. Molecular mechanisms of CRISPR–Cas spacer acquisition. Nat. Rev. Microbiol. 17 , 7–12 (2019).

Krupovic, M., Makarova, K. S., Forterre, P., Prangishvili, D. & Koonin, E. V. Casposons: a new superfamily of self-synthesizing DNA transposons at the origin of prokaryotic CRISPR–Cas immunity. BMC Biol. 12 , 36 (2014).

Article   PubMed   PubMed Central   Google Scholar  

Hickman, A. B., Kailasan, S., Genzor, P., Haase, A. D. & Dyda, F. Casposase structure and the mechanistic link between DNA transposition and spacer acquisition by CRISPR–Cas. eLife 9 , e50004 (2020).

Klompe, S. E., Vo, P. L. H., Halpin-Healy, T. S. & Sternberg, S. H. Transposon-encoded CRISPR–Cas systems direct RNA-guided DNA integration. Nature 571 , 219–225 (2019).

Strecker, J. et al. RNA-guided DNA insertion with CRISPR-associated transposases. Science 365 , 48–53 (2019).

Faure, G. et al. CRISPR–Cas in mobile genetic elements: counter-defence and beyond. Nat. Rev. Microbiol. 17 , 513–525 (2019).

Seed, K. D., Lazinski, D. W., Calderwood, S. B. & Camilli, A. A bacteriophage encodes its own CRISPR/Cas adaptive response to evade host innate immunity. Nature 494 , 489–491 (2013).

Al-Shayeb, B. et al. Clades of huge phages from across Earth’s ecosystems. Nature 578 , 425–431 (2020).

Frost, L. S., Leplae, R., Summers, A. O. & Toussaint, A. Mobile genetic elements: the agents of open source evolution. Nat. Rev. Microbiol. 3 , 722–732 (2005).

Bao, W. & Jurka, J. Homologues of bacterial TnpB_IS605 are widespread in diverse eukaryotic transposable elements. Mob. DNA 4 , 12 (2013).

Jiang, K. et al. Programmable RNA-guided DNA endonucleases are widespread in eukaryotes and their viruses. Sci. Adv. 9 , eadk0171 (2023).

Saito, M. et al. Fanzor is a eukaryotic programmable RNA-guided endonuclease. Nature 620 , 660–668 (2023).

Siguier, P., Gourbeyre, E., Varani, A., Ton-Hoang, B. & Chandler, M. Everyman’s guide to bacterial insertion sequences. Microbiol. Spectr. 3 , MDNA3-0030-2014 (2015).

Altae-Tran, H. et al. Diversity, evolution, and classification of the RNA-guided nucleases TnpB and Cas12. Proc. Natl Acad. Sci. USA 120 , e2308224120 (2023).

Makarova, K. S. et al. Evolutionary classification of CRISPR–Cas systems: a burst of class 2 and derived variants. Nat. Rev. Microbiol. 18 , 67–83 (2020).

Huang, C. J., Adler, B. A. & Doudna, J. A. A naturally DNase-free CRISPR–Cas12c enzyme silences gene expression. Mol. Cell 82 , 2148–2160.e4 (2022).

Wu, W. Y. et al. The miniature CRISPR–Cas12m effector binds DNA to block transcription. Mol. Cell 82 , 4487–4502.e7 (2022).

Peters, J. E., Makarova, K. S., Shmakov, S. & Koonin, E. V. Recruitment of CRISPR–Cas systems by Tn7-like transposons. Proc. Natl Acad. Sci. USA 114 , E7358–E7366 (2017).

Nakamura, M., Gao, Y., Dominguez, A. A. & Qi, L. S. CRISPR technologies for precise epigenome editing. Nat. Cell Biol. 23 , 11–22 (2021).

Zetsche, B. et al. Cpf1 is a single RNA-guided endonuclease of a class 2 CRISPR–Cas system. Cell 163 , 759–771 (2015).

He, S. et al. The IS200/IS605 family and “peel and paste” single-strand transposition mechanism. Microbiol. Spectr. https://doi.org/10.1128/microbiolspec.MDNA3-0039-2014 (2015).

Doeven, M. K., van den Bogaart, G., Krasnikov, V. & Poolman, B. Probing receptor–translocator interactions in the oligopeptide ABC transporter by fluorescence correlation spectroscopy. Biophys. J. 94 , 3956–3965 (2008).

Biemans-Oldehinkel, E., Doeven, M. K. & Poolman, B. ABC transporter architecture and regulatory roles of accessory domains. FEBS Lett. 580 , 1023–1035 (2006).

Mukherjee, S. et al. CsrA–FliW interaction governs flagellin homeostasis and a checkpoint on flagellar morphogenesis in Bacillus subtilis . Mol. Microbiol. 82 , 447–461 (2011).

Lawrence, J. G. Shared strategies in gene organization among prokaryotes and eukaryotes. Cell 110 , 407–413 (2002).

Siguier, P., Perochon, J., Lestrade, L., Mahillon, J. & Chandler, M. ISfinder: the reference centre for bacterial insertion sequences. Nucleic Acids Res. 34 , D32–D36 (2006).

Leenay, R. T. & Beisel, C. L. Deciphering, communicating, and engineering the CRISPR PAM. J. Mol. Biol. 429 , 177–191 (2017).

Michaux, C. et al. Single-nucleotide RNA maps for the two major nosocomial pathogens Enterococcus faecalis and Enterococcus faecium . Front. Cell. Infect. Microbiol. 10 , 600325 (2020).

Nety, S. P. et al. The transposon-encoded protein TnpB processes its own mRNA into ωRNA for guided nuclease activity. CRISPR J. 6 , 232–242 (2023).

Ohnishi, K., Kutsukake, K., Suzuki, H. & Iino, T. Gene fliA encodes an alternative sigma factor specific for flagellar operons in Salmonella typhimurium . Mol. Gen. Genet. 221 , 139–147 (1990).

Ide, N., Ikebe, T. & Kutsukake, K. Reevaluation of the promoter structure of the class 3 flagellar operons of Escherichia coli and Salmonella . Genes Genet. Syst. 74 , 113–116 (1999).

Klepsch, M. M. et al. Escherichia coli peptide binding protein OppA has a preference for positively charged peptides. J. Mol. Biol. 414 , 75–85 (2011).

Solovyev, V. A. S. in Metagenomics and its Applications in Agriculture , Biomedicine and Environmental Studies (ed Li, R. W.) 61–78 (Nova Science Publishers, 2011).

Swarts, D. C., van der Oost, J. & Jinek, M. Structural basis for guide RNA processing and seed-dependent DNA targeting by CRISPR–Cas12a. Mol. Cell 66 , 221–233.e4 (2017).

Qi, L. S. et al. Repurposing CRISPR as an RNA-guided platform for sequence-specific control of gene expression. Cell 152 , 1173–1183 (2013).

Zhang, X. et al. Multiplex gene regulation by CRISPR–ddCpf1. Cell Discov. 3 , 17018 (2017).

Kim, S. K. et al. Efficient transcriptional gene repression by type V-A CRISPR–Cpf1 from Eubacterium eligens . ACS Synth. Biol. 6 , 1273–1282 (2017).

Samatey, F. A. et al. Structure of the bacterial flagellar protofilament and implications for a switch for supercoiling. Nature 410 , 331–337 (2001).

Yonekura, K., Maki-Yonekura, S. & Namba, K. Complete atomic model of the bacterial flagellar filament by electron cryomicroscopy. Nature 424 , 643–650 (2003).

Reid, S. D., Selander, R. K. & Whittam, T. S. Sequence diversity of flagellin (fliC) alleles in pathogenic Escherichia coli . J. Bacteriol. 181 , 153–160 (1999).

Esteves, N. C., Bigham, D. N. & Scharf, B. E. Phages on filaments: a genetic screen elucidates the complex interactions between Salmonella enterica flagellin and bacteriophage Chi. PLoS Pathog. 19 , e1011537 (2023).

Guttenplan, S. B. & Kearns, D. B. Regulation of flagellar motility during biofilm formation. FEMS Microbiol. Rev. 37 , 849–871 (2013).

Dacquay, L. C. et al. E. coli nissle increases transcription of flagella assembly and formate hydrogenlyase genes in response to colitis. Gut Microbes 13 , 1994832 (2021).

Kim, M. J., Lim, S. & Ryu, S. Molecular analysis of the Salmonella typhimurium tdc operon regulation. J. Microbiol. Biotechnol. 18 , 1024–1032 (2008).

CAS   PubMed   Google Scholar  

Esteves, N. C. & Scharf, B. E. Flagellotropic bacteriophages: opportunities and challenges for antimicrobial applications. Int. J. Mol. Sci. 23 , 7084 (2022).

Yoon, S. I. et al. Structural basis of TLR5-flagellin recognition and signaling. Science 335 , 859–864 (2012).

Tenthorey, J. L. et al. The structural basis of flagellin detection by NAIP5: a strategy to limit pathogen immune evasion. Science 358 , 888–893 (2017).

Wang, L., Rothemund, D., Curd, H. & Reeves, P. R. Species-wide variation in the Escherichia coli flagellin (H-antigen) gene. J. Bacteriol. 185 , 2936–2943 (2003).

Cullender, T. C. et al. Innate and adaptive immunity interact to quench microbiome flagellar motility in the gut. Cell Host Microbe 14 , 571–581 (2013).

Brussow, H., Canchaya, C. & Hardt, W. D. Phages and the evolution of bacterial pathogens: from genomic rearrangements to lysogenic conversion. Microbiol. Mol. Biol. Rev. 68 , 560–602 (2004).

Rees, D. C., Johnson, E. & Lewinson, O. ABC transporters: the power to change. Nat. Rev. Mol. Cell Biol. 10 , 218–227 (2009).

Holtzman, L. & Gersbach, C. A. Editing the epigenome: reshaping the genomic landscape. Annu. Rev. Genomics Hum. Genet. 19 , 43–71 (2018).

Bikard, D. et al. Programmable repression and activation of bacterial gene expression using an engineered CRISPR–Cas system. Nucleic Acids Res. 41 , 7429–7437 (2013).

Cui, L. et al. A CRISPRi screen in E. coli reveals sequence-specific toxicity of dCas9. Nat. Commun. 9 , 1912 (2018).

Article   ADS   PubMed   PubMed Central   Google Scholar  

Vigouroux, A., Oldewurtel, E., Cui, L., Bikard, D. & van Teeffelen, S. Tuning dCas9’s ability to block transcription enables robust, noiseless knockdown of bacterial genes. Mol. Syst. Biol. 14 , e7899 (2018).

Workman, R. E. et al. A natural single-guide RNA repurposes Cas9 to autoregulate CRISPR–Cas expression. Cell 184 , 675–688.e19 (2021).

Ratner, H. K. et al. Catalytically active Cas9 mediates transcriptional interference to facilitate bacterial virulence. Mol. Cell 75 , 498–510.e5 (2019).

Eddy, S. R. Accelerated profile HMM searches. PLoS Comput. Biol. 7 , e1002195 (2011).

Article   ADS   MathSciNet   CAS   PubMed   PubMed Central   Google Scholar  

Huang, Y., Niu, B., Gao, Y., Fu, L. & Li, W. CD-HIT Suite: a web server for clustering and comparing biological sequences. Bioinformatics 26 , 680–682 (2010).

Katoh, K., Kuma, K., Toh, H. & Miyata, T. MAFFT version 5: improvement in accuracy of multiple sequence alignment. Nucleic Acids Res. 33 , 511–518 (2005).

Capella-Gutierrez, S., Silla-Martinez, J. M. & Gabaldon, T. trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25 , 1972–1973 (2009).

Minh, B. Q. et al. IQ-TREE 2: new models and efficient methods for phylogenetic inference in the genomic era. Mol. Biol. Evol. 37 , 1530–1534 (2020).

Letunic, I. & Bork, P. Interactive Tree of Life (iTOL) v5: an online tool for phylogenetic tree display and annotation. Nucleic Acids Res. 49 , W293–W296 (2021).

Wright, E. S. DECIPHER: harnessing local sequence context to improve protein multiple sequence alignment. BMC Bioinformatics 16 , 322 (2015).

Biostrings: String objects representing biological sequences, and matching algorithms. R package version 2.70.1 (Pagès, H. A. P., Gentleman, R. & DebRoy, S., 2023).

Camacho, C. et al. BLAST+: architecture and applications. BMC Bioinformatics 10 , 421 (2009).

Cantalapiedra, C. P., Hernandez-Plaza, A., Letunic, I., Bork, P. & Huerta-Cepas, J. eggNOG-mapper v2: functional annotation, orthology assignments, and domain prediction at the metagenomic scale. Mol. Biol. Evol. 38 , 5825–5829 (2021).

Hyatt, D. et al. Prodigal: prokaryotic gene recognition and translation initiation site identification. BMC Bioinformatics 11 , 119 (2010).

Price, M. N., Dehal, P. S. & Arkin, A. P. FastTree 2 — approximately maximum-likelihood trees for large alignments. PLoS ONE 5 , e9490 (2010).

Jumper, J. et al. Highly accurate protein structure prediction with AlphaFold. Nature 596 , 583–589 (2021).

Goddard, T. D. et al. UCSF ChimeraX: meeting modern challenges in visualization and analysis. Protein Sci. 27 , 14–25 (2018).

