@article{aimone_lavington_hoyer_deppong_mickelson-young_jacobson_kennedy_carbone_hanley-bowdoin_duffy_2021, title={Population diversity of cassava mosaic begomoviruses increases over the course of serial vegetative propagation}, volume={102}, ISSN={0022-1317 1465-2099}, url={http://dx.doi.org/10.1099/jgv.0.001622}, DOI={10.1099/jgv.0.001622}, abstractNote={Cassava mosaic disease (CMD) represents a serious threat to cassava, a major root crop for more than 300 million Africans. CMD is caused by single-stranded DNA begomoviruses that evolve rapidly, making it challenging to develop durable disease resistance. In addition to the evolutionary forces of mutation, recombination and reassortment, factors such as climate, agriculture practices and the presence of DNA satellites may impact viral diversity. To gain insight into the factors that alter and shape viral diversity in planta, we used high-throughput sequencing to characterize the accumulation of nucleotide diversity after inoculation of infectious clones corresponding to African cassava mosaic virus (ACMV) and East African cassava mosaic Cameroon virus (EACMCV) in the susceptible cassava landrace Kibandameno. We found that vegetative propagation had a significant effect on viral nucleotide diversity, while temperature and a satellite DNA did not have measurable impacts in our study. EACMCV diversity increased linearly with the number of vegetative propagation passages, while ACMV diversity increased for a time and then decreased in later passages. We observed a substitution bias toward C→T and G→A for mutations in the viral genomes consistent with field isolates. Non-coding regions excluding the promoter regions of genes showed the highest levels of nucleotide diversity for each genome component. Changes in the 5′ intergenic region of DNA-A resembled the sequence of the cognate DNA-B sequence. The majority of nucleotide changes in coding regions were non-synonymous, most with predicted deleterious effects on protein structure, indicative of relaxed selection pressure over six vegetative passages. Overall, these results underscore the importance of knowing how cropping practices affect viral evolution and disease progression.}, number={7}, journal={Journal of General Virology}, publisher={Microbiology Society}, author={Aimone, Catherine D. and Lavington, Erik and Hoyer, J. Steen and Deppong, David O. and Mickelson-Young, Leigh and Jacobson, Alana and Kennedy, George G. and Carbone, Ignazio and Hanley-Bowdoin, Linda and Duffy, Siobain}, year={2021}, month={Jul} } @article{wear_song_zynda_mickelson-young_leblanc_lee_deppong_allen_martienssen_vaughn_et al._2020, title={Comparing DNA replication programs reveals large timing shifts at centromeres of endocycling cells in maize roots}, volume={16}, ISSN={["1553-7404"]}, DOI={10.1371/journal.pgen.1008623}, abstractNote={Plant cells undergo two types of cell cycles–the mitotic cycle in which DNA replication is coupled to mitosis, and the endocycle in which DNA replication occurs in the absence of cell division. To investigate DNA replication programs in these two types of cell cycles, we pulse labeled intact root tips of maize (Zea mays) with 5-ethynyl-2’-deoxyuridine (EdU) and used flow sorting of nuclei to examine DNA replication timing (RT) during the transition from a mitotic cycle to an endocycle. Comparison of the sequence-based RT profiles showed that most regions of the maize genome replicate at the same time during S phase in mitotic and endocycling cells, despite the need to replicate twice as much DNA in the endocycle and the fact that endocycling is typically associated with cell differentiation. However, regions collectively corresponding to 2% of the genome displayed significant changes in timing between the two types of cell cycles. The majority of these regions are small with a median size of 135 kb, shift to a later RT in the endocycle, and are enriched for genes expressed in the root tip. We found larger regions that shifted RT in centromeres of seven of the ten maize chromosomes. These regions covered the majority of the previously defined functional centromere, which ranged between 1 and 2 Mb in size in the reference genome. They replicate mainly during mid S phase in mitotic cells but primarily in late S phase of the endocycle. In contrast, the immediately adjacent pericentromere sequences are primarily late replicating in both cell cycles. Analysis of CENH3 enrichment levels in 8C vs 2C nuclei suggested that there is only a partial replacement of CENH3 nucleosomes after endocycle replication is complete. The shift to later replication of centromeres and possible reduction in CENH3 enrichment after endocycle replication is consistent with a hypothesis that centromeres are inactivated when their function is no longer needed.