16
Comparative transcriptomics of a monocotyledonous geophyte reveals shared molecular mechanisms of underground storage organ formation Carrie M. Tribble 1, * , Jes ´ us Mart´ ınez-G ´ omez 1, 2 , Fernando Alzate-Guarin 3 , Carl J. Rothfels 1 , and Chelsea D. Specht 2 1 University Herbarium and Department of Integrative Biology, University of California, Berkeley, Berkeley, CA 94720 2 School of Integrative Plant Sciences and L.H. Bailey Hortorium, Cornell University, Ithaca, NY 14853 USA 3 Grupo de Estudios Bot´ anicos (GEOBOTA) and Herbario Universidad de Antioquia (HUA), Instituto de Biolog´ ıa, Facultad de Ciencias Exactas y Naturales, Universidad de Antioquia, Calle 67 N 53-108, Medell´ ın, Colombia * E-mail: [email protected] September 20, 2020 Abstract Many species from across the vascular plant tree-of-life have modified standard plant tissues into tu- bers, bulbs, corms, and other underground storage organs (USOs), unique innovations which allow these plants to retreat underground. Our ability to understand the developmental and evolutionary forces that shape these mor- phologies is limited by a lack of studies on certain USOs and plant clades. Bomarea multiflora (Alstroemeriaceae) is a monocot with tuberous roots, filling a key gap in our understanding of USO development. We take a compara- tive transcriptomics approach to characterizing the molecular mechanisms of tuberous root formation in B. multiflora and compare these mechanisms to those identified in other USOs across diverse plant lineages. We sequenced tran- scriptomes from the growing tip of four tissue types (aerial shoot, rhizome, fibrous root, and root tuber) of three individuals of B. multiflora. We identify differentially expressed isoforms between tuberous and non-tuberous roots and test the expression of a priori candidate genes implicated in underground storage in other taxa. We identify 271 genes that are differentially expressed in root tubers versus non-tuberous roots, including genes implicated in cell wall modification, defense response, and starch biosynthesis. We also identify a phosphatidylethanolamine-binding protein (PEBP), which has been implicated in tuberization signalling in other taxa and, through gene-tree analysis, place this copy in a phylogenytic context. These findings suggest that some similar molecular processes underlie the formation of underground storage structures across flowering plants despite the long evolutionary distances among taxa and non-homologous morphologies (e.g., bulbs versus tubers). [Plant development, tuberous roots, comparative transcriptomics, geophytes] 1 Introduction Scientific attention in botanical fields focuses almost ex- 1 clusively on aboveground organs and biomass. How- 2 ever, a holistic understanding of land plant evolution, 3 morphology, and ecology requires a comprehensive un- 4 derstanding of belowground structures: on average 50% 5 of an individual plant’s biomass lies beneath the ground 6 (Niklas, 2005), and these portions of a plant are critical for 7 resource acquisition, resource storage, and mediating the 8 plant’s interactions with its environment. Often, below- 9 ground biomass is thought to consist solely of standard 10 root tissue, but in some cases, plants modify “ordinary” 11 structures for specialized underground functions. Plants 12 called geophytes fall toward the extreme end of this be- 13 lowground/aboveground allocation spectrum. In a re- 14 markable example of convergent evolution of an innova- 15 tive life history strategy, geophytes retreat underground 16 by producing the buds of new growth on structures be- 17 low the soil surface, while also storing nutrients to fuel 18 this growth in highly modified, specialized underground 19 storage organs (USOs) (Raunkiaer et al., 1934; Dafni et al., 20 1981b,a; Al-Tardeh et al., 2008; Vesel´ y et al., 2011). Many 21 geophytes also have the capacity to reproduce asexually 22 through underground offshoots in addition to sexual re- 23 production. Geophytes are ecologically and economically 24 important, morphologically diverse, and have evolved 25 independently in all major groups of vascular plants ex- 26 cept gymnosperms (Howard et al., 2019, 2020). These 27 plants and their associated underground structures are 28 1 not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which was this version posted September 20, 2020. ; https://doi.org/10.1101/845602 doi: bioRxiv preprint

Comparative transcriptomics of a monocotyledonous …Comparative transcriptomics of a monocotyledonous geophyte reveals shared molecular mechanisms of underground storage organ formation

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Comparative transcriptomics of a monocotyledonous …Comparative transcriptomics of a monocotyledonous geophyte reveals shared molecular mechanisms of underground storage organ formation

Comparative transcriptomics of a monocotyledonous geophytereveals shared molecular mechanisms of underground storage

organ formation

Carrie M. Tribble1, ∗, Jesus Martınez-Gomez1, 2, Fernando Alzate-Guarin3, Carl J. Rothfels1, andChelsea D. Specht2

1University Herbarium and Department of Integrative Biology, University of California, Berkeley, Berkeley, CA 947202School of Integrative Plant Sciences and L.H. Bailey Hortorium, Cornell University, Ithaca, NY 14853 USA

3Grupo de Estudios Botanicos (GEOBOTA) and Herbario Universidad de Antioquia (HUA), Instituto de Biologıa, Facultad de CienciasExactas y Naturales, Universidad de Antioquia, Calle 67 N◦ 53-108, Medellın, Colombia

∗E-mail: [email protected]

September 20, 2020

Abstract Many species from across the vascular plant tree-of-life have modified standard plant tissues into tu-bers, bulbs, corms, and other underground storage organs (USOs), unique innovations which allow these plants toretreat underground. Our ability to understand the developmental and evolutionary forces that shape these mor-phologies is limited by a lack of studies on certain USOs and plant clades. Bomarea multiflora (Alstroemeriaceae) isa monocot with tuberous roots, filling a key gap in our understanding of USO development. We take a compara-tive transcriptomics approach to characterizing the molecular mechanisms of tuberous root formation in B. multifloraand compare these mechanisms to those identified in other USOs across diverse plant lineages. We sequenced tran-scriptomes from the growing tip of four tissue types (aerial shoot, rhizome, fibrous root, and root tuber) of threeindividuals of B. multiflora. We identify differentially expressed isoforms between tuberous and non-tuberous rootsand test the expression of a priori candidate genes implicated in underground storage in other taxa. We identify 271genes that are differentially expressed in root tubers versus non-tuberous roots, including genes implicated in cellwall modification, defense response, and starch biosynthesis. We also identify a phosphatidylethanolamine-bindingprotein (PEBP), which has been implicated in tuberization signalling in other taxa and, through gene-tree analysis,place this copy in a phylogenytic context. These findings suggest that some similar molecular processes underlie theformation of underground storage structures across flowering plants despite the long evolutionary distances amongtaxa and non-homologous morphologies (e.g., bulbs versus tubers). [Plant development, tuberous roots, comparativetranscriptomics, geophytes]

1 Introduction

Scientific attention in botanical fields focuses almost ex-1

clusively on aboveground organs and biomass. How-2

ever, a holistic understanding of land plant evolution,3

morphology, and ecology requires a comprehensive un-4

derstanding of belowground structures: on average 50%5

of an individual plant’s biomass lies beneath the ground6

(Niklas, 2005), and these portions of a plant are critical for7

resource acquisition, resource storage, and mediating the8

plant’s interactions with its environment. Often, below-9

ground biomass is thought to consist solely of standard10

root tissue, but in some cases, plants modify “ordinary”11

structures for specialized underground functions. Plants12

called geophytes fall toward the extreme end of this be-13

lowground/aboveground allocation spectrum. In a re- 14

markable example of convergent evolution of an innova- 15

tive life history strategy, geophytes retreat underground 16

by producing the buds of new growth on structures be- 17

low the soil surface, while also storing nutrients to fuel 18

this growth in highly modified, specialized underground 19

storage organs (USOs) (Raunkiaer et al., 1934; Dafni et al., 20

1981b,a; Al-Tardeh et al., 2008; Vesely et al., 2011). Many 21

geophytes also have the capacity to reproduce asexually 22

through underground offshoots in addition to sexual re- 23

production. Geophytes are ecologically and economically 24

important, morphologically diverse, and have evolved 25

independently in all major groups of vascular plants ex- 26

cept gymnosperms (Howard et al., 2019, 2020). These 27

plants and their associated underground structures are 28

1

not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted September 20, 2020. ; https://doi.org/10.1101/845602doi: bioRxiv preprint

Page 2: Comparative transcriptomics of a monocotyledonous …Comparative transcriptomics of a monocotyledonous geophyte reveals shared molecular mechanisms of underground storage organ formation