Waterhouse, A. M., Procter, J. B., Martin, D. M., Clamp, M. & Barton, G. J. Jalview version 2 — a multiple sequence alignment editor and analysis workbench. Bioinformatics 25 , 1189–1191 (2009).

Mirdita, M. et al. ColabFold: making protein folding accessible to all. Nat. Methods 19 , 679–682 (2022).

Pei, J., Kim, B. H. & Grishin, N. V. PROMALS3D: a tool for multiple protein sequence and structure alignments. Nucleic Acids Res. 36 , 2295–2300 (2008).

Nawrocki, E. P. & Eddy, S. R. Infernal 1.1: 100-fold faster RNA homology searches. Bioinformatics 29 , 2933–2935 (2013).

Will, S., Joshi, T., Hofacker, I. L., Stadler, P. F. & Backofen, R. LocARNA-P: accurate boundary prediction and improved detection of structural RNAs. RNA 18 , 900–914 (2012).

Vasimuddin M., Misra, S., Li, H. & Aluru, S. Efficient architecture-aware acceleration of BWA-MEM for multicore systems. In 2019 IEEE International Parallel and Distributed Processing Symposium (IPDPS) https://doi.org/10.1109/IPDPS.2019.00041 (IEEE, 2019).

Danecek, P. et al. Twelve years of SAMtools and BCFtools. Gigascience 10 , giab008 (2021).

Ramirez, F. et al. deepTools2: a next generation web server for deep-sequencing data analysis. Nucleic Acids Res. 44 , W160–W165 (2016).

Robinson, J. T. et al. Integrative genomics viewer. Nat. Biotechnol. 29 , 24–26 (2011).

Hoffmann, F. T. et al. Selective TnsC recruitment enhances the fidelity of RNA-guided transposition. Nature 609 , 384–393 (2022).

Zhang, Y. et al. Model-based analysis of ChIP–seq (MACS). Genome Biol. 9 , R137 (2008).

Quinlan, A. R. & Hall, I. M. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26 , 841–842 (2010).

Bailey, T. L. et al. MEME SUITE: tools for motif discovery and searching. Nucleic Acids Res. 37 , W202–W208 (2009).

Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet.journal https://doi.org/10.14806/ej.17.1.200 (2011).

Liao, Y., Smyth, G. K. & Shi, W. featureCounts: an efficient general purpose program for assigning sequence reads to genomic features. Bioinformatics 30 , 923–930 (2014).

Love, M. I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15 , 550 (2014).

Sharan, S. K., Thomason, L. C., Kuznetsov, S. G. & Court, D. L. Recombineering: a homologous recombination-based method of genetic engineering. Nat. Protoc. 4 , 206–223 (2009).

Luo, G. et al. flrA, flrB and flrC regulate adhesion by controlling the expression of critical virulence genes in Vibrio alginolyticus . Emerg. Microbes Infect. 5 , e85 (2016).

Kreutzberger, M. A. B. et al. Flagellin outer domain dimerization modulates motility in pathogenic and soil bacteria from viscous environments. Nat. Commun. 13 , 1422 (2022).

Shevchenko, A., Tomas, H., Havlis, J., Olsen, J. V. & Mann, M. In-gel digestion for mass spectrometric characterization of proteins and proteomes. Nat. Protoc. 1 , 2856–2860 (2006).

Kulak, N. A., Pichler, G., Paron, I., Nagaraj, N. & Mann, M. Minimal, encapsulated proteomic-sample processing applied to copy-number estimation in eukaryotic cells. Nat. Methods 11 , 319–324 (2014).

Meier, F. et al. Online parallel accumulation-serial fragmentation (PASEF) with a novel trapped ion mobility mass spectrometer. Mol. Cell. Proteomics 17 , 2534–2545 (2018).

Cox, J. et al. Andromeda: a peptide search engine integrated into the MaxQuant environment. J. Proteome Res. 10 , 1794–1805 (2011).

Download references

Acknowledgements

We thank S. R. Pesari and Z. Akhtar for laboratory support; G. D. Lampe for suggesting the TldR moniker; A. Bernheim for helpful discussions; F. Tesson, A. Bernheim, A. M. Earl and D. Gray for sharing E. coli and Enterobacter strains; C. Lu for Covaris sonicator access; R. K. Soni for mass spectrometry support; L. F. Landweber for qPCR instrument access; and the JP Sulzberger Columbia Genome Center for next-generation sequencing support. S.T. was supported by a Medical Scientist Training Program grant (5T32GM145440-02) from the NIH. M.W.G.W. was supported by a National Science Foundation Graduate Research Fellowship. C.M. was supported by the NIH Postdoctoral Fellowship F32 GM143924-01A1. S.H.S. was supported by the NSF Faculty Early Career Development Program (CAREER) Award 2239685, a Pew Biomedical Scholarship, an Irma T. Hirschl Career Scientist Award, and a startup package from the Columbia University Irving Medical Center Dean’s Office and the Vagelos Precision Medicine Fund.

Author information

Authors and affiliations.

Department of Biochemistry and Molecular Biophysics, Columbia University, New York, NY, USA

Tanner Wiegand, Florian T. Hoffmann, Stephen Tang, Egill Richard, Hoang C. Le, Chance Meers & Samuel H. Sternberg

Department of Biological Sciences, Columbia University, New York, NY, USA

Matt W. G. Walker

You can also search for this author in PubMed   Google Scholar

Contributions

T.W., C.M., and S.H.S. conceived and designed the project. T.W. performed all of the bioinformatics experiments and aided in the design of the experimental assays. F.T.H. performed plasmid interference, ChIP–seq, and the RFP repression assays. M.W.G.W. designed and generated the E. coli strains and plasmids for the RFP repression assays, fragments for Enterobacter recombineering, conducted the motility assays and isolated flagella for liquid chromatography with tandem mass spectrometry. S.T. performed and analysed the RNA-seq and RIP-seq experiments. E.R. cultured Enterobacter strains, extracted RNA for RNA-seq, and performed the RT–qPCR and recombineering experiments. C.M. performed the preliminary TnpB bioinformatics and neighbourhood analyses, together with H.C.L., and helped design the ChIP–seq and RFP repression assays. T.W. and S.H.S. discussed the data and wrote the manuscript, with input from all authors.

Corresponding author

Correspondence to Samuel H. Sternberg .

Ethics declarations

Competing interests.

Columbia University has filed a patent application related to this work. M.W.G.W. is a co-founder of Can9 Bioengineering. S.H.S. is a co-founder and scientific advisor to Dahlia Biosciences, a scientific advisor to CrisprBits and Prime Medicine, and an equity holder in Dahlia Biosciences and CrisprBits. All other authors declare no competing interests.

Peer review

Peer review information.

Nature thanks Wen Wu and the other, anonymous reviewer(s) for their contribution to the peer review of this work.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data figures and tables

Extended data fig. 1 phylogeny and ruvc nuclease domain analysis of oppf -associated tldrs..

a , Phylogenetic tree of oppF- associated TldR proteins from Fig. 2a , together with closely related TnpB proteins that contain intact RuvC active sites. The rings indicate RuvC DED active site intactness (inner) and TldR/TnpB domain composition (outer). Homologs marked with an orange square (TnpB) or green circle (TldR) were tested in heterologous experiments. b , Multiple sequence alignment of representative TnpB and TldR sequences from a , highlighting deterioration of RuvC active site motifs (shaded in red) and loss of the C-terminal zinc-finger (ZnF)/RuvC domain. Highly conserved residues are shaded in grey. c , Empirical ( Dra TnpB) and predicted AlphaFold structures of TnpB and TldR homologs marked with an asterisk in b , showing progressive loss of the active site catalytic triad.

Extended Data Fig. 2 Diverse prophages encode fliC P -associated tldR genes .

a , Genomic architecture of representative prophage elements whose boundaries could be identified by comparison to closely related isogenic strains. In each example, the prophage-containing strain is shown above the prophage-lacking strain, with species/strain names and NCBI genomic accession IDs indicated. Sequences flanking the left (5′) and right (3′) ends are highlighted in purple and yellow, respectively, together with their percentage sequence identities calculated using BLASTn. b , Alignment of distinct prophage elements, constructed using Mauve. Empty boxes represent open reading frames, and windows show sequence conservation for regions compared between prophage genomes with lines. Putative gene functions are shown below sequence conservation windows for the fliC P - tldR -encoding prophage from Enterobacter AR_163 (bottom). c , DNA sequence identities between the prophages in a , calculated with BLASTn. Identities were calculated as total matching nucleotides across the two genomes being compared, divided by the length of the query prophage genome.

Extended Data Fig. 3 RIP-seq reveals that some oppF -associated TldR proteins use short, 9–11-nt guides.

a , RNA immunoprecipitation sequencing (RIP-seq) data for an oppF- associated TldR homolog from Enterococcus faecalis ( Efa1 TldR) reveals the boundaries of a mature gRNA containing a 9-nt guide sequence. Reads were mapped to the TldR-gRNA expression plasmid; an input control is shown. b , Published RNA-seq data for Enterococcus faecalis V583 reveals similar gRNA boundaries, including an approximately 11-nt guide. c , RIP-seq data as in a for a second biological replicate of Efa1 TldR, further corroborating the observed 9–11-nt guide length.

Extended Data Fig. 4 oppF -associated TldRs target conserved genomic sequences that overlap with promoter elements driving oppA expression .

a , Schematic of original (left) and improved (right) search strategy to identify putative targets of gRNAs used by oppF -associated TldRs. Key insights resulted from the use of TAM and a shorter, 9-nt guide. b , Analysis of the guide sequence from the Efa1 TldR-associated gRNA in Extended Data Fig. 3 revealed a putative genomic target near the predicted promoter of oppA encoded within the same ABC transporter operon immediately adjacent to the tldR gene. The magnified schematics at the bottom show the predicted TAM and gRNA-target DNA base-pairing interactions for two representatives ( Efa1 TldR and Ece TldR), in which the gRNAs target opposite strands. Promoter elements predicted with BPROM are shown as brown squares. c , WebLogos of predicted guides and genomic targets associated with diverse oppF -associated TldRs highlighted in Extended Data Fig. 1 . d , Schematic of the oppF - tldR genomic locus (left) alongside the predicted function of OppA as a solute binding protein that facilitates transport of polypeptide substrates from the periplasm to the cytoplasm, in complex with the remainder of the ABC transporter apparatus. e , Published RNA-seq data for Enterococcus faecium AUS0004 45 , highlighting the oppA transcription start site (TSS). The predicted gRNA guide sequence (grey) is shown beneath the putative TAM (yellow) and target (purple) sequences, with guide-target complementarity represented by grey circles.

Extended Data Fig. 5 oppF -associated TldR homologs may target additional sites across the genome.

Schematic of Enterococcus cecorum genome and inset showing the oppF - tldR locus (top), with additional putative targets of the gRNA, other than the oppA promoter, numbered and highlighted in yellow along the genomic coordinate. A magnified view for each numbered target is shown below, with TAMs in yellow, prospective targets in purple, and TldR gRNA guide sequences in grey. Grey circles (right) represent positions of expected guide-target complementarity.

Extended Data Fig. 6 Genome-wide binding data from ChIP-seq experiments suggest a high mismatch tolerance for some TldR homologs.

a , Genome-wide ChIP–seq profiles for the indicated fliC P -associated TldR homologs, normalized to the highest peak within each dataset. The magnified insets at the bottom show the off-target sequences (grey) compared to the intended (engineered) on-target sequence (purple), with TAMs in yellow. Off-target #3 has no clear TAM-flanked off-target sequence but is intriguingly located at a tRNA locus, and binding was observed for diverse fliC P - and oppF- associated TldRs that recognized distinct TAMs. The phylogenetic tree at right indicates the relatedness of the tested and labeled homologs. b , Results for the indicated oppF -associated TldR homologs, shown as in a .

Extended Data Fig. 7 Plasmid interference assays confirm that TldR homologs lack detectable nuclease activity.

a , Schematic of E. coli -based plasmid interference assay using pEffector and pTarget. b , Representative dilution spot assays for Gst TnpB3 and synthetically inactivated RuvC mutant (D196A), showing the entire plate (left) and the magnified area of plating. Transformants were serially diluted, plated on selective media, and cultured at 37 °C for 16 h. Colony visibility was enhanced by inverting the colors and increasing contrast/brightness. c , Dilution spot assays for the indicated fliC -associated TldR homologs (left) and closely related TnpB homologs (right). Non-targeting (NT) gRNA controls are shown at the bottom, and the phylogenetic tree indicates the relatedness of the tested proteins. d , Results for the indicated oppF -associated TldR and TnpB homologs, shown as in c .

Extended Data Fig. 8 RFP repression assays reveal variable abilities of TldR homologs to block transcription elongation.

a , RFP repression activity was measured (right) as in Fig. 4f,g using modified gRNAs exhibiting variable complementarity to the target site, as schematized in the grid (left). A gRNA was also tested that lacked the extra 5′ sequence which was absent in RIP-seq reads of mature gRNAs (20 nt no 5′ seq). Bars indicate mean ± s.d. (n = 3 biological replicates). b , Schematic of RFP repression assay in which gRNAs were designed to target either the top or bottom strand within the 5′ UTR of RFP , downstream of the promoter. The phylogenetic trees (right) indicate the relatedness of the tested and labeled homologs. c , Bar graphs plotting normalized RFP fluorescence for the indicated conditions and TldR homologs. EV, empty vector; NT, non-targeting guide. Results with nuclease-dead dCas12 and dCas9 are shown for comparison. Bars indicate mean ± s.d. (n = 3 biological replicates for TldR; n = 6 biological replicates for dCas12/dCas9).