}, number={10}, journal={PLOS GENETICS}, author={Wear, Emily E. and Song, Jawon and Zynda, Gregory J. and Mickelson-Young, Leigh and LeBlanc, Chantal and Lee, Tae-Jin and Deppong, David O. and Allen, George C. and Martienssen, Robert A. and Vaughn, Matthew W. and et al.}, year={2020}, month={Oct} } @article{turpin_vera_savadel_lung_wear_mickelson-young_thompson_hanley-bowdoin_dennis_zhang_et al._2018, title={Chromatin structure profile data from DNS-seq: Differential nuclease sensitivity mapping of four reference tissues of B73 maize (Zea mays L)}, volume={20}, ISSN={["2352-3409"]}, DOI={10.1016/j.dib.2018.08.015}, abstractNote={Presented here are data from Next-Generation Sequencing of differential micrococcal nuclease digestions of formaldehyde-crosslinked chromatin in selected tissues of maize (Zea mays) inbred line B73. Supplemental materials include a wet-bench protocol for making DNS-seq libraries, the DNS-seq data processing pipeline for producing genome browser tracks. This report also includes the peak-calling pipeline using the iSeg algorithm to segment positive and negative peaks from the DNS-seq difference profiles. The data repository for the sequence data is the NCBI SRA, BioProject Accession PRJNA445708.}, journal={DATA IN BRIEF}, author={Turpin, Zachary M. and Vera, Daniel L. and Savadel, Savannah D. and Lung, Pei-Yau and Wear, Emily E. and Mickelson-Young, Leigh and Thompson, William F. and Hanley-Bowdoin, Linda and Dennis, Jonathan H. and Zhang, Jinfeng and et al.}, year={2018}, month={Oct}, pages={358–363} } @article{wear_song_zynda_leblanc_lee_mickelson-young_concia_mulvaney_szymanski_allen_et al._2017, title={Genomic Analysis of the DNA Replication Timing Program during Mitotic S Phase in Maize (Zea mays) Root Tips}, volume={29}, ISSN={["1532-298X"]}, url={http://europepmc.org/abstract/med/28842533}, DOI={10.1105/tpc.17.00037}, abstractNote={The time during S phase at which different maize DNA sequences replicate reveals a complex temporal program influenced by genomic features, transcriptional activity, and chromatin structure. All plants and animals must replicate their DNA, using a regulated process to ensure that their genomes are completely and accurately replicated. DNA replication timing programs have been extensively studied in yeast and animal systems, but much less is known about the replication programs of plants. We report a novel adaptation of the “Repli-seq” assay for use in intact root tips of maize (Zea mays) that includes several different cell lineages and present whole-genome replication timing profiles from cells in early, mid, and late S phase of the mitotic cell cycle. Maize root tips have a complex replication timing program, including regions of distinct early, mid, and late S replication that each constitute between 20 and 24% of the genome, as well as other loci corresponding to ∼32% of the genome that exhibit replication activity in two different time windows. Analyses of genomic, transcriptional, and chromatin features of the euchromatic portion of the maize genome provide evidence for a gradient of early replicating, open chromatin that transitions gradually to less open and less transcriptionally active chromatin replicating in mid S phase. Our genomic level analysis also demonstrated that the centromere core replicates in mid S, before heavily compacted classical heterochromatin, including pericentromeres and knobs, which replicate during late S phase.}, number={9}, journal={PLANT CELL}, author={Wear, Emily E. and Song, Jawon and Zynda, Gregory J. and LeBlanc, Chantal and Lee, Tae-Jin and Mickelson-Young, Leigh and Concia, Lorenzo and Mulvaney, Patrick and Szymanski, Eric S. and Allen, George C. and et al.}, year={2017}, month={Sep}, pages={2126–2149} } @article{mickelson-young_wear_mulvaney_lee_szymanski_allen_hanley-bowdoin_thompson_2016, title={A flow cytometric method for estimating S-phase duration in plants}, volume={67}, ISSN={["1460-2431"]}, url={http://europepmc.org/abstract/med/27697785}, DOI={10.1093/jxb/erw367}, abstractNote={Highlight We estimated S-phase duration for several plant species by following EdU-labeled nuclei from G1 to G2 using bivariate flow cytometry. S-phase duration is relatively consistent over a range of genome sizes.}, number={21}, journal={JOURNAL OF EXPERIMENTAL BOTANY}, author={Mickelson-Young, Leigh and Wear, Emily and Mulvaney, Patrick and Lee, Tae-Jin and Szymanski, Eric S. and Allen, George and Hanley-Bowdoin, Linda and Thompson, William}, year={2016}, month={Nov}, pages={6077–6087} }