Underground storage organ evolutionary development

a compelling example of evolutionary convergence; di-29

verse taxa produce a variety of structures, often from dif-30

ferent tissues, that serve the analogous function of under-31

ground nutrient storage. However, our understanding of32

the molecular processes that drive this convergence, and33

the extent to which these processes are themselves paral-34

lel, remains limited, due in part to the lack of molecular35

studies in diverse geophyte lineages. This lack of study36

is particularly true for monocotyledonous geophytic taxa,37

which comprise the majority of ecologically and econom-38

ically important geophyte diversity, but have not be sub-39

ject to wide scientific attention beyond a select few crops.40

Some of the world’s most important crop plants have41

underground storage organs, including potato (stem tu-42

ber, Solanum tuberosum), sweet potato (tuberous root,43

Ipomoea batatas), yam (epicotyl- and hypocotyl-derived44

tubers, Dioscorea spp.), cassava (tuberous root, Mani-45

hot esculenta), radish (swollen hypocotyl and taproot,46

Raphanus raphanistrum), onion (bulb, Allium cepa), lotus47

(rhizome, Nelumbo nucifera), various Brassica crops in-48

cluding kohlrabi and turnip (Hearn et al., 2018), and49

more. While several of these crops are well studied and50

have sequenced genomes or other genetic or genomic51

data that may inform the molecular mechanisms under-52

lying underground storage organ development, most de-53

tailed research has focused on a select few, which that54

do not represent the diversity of geophyte morphol-55

ogy, phylogeny, or ecology. Hearn (2006, among oth-56

ers) has proposed that “switches” in existing develop-57

mental programs can explain transitions between major58

growth forms; such a hypothesis requires broad sampling59

across the evolutionary breadth of taxa demonstrating the60

growth form. In particular, most genetic research on geo-61

phytes and their associated underground storage organs62

has been conducted in eudicots such as potato (Hannapel63

et al., 2017), sweet potato (Eserman et al., 2018; Li et al.,64

2019), cassava (Sojikul et al., 2010, 2015; Chaweewan and65

Taylor, 2015), Brassica (Hearn et al., 2018), and Adenia66

(Hearn, 2009). Fewer studies have focused on monocots67

(but see important studies in onion, such as in Lee et al.,68

2013), and these studies focus solely on bulbs; to date no69

study has characterized the molecular underpinnings of70

tuber formation in a monocotyledenous taxon.71

This limited phylogentic breadth is particularly impor-72

tant in light of the findings of Hearn et al. (2018). This73

study provide compelling evidence that within closely74

related Brassica taxa, molecular mechanisms are shared75

between stem and hypocotyl/ root modifications. They76

also demonstrate that these mechanisms have been im-77

plicated in the development of other USOs, namely the78

eudicot lineages potato and sweet potato, and increased79

phylogenetic sampling could confirm and expand these80

findings or suggest that such results are clade-specific.81

Underground storage organs originate from all major82

types of plant vegetative tissue: roots, stems, leaves, and 83

hypocotyls. Bulbs (leaf tissue), corms (stem), rhizomes 84

(stem), and tubers (stem or root) are some of the most 85

common underground storage organ morphologies (Pate 86

and Dixon, 1982), but the full breadth of morphologi- 87

cal variation in USOs includes various root modifications 88

(tuberous roots, taproots, etc.), swollen hypocotyls that 89

merge with swollen root tissue (e.g., Adenia; Hearn, 2009), 90

and intermediate structures such as rhizomes where the 91

terminal end of the rhizome forms a bulb from which 92

aerial shoots emerge (e.g., Iris; Wilson, 2006). Despite this 93

morphological complexity, USOs all develop through the 94

expansion of standard plant tissue, either derived from 95

the root or shoot, into swollen, discrete storage organs. 96

These storage organs also serve similar functions as be- 97

lowground nutrient reserves (Vesely et al., 2011), often 98

containing starch or other non-structural carbohydrates, 99

storage proteins, and water. The functional and physi- 100

ological similarities of underground storage organs may 101

drive or be driven by deep molecular homology with par- 102

allel evolution in the underlying genetic architecture of 103

storage organ development, despite differences in organ- 104

ismal level morphology and anatomy, as is suggested in 105

Hearn et al. (2018). 106

The economic importance of some geophytes and the 107

relevance of understanding the formation of storage or- 108

gans for crop improvement have motivated studies on 109

the genetic basis for storage organ development in se- 110

lect taxa. Potato has become a model system for under- 111

standing the molecular basis of USO development, and 112

numerous studies have demonstrated the complex roles 113

of plant hormones such as auxin, abscisic acid, cytokinin, 114

and gibberellin on the tuber induction process (reviewed 115

in Hannapel et al., 2017). These hormones have been 116

additionally identified in USO formation in other tuber- 117

ous root crops including sweet potato (Noh et al., 2010; 118

Dong et al., 2019) and cassava (Melis and van Staden, 119

1985; Sojikul et al., 2015), in rhizome formation in Panax 120

japonicus (Tang et al., 2019) and Nelumbo nucifera (Cheng 121

et al., 2013b; Yang et al., 2015), and in corm formation 122

in Sagittaria trifolia (Cheng et al., 2013a), suggesting that 123

parallel processes trigger tuberization in both root- and 124

stem-originating USOs. FT-like genes, members of the 125

phosphatidylethanolamine-binding protein (PEBP) fam- 126

ily, have been implicated in USO formation in potato 127

(Navarro et al., 2011; Hannapel et al., 2017), Dendrobium 128

(Wang et al., 2017), Callerya speciosa (Xu et al., 2016), trop- 129

ical lotus (Nelumbo nucifera; (Yang et al., 2015), and onion 130

(Allium cepa; (Lee et al., 2013), indicating either deep ho- 131

mology of FT involvement in USO formation across an- 132

giosperms or multiple independent involvements of FT 133

orthologues in geophytic taxa. 134

The lateral expansion of roots into tuberous roots may 135

be driven by cellular proliferation, by cellular expansion, 136

2

not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted September 20, 2020. ; https://doi.org/10.1101/845602doi: bioRxiv preprint

Page 3: Comparative transcriptomics of a monocotyledonous …Comparative transcriptomics of a monocotyledonous geophyte reveals shared molecular mechanisms of underground storage organ formation

Underground storage organ evolutionary development

or by a combination of these processes. Expansion in137

plant cells requires modification of the rigid plant cell138

wall to accommodate increases in cellular volume (Dolan139

and Davies, 2004; Humphrey et al., 2007), and genes140

such as expansins have been implicated in cellular expan-141

sion during tuberous root development in cassava and142

Callerya speciosa (Sojikul et al., 2015; Xu et al., 2016). Re-143

cent studies of the tuberous roots of sweet potato (Ipo-144

moea batatas) and related species indicate that USO for-145

mation in these taxa involves a MADS-box gene impli-146

cated in the vascular cambium (SRD1; Noh et al., 2010)147

and a WUSCHEL-related homeobox gene (WOX4; Eser-148

man et al., 2018), also involved in vascular cambium de-149

velopment. Additional work on cassava (Manihot escu-150

lenta) also suggests that tuberous root enlargement is due151

to secondary thickening growth originating in the vascu-152

lar cambium (Chaweewan and Taylor, 2015). However,153

geophytes are especially common in monocotyledonous154

plants (Howard et al., 2019, 2020), which lack a vascu-155

lar cambium entirely. No previous study has addressed156

the molecular mechanisms of root tuber development in157

this major clade, so the causes of root thickening are par-158

ticularly enigmatic. Do monocots form tuberous roots159

through genetic machinery that shares deep homology160

with the eudicot vascular-cambium-related pathways, or161

have they evolved an entirely independent mechanism?162

Aerial shootmeristem

Rhizome meristem

Fibrous root tip

Tuberous root tip

Figure 1: Sampling scheme of tissue types. Bomarea multiflora has modi-fied underground stems (rhizomes) and modified roots (tuberous roots).We extracted RNA from the aerial shoot meristem, rhizome meristem,fibrous root tip, and tuberous root tip.

Bomarea multiflora (L. f.) Mirb. is a climbing mono-163

cotyledonous geophyte native to Venezuela, Colombia,164

and Ecuador, where it typically grows in moist cloud165

forests between 1800 − 3800 meters elevation (Hofreiter,166

2008). Bomarea multiflora is an excellent model in which167

to study the molecular mechanisms underlying under-168

ground storage organ formation in the monocots because 169

it has two types of underground modifications: tuber- 170

ous roots and rhizomes. However, prior to this study, 171

no genomic or transcriptomic data was available for any 172

species of Bomarea. Comparative transcriptomics per- 173

mits comprehensive examination of the molecular basis 174

of development, tissue differentiation, and physiology by 175

comparing the genes expressed in different organs, de- 176

velopmental stages, or ecological conditions (Ekblom and 177

Galindo, 2011; Oppenheim et al., 2015). Because no prior 178

genomic or transcriptomic data is needed for compara- 179

tive transcriptomic studies, this method is especially ap- 180

propriate for studies of non-model organisms. 181

In this study, we investigate the molecular mechanisms 182

underlying the formation of tuberous roots in Bomarea 183

multiflora, the first in a monocotyledonous taxon, using a 184

comparative transcriptomics approach and quantify the 185

extent to which these mechanisms are shared across the 186

taxonomic and morphological breadth of geophytic taxa. 187

Specifically, we ask by which developmental mechanisms 188

does the plant modify fibrous roots into tuberous roots: 189

(1) how expansion occurs, (2) when tuberization is trig- 190

gered, and (3) what the tuberous roots store. 191

2 Materials and Methods 192

2.1 Greenhouse and Laboratory Procedures 193

We collected seeds from a single inflorescence of Bomarea 194

multiflora in Antioquia, Colombia [vouchered as Tribble 195

194, deposited at UC (University Herbarium at UC Berke- 196

ley)] and germinated them in greenhouse conditions at 197

the University of California, Berkeley designed to repli- 198

cate native conditions for emphB. multiflora (70◦F− 85◦F 199

and 50% humidity). Six months after germination, we 200

harvested three sibling individuals as biological repli- 201

cates. We dissected a single aerial shoot apical meristem 202

(SAM), the rhizome apical meristem (RHI), root apical 203

meristems (RAM) of several fibrous roots, and the grow- 204

ing tip of a tuberous root (TUB; Figure 1) from each of the 205

three individuals for a total of 12 tissue samples. 206

We immediately froze samples in liquid nitrogen and 207

maintained them at −80◦C until extraction. We extracted 208

total RNA from all samples using the Agilent Plant RNA 209

Isolation Mini Kit (Agilent, Santa Clara, Ca), optimized 210

for non-standard plant tissues, especially those that may 211

be high in starch. Quality of total RNA was measured 212

with Qubit (ThermoFisher, Waltham MA) and Bioana- 213

lyzer 2100 (Agilent Technologies, Santa Clara, CA); if 214

needed, we used a Sera-Mag bead clean-up to further 215

clean extracted RNA (Yockteng et al., 2013). Two SAM 216

samples failed to extract at sufficient concentrations, so 217

we harvested the SAMs of two additional individuals and 218

extracted using the Yockteng et al. (2013) protocol. Sam- 219

3

not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted September 20, 2020. ; https://doi.org/10.1101/845602doi: bioRxiv preprint

Page 4: Comparative transcriptomics of a monocotyledonous …Comparative transcriptomics of a monocotyledonous geophyte reveals shared molecular mechanisms of underground storage organ formation

Underground storage organ evolutionary development

B

C

A

D

A

E

E

A

DD

A

E

PC1: 45% variance

PC

2: 2

2%

va

ria

nce

Tissue Type

Aerial ShootRhizomeFibrous RootTuberous Root

Figure 2: Principal component analysis of VST-transformed transcript counts from all samples. We performed a principal component analysis ofvariance-stabilizing-transformed (VST) transcript counts from all biological replicates of all tissue types. Points are colored by tissue type, andletters correspond to the individual plants sampled.

ples with an RNA integrity (RIN) score >7 proceeded220

directly to library prep. We used the KAPA Stranded221

mRNA-Seq Kit (Kapa Biosystems, Waltham MA) proto-222

col for library prep, applying half reactions with an input223

of at least 500 ng of RNA; however, all but two samples224

had 1 ug of RNA. RNA fragmentation time depended on225

RIN score (7 <RIN <8: 4 min; 8 <RIN <9: 5 min; 9 <RIN:226

6 min). We split samples in half after the second post-227

ligation clean up (Step 10 in the Kapa protocol) in order228

to fine-tune the enrichment step. We amplified the first229

half of the samples with 12 PCR cycles; this proved too230

low and we increased to 15 cycles for the second half of231

the samples. We combined samples and assessed library232

quality with a Bioanalyzer 2100 using the DNA 1000 kit.233

We performed a bead clean-up on libraries showing sig-234

nificant adaptor peaks (Yockteng et al., 2013). We cleaned,235

multiplexed, and sequenced samples on a single lane of236

HiSeq4000 at the California Institute for Quantitative Bio-237

sciences (QB3) Vincent J. Coates Genomics Sequencing238

Lab.239

2.2 Transcriptome Assembly, Annotation,240

and Quantification241

We cleaned, processed, and assembled the raw reads242

using the Trinity RNA-Seq De novo Assembly pipeline243

(Grabherr et al., 2011) under the default settings un-244

less otherwise stated in associated scripts. We ran245

all analyses using the Savio supercomputing resource 246

from the Berkeley Research Computing program at 247

UC Berkeley. We cleaned reads with Trim Ga- 248

lore! (https://www.bioinformatics.babraham.ac.uk/ 249

projects/trim_galore/), keeping unpaired reads and 250

using a minimum fragment length of 36 base pairs. We 251

used all reads from all samples to generate a consen- 252

sus transcriptome, assembled de novo from the concate- 253

nated data using Trinity. We aligned each sample back 254

to the assembled consensus transcriptome using Bowtie2 255

(Langmead and Salzberg, 2012), and quantified isoform 256

abundances using RSEM (RNASeq by Expectation Max- 257

imization; Li and Dewey, 2011). We assessed transcrip- 258

tome assembly quality by comparing the percentages 259

of reads that mapped back to the assembled transcrip- 260

tome using Bowtie2 (Langmead and Salzberg, 2012) to 261

align reads, among other standard metrics. We annotated 262

the consensus transcriptome with a standard Trinotate 263

pipeline (https://trinotate.github.io/), comparing 264

assembled isoforms to SWISS-PROT (Boeckmann et al., 265

2003), RNAmmer (Lagesen et al., 2007), Pfam (Finn et al., 266

2014), eggNOG (Powell et al., 2014), KEGG (Tanabe and 267

Kanehisa, 2012), and Gene Ontology (Gene Ontology 268

Consortium, 2004) databases. We tested the concordance 269

of biological replicates by looking for significant differ- 270

ences between the total number of fragments per repli- 271

cate, by comparing the transcript quantities of all repli- 272

cates to each other, and by checking the correlations be- 273

4

not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted September 20, 2020. ; https://doi.org/10.1101/845602doi: bioRxiv preprint