Extended Data Fig. 9 Enterobacter RNA-seq data confirm the native expression of gRNAs from fliC P - tldR loci.

a , RNA-seq read coverage from three Enterobacter strains that natively encode fliC P - tldR loci, revealing clear peaks associated with mature gRNAs containing ~95–97-nt scaffolds and 16-nt guides. Data from three biological replicates are overlaid. b , Predicted secondary structure and sequence of the gRNA associated with Eho TldR. c , Multiple sequence alignment of the DNA encoding gRNA scaffold sequences for representative fliC P -associated TldRs, with conserved positions colored in darker blue.

Extended Data Fig. 10 FliC P is expressed and incorporated into Enterobacter flagella, concomitantly with host FliC repression.

a , RNA-seq read coverage across the tldR -encoding prophage of Enterobacter sp. BIDMC93, demonstrating strong expression of fliC P , tldR , and the gRNA, alongside other genes involved in lysogeny maintenance (e.g. CI). b , Motility assays (left) with wild-type (WT) and Enterobacter deletion strains reveal similar motility phenotypes, as visualized with LB-agar plate images (middle) and a bar graph quantifying motility via halo size (right). Plate images and bar graphs represent three biological replicates; bars indicate mean ± s.d. c , Schematic representation of FliC/FliC P homologs encoded by Enterobacter sp. BIDMC93, with relative genomic positions indicated. FliC 2 is a second host flagellin gene copy encoded at an alternate flagellar assembly locus within this strain, which is not targeted by TldR and not commonly present in other Enterobacter strains. d , Results from liquid chromatography with tandem mass spectrometry (LC–MS/MS) analyses performed on digested peptides from purified flagellar filaments, isolated from the three indicated Enterobacter sp. BIDMC93 strains. The WT ( + CmR) strain encodes the cmR gene downstream of the tldR-gRNA locus (as in Fig. 5e ). Data represent the label free quantification (LFQ) intensities reflecting the variable D2-3 regions of FliC, FliC P , or FliC 2 . Although the FliC 2 appears to be the most dominant flagellin component, the relevant amounts of host FliC and FliC P demonstrate that prophage-encoded FliC P readily assembles into extracellular flagellar filaments, and that host FliC production is de-repressed upon prophage deletion. e , Quantification of changes in the expression profiles of Enterobacter FliC homologs, measured from RNA-seq data of three biological replicates depicted in Fig. 5f,g . TPM, transcripts per million. f , Alignment of fliC / fliC P / fliC 2 promoters indicates that guide RNA-target DNA mismatches prevent TldR-targeting of fliC 2 and fliC P in Enterobacter sp. BIDMC93. g , RNA-seq read coverage in the host fliC promoter/5′-UTR region overlayed for three biological replicates of four Enterobacter strains, with labeled TAM and target sequences highlighted upstream of the TSS. Strain AR136 (top) does not encode a fliC P - tldR locus; note the distinct expression levels, measured via relative counts per million (CPM). h , Alignment of host fliC promoter regions for the strains shown in g compared to E. coli K12, with percent sequence identities indicated on the right. Reported FliA/σ 28 promoter elements from E. coli K12 are shown below the alignment. i , RNA-seq read coverage in the prophage-encoded fliC P promoter/5′-UTR region overlayed for three biological replicates of two representative Enterobacter strains, confirming the predicted TSS. j , Schematic of multiple sequence alignment of the promoter region driving fliC P gene expression, across six verified prophages described in Extended Data Fig. 2 , highlighting the region that was queried for MEME motif detection.

Supplementary information

Supplementary information.

This file contains Supplementary Figs. 1–6.

Reporting Summary

Supplementary tables.

Supplementary Tables 1–8.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Cite this article.

Wiegand, T., Hoffmann, F.T., Walker, M.W.G. et al. TnpB homologues exapted from transposons are RNA-guided transcription factors. Nature (2024). https://doi.org/10.1038/s41586-024-07598-4

Download citation

Received : 27 November 2023

Accepted : 23 May 2024

Published : 26 June 2024

DOI : https://doi.org/10.1038/s41586-024-07598-4

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

By submitting a comment you agree to abide by our Terms and Community Guidelines . If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing: Microbiology newsletter — what matters in microbiology research, free to your inbox weekly.

transcription factor thesis

Information

  • Author Services

Initiatives

You are accessing a machine-readable page. In order to be human-readable, please install an RSS reader.

All articles published by MDPI are made immediately available worldwide under an open access license. No special permission is required to reuse all or part of the article published by MDPI, including figures and tables. For articles published under an open access Creative Common CC BY license, any part of the article may be reused without permission provided that the original article is clearly cited. For more information, please refer to https://www.mdpi.com/openaccess .

Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications.

Feature papers are submitted upon individual invitation or recommendation by the scientific editors and must receive positive feedback from the reviewers.

Editor’s Choice articles are based on recommendations by the scientific editors of MDPI journals from around the world. Editors select a small number of articles recently published in the journal that they believe will be particularly interesting to readers, or important in the respective research area. The aim is to provide a snapshot of some of the most exciting work published in the various research areas of the journal.

Original Submission Date Received: .

  • Active Journals
  • Find a Journal
  • Proceedings Series
  • For Authors
  • For Reviewers
  • For Editors
  • For Librarians
  • For Publishers
  • For Societies
  • For Conference Organizers
  • Open Access Policy
  • Institutional Open Access Program
  • Special Issues Guidelines
  • Editorial Process
  • Research and Publication Ethics
  • Article Processing Charges
  • Testimonials
  • Preprints.org
  • SciProfiles
  • Encyclopedia

ijms-logo

Article Menu

transcription factor thesis

  • Subscribe SciFeed
  • Recommended Articles
  • Author Biographies
  • Google Scholar
  • on Google Scholar
  • Table of Contents

Find support for a specific problem in the support section of our website.

Please let us know what you think of our products and services.

Visit our dedicated information section to learn more about MDPI.

JSmol Viewer

The characterization of a novel prmads11 transcription factor from pinus radiata induced early in bent pine stem.

transcription factor thesis

1. Introduction

2.1. characterization of prmads11, 2.2. differential gene expression of prmads11 in young pine seedlings, 2.3. gene modulation in arabidopsis plants over-expressing prmads11, 2.4. rt-qpcr on selected genes, 2.5. protein–protein interactions using string, 2.6. protein–dna interaction analysis using emsa assays, 3. discussion, 4. materials and methods, 4.1. pine seedlings subjected to inclination stimuli, 4.2. gene cloning and vector construction, 4.3. phylogenetic analysis, 4.4. stable transformation of arabidopsis with truncated prmads11, 4.5. electrophoretic mobility shift assay (emsa), 4.6. rna extraction and quantitative rt-pcr (rt-qpcr), 4.7. arabidopsis promotor analysis, 4.8. lignin quantitation, 4.9. microarray data, 4.10. functional classification based on mapman, 4.11. string interaction network, 5. conclusions, supplementary materials, author contributions, institutional review board statement, informed consent statement, data availability statement, acknowledgments, conflicts of interest.

  • Plomion, C.; Leprovost, G.; Stokes, A. Wood Formation in Trees. Plant Physiol. 2001 , 127 , 1513–1523. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Demura, T.; Fukuda, H. Transcriptional regulation in wood formation. Trends Plant Sci. 2007 , 12 , 64–70. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Wang, H.-Z.; Dixon, R.A. On-off switches for secondary cell wall biosynthesis. Mol. Plant 2012 , 5 , 297–303. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Messenguy, F.; Dubois, E. Role of MADS box proteins and their cofactors in combinatorial control of gene expression and cell development. Gene 2003 , 316 , 1–21. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Aswath, C.R.; Kim, S.H. Another Story of MADS-Box Genes—Their Potential in Plant Biotechnology. Plant Growth Regul. 2005 , 46 , 177–188. [ Google Scholar ] [ CrossRef ]
  • Honma, T.; Goto, K. Complexes of MADS-box proteins are sufficient to convert leaves into floral organs. Nature 2001 , 409 , 525–529. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Theissen, G.; Saedler, H. Plant biology. Floral quartets. Nature 2001 , 409 , 469–471. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Manning, K.; Tör, M.; Poole, M.; Hong, Y.; Thompson, A.J.; King, G.J.; Giovannoni, J.J.; Seymour, G.B. A naturally occurring epigenetic mutation in a gene encoding an SBP-box transcription factor inhibits tomato fruit ripening. Nat. Genet. 2006 , 38 , 948–952. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Jaakola, L.; Poole, M.; Jones, M.O.; Kämäräinen-Karppinen, T.; Koskimäki, J.J.; Hohtola, A.; Häggman, H.; Fraser, P.D.; Manning, K.; King, G.J.; et al. A SQUAMOSA MADS Box Gene Involved in the Regulation of Anthocyanin Accumulation in Bilberry Fruits. Plant Physiol. 2010 , 153 , 1619–1629. [ Google Scholar ] [ CrossRef ]
  • Shore, P.; Sharrocks, A.D. The MADS-box family of transcription factors. Eur. J. Biochem. 1995 , 229 , 1–13. [ Google Scholar ] [ CrossRef ]
  • Cseke, L.J.; Zheng, J.; Podila, G.K. Characterization of PTM5 in aspen trees: A MADS-box gene expressed during woody vascular development. Gene 2003 , 318 , 55–67. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Dias, B.F.d.O.; Simões-Araújo, J.L.; Russo, C.A.M.; Margis, R.; Alves-Ferreira, M. Unravelling MADS-box gene family in Eucalyptus spp.: A starting point to an understanding of their developmental role in trees. Genet. Mol. Biol. 2005 , 28 , 501–510. [ Google Scholar ] [ CrossRef ]
  • Mouradov, A.; Glassick, T.V.; Hamdorf, B.A.; Murphy, L.C.; Marla, S.S.; Yang, Y.; Teasdale, R.D. Family of MADS-Box Genes Expressed Early in Male and Female Reproductive Structures of Monterey Pine. Plant Physiol. 1998 , 117 , 55–62. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Walden, A.R.; Wang, D.Y.; Walter, C.; Gardner, R.C. A large family of TM3 MADS-box cDNAs in Pinus radiata includes two members with deletions of the conserved K domain1GenBank accession numbers: PrMADS4; U90345; PrMADS5; U90346; PrMADS6; U90347; PrMADS7; U90348; PrMADS8; U90349.1. Plant Sci. 1998 , 138 , 167–176. [ Google Scholar ] [ CrossRef ]
  • Cruz, N.; Méndez, T.; Ramos, P.; Urbina, D.; Vega, A.; Gutiérrez, R.A.; Moya-León, M.A.; Herrera, R. Induction of PrMADS10 on the lower side of bent pine tree stems: Potential role in modifying plant cell wall properties and wood anatomy. Sci. Rep. 2019 , 9 , 18981. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Cseke, L.J.; Cseke, S.B.; Ravinder, N.; Taylor, L.C.; Shankar, A.; Sen, B.; Thakur, R.; Karnosky, D.F.; Podila, G.K. SEP-class genes in Populus tremuloides and their likely role in reproductive survival of poplar trees. Gene 2005 , 358 , 18981. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Liljegren, S.J.; Ditta, G.S.; Eshed, Y.; Savidge, B.; Bowman, J.L.; Yanofsky, M.F. SHATTERPROOF MADS-box genes control seed dispersal in Arabidopsis. Nature 2000 , 404 , 766–770. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Ramos, P.; Valenzuela, C.; le Provost, G.; Plomion, C.; Gantz, C.; Moya-León, M.A.; Herrera, R. ACC Oxidase and ACC Synthase Expression Profiles after Leaning of Young Radiata ( P. radiata D. Don) and Maritime Pine ( P. pinaster Ait.) Seedlings. J. Plant Growth Regul. 2012 , 31 , 382–391. [ Google Scholar ] [ CrossRef ]
  • Ramos, P.; Provost, G.L.; Gantz, C.; Plomion, C.; Herrera, R. Transcriptional analysis of differentially expressed genes in response to stem inclination in young seedlings of pine. Plant Biol. 2012 , 14 , 923–933. [ Google Scholar ] [ CrossRef ]
  • Bechtold, U.; Penfold, C.A.; Jenkins, D.J.; Legaie, R.; Moore, J.D.; Lawson, T.; Matthews, J.S.A.; Vialet-Chabrand, S.R.M.; Baxter, L.; Subramaniam, S.; et al. Time-Series Transcriptomics Reveals That AGAMOUS-LIKE22 Affects Primary Metabolism and Developmental Processes in Drought-Stressed Arabidopsis. Plant Cell 2016 , 28 , 345–366. [ Google Scholar ] [ CrossRef ]
  • Mou, Y.; Yuan, C.; Sun, Q.; Yan, C.; Zhao, X.; Wang, J.; Wang, Q.; Shan, S.; Li, C. MIKC-type MADS-box transcription factor gene family in peanut: Genome-wide characterization and expression analysis under abiotic stress. Front. Plant Sci. 2022 , 13 , 980933. [ Google Scholar ] [ CrossRef ]
  • Zhang, B.; Su, X.; Zhou, X. A MADS-box gene of Populus deltoides expressed during flower development and in vegetative organs. Tree Physiol. 2008 , 28 , 929–934. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Zhang, Z.; Xiao, M.; Song, S.; Jiang, Y.; Zhu, X.; Shi, L.; Zheng, X.; Jiang, J.; Miao, M. Genome-Wide Identification, Classification and Expression Analyses of MADS-Box Genes Reveal Their Role in Stem Gall Formation and Expansion of Zizania latifolia. Agronomy 2023 , 13 , 1758. [ Google Scholar ] [ CrossRef ]
  • Minic, Z.; Jamet, E.; San-Clemente, H.; Pelletier, S.; Renou, J.-P.; Rihouey, C.; Okinyo, D.P.; Proux, C.; Lerouge, P.; Jouanin, L. Transcriptomic analysis of Arabidopsis developing stems: A close-up on cell wall genes. BMC Plant Biol. 2009 , 9 , 6. [ Google Scholar ] [ CrossRef ]
  • Taylor-Teeples, M.; Lin, L.; de Lucas, M.; Turco, G.; Toal, T.W.; Gaudinier, A.; Young, N.F.; Trabucco, G.M.; Veling, M.T.; Lamothe, R.; et al. An Arabidopsis gene regulatory network for secondary cell wall synthesis. Nature 2015 , 517 , 571–575. [ Google Scholar ] [ CrossRef ]
  • Zhou, P.; Pu, T.; Gui, C.; Zhang, X.; Gong, L. Transcriptome Analysis Reveals Biosynthesis of Important Bioactive Constituents and Mechanism of Stem Formation of Dendrobium huoshanense. Sci. Rep. 2020 , 10 , 2857. [ Google Scholar ] [ CrossRef ]
  • Edmunds, C.W.; Peralta, P.; Kelley, S.S.; Chiang, V.L.; Sharma-Shivappa, R.R.; Davis, M.F.; Harman-Ware, A.E.; Sykes, R.W.; Gjersing, E.; Cunningham, M.W.; et al. Characterization and enzymatic hydrolysis of wood from transgenic Pinus taeda engineered with syringyl lignin or reduced lignin content. Cellulose 2017 , 24 , 1901–1914. [ Google Scholar ] [ CrossRef ]
  • Chanoca, A.; de Vries, L.; Boerjan, W. Lignin Engineering in Forest Trees. Front. Plant Sci. 2019 , 10 , 912. [ Google Scholar ] [ CrossRef ]
  • McCarthy, R.L.; Zhong, R.; Ye, Z.-H. MYB83 Is a Direct Target of SND1 and Acts Redundantly with MYB46 in the Regulation of Secondary Cell Wall Biosynthesis in Arabidopsis. Plant Cell Physiol. 2009 , 50 , 1950–1964. [ Google Scholar ] [ CrossRef ]
  • Turco, G.M.; Rodriguez-Medina, J.; Siebert, S.; Han, D.; Valderrama-Gómez, M.Á.; Vahldick, H.; Shulse, C.N.; Cole, B.J.; Juliano, C.E.; Dickel, D.E.; et al. Molecular Mechanisms Driving Switch Behavior in Xylem Cell Differentiation. Cell Rep. 2019 , 28 , 342–351.e4. [ Google Scholar ] [ CrossRef ]
  • Gigolashvili, T.; Berger, B.; Flügge, U.-I. Specific and coordinated control of indolic and aliphatic glucosinolate biosynthesis by R2R3-MYB transcription factors in Arabidopsis thaliana. Phytochem. Rev. 2009 , 8 , 3–13. [ Google Scholar ] [ CrossRef ]
  • Zhou, M.; Zhang, K.; Sun, Z.; Yan, M.; Chen, C.; Zhang, X.; Tang, Y.; Wu, Y. LNK1 and LNK2 Corepressors Interact with the MYB3 Transcription Factor in Phenylpropanoid Biosynthesis. Plant Physiol. 2017 , 174 , 1348–1358. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Lin, Y.; Laosatit, K.; Liu, J.; Chen, J.; Yuan, X.; Somta, P.; Chen, X. The mungbean VrP locus encoding MYB90, an R2R3-type MYB protein, regulates anthocyanin biosynthesis. Front. Plant Sci. 2022 , 13 , 895634. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Zhong, R.; Lee, C.; Zhou, J.; McCarthy, R.L.; Ye, Z.-H. A Battery of Transcription Factors Involved in the Regulation of Secondary Cell Wall Biosynthesis in Arabidopsis. Plant Cell 2008 , 20 , 2763–2782. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Lam, P.Y.; Wang, L.; Lui, A.C.W.; Liu, H.; Takeda-Kimura, Y.; Chen, M.-X.; Zhu, F.-Y.; Zhang, J.; Umezawa, T.; Tobimatsu, Y.; et al. Deficiency in flavonoid biosynthesis genes CHS, CHI, and CHIL alters rice flavonoid and lignin profiles. Plant Physiol. 2021 , 188 , 1993–2011. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Kim, S.H.; Mizuno, K.; Fujimura, T. Isolation of MADS-box genes from sweet potato ( Ipomoea batatas (L.) Lam.) expressed specifically in vegetative tissues. Plant Cell Physiol. 2002 , 43 , 314–322. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Yu, H.; Xu, Y.; Tan, E.L.; Kumar, P.P. AGAMOUS-LIKE 24, a dosage-dependent mediator of the flowering signals. Proc. Natl. Acad. Sci. USA 2002 , 99 , 16336–16341. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Liu, Y.; Jia, Z.; Li, X.; Wang, Z.; Chen, F.; Mi, G.; Forde, B.; Takahashi, H.; Yuan, L. Involvement of a truncated MADS-box transcription factor ZmTMM1 in root nitrate foraging. J. Exp. Bot. 2020 , 71 , 4547–4561. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Blaschek, L.; Murozuka, E.; Serk, H.; Ménard, D.; Pesquet, E. Different combinations of laccase paralogs nonredundantly control the amount and composition of lignin in specific cell types and cell wall layers in Arabidopsis. Plant Cell 2023 , 35 , 889–909. [ Google Scholar ] [ CrossRef ]
  • Lima, T.R.A.; Carvalho, E.C.D.; Martins, F.R.; Oliveira, R.S.; Miranda, R.S.; Müller, C.S.; Pereira, L.; Bittencourt, P.R.L.; Sobczak, J.C.M.S.M.; Gomes-Filho, E.; et al. Lignin composition is related to xylem embolism resistance and leaf life span in trees in a tropical semiarid climate. New Phytol. 2018 , 219 , 1252–1262. Available online: https://nph.onlinelibrary.wiley.com/doi/10.1111/nph.15211 (accessed on 1 May 2024). [ CrossRef ]
  • Ménard, D.; Blaschek, L.; Kriechbaum, K.; Lee, C.C.; Serk, H.; Zhu, C.; Lyubartsev, A.; Nuoendagula; Bacsik, Z.; Bergström, L.; et al. Plant biomechanics and resilience to environmental changes are controlled by specific lignin chemistries in each vascular cell type and morphotype. Plant Cell 2022 , 34 , 4877–4896. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Zheng, X.; Li, Y.; Ma, C.; Chen, B.; Sun, Z.; Tian, Y.; Wang, C. A mutation in the promoter of the arabinogalactan protein 7-like gene PcAGP7-1 affects cell morphogenesis and brassinolide content in pear ( Pyrus communis L.) stems. Plant J. 2022 , 109 , 47–63. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Blanc, G.; Wolfe, K.H. Functional divergence of duplicated genes formed by polyploidy during Arabidopsis evolution. Plant Cell 2004 , 16 , 1679–1691. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Ganko, E.W.; Meyers, B.C.; Vision, T.J. Divergence in Expression between Duplicated Genes in Arabidopsis. Mol. Biol. Evol. 2007 , 24 , 2298–2309. [ Google Scholar ] [ CrossRef ]
  • Kumar, S.; Stecher, G.; Li, M.; Knyaz, C.; Tamura, K. MEGA X: Molecular Evolutionary Genetics Analysis across Computing Platforms. Mol. Biol. Evol. 2018 , 35 , 1547–1549. [ Google Scholar ] [ CrossRef ]
  • Hall, T. Bioedit: A User-Friendly Biological Sequence Alignment Editor and Analysis Program for Windows 95/98/ NT. Available online: https://www.semanticscholar.org/paper/BIOEDIT%3A-A-USER-FRIENDLY-BIOLOGICAL-SEQUENCE-EDITOR-Hall/0ae262d9cf78536754bc064e07113ab5e978f208 (accessed on 7 June 2024).
  • Clough, S.J.; Bent, A.F. Floral dip: A simplified method for Agrobacterium-mediated transformation of Arabidopsis thaliana. Plant J. Cell Mol. Biol. 1998 , 16 , 735–743. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • le Provost, G.; Herrera, R.; Paiva, J.A.; Chaumeil, P.; Salin, F.; Plomion, C. A micromethod for high throughput RNA extraction in forest trees. Biol. Res. 2007 , 40 , 291–297. [ Google Scholar ] [ CrossRef ]
  • Vandesompele, J.; De Preter, K.; Pattyn, F.; Poppe, B.; Van Roy, N.; De Paepe, A.; Speleman, F. Accurate normalization of real-time quantitative RT-PCR data by geometric averaging of multiple internal control genes. Genome Biol. 2002 , 3 , research0034.1. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Lescot, M.; Déhais, P.; Thijs, G.; Marchal, K.; Moreau, Y.; Van de Peer, Y.; Rouzé, P.; Rombauts, S. PlantCARE, a database of plant cis-acting regulatory elements and a portal to tools for in silico analysis of promoter sequences. Nucleic Acids Res. 2002 , 30 , 325–327. [ Google Scholar ] [ CrossRef ]
  • Irizarry, R.A.; Bolstad, B.M.; Collin, F.; Cope, L.M.; Hobbs, B.; Speed, T.P. Summaries of Affymetrix GeneChip probe level data. Nucleic Acids Res. 2003 , 31 , e15. [ Google Scholar ] [ CrossRef ]
  • Gautier, L.; Cope, L.; Bolstad, B.M.; Irizarry, R.A. affy--analysis of Affymetrix GeneChip data at the probe level. Bioinforma. Oxf. Engl. 2004 , 20 , 307–315. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Moyano, T.C.; Vidal, E.A.; Contreras-López, O.; Gutiérrez, R.A. Constructing simple biological networks for understanding complex high-throughput data in plants. Methods Mol. Biol. Clifton NJ 2015 , 1284 , 503–526. [ Google Scholar ] [ CrossRef ]
  • Szklarczyk, D.; Gable, A.L.; Lyon, D.; Junge, A.; Wyder, S.; Huerta-Cepas, J.; Simonovic, M.; Doncheva, N.T.; Morris, J.H.; Bork, P.; et al. STRING v11: Protein–protein association networks with increased coverage, supporting functional discovery in genome-wide experimental datasets. Nucleic Acids Res. 2019 , 47 , D607–D613. [ Google Scholar ] [ CrossRef ] [ PubMed ]

Click here to enlarge figure

The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

Méndez, T.; Guajardo, J.; Cruz, N.; Gutiérrez, R.A.; Norambuena, L.; Vega, A.; Moya-León, M.A.; Herrera, R. The Characterization of a Novel PrMADS11 Transcription Factor from Pinus radiata Induced Early in Bent Pine Stem. Int. J. Mol. Sci. 2024 , 25 , 7245. https://doi.org/10.3390/ijms25137245

Méndez T, Guajardo J, Cruz N, Gutiérrez RA, Norambuena L, Vega A, Moya-León MA, Herrera R. The Characterization of a Novel PrMADS11 Transcription Factor from Pinus radiata Induced Early in Bent Pine Stem. International Journal of Molecular Sciences . 2024; 25(13):7245. https://doi.org/10.3390/ijms25137245

Méndez, Tamara, Joselin Guajardo, Nicolás Cruz, Rodrigo A. Gutiérrez, Lorena Norambuena, Andrea Vega, María A. Moya-León, and Raúl Herrera. 2024. "The Characterization of a Novel PrMADS11 Transcription Factor from Pinus radiata Induced Early in Bent Pine Stem" International Journal of Molecular Sciences 25, no. 13: 7245. https://doi.org/10.3390/ijms25137245

Article Metrics

Article access statistics, supplementary material.

ZIP-Document (ZIP, 452 KiB)

Further Information

Mdpi initiatives, follow mdpi.

MDPI

Subscribe to receive issue release notifications and newsletters from MDPI journals

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • Front Immunol
  • PMC10034027

Transcription factors in megakaryocytes and platelets

Hengjie yuan.

1 Tianjin Institute of Neurology, Department of Neurosurgery, Tianjin Medical University General Hospital, Tianjin, China

2 BloodWorks Research Institute, Seattle, WA, United States

Jianning Zhang

Jing-fei dong.

3 Division of Hematology, Department of Medicine, University of Washington, School of Medicine, Seattle, WA, United States

Zilong Zhao

Transcription factors bind promoter or regulatory sequences of a gene to regulate its rate of transcription. However, they are also detected in anucleated platelets. The transcription factors RUNX1, GATA1, STAT3, NFκB, and PPAR have been widely reported to play key roles in the pathophysiology of platelet hyper-reactivity, thrombosis, and atherosclerosis. These non-transcriptional activities are independent of gene transcription or protein synthesis but their underlying mechanisms of action remain poorly defined. Genetic and acquired defects in these transcription factors are associated with the production of platelet microvesicles that are known to initiate and propagate coagulation and to promote thrombosis. In this review, we summarize recent developments in the study of transcription factors in platelet generation, reactivity, and production of microvesicles, with a focus on non-transcriptional activities of selected transcription factors.