Page 5: Comparative transcriptomics of a monocotyledonous …Comparative transcriptomics of a monocotyledonous geophyte reveals shared molecular mechanisms of underground storage organ formation

Underground storage organ evolutionary development

tween replicates. Isoforms with fewer than 10 total counts274

were discarded prior to subsequent analyses. We trans-275

formed transcript counts using the variance-stabilized276

transformation (VST) and compared all 12 samples using277

a principal components analysis.278

2.3 Differential Expression279

We identified differentially expressed isoforms (hereafter280

referred to as DEGs) between fibrous (FR) and tuberous281

(TR) roots with the R (R Core Team, 2013) DESeq2 pack-282

age (Love et al., 2014). We used a p-adjusted cut-off (padj,283

using a Benjamini-Hochberg correction for false discov-284

ery rate) of 0.01 and a log2-fold change cut-off of 2 to285

determine statistically significant and sufficiently differ-286

entially expressed isoforms for downstream analyses. To287

test if the distribution of functional annotations for the288

DEGs is statistically different from the overall pool of an-289

notated isoforms, we performed Fisher’s exact tests in R290

(R Core Team, 2013) to compare the distributions of num-291

ber of isoforms associated with (1) biological process, (2)292

molecular function, and (3) cellular component GO anno-293

tations. If the overall distribution of annotations differed,294

we identified the specific GO terms that are enriched in295

the DEG dataset relative to the pool of all annotated iso-296

forms. To identify enriched GO terms, we consider the297

number of isoforms associated with a particular GO cate-298

gory to be drawn from a binomial distribution:299

gi ∼ Binomial(n, pi),

where gi is the number of non-differentially expressed300

isoforms in GO category i, n is the total number of non-301

differentially expressed isoforms, and pi is the probability302

that a given isoform is in GO category i. Thus, the maxi-303

mum likelihood estimator of pi is given by:304

pi = gi/n

Our null hypothesis is that the probability that a given305

transcript is in GO category i is not greater in the pool306

of differentially expressed isoforms than the pool of non-307

differentially expressed isoforms. Under our null, the ex-308

pected number of differentially expressed isoforms asso-309

ciated with GO category i (ki) is also defined by a bino-310

mial distribution:311

ki ∼ Binomial(m, pi),

where m is the total number of differentially expressed312

isoforms and pi is defined above. To test our null hy-313

pothesis, we compute the probability that the observed314

value is greater than ki, conditioning on the probability315

of at least one transcript count per GO category (as GO316

categories with zero differentially expressed transcript317

counts are not represented in the differentially expressed318

dataset) and using a Bonferroni correction (Bonferroni,319

1935) for multiple comparisons.320

2.4 Parallel Processes Across Taxa 321

To complement the broad survey of expression patterns, 322

we additionally identified specific candidate genes, gene 323

families, and molecular processes that might be involved 324

in the development of underground storage organs via a 325

survey of the recent literature on the molecular basis of 326

USO formation. For each group of genes hypothesized 327

to be involved in USO formation (either gene families or 328

molecular/physiological processes), we first queried the 329

annotated transcriptome for isoforms with annotations 330

matching the associated process or family (see Supple- 331

mental Table 2 for the specific search terms used), and 332

then tested if that group was more or less differentially 333

expressed than expected by chance. Specifically, for each 334

focal group of n isoforms, we randomly sampled n iso- 335

forms 10,000 times from the pool of all isoforms. For each 336

of the 10,000 samples, we (1) compared the distribution 337

of absolute log2-fold change values of the sampled iso- 338

forms to the distribution of log2-fold change values of 339

the full dataset and calculated the effect size of a non- 340

parametric Mann–Whitney–Wilcoxon test, generating an 341

expected distribution of effect sizes for a random group of 342

genes of size n, and (2) counted the number of significant 343

DEGs. Under our null model, we expect the effect sizes of 344

the focal group and the randomly sampled groups to be 345

the same. To determine significance of the overall distri- 346

butions of log2-fold change values, we compared the ef- 347

fect size of the focal group to the distribution of simulated 348

effect sizes and determined if the focal group effect size 349

fell within the 95% credible interval of the simulated dis- 350

tribution. Focal effect sizes that were larger than 97.5% of 351

the simulated effect sizes indicated that the focal group is 352

generally more differentially expressed than the null ex- 353

pectation; similarly, focal effect sizes smaller than 97.5% 354

of the simulated effect sizes indicate that the focal group 355

is less differentially expressed than the null expectation. 356

Note that as we performed this analysis on the absolute 357

values of log2-fold changes, less differentially expressed 358

refers to expression levels that are more similar between 359

the focal groups and the null than expected by chance, 360

rather than negative log2-fold change values. Our null 361

model also predicts that the number of significant DEGs 362

in each focal group falls within the 95% credible set of 363

the distributions of number of significant DEGs. We com- 364

pared the number of DEGs in the focal group to distribu- 365

tion of number of DEGs from our 10,000 random sam- 366

ples. We determined significance by identifying groups 367

that contain more or fewer DEGs than the 95% credible 368

set of the simulated distribution. 369

For all targeted candidate genes, we blasted the amino 370

acid sequence of the candidate gene to the assembled con- 371

sensus transcriptome (see Supplemental Table 3 for the 372

blasted sequence specifications) using an e-value cut-off 373

of 0.01 to assess if the identified homologs were differen- 374

5

not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted September 20, 2020. ; https://doi.org/10.1101/845602doi: bioRxiv preprint

Page 6: Comparative transcriptomics of a monocotyledonous …Comparative transcriptomics of a monocotyledonous geophyte reveals shared molecular mechanisms of underground storage organ formation

Underground storage organ evolutionary development

Table 1: Top ten most differentially expressed isoforms (with padj < 0.01) and their corresponding annotations.

Gene Ontology

Transcript ID: AnnotatedName (SPROT)

Log2 FoldChange

Cellular Components Molecular Functions Biological Processes

TRINITY DN116220 c0 g1 i4:Zinc finger CCCHdomain-containingprotein 55

40.25 DNA binding, metal ion binding

TRINITY DN128685 c1 g3 i4:Callose synthase 3

33.85 1,3-beta-D-glucan synthasecomplex, integral component ofmembrane, plasma membrane

1,3-beta-D-glucan synthaseactivity

(1-¿3)-beta-D-glucan biosyntheticprocess, cell wall organization,regulation of cell shape

TRINITY DN121298 c2 g2 i5:Heat shock 70 kDaprotein 15

33.52 cell wall, cytosol, nucleus, plasmamembrane, plasmodesma

ATP binding

TRINITY DN115892 c0 g1 i3:Pre-mRNA-splicing factorATP-dependent RNAhelicase DEAH1

33.52 membrane, spliceosomal complex ATP binding, ATP-dependent3’-5’ RNA helicase activity, RNAbinding

mRNA processing,posttranscriptional gene silencingby RNA, RNA splicing

TRINITY DN127064 c0 g3 i1:LRR receptor-likeserine/threonine-proteinkinase HSL2

33.32 integral component ofmembrane, plasma membrane

ATP binding, proteinserine/threonine kinase activity

defense response toGram-negative bacterium, lateralroot morphogenesis, leafabscission, regulation of geneexpression

TRINITY DN128839 c3 g1 i6:Palmitoyl-acyl carrierprotein thioesterase,chloroplastic

33.01 chloroplast thiolester hydrolase activity fatty acid biosynthetic process

TRINITY DN121430 c10 g2 i1:Sucrose nonfermenting4-like protein

32.64 chloroplast, chloroplast starchgrain, cytoplasm, nucleus

kinase binding, maltose binding,protein kinase activator activity,protein kinase regulator activity,protein serine/threonine kinaseactivity

carbohydrate metabolic process,cellular response to glucosestarvation, mitochondrial fission,peroxisome fission, pollenhydration, proteinautophosphorylation, regulationof protein kinase activity,regulation of reactive oxygenspecies metabolic process

TRINITY DN124527 c1 g1 i5:Bifunctional TH2 protein,mitochondrial

32.64 cytosol, mitochondrion thiaminase activity, thiaminephosphate phosphatase activity

thiamine biosynthetic process

TRINITY DN120224 c4 g1 i1:Enhancer of polycombhomolog 2

32.25 Piccolo NuA4 histoneacetyltransferase complex

DNA repair, histone acetylation,regulation of transcription byRNA polymerase II

TRINITY DN122787 c0 g1 i1:Protein IQ-DOMAIN 32

-32.16 chloroplast envelope, cytosol,microtubule associated complex,nucleus, plasma membrane

response to abscisic acid

tially expressed.375

2.5 PEBP Gene Family Evolution376

We reconstructed the evolutionary history of the377

phosphatidylethanolamine-binding protein (PEBP) gene378

family by combining amino acid sequences from an379

extensive previously published alignment (Liu et al.,380

2016) with the addition of sequences specifically im-381

plicated in USO formation in onion and potato or382

from geophytic taxa such as Narcissus tazetta (accession383

AFS50164.1), Tulipa gesneriana (accessions MG121853,384

MG121854, and MG121855), Crocus sativa (saffron, ac-385

cession ACX53295.1), and Lilium longiflorum (accessions386

MG121858, MG121857, MG121859) (Navarro et al., 2011;387

Tsaftaris et al., 2012; Lee et al., 2013; Li et al., 2013; Leeg-388

gangers et al., 2017) and with copies identified in our389

transcriptome. For copies from B. multiflora, we selected390

the longest isoform per gene to include in the alignment.391

We translated coding sequences from the Bomarea multi-392

flora transcriptome to amino acid sequences using Trans- 393

Decoder v5.5.0 (Haas et al., 2013), removing isoforms 394

that failed to align properly. We aligned amino acids 395

with MAFFT as implemented in AliView v1.18 (Lars- 396

son, 2014), using trimAl v1.4.rev15 (Capella-Gutierrez 397

et al., 2009) with the -gappyout option. We selected the 398

best evolutionary model with ModelTest-NG v0.1.5 (Dar- 399

riba et al., 2016) (JTT+G4 amino acid substitution model 400

(Jones et al., 1992)) and reconstructed unrooted gene trees 401

under maximum likelihood as implemented in IQtree 402

(Nguyen et al., 2014), run on XSEDE using the CIPRES 403

portal (Miller et al., 2010). Using the amino acid align- 404

ment, we compared amino acid residues from Bomarea 405

multiflora orthologs to those that have been functionally 406

characterized in the Arabidopsis FT orthologs (reviewed 407

in Ho and Weigel, 2014). 408

All scripts used in Sections 2.2, 2.3, 2.4, and 2.5 are 409

available on GitHub at (github.com/cmt2/bomTubers) 410

6

not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted September 20, 2020. ; https://doi.org/10.1101/845602doi: bioRxiv preprint