1. Introduction

Transcription factors (TFs) are a group of mediators that bind the promoter or regulatory sequence of a gene to control its rate of transcribing genetic information from DNA to messenger RNA ( 1 ). This transcription control is key to ensuring an adequate level of expression of a given protein in targeted cells at a particular developmental stage. It not only directs the processes of proliferation, growth, and death of a cell, but also controls the rate of cell migration and organizational development during embryonic development, as well as regulating cellular response to the extracellular matrices. Thus far, more than 1600 transcription factors have so far been identified ( 2 , 3 ), and they work in a coordinated fashion to down- as well as up-regulate target genes. The activation of a given gene can be regulated by multiple transcriptional factors and one transcription factor can regulate multiple genes. Such a multivalent activity is possible because of the modular structure of a transcriptional factor, which typically includes a DNA-binding domain, signal-sensing domain that contains binding sites for transcription co-regulators, and an optional transactivation domain, which senses external signals and transmits them to the rest of the transcription complex ( 4 , 5 ). Because of their roles in regulating gene transcription, the activation and suppression of transcription factors is extensively reported in cancer development ( 6 ).

Paradoxically, multiple transcription factors have been reported to express and be active in platelets ( Table 1 ), the anucleated offspring of megakaryocytes with a very limited capacity for protein synthesis ( 7 ). An obvious question is whether these transcription factors are merely leftover from parental megakaryocytes or have unique activities in platelets. Reports from studies on platelet transcription factors have been scarce in the literature, but increasing evidence suggests that transcription factors in platelets have unique activities of their own independent of their transcriptional activities ( 8 – 10 ). However, past research on transcription factors in platelets is often limited to reporting their presence and activation status, without further investigation of their activities in regulating platelet functions and, more importantly the underlying mechanism of their regulatory activities.

Table 1

Roles of transcription factors in platelets.

Transcription factorRoles under activation or mutationAssociated hematologic abnormalities
RUNX1Platelet granule development, platelet activationMDS/AML
GATA1Inhibit aggregationDyserythropoiesis
STAT3Increase aggregation, P-selectin,
thrombosis
Coronary artery diseases
NFkBIncrease aggregation, spreading, clot
retraction, GPIBa shedding
Cardiovascular diseases
PPARInhibition of platelet functionCardiovascular diseases

AML, acute myeloid leukemia; MDS, myelodysplastic syndrome.

Platelets circulate along the vessel wall and act to stop bleeding at sites of vessel injury. This hemostatic process requires multiple ligand-receptor interactions to tether, activate, and aggregate platelets. The tightly controlled platelet activation and aggregation that occurs at the site of vascular injury during hemostasis can become dysregulated in pathological conditions, promoting thrombosis and inflammation. For example, platelets promote arterial thrombosis or thromboembolism when activated either on the surface of a ruptured atherosclerotic plaque or by pathological levels of high fluid shear stress in the area of arterial stenosis, leading to acute thrombotic events such as ischemic stroke and myocardial infarction ( 11 ). Emerging evidence further suggests that platelets also act as a cellular mediator in a variety of pathophysiological conditions such as cancer, rheumatoid arthritis, atherosclerosis, trauma, and immune response ( 12 – 14 ). How transcription factors regulate platelet production from megakaryocytes has been extensively reported, but their non-transcriptional activities (i.e., activity independent of gene regulations) have only begun to be recognized. Here, we discuss several transcription factors that have been reported to regulate platelet production and function.

2. Transcription factors in platelet production

2.1. runt-related transcription factor 1.

In 1969, Weiss, et al. identified a family with an autosomal dominant inherited thrombocytopenia, caused primarily by decreased dense granule contents ( 15 ). A heterozygous Y260X mutation in the RUNX1 gene was subsequently shown to be the genetic basis of this inherited platelet defect ( 15 , 16 ). To date, more than 200 families with RUNX1 variants have been reported ( 17 ). RUNX1/AML1 (also known as CBFA2 and PEBP2αB) is a member of the Runt family, which has three known transcription factors (RUNX1, RUNX2, and RUNX3), which share the Runt homology domain near the N-terminus. This domain interacts with CBFb to bind specific sequences of DNA to regulate its transcription ( 18 ).

RUNX1 regulates several genes that control platelet production, structure, function, and intracellular signaling. One report found that 22 patients in a family with autosomal dominant thrombocytopenia had mutations in the RUNX1 gene ( 19 ) and 6 of them developed hematologic malignancies ( 20 ). RUNX1-deficient mice die in uterus due to defective hematopoiesis and resultant severe bleeding ( 21 , 22 ). Mice with the conditional knockout survive but have an impaired megakaryocyte maturation with a significant reduction in megakaryocyte polyploidization ( 23 ). Variations in the RUNX1 gene often result in bleeding diathesis, primarily because of defective platelet granules ( 15 , 16 ), which reduce platelet activation and aggregation ( 24 ). For example, mice carrying the RUNX1 p.Leu43Ser variant (equivalent to human p.Leu56Ser) exhibit a prolonged bleeding time because of defective α-granule secretion and platelet spreading ( 25 ). RUNX1 deficiency can result in pallidin dysregulation and deficient dense granules in platelets ( 26 ) as well as the Ras-related protein RAB31-mediated early endosomal trafficking of von Willebrand factor (VWF) and epidermal growth factor receptor (EGFR) in megakaryocytes ( 27 ). RUNX1 regulates the development of platelet granules through interaction with genes involved in the biogenesis of platelet granules such as the nuclear factor erythroid 2 (NF-E2).

In addition, RUNX1 can also regulate genes related to platelet functions. For example, it regulates the transcription of the non-muscle myosin IIA (MYH9) and IIB (MYH10) genes, which encode non-muscle myosin II heavy chains; RUNX1 mutations are associated with dysregulated expression of MYH10 in platelets ( 28 ); and the expression level of non-muscle myosin is used as a marker for changes in transcriptional activity of RUNX1 as well as friend leukemia integration 1 transcription factor (FLI1) ( 29 ). RUNX1 also regulates the expression of the arachidonate 12-lipoxygenase gene (ALOX12) ( 30 ), which encodes the enzyme that acts on polyunsaturated fatty acid substrates to generate bioactive lipid mediators to regulate platelet function ( 30 ). PCTP (phosphatidylcholine transfer protein) regulates the intermembrane transfer of phosphatidylcholine and its upregulation by RUNX1 sensitizes platelet response to thrombin through protease-activated receptor 4 ( 31 ). RUNX1 also regulates the expression of platelet factor 4 through coordination with transcription factors in the ETS family that share a conserved winged helix-turn-helix DNA binding domain that recognizes unique DNA sequences containing GGAA/T ( 32 ). Platelet factor 4 belongs to the CXC chemokine family and is released from α-granules of activated platelets to promote coagulation and to participate in heparin-induced thrombocytopenia ( 33 , 34 ). A recent report shows that RUNX-1 haploinsufficiency inhibits the differentiation of hematopoietic progenitor cells (HPCs) into megakaryocytes ( 35 ).

2.2. GATA-binding protein 1

GATA-binding protein 1 (GATA1) is a transcription factor that contains two zinc finger domains: a C-terminal zinc finger that binds the (T/A) GATA(A/G) motif of DNA and an N-terminal zinc finger that is required for stabilizing the C-terminal structure and also interacts with a nuclear co-factor protein called friend for GATA1 (FOG1), which stabilizes GATA1 binding ( 36 , 37 ). GATA plays a pivotal role in hematopoietic development and is found in megakaryocytes ( 38 ). GATA1-deficient mice die before birth at approximately embryonic day 10, primarily because of severe anemia ( 39 ). However, mutations in the N-terminal zinc finger domain, which reduces the transcriptional activation of GATA1 ( 36 , 40 ), are found in patients with myeloproliferative disorders and acute megakaryoblastic leukemia ( 41 ), suggesting that GATA1-FOG1 interaction is essential for the development and maturation of megakaryocytes, the parental cells of platelets. Decreased GATA-1 expression has also been reported in patients with myelodysplastic syndrome ( 42 ).

Embryonic stem cells from GATA1-deficient mice are smaller and show low expression of megakaryocytic markers, but have a high rate of proliferation ( 43 ). Complementation of these cells with a wild-type GATA1 gene allows megakaryocytes and erythrocytes to develop in response to a variety of cytokines. Additionally, cell division is attenuated in the megakaryocytic progenitor G1ME cells that overexpress GATA1. A recent report further shows that impaired MYH10 silencing causes GATA1-related polyploidization defect during megakaryocyte differentiation ( 44 ).

Furthermore, platelet aggregation induced by collagen is inhibited in GATA1 - deficient mice ( 45 ), primarily due to reduced expression of the collagen receptor GPVI. Platelet adhesion and aggregation induced by shear stress are also reduced in GATA1 - deficient mice ( 45 ). How a GATA1 deficiency causes these changes in platelet reactivity remains unknown, but these phenotypic changes in the mice provide the first indication that transcription factors could perform non-transcriptional activities in anucleated platelets.

3. Non-transcriptional activity in platelets

3.1. signal transducer and activator of transcription 3.

STAT includes a family of transcription factors critical for inflammatory and acute-phase reactions ( 46 , 47 ). They also play vital roles in cancer development and hematopoiesis ( 48 ). The homologous STAT1, STAT3, and STAT5 are expressed in human platelets and are reported to regulate platelet reactivity through residual or mitochondrial transcriptional activity in platelets. For example, STAT3 affects mitochondrial transcription by binding to the regulatory D-loop region of mitochondrial DNA upon platelet activation ( 49 ).

However, STAT3 can also be activated (phosphorylated) and dimerized in platelets stimulated with thrombopoietin ( 49 , 50 ), suggesting that STAT3 can also regulate platelet reactivity through non-transcriptional means. We have shown that STAT3 is activated and dimerized in collagen-stimulated platelets to serve as a protein scaffold that facilitates the catalytic interaction between spleen tyrosine kinase (Syk) and its substrate, PLCγ to enhance collagen-induced calcium mobilization and platelet activation ( 8 ). More importantly, STAT3 is activated to form dimers by a complex of IL-6 with its soluble receptor IL-6Rα, which activates JAK2 ( 51 ). The pharmacological inhibition of platelet STAT3 reduces collagen-induced platelet aggregation and thrombus formation on the collagen matrix ( 8 , 52 ). Platelets from STAT3-deficient mice or mice infused with a STAT3 inhibitor have reduced collagen-induced aggregation. This non-transcriptional activity of STAT3 may be critical for the development of platelet hyper-reactivity, which has been widely associated with inflammation, especially that related to the activity of the proinflammatory cytokine IL-6 ( 8 ). We have also shown that the piper longum derivative piperlongumine (PL) blocks collagen-induced platelet reactivity in a dose-dependent manner by targeting STAT3 ( 53 ). Consistent with our observations, the small molecular STAT3 inhibitor SC99 has been shown to reduce platelet activation and aggregation induced by collagen and thrombin ( 54 ). These findings offer a new pathway for reducing platelet hyper-reactivity in conditions of inflammation and in prothrombotic states associated with trauma, cancer, autoimmune diseases, and severe infection.

3.2. Nuclear factor kappa β

Nuclear factor kappa β (NFκB) is a well-defined redox-sensitive transcription factor that regulates the immune response and inflammation by controlling the expression of multiple genes activated by inflammatory mediators ( 55 – 57 ). Blocking NFκB can therefore improve outcomes of inflammatory diseases ( 58 ). NFκB is composed of p50 and p65 subunits, normally as an inactive cytoplasmic complex. The inhibitory proteins of the IκB family tightly bind the subunits of NFκB ( 59 ). Upon activation, the IκK complex phosphorylates IκBα, thus activating NFκB by detaching it from IkBα ( 60 – 62 ). Three IκK family members, α, β, and γ, are expressed in platelets, with β being the most abundant, and are reported to regulate platelet reactivity through non-transcriptional activity ( 9 , 10 , 63 ). For example, the pharmacological inhibition of IκKβ leads to reduced agonist-induced platelet activation, increased bleeding time, and prolonged thrombus formation in a mouse model ( 64 ). NF-κB has also been reported to be partially involved in the regulation of SERCA activity to regulate calcium homeostasis in platelets ( 65 ). IκKβ-deficient platelets lose the ability to shed the ectodomain of GP Ibα in response to ADP or collagen stimulations ( 66 ) but preserve thrombin-induced GP Ibα shedding ( 67 ). Collagen-induced p65 and IκKβ phosphorylation is blocked by inhibition of MAP kinase, but not by inhibition of ERK in platelets ( 68 ). The thrombin-induced GP Ibα shedding requires p38 mitogen-activated protein kinase (MAPK) and extracellular signal-regulated kinase (ERK) as its upstream and downstream molecules ( 68 , 69 ).

3.3. Peroxisome proliferator-activated receptors

The peroxisome proliferator-activated receptors (PPARs) are ligand-activated receptors in the nuclear hormone receptor family. They contain three subtypes (PPARα, PPARβ/δ, and PPARγ), which are essential in the regulation of cell differentiation, development, and metabolism ( 70 – 72 ). All PPARs heterodimerize with retinoid X receptor (RXR) and subsequently bind to a specific region of target genes called a peroxisome proliferator response element (PPRE) ( 73 ). PPARγ plays a transcription factor role in regulating platelet production from megakaryocytes, but the PPARγ ligand thiazolidinedione inhibits platelet aggregation induced by ADP under hydrostatic pressure and in diabetic mice ( 74 – 76 ). Similarly, activating PPARβ/δ also reduces platelet reactivity to ADP, thrombin, and collagen ( 77 , 78 ). However, PPARα is also required for platelet activation and thrombus formation, in which it regulates the dense granule secretion of platelets in hyperlipidemic mice ( 79 ). The reason for this apparent contradiction remains to be further investigated. PPARγ is recruited and phosphorylated by Syk to promote the recruitment of the protein called Linker for the Activation of T cells (LAT), which is necessary for collagen-induced platelet activation through glycoprotein VI ( 80 ).

While transcription factors are critically involved in megakaryocyte development and platelet production, they may also regulate platelet reactivity to conventional and specific platelet agonists ( Figure 1 ). The latter is independent of transcriptional activity, for which it is present but at a residual level. This non-transcriptional activity remains poorly understood and requires further investigation because it helps understanding how platelets are activated either by conventional agonists for hemostasis or as complications found in patients treated with drugs that block transcriptional activity of cells (e.g., cancer treatments). Such research will also play an important role in developing new therapeutics targeting these transcription factors to enhance or reduce platelet reactivity.

An external file that holds a picture, illustration, etc.
Object name is fimmu-14-1140501-g001.jpg

Transcription factors regulate platelet aggregation through non-transcriptional activities. (A) PPARγ is recruited and phosphorylated by Syk to promote the recruitment of LAT and enhance platelet aggregation; (B) NFκB is activated by upstream p38 mitogen-activated protein kinase (MAPK) and promotes platelet aggregation by regulating downstream extracellular signal-regulated kinase (ERK); (C) A complex of IL-6 with its soluble receptor IL-6R activates JAK2 to phosphorylate and dimerize STAT3, then the activated STAT3 serves as a protein scaffold to facilitate the catalytic interaction between the spleen tyrosine kinase (Syk) and its substrate PLCγ2 to promote platelet aggregation.

4. Transcription factors in extracellular vesicles released from platelets

Extracellular vesicles (EVs) are shed membrane fragments, intracellular organelles, and nuclear components from cells undergoing active microvesiculation ( 81 – 84 ) or apoptosis ( 85 – 87 ). The former is triggered by the activation of the cysteine protease calpain, which disrupts the membrane-cytoskeleton association ( 88 – 91 ). Platelets are the primary source of EVs circulating in blood, accounting for approximately 80% of total EVs ( 92 – 94 ). The subcellular size of EVs allows them to travel to areas where parental cells are unable to go. In additional to inherent functions from their parental cells, EVs also perform unique activities of their own because of molecules expressed on their surface and carried by them, the latter of which include transcription factors such as STAT3, STAT5, and PPARγ ( 95 ) as well as regulators of transcription factors ( 96 , 97 ). This EV-derived transcriptional activity has been scarcely reported but hold greats potential for influencing biological activities of target cells. For example, PPARγ in platelet EVs is taken up by monocytic THP-1 cells, where it induces the expression of fatty acid-binding protein-4 (FABP4). Monocytes receiving PPARγ-containing platelet EVs produce less inflammatory mediators and become more adherent through increased fibronectin production ( 95 ). Although reports on platelet-derived transcription factors remain very limited, a large body of evidence in the literature shows that platelet-derived EVs, especially EV-carried microRNAs, can change transcriptional activities, thus regulating the function of target cells. Platelet EV-carried NLR family pyrin domain containing 3 (NLRP3) stimulates endothelial cells to undergo pyroptosis through the NLRP3/nuclear factor (NF)-κB pathway ( 98 ). EVs from platelets stimulated with bacteria provoke proinflammatory activity of monocytes through the TRAF6/NFκB pathway ( 99 ). MicroRNA-142-3p carried by platelet-derived EVs promotes the proliferation of endothelial cells ( 100 ), whereas microRNA-126-3p-carrying platelet EVs can be internalized by macrophages to dose-dependently downregulate expression of target mRNA ( 101 ). These observations mostly pertain to phenotypic characterization with less information regarding the underlying pathways involved. Systemic studies of EV-carrying transcription factors and related mediators are therefore urgently needed.

5. Conclusion

Platelets lack a nucleus and de novo transcription, but a number of transcription factors are found in platelets and may have non-transcriptional activities that regulate platelet function. Transferring transcription factors between platelets and target cells through platelet EVs could also be a novel regulatory mechanism of cell-cell communications and a potential therapeutic target for a variety of pathologies.

Author contributions

HY and YL performed the literature search and compiled all the information from the researched articles and wrote the manuscript. ZZ, J-FD and JZ formulated, proposed, guided and wrote the manuscript. All authors contributed to the article and approved the submitted version.

Funding Statement

This study is supported by Young Scientists Award 82022020 from the National Natural Science Foundation of China (ZZ), National Natural Science Foundation of China 81971176 (ZZ), 81271361, 81271359 (JZ), 81102447 (HY), National Natural Science Foundation of China State Key Program Grant 81330029, National Natural Science Foundation of China Major International Joint Research Project 81720108015 (JZ), and Postdoctoral Science Foundation of China Grants 2013M541190 (HY).

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Support Biology

Dei council and dei faculty committee, biology diversity community, mit biology catalyst symposium, honors and awards, employment opportunities, faculty and research, current faculty, in memoriam, areas of research, biochemistry, biophysics, and structural biology, cancer biology, cell biology, computational biology, human disease, microbiology, neurobiology, stem cell and developmental biology, core facilities, video gallery, faculty resources, undergraduate, why biology, undergraduate testimonials, major/minor requirements, general institute requirement, advanced standing exam, transfer credit, current students, subject offerings, research opportunities, biology undergraduate student association, career development, why mit biology, diversity in the graduate program, nih training grant, career outcomes, graduate testimonials, prospective students, application process, interdisciplinary and joint degree programs, living in cambridge, graduate manual: key program info, graduate teaching, career development resources, biology graduate student council, biopals program, postdoctoral, life as a postdoc, postdoc associations, postdoc testimonials, workshops for mit biology postdocs entering the academic job market, responsible conduct of research, postdoc resources, non-mit undergraduates, bernard s. and sophie g. gould mit summer research program in biology (bsg-msrp-bio), bsg-msrp-bio gould fellows, quantitative methods workshop, high school students and teachers, summer workshop for teachers, mit field trips, leah knox scholars program, additional resources, mitx biology, biogenesis podcast, biology newsletter, department calendar, ehs and facilities, graduate manual, resources for md/phd students, preliminary exam guidelines, thesis committee meetings, guidelines for graduating, mentoring students and early-career scientists, remembering stephen goldman (1962 – 2022), it takes three to tango: transcription factors bind dna, protein, and rna.

It takes three to tango: transcription factors bind DNA, protein, and RNA

Greta Friar | Whitehead Institute

July 7, 2023.

Transcription factors could be the Swiss Army knives of gene regulation; they are versatile proteins containing multiple specialized regions. On one end they have a region that can bind to DNA. On the other end they have a region that can bind to proteins. Transcription factors help to regulate gene expression—turning genes on or off and dialing up or down their level of activity—often in partnership with the proteins that they bind. They anchor themselves and their partner proteins to DNA at binding sites in genetic regulatory sequences, bringing together the components that are needed to make gene expression happen.

Transcription factors are a well-known family of proteins, but new research from Whitehead Institute Member Richard Young and colleagues shows that the picture we have had of them is incomplete. In a  paper published in  Molecular Cell  on July 3 , Young and postdocs Ozgur Oksuz and Jonathan Henninger reveal that along with DNA and protein, many transcription factors can also bind RNA. The researchers found that RNA binding keeps transcription factors near their DNA binding sites for longer, helping to fine tune gene expression. This rethinking of how transcription factors work may lead to a better understanding of gene regulation, and may provide new targets for RNA-based therapeutics.

“It’s as if, after carrying around a Swiss Army knife all your life for its blade and scissors, you suddenly realize that the odd, small piece in the back of the knife is a screwdriver,” Young says. “It’s been staring you in the face this whole time, and now that you finally see it, it becomes clear how many more uses there are for the knife than you had realized.”

How transcription factors’ RNA binding went overlooked

A few papers, including one from Young’s lab, had previously identified individual transcription factors as being able to bind RNA, but researchers thought that this was a quirk of the specific transcription factors. Instead, Young, Oksuz, Henninger and collaborators have shown that RNA-binding is in fact a common feature present in at least half of transcription factors.

“We show that RNA binding by transcription factors is a general phenomenon,” Oksuz says. “Individual examples in the past were thought to be exceptions to the rule. Other studies dismissed signs of RNA binding in transcription factors as an artifact—an accident of the experiment rather than a real finding. The clues have been there all along, but I think earlier work was so focused on the DNA and protein interactions that they didn’t consider RNA.”

The reason that researchers had not recognized transcription factors’ RNA binding region as such is because it is not a typical RNA binding domain. Typical RNA binding domains form stable structures that researchers can detect or predict with current technologies. Transcription factors do not contain such structures, and so standard searches for RNA binding domains had not identified them in transcription factors.

Young, Oksuz and Henninger got their biggest clue that researchers might be overlooking something from the human immunodeficiency virus (HIV), which produces a transcription factor-like protein called Tat. Tat increases the transcription of HIV’s RNA genome by binding to the virus’ RNA and then recruiting cellular machinery to it. However, Tat does not contain a structured RNA binding site; instead, it binds RNA from a region called an arginine-rich motif (ARM) that is unstructured but has a high affinity for RNA. When the ARM binds to HIV RNA, the two molecules form a more stable structure together.

The researchers wondered if Tat might be more similar to human transcription factors than anyone had realized. They went through the list of transcription factors, and instead of looking for structured RNA binding domains, they looked for ARMs. They found them in abundance; the majority of human transcription factors contain an ARM-like region between their DNA and protein binding regions, and these sequences were conserved across animal species. Further testing confirmed that many transcription factors do in fact use their ARMs to bind RNA.

RNA binding fine tunes gene expression

Next, the researchers tested to see if RNA binding affected the transcription factors’ function. When transcription factors had their ARMs mutated so they couldn’t bind RNA, those transcription factors were less effective in finding their target sites, remaining at those sites and regulating genes. The mutations did not prevent transcription factors from functioning altogether, suggesting that RNA binding contributes to fine-tuning of gene regulation.

Further experiments confirmed the importance of RNA binding to transcription factor function. The researchers mutated the ARM of a transcription factor important to embryonic development, and found that this led to developmental defects in zebrafish. Additionally, they looked through a list of genetic mutations known to contribute to cancer and heritable diseases, and found that a number of these occur in the RNA binding regions of transcription factors. All of these findings point to RNA binding playing an important role in transcription factors’ regulation of gene expression.

They may also provide therapeutic opportunities. The transcription factors studied by the researchers were found to bind RNA molecules that are produced in the regulatory regions of the genome where the transcription factors bind DNA. This set of transcription factors includes factors that can increase or decrease gene expression. “With evidence that RNAs can tune gene expression through their interaction with positive and negative transcription factors,” says Henninger, “we can envision using existing RNA-based technologies to target RNA molecules, potentially increasing or decreasing expression of specific genes in disease settings.”

Ozgur Oksuz, Jonathan E. Henninger, Robert Warneford-Thomson, Ming M. Zheng, Hailey Erb, Adrienne Vancura, Kalon J. Overholt, Susana Wilson Hawken, Salman F. Banani, Richard Lauman, Lauren N. Reich, Anne L. Robertson, Nancy M. Hannett, Tong I. Lee, Leonard I. Zon, Roberto Bonasio, Richard A. Young. “Transcription factors interact with RNA to regulate genes.”  Molecular Cell , July 3, 2023.  https://doi.org/10.1016/j.molcel.2023.06.012 .

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings
  • My Bibliography
  • Collections
  • Citation manager

Save citation to file

Email citation, add to collections.

  • Create a new collection
  • Add to an existing collection

Add to My Bibliography

Your saved search, create a file for external citation management software, your rss feed.

  • Search in PubMed
  • Search in NLM Catalog
  • Add to Search

The AP2/ERF transcription factor SmERF1L1 regulates the biosynthesis of tanshinones and phenolic acids in Salvia miltiorrhiza

Affiliations.