Page 7: Comparative transcriptomics of a monocotyledonous …Comparative transcriptomics of a monocotyledonous geophyte reveals shared molecular mechanisms of underground storage organ formation

Underground storage organ evolutionary development

3 Results411

3.1 Transcriptome Assembly, Annotation,412

and Quantification413

We recovered a total of 359 M paired-end 100 bp reads414

from the single HiSeq 4000 lane for the multiplexed 12415

samples (NCBI BioProject #####). The assembled consen-416

sus transcriptome consists of 370,672 loci, corresponding417

to 224,661 Trinity “genes” (Dryad #####). The consensus418

transcriptome has a GC content of 45.14%, N50 of 1191419

bp, median transcript length of 317 bp, and mean tran-420

script length of 556.95 bp (see also Supplemental Mate-421

rials Section 1.1). Of all reads, 85.54% aligned concor-422

dantly (in a way which matches Bowtie2’s (Langmead423

and Salzberg, 2012) expectation for paired-end reads) to424

the assembled transcriptome, indicating sufficient assem-425

bly quality to proceed with downstream analyses (see426

Supplemental Materials section 1.2). 8.15% of genes had427

at least 10% sequence identity with the UniProt database428

(See Supplemental Material section 1.3, Table 1 and Fig-429

ure 1; Consortium, 2019). Of those, 53.49% blasted with at430

least 80% sequence identity. The Trinotate pipeline anno-431

tated 1.70% of all isoforms. All four tissue types showed432

concordance between the three biological replicates with433

generally 1:1 ratios of transcript quantities to each other434

(see Supplemental Materials section 2: Figures 2 - 5) so435

we proceeded with analyses using data from all three bi-436

ological replicates.437

A principal component analysis (PCA) of the VST tran-438

script counts (Figure 2) shows that the first PC axis (45%439

of the variance in samples) generally explains the varia-440

tion between tissue types. The shoot tissues (SAM and441

RHI) cluster separately, while the root tissues (ROO and442

TUB) cluster together. The underground rhizome sam-443

ples (RHI) fall out intermediate between the aerial shoot444

samples (SAM) and the underground root and tuber sam-445

ples (ROO and TUB) along this axis. The co-clustering446

of fibrous and tuberous root samples in the PCA indi-447

cates that the overwhelming, general pattern of expres-448

sion in all root samples is similar, especially in contrast to449

the very distinct expression profiles of the shoot samples.450

The second PC axis (22% of variance) generally explains451

variance among biological replicates, with Individual A452

particularly distinct from other individuals, perhaps due453

to microhabitat variation in the greenhouse, genotype dif-454

ferences, increased or decreased herbivory compared to455

other individuals.456

3.2 Differential Expression457

We recovered a total of 271 differentially expressed iso-458

forms (DEGs) between fibrous and tuberous roots (FR vs.459

TR). Of these, 226 correspond to regions of the assembled460

consensus transcriptome with functional annotations.461

Of the three types of Gene Ontology (GO) annotations, 462

we recovered significant differences in the distributions 463

of the number of isoforms associated with GO categories 464

between the differentially expressed dataset and the pool 465

of all isoforms for molecular functions (p = 1x10−4) 466

and cellular components (p = 0.031) but not for bio- 467

logical processes (p = 0.921). We recovered no signifi- 468

cantly enriched individual GO annotations for biological 469

processes. For cellular components, we found that cyto- 470

plasm, integral component of membrane, nucleus, and plasma 471

membrane were enriched in the differentially expressed 472

dataset relative to non-differentially expressed isoforms. 473

For molecular functions, we found that ATP binding and 474

metal ion binding were enriched. 475

Of the 271 DEGs, 126 (46.5%) were over-expressed 476

in tuberous roots while the remaining 145 (53.5%) were 477

under-expressed. All top ten most differentially ex- 478

pressed isoforms (the ten DEGs with the highest abso- 479

lute value log2-fold change values between fibrous and 480

tuberous roots) are implicated in cellular and biologi- 481

cal processes (Table 1). All but one of these top ten 482

DEGs are overexpressed in tuberous roots and are gen- 483

erally implicated in nucleotide and ATP binding, cell 484

wall modification, root morphogenesis, and carbohy- 485

drate and fatty acid biosynthesis. The most differen- 486

tially expressed isoform (with a 40.25 log2-fold change), 487

TRINITY DN116220 c0 g1 i4, is a Zinc finger CCCH 488

domain-containing protein 55, a possible transcription 489

factor of unknown function. Other notable top DEGs 490

include TRINITY DN128685 c1 g3 i4, callose synthase 3, 491

which regulates cell shape, TRINITY DN121298 c2 g2 i5, 492

a heat shock protein, TRINITY DN127064 c0 g3 i1, an 493

LRR receptor-like serine implicated in lateral root mor- 494

phogenesis, and TRINITY DN121430 c10 g2 i, a carbo- 495

hydrate metabolism protein. The tenth most differen- 496

tially expressed DEG, under-expressed in tuberous roots, 497

is implicated in abscisic acid signaling. 498

3.3 Parallel Processes Across Taxa 499

Based on our literature survey, we identified 12 groups 500

of genes that have been implicated in USO forma- 501

tion in other geophytes: abscisic acid response genes, 502

calcium-dependent protein kinases (CDPK), expansins, 503

lignin biosynthesis, MADS-Box genes, starch biosynthe- 504

sis, auxin response genes, cytokinin response genes, 14- 505

3-3 genes, gibberellin response genes, KNOX genes, and 506

lipoxygenases (See Figure 3 and Supplemental Table 4). 507

Of these 12 gene groups, six have significantly different 508

expression values than the overall distribution of expres- 509

sion for all isoforms. In four cases (expansins, lignin, 510

MADS-Box, and starch biosynthesis) the overall expres- 511

sion levels are significantly more differentially expressed 512

than expected by chance. Expansins and lignin are gener- 513

ally under-expressed in TR compared to FR while MADS- 514

7

not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted September 20, 2020. ; https://doi.org/10.1101/845602doi: bioRxiv preprint

Page 8: Comparative transcriptomics of a monocotyledonous …Comparative transcriptomics of a monocotyledonous geophyte reveals shared molecular mechanisms of underground storage organ formation

Underground storage organ evolutionary development

Table 2: Differentially expressed isoforms (with padj < 0.01) in specific gene groups and their corresponding annotations.

Log2Fold

ChangeTranscript ID:

Annotated Name (SPROT)

Gene Ontology

Process Group Cellular Components Molecular Functions Biological Processes

Abscisic AcidSignaling

TRINITY DN122359 c1 g1 i3:AMP deaminase

-8.67 cytosol, endoplasmic reticulum, integralcomponent of mitochondrial outermembrane, intracellularmembrane-bounded organelle, nucleus

AMP deaminase activity, ATP binding,metal ion binding, protein histidinekinase binding

embryo development ending in seeddormancy, IMP salvage, response toabscisic acid

Abscisic AcidSignaling

TRINITY DN117636 c1 g2 i8:Dual specificity proteinphosphatase PHS1

-9.32 cytoplasm kinase activity, phosphoproteinphosphatase activity, protein tyrosinephosphatase activity, proteintyrosine/serine/threonine phosphataseactivity

abscisic acid-activated signalingpathway, cortical microtubuleorganization, regulation of geneexpression, regulation of stomatalmovement, response to abscisic acid

Abscisic AcidSignaling

TRINITY DN121543 c6 g2 i1:Probable RNA-bindingprotein ARP1

-10.84 nucleus, ribonucleoprotein complex mRNA binding, RNA binding mRNA processing, regulation of seedgermination, response to abscisic acid,response to salt stress, response to waterdeprivation

Abscisic AcidSignaling

TRINITY DN122705 c2 g1 i2:ACT domain-containingprotein ACR8

-22.91 amino acid binding response to abscisic acid

Abscisic AcidSignaling

TRINITY DN122787 c0 g1 i1:Protein IQ-DOMAIN 32

-32.16 chloroplast envelope, cytosol,microtubule associated complex,nucleus, plasma membrane

response to abscisic acid

Calcium-DependentProtein Kinases

TRINITY DN124121 c3 g1 i11:Calcium-dependent proteinkinase 2

13.05 cytoplasm, nucleus ATP binding, calcium ion binding,calcium-dependent proteinserine/threonine kinase activity,calmodulin binding,calmodulin-dependent protein kinaseactivity

abscisic acid-activated signalingpathway, intracellular signaltransduction, peptidyl-serinephosphorylation, proteinautophosphorylation

CytokininSignaling

TRINITY DN128252 c1 g4 i1:Ferritin-3, chloroplastic

8.22 chloroplast, chloroplast envelope,chloroplast stroma, chloroplastthylakoid membrane, cytoplasm,membrane, mitochondrion, thylakoid

ferric iron binding, ferrous iron binding,ferroxidase activity, identical proteinbinding, iron ion binding

flower development, intracellularsequestering of iron ion, iron iontransport, leaf development,photosynthesis, response to bacterium,response to cold, response to cytokinin,response to hydrogen peroxide,response to iron ion, response to reactiveoxygen species

CytokininSignaling

TRINITY DN124688 c1 g1 i3:Temperature-inducedlipocalin-1

-9.18 chloroplast envelope, chloroplastmembrane, cytoplasm, cytoplasmic sideof plasma membrane, endoplasmicreticulum, Golgi apparatus,mitochondrion, plasma membrane,plasmodesma, vacuolar membrane,vacuole

nutrient reservoir activity, transporteractivity

cellular chloride ion homeostasis,cellular sodium ion homeostasis, heatacclimation, hyperosmotic salinityresponse, lipid metabolic process,positive regulation of response tooxidative stress, positive regulation ofresponse to salt stress, response to cold,response to cytokinin, response tofreezing, response to heat, response tohigh light intensity, response to lightstimulus, response to paraquat, responseto reactive oxygen species, response towater deprivation, seed maturation

CytokininSignaling

TRINITY DN126720 c3 g1 i6:Two-component responseregulator ARR2

-19.35 nucleus DNA binding, DNA-bindingtranscription factor activity,phosphorelay response regulatoractivity

cellular response to cytokinin stimulus,cytokinin-activated signaling pathway,ethylene-activated signaling pathway,leaf senescence, regulation of rootmeristem growth, regulation of seedgrowth, regulation of stomatalmovement, response to cytokinin,response to ethylene, root development

LigninBiosynthesis

TRINITY DN125451 c5 g1 i3:Cinnamoyl-CoAreductase-like SNL6

-20.52 3-beta-hydroxy-delta5-steroiddehydrogenase activity, oxidoreductaseactivity, oxidoreductase activity, actingon the CH-OH group of donors, NAD orNADP as acceptor

defense response to bacterium, ligninbiosynthetic process, steroidbiosynthetic process