  • 1 Laboratory of Medicinal Plant Biotechnology, College of Pharmacy, Zhejiang Chinese Medical University, Hangzhou, Zhejiang 310053, PR China; Institute of Plant Biotechnology, College of Life and Environment Sciences, Shanghai Normal University, Shanghai 200234, PR China.
  • 2 Institute of Plant Biotechnology, College of Life and Environment Sciences, Shanghai Normal University, Shanghai 200234, PR China.
  • 3 Laboratory of Medicinal Plant Biotechnology, College of Pharmacy, Zhejiang Chinese Medical University, Hangzhou, Zhejiang 310053, PR China.
  • 4 Jiangsu Provincial Key Laboratory of Coastal Wetland Biological Resources and Environmental Protection, School of Marine and Biological Engineering, Yancheng Teachers Uninversity, Yancheng, Jiangsu Province 224051, PR China.
  • 5 Laboratory of Medicinal Plant Biotechnology, College of Pharmacy, Zhejiang Chinese Medical University, Hangzhou, Zhejiang 310053, PR China; Institute of Plant Biotechnology, College of Life and Environment Sciences, Shanghai Normal University, Shanghai 200234, PR China. Electronic address: [email protected].
  • PMID: 30372953
  • DOI: 10.1016/j.foodchem.2018.08.119

Tanshinones and phenolic acids are two important metabolites synthesized by the traditional Chinese medicinal plant Salvia miltiorrhiza. There is increasing market demand for these compounds. Here, we isolated and functionally characterized SmERF1L1, a novel JA (Jasmonic acid)-responsive gene encoding AP2/ERF transcription factor, from Salvia miltiorrhiza. SmERF1L1 was responsive to methyl jasmonate (MJ), yeast extraction (YE), salicylic acid (SA) and ethylene treatments. Subcellular localization assay indicated that SmERF1L1 located in the nucleus. Overexpression of SmERF1L1 significantly increased tanshinones production in transgenic S. miltiorrhiza hairy roots by comprehensively upregulating tanshinone biosynthetic pathway genes, especially SmDXR. Yeast one-hybrid (Y1H) and electrophoretic mobility shift assay (EMSA) showed that SmERF1L1 binds to the GCC-box of SmDXR promoter while dual luciferase (Dual-LUC) assay showed that SmERF1L1 positively regulated the expression of SmDXR. Our study suggested that the SmERF1L1 may be a good potential target for further metabolic engineering of bioactive component biosynthesis in S. miltiorrhiza.

Keywords: AP2/ERF transcription factor; Biosynthesis; Caffeic acid (PubChem CID: 689043); Cryptotanshinone (PubChem CID: 160254); Dihydrotanshinone (PubChem CID: 11425923); Phenolic acids; Rosmarinic acid (PubChem CID: 5315615); Salvia miltiorrhiza; Salvianolic acid A (PubChem CID: 5281793); Salvianolic acid B (PubChem CID: 11629084); Tanshinone I (PubChem CID: 114917); Tanshinone IIA (PubChem CID: 164676); Tanshinones.

Copyright © 2018 Elsevier Ltd. All rights reserved.

PubMed Disclaimer

Similar articles

  • MAPKK2/4/5/7-MAPK3-JAZs modulate phenolic acid biosynthesis in Salvia miltiorrhiza. Xie Y, Ding M, Yin X, Wang G, Zhang B, Chen L, Ma P, Dong J. Xie Y, et al. Phytochemistry. 2022 Jul;199:113177. doi: 10.1016/j.phytochem.2022.113177. Epub 2022 Mar 28. Phytochemistry. 2022. PMID: 35358599 Review.
  • The biosynthesis of phenolic acids is positively regulated by the JA-responsive transcription factor ERF115 in Salvia miltiorrhiza. Sun M, Shi M, Wang Y, Huang Q, Yuan T, Wang Q, Wang C, Zhou W, Kai G. Sun M, et al. J Exp Bot. 2019 Jan 1;70(1):243-254. doi: 10.1093/jxb/ery349. J Exp Bot. 2019. PMID: 30299490
  • Overexpression of SmbHLH148 induced biosynthesis of tanshinones as well as phenolic acids in Salvia miltiorrhiza hairy roots. Xing B, Liang L, Liu L, Hou Z, Yang D, Yan K, Zhang X, Liang Z. Xing B, et al. Plant Cell Rep. 2018 Dec;37(12):1681-1692. doi: 10.1007/s00299-018-2339-9. Epub 2018 Sep 18. Plant Cell Rep. 2018. PMID: 30229287
  • Comprehensive transcriptome profiling of Salvia miltiorrhiza for discovery of genes associated with the biosynthesis of tanshinones and phenolic acids. Zhou W, Huang Q, Wu X, Zhou Z, Ding M, Shi M, Huang F, Li S, Wang Y, Kai G. Zhou W, et al. Sci Rep. 2017 Sep 5;7(1):10554. doi: 10.1038/s41598-017-10215-2. Sci Rep. 2017. PMID: 28874707 Free PMC article.
  • The Biosynthetic Pathways of Tanshinones and Phenolic Acids in Salvia miltiorrhiza. Ma XH, Ma Y, Tang JF, He YL, Liu YC, Ma XJ, Shen Y, Cui GH, Lin HX, Rong QX, Guo J, Huang LQ. Ma XH, et al. Molecules. 2015 Sep 8;20(9):16235-54. doi: 10.3390/molecules200916235. Molecules. 2015. PMID: 26370949 Free PMC article. Review.
  • SmEIL1 transcription factor inhibits tanshinone accumulation in response to ethylene signaling in Salvia miltiorrhiza . Li X, Xu M, Zhou K, Hao S, Li L, Wang L, Zhou W, Kai G. Li X, et al. Front Plant Sci. 2024 Apr 2;15:1356922. doi: 10.3389/fpls.2024.1356922. eCollection 2024. Front Plant Sci. 2024. PMID: 38628367 Free PMC article.
  • Overexpression of AtMYB2 Promotes Tolerance to Salt Stress and Accumulations of Tanshinones and Phenolic Acid in Salvia miltiorrhiza . Li T, Zhang S, Li Y, Zhang L, Song W, Chen C. Li T, et al. Int J Mol Sci. 2024 Apr 8;25(7):4111. doi: 10.3390/ijms25074111. Int J Mol Sci. 2024. PMID: 38612919 Free PMC article.
  • DoAP2/ERF89 activated the terpene synthase gene DoPAES in Dendrobium officinale and participated in the synthesis of β- patchoulene. Li D, Liu L, Li X, Wei G, Cai Y, Sun X, Fan H. Li D, et al. PeerJ. 2024 Jan 18;12:e16760. doi: 10.7717/peerj.16760. eCollection 2024. PeerJ. 2024. PMID: 38250724 Free PMC article.
  • The SmERF1b-like regulates tanshinone biosynthesis in Salvia miltiorrhiza hairy root. Li D, Liu Y, Chen G, Yan Y, Bai Z. Li D, et al. AoB Plants. 2023 Dec 8;16(1):plad086. doi: 10.1093/aobpla/plad086. eCollection 2024 Jan. AoB Plants. 2023. PMID: 38249522 Free PMC article.
  • Adipose tissue inflammation linked to obesity: A review of current understanding, therapies and relevance of phyto-therapeutics. Aruwa CE, Sabiu S. Aruwa CE, et al. Heliyon. 2023 Dec 2;10(1):e23114. doi: 10.1016/j.heliyon.2023.e23114. eCollection 2024 Jan 15. Heliyon. 2023. PMID: 38163110 Free PMC article. Review.
  • Search in MeSH

Related information

  • PubChem Compound (MeSH Keyword)

LinkOut - more resources

Full text sources.

  • Elsevier Science

Research Materials

  • NCI CPTC Antibody Characterization Program
  • Citation Manager

NCBI Literature Resources

MeSH PMC Bookshelf Disclaimer

The PubMed wordmark and PubMed logo are registered trademarks of the U.S. Department of Health and Human Services (HHS). Unauthorized use of these marks is strictly prohibited.

transcription factor thesis

Lab on a Chip

An integrative temperature-controlled microfluidic system for budding yeast heat shock response analysis at the single-cell level †.

ORCID logo

* Corresponding authors

a The State Key Laboratory for Artificial Microstructures and Mesoscopic Physics, School of Physics, Peking University, Beijing, China E-mail: [email protected]

b Center for Quantitative Biology, Academy for Advanced Interdisciplinary Studies, Peking University, Beijing, China

c Wenzhou Institute, University of Chinese Academy of Sciences, Wenzhou, Zhejiang, China

Cells can respond and adapt to complex forms of environmental change. Budding yeast is widely used as a model system for these stress response studies. In these studies, the precise control of the environment with high temporal resolution is most important. However, there is a lack of single-cell research platforms that enable precise control of the temperature and form of cell growth. This has hindered our understanding of cellular coping strategies in the face of diverse forms of temperature change. Here, we developed a novel temperature-controlled microfluidic platform that integrates a microheater (using liquid metal) and a thermocouple (liquid metal vs. conductive PDMS) on a chip. Three forms of temperature changes (step, gradient, and periodical oscillations) were realized by automated equipment. The platform has the advantages of low cost and a simple fabrication process. Moreover, we investigated the nuclear entry and exit behaviors of the transcription factor Msn2 in yeast in response to heat stress (37 °C) with different heating modes. The feasibility of this temperature-controlled platform for studying the protein dynamic behavior of yeast cells was demonstrated.

Graphical abstract: An integrative temperature-controlled microfluidic system for budding yeast heat shock response analysis at the single-cell level

Supplementary files

  • Supplementary information PDF (1027K)

Article information

Download citation, permissions.

transcription factor thesis

An integrative temperature-controlled microfluidic system for budding yeast heat shock response analysis at the single-cell level

J. Hong, H. He, Y. Xu, S. Wang and C. Luo, Lab Chip , 2024, Advance Article , DOI: 10.1039/D4LC00313F

To request permission to reproduce material from this article, please go to the Copyright Clearance Center request page .

If you are an author contributing to an RSC publication, you do not need to request permission provided correct acknowledgement is given.

If you are the author of this article, you do not need to request permission to reproduce figures and diagrams provided correct acknowledgement is given. If you want to reproduce the whole article in a third-party publication (excluding your thesis/dissertation for which permission is not required) please go to the Copyright Clearance Center request page .

Read more about how to correctly acknowledge RSC content .

Social activity

Search articles by author.

This article has not yet been cited.

Advertisements

  • Search Menu

Sign in through your institution

  • Advance Articles
  • Collections
  • Focus Collections
  • Browse by cover
  • High-Impact Research
  • Author Guidelines
  • Quick and Simple Author Support
  • Focus Issues Call for Papers
  • Submission Site
  • Open Access Options
  • Self-Archiving Policy
  • Why Publish with Us?
  • About Plant Physiology
  • Editorial Board
  • Advertising & Corporate Services
  • Journals on Oxford Academic
  • Books on Oxford Academic

The American Society of Plant Biologists

  • < Previous

Regulatory loops between rice transcription factors OsNAC25 and OsNAC20/26 balance starch synthesis

ORCID logo

  • Article contents
  • Figures & tables
  • Supplementary Data

Juan Wang, Haiqin Zhang, Yuanjiang Wang, Shanshan Meng, Qing Liu, Qian Li, Zhiwen Zhao, Qiaoquan Liu, Cunxu Wei, Regulatory loops between rice transcription factors OsNAC25 and OsNAC20/26 balance starch synthesis, Plant Physiology , Volume 195, Issue 2, June 2024, Pages 1365–1381, https://doi.org/10.1093/plphys/kiae139

  • Permissions Icon Permissions

Several starch synthesis regulators have been identified, but these regulators are situated in the terminus of the regulatory network. Their upstream regulators and the complex regulatory network formed between these regulators remain largely unknown. A previous study demonstrated that NAM, ATAF, and CUC (NAC) transcription factors, OsNAC20 and OsNAC26 (OsNAC20/26), redundantly and positively regulate the accumulation of storage material in rice ( Oryza sativa ) endosperm. In this study, we detected OsNAC25 as an upstream regulator and interacting protein of OsNAC20/26. Both OsNAC25 mutation and OE resulted in a chalky seed phenotype, decreased starch content, and reduced expression of starch synthesis–related genes, but the mechanisms were different. In the osnac25 mutant, decreased expression of OsNAC20 / 26 resulted in reduced starch synthesis; however, in OsNAC25 -overexpressing plants, the OsNAC25–OsNAC20/26 complex inhibited OsNAC20/26 binding to the promoter of starch synthesis–related genes. In addition, OsNAC20/26 positively regulated OsNAC25 . Therefore, the mutual regulation between OsNAC25 and OsNAC20/26 forms a positive regulatory loop to stimulate the expression of starch synthesis–related genes and meet the great demand for starch accumulation in the grain filling stage. Simultaneously, a negative regulatory loop forms among the 3 proteins to avoid the excessive expression of starch synthesis–related genes. Collectively, our findings demonstrate that both promotion and inhibition mechanisms between OsNAC25 and OsNAC20/26 are essential for maintaining stable expression of starch synthesis–related genes and normal starch accumulation.

American Society of Plant Biologists

American Society of Plant Biologists members

Personal account.

  • Sign in with email/username & password
  • Get email alerts
  • Save searches
  • Purchase content
  • Activate your purchase/trial code
  • Add your ORCID iD

Institutional access

Sign in with a library card.

  • Sign in with username/password
  • Recommend to your librarian
  • Institutional account management
  • Get help with access

Access to content on Oxford Academic is often provided through institutional subscriptions and purchases. If you are a member of an institution with an active account, you may be able to access content in one of the following ways:

IP based access

Typically, access is provided across an institutional network to a range of IP addresses. This authentication occurs automatically, and it is not possible to sign out of an IP authenticated account.

Choose this option to get remote access when outside your institution. Shibboleth/Open Athens technology is used to provide single sign-on between your institution’s website and Oxford Academic.

  • Click Sign in through your institution.
  • Select your institution from the list provided, which will take you to your institution's website to sign in.
  • When on the institution site, please use the credentials provided by your institution. Do not use an Oxford Academic personal account.
  • Following successful sign in, you will be returned to Oxford Academic.