MADS-BoxGenes

TRINITY DN127253 c5 g3 i1:MADS-box transcriptionfactor 33

6.60 nucleus DNA-binding transcription factoractivity, protein dimerization activity,RNA polymerase II regulatory regionsequence-specific DNA binding

positive regulation of transcription byRNA polymerase II

Response toGibberellin

TRINITY DN110988 c0 g1 i2:Gibberellin3-beta-dioxygenase 1

-8.35 cytoplasm dioxygenase activity, gibberellin3-beta-dioxygenase activity, metal ionbinding

gibberellic acid mediated signalingpathway, gibberellin biosyntheticprocess, response to gibberellin,response to red light, response to red orfar red light

StarchBiosynthesis

TRINITY DN119512 c6 g1 i4:Granule-bound starchsynthase 1,chloroplastic/amyloplastic

23.73 amyloplast, chloroplast alpha-1,4-glucan synthase activity,glycogen (starch) synthase activity,starch synthase activity

starch biosynthetic process

StarchBiosynthesis

TRINITY DN118583 c0 g1 i7:1,4-alpha-glucan-branchingenzyme,chloroplastic/amyloplastic

22.47 amyloplast, chloroplast 1,4-alpha-glucan branching enzymeactivity, 1,4-alpha-glucan branchingenzyme activity (using a glucosylatedglycogenin as primer for glycogensynthesis), cation binding, hydrolaseactivity, hydrolyzing O-glycosylcompounds

carbohydrate metabolic process,glycogen biosynthetic process, starchbiosynthetic process, starch metabolicprocess

StarchBiosynthesis

TRINITY DN114909 c0 g1 i2:Sedoheptulose-1,7-bisphosphatase,chloroplastic

9.19 apoplast, chloroplast envelope,chloroplast stroma, thylakoid

metal ion binding,sedoheptulose-bisphosphatase activity

defense response to bacterium,reductive pentose-phosphate cycle,starch biosynthetic process, sucrosebiosynthetic process

Box and starch are generally over-expressed in TR com-515

pared to FR. In two cases (14-3-3 and lipoxygenases), the516

overall expression levels are less differentially expressed517

than expected by chance; their expression levels are gen-518

erally conserved between TR and FR. In the remaining six519

cases (abscisic acid, auxin, CDPK, cytokinin, gibberellin,520

and knox) we recover no significant pattern in overall ex-521

pression levels. 522

We identified fifteen individual DEGs in these gene 523

groups (Table 2); interestingly, there seems to be no gen- 524

eralizable relationships between the significance and di- 525

rectionality of a particular group’s distribution with the 526

presence and directionality of expression for individual 527

DEGs. For example, the expression distribution of gib- 528

8

not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted September 20, 2020. ; https://doi.org/10.1101/845602doi: bioRxiv preprint

Page 9: Comparative transcriptomics of a monocotyledonous …Comparative transcriptomics of a monocotyledonous geophyte reveals shared molecular mechanisms of underground storage organ formation

Underground storage organ evolutionary development

Table 3: Specific candidate genes and results from blasting to assembled transcriptome. Asterisk indicates marginal significance. Isoforms foundcorresponds to the number of copies identified in the assembled Bomarea multiflora transcriptome, and Number DE corresponds to the number ofthose copies which are significantly differentially expressed.

Gene Names Original Taxon Isoforms Found Number DE

OsbHLH120 Oryza sativa 22 1*IDD5 Ipomea batatas 63 0WOX4 Ipomea batatas 31 0Sulfite reductase Manihot esculenta 9 0FT-like Solanum tuberosum, Allium cepa 37 1

berellin genes does not deviate significantly from the529

global pool of isoforms, but there is one significantly530

under-expressed DEG in the gibberellin group; similarly,531

the CDPK genes are under-expressed as a group, but532

the only CDPK-related significant DEG is over-expressed533

(Table 2).534

In addition to the gene groups, we identified five spe-535

cific candidate genes from the literature: OsbHLH120536

(qRT9) has been implicated in root thickening in rice (Li537

et al., 2015); IDD5 and WOX4 are implicated in starch538

biosynthesis and TR formation, respectively, in Con-539

volvulaceae (Eserman et al., 2018); sulfite reductase is540

associated with TR formation in Manihot esculenta (So-541

jikul et al., 2010); and FLOWERING LOCUS T (FT) has542

been implicated in signaling the timing of USO forma-543

tion in a variety of taxa, notably Allium cepa and Solanum544

tuberosum (Navarro et al., 2011; Hannapel et al., 2017).545

We recover between nine and 63 putative homologs of546

these candidates (Table 3), but only one is significantly547

differentially expressed (padj < 0.01): a putative FT548

homolog (TRINITY DN129076 c1 g1 i1), further investi-549

gated in the PEBP Gene Family Evolution section (be-550

low). One putative qRT9 homolog is marginally signif-551

icant (padj = 0.050), and the E-value from the BLAST552

result to this isoform was 0.09. Given these marginal sig-553

nificance values, it is likely that the result is spurious and554

we do not follow up with further analysis.555

3.4 PEBP Gene Family Evolution556

We identified thirty-seven Bomarea isoforms as putative557

FLOWERING LOCUS T (FT) homologs. After filtering558

for the longest isoform per gene and filtering out se-559

quences which failed to align properly, ultimately, we in-560

clude five sequences in addition to the significantly dif-561

ferentially expressed copy. We recover three major clus-562

ters in our unrooted gene tree, all with strong bootstrap563

support (Figure 4a); these correspond to the FT cluster,564

TERMINAL FLOWER 1 (TFL1) cluster, and MOTHER565

OF FT AND TFL1 (MFT) cluster recovered in previ-566

ous analyses (Liu et al., 2016). Three of the six Bo-567

marea multiflora isoforms fall out in the FT cluster and568

three in the TFL1 cluster. The Bomarea DEG homolog569

is highly supported in the TFL1 cluster with sequences 570

from other monocot taxa (Figure 4b). Members of this 571

TFL1 clade have been functionally characterized in Oryza 572

sativa, where four orthologs (OsRNC1, OsRNC2, Os- 573

RNC3 and OsRNC14) antagonize the rice ortholog of FT 574

to regulate inflorescence development (Kaneko-Suzuki 575

et al., 2018). mRNA isoforms of all OsRNC are expressed 576

in the root and transported to the SAM. Interestingly, 577

the Crocus sativus ortholog (CsatCEN/TFL1) belongs 578

to the same clade and is also expressed underground 579

(in corms; Tsaftaris et al., 2012). Amino acid analysis 580

of TRINITY DN129076 c1 g1 i1.p1 shows that it conver- 581

gently shares a glycine residue at AthFT position G137, 582

typically characteristic of FT rather than TFL genes (Sup- 583

plemental Figure 8 Pin et al., 2010). FT homologs from Al- 584

lium cepa and Solanum tuberosum that have been function- 585

ally implicated in stem tuber and bulb formation, respec- 586

tively, are in the FT cluster but do not cluster together; 587

rather all USO-implicated PEBP genes are more closely 588

related to non-USO copies than to each other. 589

4 Discussion 590

4.1 How to Make a Tuberous Root 591

Our results suggest potential developmental mechanisms 592

by which the plant modifies fibrous roots into tuberous 593

roots: 1) how expansion occurs, 2) when tuberization is 594

triggered, and 3) what the tuberous roots store. 595

Root expansion likely occurs due to primary thicken- 596

ing growth via cellular expansion B. multiflora. Due to the 597

absence of a vascular cambium, secondary growth is not 598

likely to be involved, despite the prevalence of this mech- 599

anism in other taxa such as sweet potato (Noh et al., 2010; 600

Eserman et al., 2018) and cassava (Melis and van Staden, 601

1985). 602

Cell wall-related genes include pectinesterase TRIN- 603

ITY DN122210 c6 g1 i1 (log2-fold change = 21.91; padj = 604

1.22E-8), which modifies pectin in cell walls leading to 605

cell wall softening, as demonstrated for example in Ara- 606

bidopsis (Braybrook and Peaucelle, 2013). Interestingly, 607

expansins as a group were under- rather than over over- 608

9

not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted September 20, 2020. ; https://doi.org/10.1101/845602doi: bioRxiv preprint

Page 10: Comparative transcriptomics of a monocotyledonous …Comparative transcriptomics of a monocotyledonous geophyte reveals shared molecular mechanisms of underground storage organ formation

Underground storage organ evolutionary development

Figure 3: Differential expression of candidate gene groups. We identified isoforms corresponding to specific pathways and gene families (groupsof genes) and categorize the log2-fold change of expression between tuberous and fibrous roots of those groups. Positive values correspond tooverexpression in tuberous vs. fibrous roots. Box plots correspond to the log2-fold change value for the gene groups and grey points correspond toindividual isoforms. Isoforms that are significantly differentially expressed (padj <0.01) are labeled as red asterisks within the scatter plots. Groupswith absolute value log2-fold change distributions that are significantly larger from the entire dataset (shown in All Genes) are labeled in red onthe X-axis, indicating that these groups are more differentially expressed (either generally up or down in tuberous roots) than expected:expansins,lignin, MADS box, and starch. Groups with less differential expression than expected by chance are labeled in blue: 14-3-3 and lipoxygenases.Groups with significantly more differentially expressed isoforms that expected by chance are labeled with an asterisk on the X-axis: starch.

expressed in B. multiflora, though this is the mechanism609

by which cell expansion occurs in other taxa (see Ex-610

pansins discussion below). Two of the enriched cellu-611

lar component GO categories involve modifications to612

the cell membrane (integral component of membrane and613

plasma membrane), which suggests that modifications to614

the membrane may also be necessary in cellular expan-615

sion.616

Flowering development genes such as TFL genes may617

be involved in mediating environmental signals and in-618

ducing tuber formation. Tuberization signaling may also619

be mediated by callose production, influencing symplas-620

tic signaling pathways through plasmodesmata modifi-621

cation. Callose synthase 3 is one of the most highly differ-622

entially expressed DEGs (TRINITY DN128685 c1 g3 i4,623

Table 1). Callose is a much less common component of624

cell walls than cellulose (Schneider et al., 2016), but it625

is often implicated in specialized cell walls and in root-626

specific expression (Vaten et al., 2011; Benitez-Alfonso627

et al., 2013). Callose synthase has been implicated in the628

development of other unique root-based structures such629

as root nodules (Gaudioso-Pedraza et al., 2018) and muta-630

tions in callose synthase 3 affect root morphology (Vaten631

et al., 2011), suggesting that callose synthase 3 may play632

an integral role in triggering tuberous roots development 633

in B. multiflora through symplastic signaling pathways 634

and/or in modifying cell walls to accomodate expansion. 635

Callose involvment in USO formation has not previously 636

been reported and may be unique to B. multiflora or to 637

monocots. 638

Finally, starch is thought to be the primary nutrient 639

reserve in Bomarea tubers (Kubitzki and Huber, 1998). 640

Many previous studies have found evidence of overex- 641

pression of carbohydrate and starch synthesis molecules 642

in USOs (for example in sweet potato; Eserman et al., 643

2018). Differentially expressed isoforms implicated in 644

the carbohydrate metabolic process support the pres- 645

ence of active starch synthesis in our data. One of the 646

most differentially expressed isoforms is a homolog of 647

sucrose non-fermenting 4-like protein (Table 1, TRIN- 648

ITY DN121430 c10 g2 i1) and participates in carbohy- 649

drate biosynthesis, demonstrating that B. multiflora tubers 650

were actively synthesizing starch when harvested. Ad- 651

ditionally, genes implicated in defense response, such as 652

TRINITY DN127064 c0 g3 i1 (LRR receptor-like serine/ 653

threonine-protein kinase HSL2, Table 1) may be differ- 654

entially expressed in tuberous roots to protect starch re- 655

serves against potential predation by belowground herbi- 656

10

not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted September 20, 2020. ; https://doi.org/10.1101/845602doi: bioRxiv preprint

Page 11: Comparative transcriptomics of a monocotyledonous …Comparative transcriptomics of a monocotyledonous geophyte reveals shared molecular mechanisms of underground storage organ formation

Underground storage organ evolutionary development

0.03

PviFTL1

PviFTL4

ACX53295.1

SitFTL10

SbiFTL14

PdaTFL1c

OsaRCN2

SbiFTL10

ZmaFTL10

PdaTFL1b

ZmaFTL20

MacFTL21

BdTFL1

SbiFTL13

MacFTL20

MacFTL18

MacFTL19

OsaRCN4

ZmaFTL24

PviFTL18

MacFTL17

ZmaFTL14

SitFTL12

BdRCN2

PviFTL6

TRINITY_DN129076_c1_g1_i1.p1

PviFTL19

BdRCN4

TRINITY_DN45900_c0_g1_i1.p1

SbiFTL7

SitFTL15

OsaRCN1

PviFTL17

OsaRCN3

PviFTL16

ZmaFTL22

SitFTL9

4559

72

95

92

34

98

94

94

95

47

89

65

72

35

98

95

60

26

96

30

100

73

97

95

99

74

67

97

70

70

55

34

47

99

88

Setaria italica

Panicum virgatum

Panicum virgatum

Sorghum bicolor

Zea mays

Brachypodium distachyon

Oryza sativa

Setaria italica

Panicum virgatum

Panicum virgatum

Sorghum bicolor

Zea mays

Zea mays

Brachypodium distachyon

Oryza sativa

Oryza sativa

Oryza sativa

Setaria italica

Setaria italica

Panicum virgatum

Panicum virgatum

Zea mays

Sorghum bicolor

Sorghum bicolor

Zea mays

Panicum virgatum

Musa acuminata

Musa acuminata

Musa acuminata

Musa acuminata

Musa acuminata

Bomarea multiflora

Bomarea multiflora

Crocus sativus

Phoenix dactylifera

Phoenix dactylifera

Brachypodium distachyon

TFLMFT

FT100

100

96

90

a b

Figure 4: Evolution of PEBP genes: (a) Unrooted gene tree of 540 PEBP gene copies from across land plants. Stars indicate all included copiesfrom Bomarea multiflora; the red star corresponds to the significantly differentially expressed isoform TRINITY DN129076 c1 g1 i1). Red andarrows indicate PEBP copies that have been implicated in USO formation in other taxa (Allium cepa and Solanum tuberosum). The major clusterscorrespond to the FLOWERING LOCUS T (FT, in purple), TERMINAL FLOWER 1 (TFL1, in yellow), and the MOTHER OF FLOWERING LOCUST AND TERMINAL FLOWER 1 (MFT, in red) gene groups and are labeled with high bootstrap support. (b) Detailed view of the cluster indicatedby a dashed circle in (a), including monocot-specific copies of TERMINAL FLOWER 1 genes. Line thickness corresponds to bootstrap support.)

vores. LRR receptors have been implicated in triggering657

various downstream plant immune responses (Liang and658

Zhou, 2018).659

We emphasize that more detailed work to follow up on660

these aspects of root tuber development should include661

morphological, anatomical, and developmental charac-662

terization of the tuberization process. The integration663

of these methods with genetic and molecular character-664

ization will likely provide important functional links be-665

tween the observed expression patterns we report here666

and the specific effects on plant form and function.667

4.2 Similarities in Molecular Mechanisms of668

USO Formation669

We identify four molecular processes, previously impli-670

cated in USO formation in other taxa, which are either671

over- or under-expressed in the tuberous roots of Bomarea672

multiflora (Figure 3). These processes suggest potential673

parallel function across deeply divergent evolutionary674

distances and in distinct plant structures.675

Expansins are cell wall modifying genes known to676

loosen cell walls in organ formation (Dolan and Davies,677

2004; Humphrey et al., 2007; Braybrook and Peaucelle,678

2013). Their involvement in USO formation has been doc-679

umented in the tuberous roots of cassava (Sojikul et al.,680

2015) and Callerya speciosa (Xu et al., 2016), the rhizomes681

of Nelumbo nucifera (Cheng et al., 2013b), the tuberous682

roots of various Convolvulaceae (Eserman et al., 2018),683

and the stem tubers of potato (Jung et al., 2010). As684

a group, in our data expansins are under-expressed in 685

tuberous compared to fibrous roots, but no individual 686

DEGs are statistically significant. These results suggest, 687

surprisingly, that expansins likely do not play an impor- 688

tant role in tuberous root formation in Bomarea multiflora. 689

We do identify callose synthase 3 as one of the most dif- 690

ferentially expressed genes (Table 1, so it is possible that 691

cell wall modification during tuber development is pri- 692

marily driven by callose rather expansin action in B. mul- 693

tiflora. 694

Lignin biosynthesis genes are under-expressed in sev- 695

eral geophytic taxa with tuberous roots, including cas- 696

sava (Sojikul et al., 2015), wild sweet potato (Ipomoea 697

trifida; (Li et al., 2019), and Callerya speciosa (Xu et al., 698

2016). Similarly, we find that lignin biosynthesis over- 699

all is under-expressed in tuberous compared to fibrous 700

roots, and one isoform in particular is significantly under- 701

expressed: TRINITY DN125451 c5 g1 i3: Cinnamoyl- 702

CoA reductase-like SNL6 (Table 2). This gene has been 703

found to significantly decrease lignin content without 704

otherwise affecting development in tobacco (Chabannes 705

et al., 2001). Decreased lignin in tuberous roots may fur- 706

ther allow for cell expansion and permit lateral swelling 707

of tuberous roots during development. 708

MADS-Box genes are implicated in USO formation in 709

the tuberous roots of wild sweet potato (Ipomoea trifida 710

(Li et al., 2019) and sweet potato (Ipomoea batatas (Noh 711

et al., 2010; Dong et al., 2019), the rhizomes of Nelumbo 712

nucifera (Cheng et al., 2013b), and the corms of Sagittaria 713

trifolia (Cheng et al., 2013a), indicating widespread par- 714

11

not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted September 20, 2020. ; https://doi.org/10.1101/845602doi: bioRxiv preprint

Page 12: Comparative transcriptomics of a monocotyledonous …Comparative transcriptomics of a monocotyledonous geophyte reveals shared molecular mechanisms of underground storage organ formation

Underground storage organ evolutionary development

allel use of MADS-Box genes in the formation of USOs.715

Similarly, we find that MADS-Box genes overall, and one716

DEG in particular, are over-expressed in Bomarea tuber-717

ous roots. MADS-Box genes are implicated widely as718

important transcription factors regulating plant develop-719

ment (Buylla et al., 2000). It is thus unsurprising that720

MADS-Box genes are regularly implicated in USO forma-721

tion. It remains unclear if the MADS-Box genes identified722

in the aforementioned studies represent independent in-723

volvement of MADS-Box genes in USOs from other as-724

pects of plant development, or if they form a clade of725

USO-specific copies. Follow-up phylogenetic analyses of726

these genes could prove informative in determining if727

MADS-Box genes involved in USO development form a728

clade (perhaps indicating a common origin for MADS-729

Box involvement in USOs or molecular convergence) or730

if they fall out independently (perhaps indicating mul-731

tiple events of MADS-box involvement through distinct732

molecular changes).733

Starch biosynthesis genes are very commonly identi-734

fied in the formation of USOs, including in cassava (So-735

jikul et al., 2010, 2015), Nelumbo nucifera (Cheng et al.,736

2013b; Yang et al., 2015), wild and domesticated sweet737

potatoes (Eserman et al., 2018; Li et al., 2019; Dong et al.,738

2019), and potato (Xu et al., 1998). Since starch is so ubiq-739

uitous in USOs, this is unsurprising. We also find starch740

isoforms overall to be over-expressed in Bomarea tuber-741

ous roots, and three genes in particular are significantly742

over-expressed (see Table 2).743

Unexpectedly, two of our candidate gene groups, 14-744

3-3 genes and lipoxygenases, show significantly less dif-745

ferential expression that expected by chance. This result746

suggests that some groups of genes previously identi-747

fied as involved in USO formation may be conserved in748

their expression patterns between tuberous and fibrous749

roots in Bomarea multiflora unlike their differential ex-750

pression patterns in other taxa. The evolutionary ori-751

gin of 14-3-3 and lipoxygenase gene families likely pre-752

dates the divergence of plant groups and the evolution753

of any kind of USO (both gene families are found across754

eukaryotes), implying that their involvement in USO de-755

velopment is perhaps due to neofunctionalization in spe-756

cific plant clade(s). To our knowledge, the gene families’757

involvement in USO development has so far been only758

documented in eudicots, suggesting that this neofunc-759

tionalization may have occurred in the eudicot clade and760

would thus not be documented in monocots, a counter-761

example to parallelism in USO development across land762

plants. Gene tree analysis of these families and the ho-763

mologs implicated in USO development would yield fur-764

ther insight into these potential patterns.765

The other molecular processes we tested failed to show766

group-level differences from the global distribution of767

expression levels. However, the presence of DEGs in768

some of these groups indicates that the phytohormones 769

in particular may play a role in tuberous root formation. 770

One gibberellin response isoform is significantly under- 771

expressed in tuberous roots, which aligns with previ- 772

ous research suggesting that decreased gibberellin con- 773

centrations in roots can lead to root enlargement (Tani- 774

moto, 2012) and tuber formation (Xu et al., 2016; Li et al., 775

2019; Dong et al., 2019). The lack of significant auxin- 776

related isoforms as differentially expressed is surprising, 777

as auxin has been implicated in USO formation in several 778

previous studies (Noh et al., 2010; Cheng et al., 2013b; So- 779

jikul et al., 2015; Yang et al., 2015; Xu et al., 2016; Han- 780

napel et al., 2017; Li et al., 2019; Dong et al., 2019; Ko- 781

lachevskaya et al., 2019), but it is possible this paritcular 782

role of auxin is not part of tuberous root development in 783

monocot taxa, or that it is simply not identified in this 784

study. 785

4.3 PEBP and FT-Like Gene Evolution in 786

Geophytic Taxa 787

Gene tree analysis of PEBPs indicates that copies of these 788

genes have independently evolved several times to sig- 789

nal USO formation in diverse angiosperms (including 790

monocots and eudicots) and in diverse USO morpholo- 791

gies (including tuberous roots, bulbs, and stem tubers). 792

Furthermore, the presence of TFL1 and FT homologs in 793

gymnosperms and other non-flowering plants (Liu et al., 794

2016) suggests that the origin of these genes predates 795

the evolution of the flower, their name notwithstand- 796

ing. Instead, it seems likely that these genes originally 797

evolved as environmental signaling genes with wider in- 798

volvement in triggering the seasonality of various as- 799

pects of plant development. Subsequently, it is possi- 800

ble that gene duplication followed by neofunctionaliza- 801

tion caused many copies in flowering plants to signal 802

flower develepment and other copies to signal USO de- 803

velopment. Given our results, it is possible that USO- 804

specialized FT and TFL1 genes arose at least four times 805

independently, suggesting that broadly parallel molecu- 806

lar evolution may underlie the convergent morphologi- 807

cal evolution of USOs. However, without further work 808

to quantify and characterize the differentially expressed 809

PEBP gene in Bomarea multiflora, it is unclear if this gene 810

is involved in tuberization signalling. Previously verified 811

USO-specialized PEBP genes are clearly FT genes, but the 812

Bomarea multiflora copy is a TFL1 ortholog. In the Beta vul- 813

garis FT ortholog BvFT1, a mutation from glycine to glu- 814

tamine along with two other mutations are sufficient to 815

turn FT into an antagonist of the functional FT orthology 816

BvFT2 (Pin et al., 2010). None of the other TFL1 orthologs 817

in the monocot clade share this mutation. While the B. 818

multiflora ortholog does show signs of molecular conver- 819

gence with FT genes at one residue known to induce TFL- 820

like function in FT genes 8, it is unknown if function re- 821

12

not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted September 20, 2020. ; https://doi.org/10.1101/845602doi: bioRxiv preprint