If your institution is not listed or you cannot sign in to your institution’s website, please contact your librarian or administrator.

Enter your library card number to sign in. If you cannot sign in, please contact your librarian.

Society Members

Society member access to a journal is achieved in one of the following ways:

Sign in through society site

Many societies offer single sign-on between the society website and Oxford Academic. If you see ‘Sign in through society site’ in the sign in pane within a journal:

  • Click Sign in through society site.
  • When on the society site, please use the credentials provided by that society. Do not use an Oxford Academic personal account.

If you do not have a society account or have forgotten your username or password, please contact your society.

Sign in using a personal account

Some societies use Oxford Academic personal accounts to provide access to their members. See below.

A personal account can be used to get email alerts, save searches, purchase content, and activate subscriptions.

Some societies use Oxford Academic personal accounts to provide access to their members.

Viewing your signed in accounts

Click the account icon in the top right to:

  • View your signed in personal account and access account management features.
  • View the institutional accounts that are providing access.

Signed in but can't access content

Oxford Academic is home to a wide variety of products. The institutional subscription may not cover the content that you are trying to access. If you believe you should have access to that content, please contact your librarian.

For librarians and administrators, your personal account also provides access to institutional account management. Here you will find options to view and activate subscriptions, manage institutional settings and access options, access usage statistics, and more.

Short-term Access

To purchase short-term access, please sign in to your personal account above.

Don't already have a personal account? Register

Month: Total Views:
March 2024 338
April 2024 259
May 2024 126
June 2024 123

Email alerts

Citing articles via.

  • Recommend to Your Librarian
  • Advertising & Corporate Services
  • Awards & Funding
  • Plant Science Today
  • Plant Biology Meeting
  • Meeting Management Services
  • Plant Science Research Weekly
  • Taproot: A Plantae Podcast

Affiliations

  • Online ISSN 1532-2548
  • Print ISSN 0032-0889
  • Copyright © 2024 American Society of Plant Biologists
  • About Oxford Academic
  • Publish journals with us
  • University press partners
  • What we publish
  • New features  
  • Open access
  • Rights and permissions
  • Accessibility
  • Advertising
  • Media enquiries
  • Oxford University Press
  • Oxford Languages
  • University of Oxford

Oxford University Press is a department of the University of Oxford. It furthers the University's objective of excellence in research, scholarship, and education by publishing worldwide

  • Copyright © 2024 Oxford University Press
  • Cookie settings
  • Cookie policy
  • Privacy policy
  • Legal notice

This Feature Is Available To Subscribers Only

Sign In or Create an Account

This PDF is available to Subscribers Only

For full access to this pdf, sign in to an existing account, or purchase an annual subscription.

IMAGES

  1. Transcription Factor Interplay

    transcription factor thesis

  2. What are Transcription factors

    transcription factor thesis

  3. | Transcription factors analyses. (A) Families of all transcription

    transcription factor thesis

  4. Figure 2 from Transcription factor heterogeneity in pluripotent stem

    transcription factor thesis

  5. 16.5C: Cancer and Transcriptional Control

    transcription factor thesis

  6. What are Transcription factors

    transcription factor thesis

VIDEO

  1. Medical vocabulary: What does E2F1 Transcription Factor mean

  2. TRANSCRIPTION FACTORS: DIFFERENTIATION

  3. Transcription factor activation & inhibition in microRNA, mRNA & RNA-Seq cancer-related experiments

  4. Master Transcription Factor Binding Regulates Early...

  5. Transcriptional Unit

  6. Prediction of transcription factor binding sites...

COMMENTS

  1. Mechanisms and biotechnological applications of transcription factors

    1. Introduction. Transcription is a crucial component of the central dogma of molecular biology (DNA-RNA-protein) [], serving as a bridge that translates genetic information into diverse forms at the individual level.Transcription factors (TFs) play a pivotal role in regulating the transcription of target genes by selectively recognizing and binding specific DNA regions known as TF binding ...

  2. Predicting Transcription Factor Binding Using Neural Structured Learning

    PREDICTING TRANSCRIPTION FACTOR BINDING USING NEURAL STRUCTURED LEARNING A Thesis in Bioinformatics and Genomics by Natalie Zesati Submitted in Partial Fulfillment of the Requirements for the Degree of Master of Science December 2020. ii The thesis of Natalie Zesati was reviewed and approved by the following: Shaun Mahony Assistant Professor of ...

  3. Characterization of A Putative Transcription Factor

    Transcription factors are proteins that initiate and modulate transcription rate by interacting with specific DNA recognition sequences in the target genes. As shown in Fig. 1, these DNA-binding transcription factors are structurally classified into four major classes: Helix-turn-helix homeodomain (e.g. PBX1 ), C. 2. H. 2 . zinc

  4. Investigating the Localization of FOXO Transcription Factors in

    Hinojosa, Leetoria, Investigating the Localization of FOXO Transcription Factors in. Glioblastoma. Master of Sciences (MS), May, 2020, 32pp., 1 table, 7 figures, 17 references. The Phosphatidylinositol 3 Kinase (PI3K) pathway is an essential intracellular signaling. pathway that regulates cellular growth, survival, and fate.

  5. The Human Transcription Factors

    The Human Transcription Factors. Transcription factors (TFs) recognize specific DNA sequences to control chromatin and transcription, forming a complex system that guides expression of the genome. Despite keen interest in understanding how TFs control gene expression, it remains challenging to determine how the precise genomic binding sites of ...

  6. PDF Thesis proposal Functional Validation of Transcription Factor to Gene

    Thesis proposal Functional Validation of Transcription Factor to Gene Interactions by Statistical Learning of Gaussian Bayesian networks from SNP and Expression data. Jing Xiang Machine Learning Department Carnegie Mellon University [email protected] Committee members: Seyoung Kim Geoff Gordon Carl Kingsford Steffi Oesterreich January 23, 2017

  7. PDF Production and Characterization of Recombinant Transducible

    The overall objective of this thesis was to produce and characterize recombinant transducible (able to enter cells) transcription factors (TFs). TFs are complex, difficult-to-make proteins that regulate gene expression and cell fate. Thus, we wanted to deliver transducible TFs exogenously to cells to change gene expression and alter cell fate.

  8. Refactoring transcription factors for metabolic engineering

    The transcription factors binding to a specific DNA region is a key in transcription regulation. Therefore, identifying transcription factor DBSs and verifying the interactions between transcription factors are key to understanding transcriptional regulatory mechanisms and networks (Fig. 2).Download : Download high-res image (438KB) Download : Download full-size image

  9. Proton-pumping photoreceptor controls expression of ABC ...

    The same concentration of GR as HTH transcription factor protein and BSA was prepared in 50 mM Tris and 150 mM NaCl at pH 7.0. The average of each dataset was used for the fitting process. Based ...

  10. STAN, a computational framework for inferring spatially ...

    Transcription factors (TFs) drive significant cellular changes in response to environmental cues and intercellular signaling. Neighboring cells influence TF activity and, consequently, cellular fate and function. Spatial transcriptomics (ST) captures mRNA expression patterns across tissue samples, enabling characterization of the local microenvironment. However, these datasets have not been ...

  11. ROLE OF TRANSCRIPTION FACTORS IN SENSORY NEURON SPECIFICATION

    This thesis explores the role of transcription factors in sensory neuron specification. We describe the transcription factor Foxs1 as an early sensory neuronal marker and use it to

  12. The Human Transcription Factors

    Abstract. Transcription factors (TFs) recognize specific DNA sequences to control chromatin and transcription, forming a complex system that guides expression of the genome. Despite keen interest in understanding how TFs control gene expression, it remains challenging to determine how the precise genomic binding sites of TFs are specified and ...

  13. PDF NUCLEIC ACID STRUCTURAL RECOGNITION BY TRANSCRIPTION FACTOR By A Thesis

    transcription factors to bind to DNA and slide until they find their consensus sequence. One transcription factor, AmrZ, functions as both an activator and repressor and studying ... thesis. The changes in conformation can also be due to the specific conformation of DNA. There are three main forms that DNA can take: A-, B-, and Z-form DNA ...

  14. PDF Transcription factor occupancy in differentiating skeletal muscle

    in the thesis, they were very helpful to the overall understanding of the role and function of bHLH transcription factors in various developmental lineages. Dr. Fancis Collins has given me a superb opportunity to pursue a postdoctoral project in the genomics of type 2 diabetes, and showed a lot of

  15. TnpB homologues exapted from transposons are RNA-guided transcription

    RNA-guided transcription factors arose repeatedly via the domestication of transposon-encoded tnpB genes, representing a parallel evolutionary path to CRISPR-Cas adaptive immunity.

  16. PDF Predicting enhancer regions and transcription factor binding sites in D

    I have also applied supervised learning methods for predicting transcription factor binding locations based on combinations of regulatory motifs. For each experiment in a compendium of ChIP-chip studies, I constructed a classifier to distinguish between regions bound by the given factor and regions bound by any other factor. For each

  17. PDF Investigating the role of transcription factor, Trl, during germline

    Investigating the role of transcription factor, Trl, during germline development in the Drosophila ovary by Lindsay L. Davenport July 2019 Director of Thesis: Elizabeth T. Ables, Ph. D. Major Department: Biology Oogenesis is the process by which an egg develops from undifferentiated cells in the ovary.

  18. The Characterization of a Novel PrMADS11 Transcription Factor from

    A novel MADS-box transcription factor from Pinus radiata D. Don was characterized. PrMADS11 encodes a protein of 165 amino acids for a MADS-box transcription factor belonging to group II, related to the MIKC protein structure. PrMADS11 was differentially expressed in the stems of pine trees in response to 45° inclination at early times (1 h). Arabidopsis thaliana was stably transformed with a ...

  19. Deciphering the Role of the Ser-Phosphorylation Pattern on the ...

    The chemical synthesis of site-specifically modified transcription factors (TFs) is a powerful method to investigate how post-translational modifications (PTMs) influence TF-DNA interactions and impact gene expression. Among these TFs, Max plays a pivotal role in controlling the expression of 15 % of the genome. The activity of Max is regulated ...

  20. Graduate Thesis Or Dissertation

    These programs are fundamentally regulated at transcription and are orchestrated by sequence-specific transcription factors (TFs). The work presented here focuses on the wide-spread survey of TF activity, as well as an in depth study of a single TF, p53. ... The focus of the first part of this thesis is to define a computational model to assess ...

  21. Comprehensive analysis of Transcription Factors identified novel

    Background: Transcriptional factors (TFs) are responsible for regulating the transcription of pro-oncogenes and tumor suppressor genes in the process of tumor development. However, the role of these transcription factors in Bladder cancer (BCa) remains unclear. And the main purpose of this research is to explore the possibility of these TFs serving as biomarkers for BCa.

  22. Transcription factor

    Illustration of an activator. In molecular biology, a transcription factor (TF) (or sequence-specific DNA-binding factor) is a protein that controls the rate of transcription of genetic information from DNA to messenger RNA, by binding to a specific DNA sequence. The function of TFs is to regulate—turn on and off—genes in order to make sure that they are expressed in the desired cells at ...

  23. Transcription factors in megakaryocytes and platelets

    1. Introduction. Transcription factors (TFs) are a group of mediators that bind the promoter or regulatory sequence of a gene to control its rate of transcribing genetic information from DNA to messenger RNA ().This transcription control is key to ensuring an adequate level of expression of a given protein in targeted cells at a particular developmental stage.

  24. It takes three to tango: transcription factors bind DNA, protein, and

    Transcription factors could be the Swiss Army knives of gene regulation; they are versatile proteins containing multiple specialized regions. On one end they have a region that can bind to DNA. On the other end they have a region that can bind to proteins. Transcription factors help to regulate gene expression—turning genes on or off […]

  25. The AP2/ERF transcription factor SmERF1L1 regulates the ...

    Here, we isolated and functionally characterized SmERF1L1, a novel JA (Jasmonic acid)-responsive gene encoding AP2/ERF transcription factor, from Salvia miltiorrhiza. SmERF1L1 was responsive to methyl jasmonate (MJ), yeast extraction (YE), salicylic acid (SA) and ethylene treatments. Subcellular localization assay indicated that SmERF1L1 ...

  26. An integrative temperature-controlled microfluidic system for budding

    Moreover, we investigated the nuclear entry and exit behaviors of the transcription factor Msn2 in yeast in response to heat stress (37 °C) with different heating modes. ... If you want to reproduce the whole article in a third-party publication (excluding your thesis/dissertation for which permission is not required) please go to the ...

  27. Regulatory loops between rice transcription factors OsNAC25 and OsNAC20

    A previous study demonstrated that NAM, ATAF, and CUC (NAC) transcription factors, OsNAC20 and OsNAC26 (OsNAC20/26), redundantly and positively regulate the accumulation of storage material in rice (Oryza sativa) endosperm. In this study, we detected OsNAC25 as an upstream regulator and interacting protein of OsNAC20/26.

  28. PDF Identification of Transcription Factor Binding Sites

    TRANSCRIPTION FACTOR BINDING SITES By LIANG ZHAO Bachelor of Science Zhejiang University Hangzhou, China 1992 . . Master of Engineering BetJmg Research Institute of Chemical Industry Beijing, China 1995 Submitted to the Faculty of the Graduate College of the Oklahoma State University in partial fulfillment of the requirements for the degree of