Page 13: Comparative transcriptomics of a monocotyledonous …Comparative transcriptomics of a monocotyledonous geophyte reveals shared molecular mechanisms of underground storage organ formation

Underground storage organ evolutionary development

covery can occur from TFL to FT and if only one muta-822

tion out of three is sufficient to recover function. Follow823

up studies to characterize the functions of various TFL1824

and FT orthologs in Bomarea multiflora are necessary un-825

derstand the role this DEG plays in tuber development.826

Furthermore, the identification of additional USO-827

specific PEBP genes would shed more light on patterns828

of PEBP family involvement in USOs, but the dearth of829

studies on the molecular basis of USO development im-830

pedes such analyes. With increased sampling, follow-831

up studies could identify unique patterns of convergent832

molecular evolution on the USO-specific FT genes, such833

as tests for selection or further characterization of inde-834

pendently derived subsequences or motifs that could re-835

flect or cause shared function.836

4.4 Conclusions837

We provide the first evidence of the molecular mecha-838

nisms of tuberous root formation in a monocotyledonous839

taxon, filling a key gap in understanding the commonali-840

ties of storage organ formation across taxa. We demon-841

strate that several groups of genes shared across geo-842

phytic taxa are implicated in tuberous root development843

in Bomarea multiflora, patterns which suggest that deep844

parallel evolution at the molecular level may underlie the845

convergent evolution of an adaptive trait. In particular,846

we demonstrate that PEBP genes previously implicated847

in underground storage organ formation appear multiple848

times across the gene tree, and we suggest that a TFL1849

gene in Bomarea multiflora may also be involved in USO850

development. These patterns imply that repeated mor-851

phological convergence may be matched by independent852

evolutions of similar molecular mechanisms. However,853

we also demonstrate two counter examples to this pat-854

tern: groups of genes previously implicated in USO de-855

velopment whose patterns of expression are more simi-856

lar in tuberous and fibrous roots than expected by chance857

(14-3-3 genes and lipoxygenases). These findings sug-858

gest further avenues for research on the molecular mech-859

anisms of how plants retreat underground and evolve860

strategies enabling adaptation to environmental stresses.861

More molecular studies on diverse, non-model taxa and862

more thorough sampling of underground morphological863

diversity will enhance our understanding of the full ex-864

tent of these convergences and add to our general under-865

standing of the molecular basis for adaptive, convergent866

traits.867

4.5 Acknowledgements868

This research used the Savio computational cluster re-869

source provided by the Berkeley Research Computing870

program at the University of California, Berkeley (sup-871

ported by the UC Berkeley Chancellor, Vice Chancellor872

for Research, and Chief Information Officer). We addi- 873

tionally thank Lydia Smith (Evolutionary Genetics Lab, 874

UC Berkeley) for training and sharing her expertise on 875

RNA-Seq, NSF GRFP, SSB, ASPT, Pacific Bulb Society, UC 876

Berkeley’s Integrative Biology Department, and the Tin- 877

ker Foundation for support to CMT, and UC Berkeley 878

CNR and the University and Jepson Herbaria for sup- 879

porting sequencing costs. Michael R. May, David D. Ack- 880

erly, and Benjamin K. Blackman, and two anonymous re- 881

viewers provided feedback on the manuscript. 882

13

not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted September 20, 2020. ; https://doi.org/10.1101/845602doi: bioRxiv preprint

Page 14: Comparative transcriptomics of a monocotyledonous …Comparative transcriptomics of a monocotyledonous geophyte reveals shared molecular mechanisms of underground storage organ formation

Underground storage organ evolutionary development

References883

Al-Tardeh, S., Sawidis, T., Diannelidis, B.-E., and Delivopoulos, S.884

(2008). Water content and reserve allocation patterns within the885

bulb of the perennial geophyte red squill (Liliaceae) in relation to the886

Mediterranean climate. Botany, 86(3):291–299.887

Benitez-Alfonso, Y., Faulkner, C., Pendle, A., Miyashima, S., Helariutta,888

Y., and Maule, A. (2013). Symplastic intercellular connectivity regu-889

lates lateral root patterning. Developmental cell, 26(2):136–147.890

Boeckmann, B., Bairoch, A., Apweiler, R., Blatter, M.-C., Estreicher, A.,891

Gasteiger, E., Martin, M. J., Michoud, K., O’Donovan, C., Phan, I.,892

et al. (2003). The SWISS-PROT protein knowledgebase and its sup-893

plement TrEMBL in 2003. Nucleic acids research, 31(1):365–370.894

Bonferroni, C. E. (1935). Il calcolo delle assicurazioni su gruppi di teste.895

Studi in onore del professore salvatore ortu carboni, pages 13–60.896

Braybrook, S. A. and Peaucelle, A. (2013). Mechano-chemical aspects897

of organ formation in Arabidopsis thaliana: the relationship between898

auxin and pectin. PloS one, 8(3):e57813.899

Buylla, E. R. A., Liljegren, S. J., Pelaz, S., Gold, S. E., Burgeff, C., Ditta,900

G. S., Silva, F. V., and Yanofsky, M. F. (2000). MADS-box gene evo-901

lution beyond flowers: expression in pollen, endosperm, guard cells,902

roots and trichomes. The Plant Journal, 24(4):457–466.903

Capella-Gutierrez, S., Silla-Martınez, J. M., and Gabaldon, T. (2009). tri-904

mAl: a tool for automated alignment trimming in large-scale phylo-905

genetic analyses. Bioinformatics, 25(15):1972–1973.906

Chabannes, M., Barakate, A., Lapierre, C., Marita, J. M., Ralph, J.,907

Pean, M., Danoun, S., Halpin, C., Grima-Pettenati, J., and Boudet,908

A. M. (2001). Strong decrease in lignin content without significant909

alteration of plant development is induced by simultaneous down-910

regulation of cinnamoyl CoA reductase (CCR) and cinnamyl alcohol911

dehydrogenase (CAD) in tobacco plants. The Plant Journal, 28(3):257–912

270.913

Chaweewan, Y. and Taylor, N. (2015). Anatomical assessment of root914

formation and tuberization in cassava (Manihot esculenta Crantz).915

Tropical Plant Biology, 8(1-2):1–8.916

Cheng, L., Li, S., Xu, X., Hussain, J., Yin, J., Zhang, Y., Li, L., and Chen,917

X. (2013a). Identification of differentially expressed genes relevant to918

corm formation in Sagittaria trifolia. PLoS One, 8(1):e54573.919

Cheng, L., Li, S., Yin, J., Li, L., and Chen, X. (2013b). Genome-wide anal-920

ysis of differentially expressed genes relevant to rhizome formation921

in lotus root (Nelumbo nucifera Gaertn). PloS one, 8(6):e67116.922

Consortium, U. (2019). Uniprot: a worldwide hub of protein knowl-923

edge. Nucleic acids research, 47(D1):D506–D515.924

Dafni, A., Cohen, D., and Noy-Mier, I. (1981a). Life-cycle variation in925

geophytes. Annals of the Missouri Botanical Garden, pages 652–660.926

Dafni, A., Shmida, A., and Avishai, M. (1981b). Leafless927

autumnal-flowering geophytes in the Mediterranean re-928

gion—phytogeographical, ecological and evolutionary aspects.929

Plant Systematics and Evolution, 137(3):181–193.930

Darriba, D., Weiß, M., and Stamatakis, A. (2016). Prediction of missing931

sequences and branch lengths in phylogenomic data. Bioinformatics,932

32(9):1331–1337.933

Dolan, L. and Davies, J. (2004). Cell expansion in roots. Current opinion934

in plant biology, 7(1):33–39.935

Dong, T., Zhu, M., Yu, J., Han, R., Tang, C., Xu, T., Liu, J., and Li, Z. 936

(2019). RNA-Seq and iTRAQ reveal multiple pathways involved in 937

storage root formation and development in sweet potato (Ipomoea 938

batatas L.). BMC plant biology, 19(1):1–16. 939

Ekblom, R. and Galindo, J. (2011). Applications of next generation se- 940

quencing in molecular ecology of non-model organisms. Heredity, 941

107(1):1–15. 942

Eserman, L. A., Jarret, R. L., and Leebens-Mack, J. H. (2018). Parallel 943

evolution of storage roots in morning glories (Convolvulaceae). BMC 944

plant biology, 18(1):95. 945

Finn, R. D., Bateman, A., Clements, J., Coggill, P., Eberhardt, R. Y., 946

Eddy, S. R., Heger, A., Hetherington, K., Holm, L., Mistry, J., et al. 947

(2014). Pfam: the protein families database. Nucleic acids research, 948

42(D1):D222–D230. 949

Gaudioso-Pedraza, R., Beck, M., Frances, L., Kirk, P., Ripodas, C., 950

Niebel, A., Oldroyd, G. E., Benitez-Alfonso, Y., and de Carvalho- 951

Niebel, F. (2018). Callose-regulated symplastic communication co- 952

ordinates symbiotic root nodule development. Current Biology, 953

28(22):3562–3577. 954

Gene Ontology Consortium (2004). The Gene Ontology (GO) database 955

and informatics resource. Nucleic acids research, 32(suppl 1):D258– 956

D261. 957

Grabherr, M. G., Haas, B. J., Yassour, M., Levin, J. Z., Thompson, D. A., 958

Amit, I., Adiconis, X., Fan, L., Raychowdhury, R., Zeng, Q., et al. 959

(2011). Full-length transcriptome assembly from RNA-Seq data with- 960

out a reference genome. Nature biotechnology, 29(7):644–652. 961

Haas, B. J., Papanicolaou, A., Yassour, M., Grabherr, M., Blood, P. D., 962

Bowden, J., Couger, M. B., Eccles, D., Li, B., Lieber, M., et al. (2013). 963

De novo transcript sequence reconstruction from RNA-seq using the 964

Trinity platform for reference generation and analysis. Nature proto- 965

cols, 8(8):1494–1512. 966

Hannapel, D. J., Sharma, P., Lin, T., and Banerjee, A. K. (2017). The 967

multiple signals that control tuber formation. Plant Physiology, 968

174(2):845–856. 969

Hearn, D. J. (2006). Adenia (passifloraceae) and its adaptive radia- 970

tion: phylogeny and growth form diversification. Systematic Botany, 971

31(4):805–821. 972

Hearn, D. J. (2009). Descriptive anatomy and evolutionary patterns of 973

anatomical diversification in Adenia Passifloraceae). Aliso: A journal 974

of systematic and evolutionary botany, 27(1):13–38. 975

Hearn, D. J., O’Brien, P., and Poulsen, T. M. (2018). Comparative tran- 976

scriptomics reveals shared gene expression changes during indepen- 977

dent evolutionary origins of stem and hypocotyl/root tubers in Bras- 978

sica (Brassicaceae). PLOS ONE, 13(6):e0197166–25. 979

Ho, W. W. H. and Weigel, D. (2014). Structural features determining 980

flower-promoting activity of arabidopsis flowering locus t. The Plant 981

Cell, 26(2):552–564. 982

Hofreiter, A. (2008). A revision of Bomarea subgenus Bomarea s. str. sec- 983

tion Multiflorae (alstroemeriaceae). Systematic Botany, 33(4):661–684. 984

Howard, C. C., Folk, R. A., Beaulieu, J. M., and Cellinese, N. (2019). The 985

monocotyledonous underground: global climatic and phylogenetic 986

patterns of geophyte diversity. American journal of botany, 106(6):850– 987

863. 988

Howard, C. C., Landis, J. B., Beaulieu, J. M., and Cellinese, N. (2020). 989

Geophytism in monocots leads to higher rates of diversification. New 990

Phytologist, 225(2):1023–1032. 991

14

not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted September 20, 2020. ; https://doi.org/10.1101/845602doi: bioRxiv preprint

Page 15: Comparative transcriptomics of a monocotyledonous …Comparative transcriptomics of a monocotyledonous geophyte reveals shared molecular mechanisms of underground storage organ formation

Underground storage organ evolutionary development

Humphrey, T. V., Bonetta, D. T., and Goring, D. R. (2007). Sentinels at992

the wall: cell wall receptors and sensors. New Phytologist, 176(1):7–21.993

Jones, D. T., Taylor, W. R., and Thornton, J. M. (1992). The rapid genera-994

tion of mutation data matrices from protein sequences. Bioinformatics,995

8(3):275–282.996

Jung, J., O’Donoghue, E. M., Dijkwel, P. P., and Brummell, D. A. (2010).997

Expression of multiple expansin genes is associated with cell expan-998

sion in potato organs. Plant Science, 179(1-2):77–85.999

Kaneko-Suzuki, M., Kurihara-Ishikawa, R., Okushita-Terakawa, C., Ko-1000

jima, C., Nagano-Fujiwara, M., Ohki, I., Tsuji, H., Shimamoto, K., and1001

Taoka, K.-I. (2018). Tfl1-like proteins in rice antagonize rice ft-like1002

protein in inflorescence development by competition for complex1003

formation with 14-3-3 and fd. Plant and Cell Physiology, 59(3):458–468.1004

Kolachevskaya, O. O., Lomin, S. N., Arkhipov, D. V., and Romanov,1005

G. A. (2019). Auxins in potato: molecular aspects and emerging roles1006

in tuber formation and stress resistance. Plant cell reports, 38(6):681–1007

698.1008

Kubitzki, K. and Huber, H. (1998). Flowering plants, monocotyledons: Lil-1009

ianae (except Orchidaceae). Springer.1010

Lagesen, K., Hallin, P., Rødland, E. A., Stærfeldt, H.-H., Rognes, T., and1011

Ussery, D. W. (2007). RNAmmer: consistent and rapid annotation of1012

ribosomal RNA genes. Nucleic acids research, 35(9):3100–3108.1013

Langmead, B. and Salzberg, S. L. (2012). Fast gapped-read alignment1014

with Bowtie 2. Nature methods, 9(4):357.1015

Larsson, A. (2014). AliView: a fast and lightweight alignment viewer1016

and editor for large datasets. Bioinformatics, 30(22):3276–3278.1017

Lee, R., Baldwin, S., Kenel, F., McCallum, J., and Macknight, R. (2013).1018

FLOWERING LOCUS T genes control onion bulb formation and1019

flowering. Nature communications, 4(1):1–9.1020

Leeggangers, H. A., Nijveen, H., Bigas, J. N., Hilhorst, H. W., and Im-1021

mink, R. G. (2017). Molecular Regulation of Temperature-Dependent1022

Floral Induction in Tulipa gesneriana. Plant Physiology, 173(3):1904–1023

1919.1024

Li, B. and Dewey, C. N. (2011). RSEM: accurate transcript quantifica-1025

tion from RNA-Seq data with or without a reference genome. BMC1026

bioinformatics, 12(1):323.1027

Li, J., Han, Y., Liu, L., Chen, Y., Du, Y., Zhang, J., Sun, H., and Zhao, Q.1028

(2015). qRT9, a quantitative trait locus controlling root thickness and1029

root length in upland rice. Journal of experimental botany, 66(9):2723–1030

2732.1031

Li, M., Yang, S., Xu, W., Pu, Z., Feng, J., Wang, Z., Zhang, C., Peng, M.,1032

Du, C., Lin, F., Wei, C., Qiao, S., Zou, H., Zhang, L., Li, Y., Yang, H.,1033

Liao, A., Song, W., Zhang, Z., Li, J., Wang, K., Zhang, Y., Lin, H.,1034

Zhang, J., and Tan, W. (2019). The wild sweetpotato (Ipomoea trifida)1035

genome provides insights into storage root development. BMC Plant1036

Biology, 19(1):119–17.1037

Li, X.-F., Jia, L.-Y., Xu, J., Deng, X.-J., Wang, Y., Zhang, W., Zhang, X.-1038

P., Fang, Q., Zhang, D.-M., Sun, Y., and Xu, L. (2013). FT-Like NFT11039

Gene May Play a Role in Flower Transition Induced by Heat Accu-1040

mulation in Narcissus tazetta var. chinensis. Plant & cell physiology,1041

54(2):270–281.1042

Liang, X. and Zhou, J.-M. (2018). Receptor-like cytoplasmic kinases:1043

central players in plant receptor kinase–mediated signaling. Annual1044

review of plant biology, 69:267–299.1045

Liu, Y.-Y., Yang, K.-Z., Wei, X.-X., and Wang, X.-Q. (2016). Revisit- 1046

ing the phosphatidylethanolamine-binding protein (PEBP) gene fam- 1047

ily reveals cryptic FLOWERING LOCUS Tgene homologs in gym- 1048

nosperms and sheds new light on functional evolution. New Phytolo- 1049

gist, 212(3):730–744. 1050

Love, M. I., Huber, W., and Anders, S. (2014). Moderated estimation of 1051

fold change and dispersion for RNA-seq data with DESeq2. Genome 1052

Biology, 15(12):31–21. 1053

Melis, R. J. and van Staden, J. (1985). Tuberization in Cassava (Manihot 1054

esculenta): Cytokinin and Abscisic Acid Activity in Tuberous Roots. 1055

Journal of plant physiology, 118(4):357–366. 1056

Miller, M. A., Pfeiffer, W., and Schwartz, T. (2010). Creating the cipres 1057

science gateway for inference of large phylogenetic trees. In 2010 1058

gateway computing environments workshop (GCE), pages 1–8. Ieee. 1059

Navarro, C., Abelenda, J. A., Cruz-Oro, E., Cuellar, C. A., Tamaki, S., 1060

Silva, J., Shimamoto, K., and Prat, S. (2011). Control of flowering 1061

and storage organ formation in potato by FLOWERING LOCUS T. 1062

Nature, 478(7367):119–122. 1063

Nguyen, L.-T., Schmidt, H. A., von Haeseler, A., and Minh, B. Q. (2014). 1064

IQ-TREE: A fast and effective stochastic algorithm for estimating 1065

maximum-likelihood phylogenies. Molecular Biology and Evolution, 1066

32(1):268–274. 1067

Niklas, K. J. (2005). Modelling below- and above-ground biomass for 1068

non-woody and woody plants. Annals of Botany, 95(2):315–321. 1069

Noh, S. A., Lee, H.-S., Huh, E. J., Huh, G. H., Paek, K.-H., Shin, J. S., 1070

and Bae, J. M. (2010). SRD1 is involved in the auxin-mediated ini- 1071

tial thickening growth of storage root by enhancing proliferation of 1072

metaxylem and cambium cells in sweetpotato (Ipomoea batatas). Jour- 1073

nal of Experimental Botany, 61(5):1337–1349. 1074

Oppenheim, S. J., Baker, R. H., Simon, S., and DeSalle, R. (2015). We 1075

can’t all be supermodels: the value of comparative transcriptomics 1076

to the study of non-model insects. Insect molecular biology, 24(2):139– 1077

154. 1078

Pate, J. S. and Dixon, K. W. (1982). Tuberous, cormous and bulbous plants. 1079

International Scholarly Book Services Inc.[distributor]. 1080

Pin, P. A., Benlloch, R., Bonnet, D., Wremerth-Weich, E., Kraft, T., Gie- 1081

len, J. J., and Nilsson, O. (2010). An antagonistic pair of ft ho- 1082

mologs mediates the control of flowering time in sugar beet. Science, 1083

330(6009):1397–1400. 1084

Powell, S., Forslund, K., Szklarczyk, D., Trachana, K., Roth, A., Huerta- 1085

Cepas, J., Gabaldon, T., Rattei, T., Creevey, C., Kuhn, M., Jensen, L. J., 1086

von Mering, C., and Bork, P. (2014). eggNOG v4.0: nested orthology 1087

inference across 3686 organisms. Nucleic acids research, 42(Database 1088

issue):D231–9. 1089

R Core Team (2013). R: A Language and Environment for Statistical Com- 1090

puting. R Foundation for Statistical Computing, Vienna, Austria. 1091

Raunkiaer, C. et al. (1934). The life forms of plants and statistical plant 1092

geography; being the collected papers of C. Raunkiaer. The life forms 1093

of plants and statistical plant geography; being the collected papers of C. 1094

Raunkiaer. 1095

Schneider, R., Hanak, T., Persson, S., and Voigt, C. A. (2016). Cellulose 1096

and callose synthesis and organization in focus, what’s new? Current 1097

Opinion in Plant Biology, 34:9–16. 1098

Sojikul, P., Kongsawadworakul, P., Viboonjun, U., Thaiprasit, J., Inta- 1099

wong, B., Narangajavana, J., and Svasti, M. R. J. (2010). AFLP-based 1100

transcript profiling for cassava genome-wide expression analysis in 1101

the onset of storage root formation. Physiologia Plantarum, 140(2):189– 1102

198. 1103

15

not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted September 20, 2020. ; https://doi.org/10.1101/845602doi: bioRxiv preprint

Page 16: Comparative transcriptomics of a monocotyledonous …Comparative transcriptomics of a monocotyledonous geophyte reveals shared molecular mechanisms of underground storage organ formation

Underground storage organ evolutionary development

Sojikul, P., Saithong, T., Kalapanulak, S., Pisuttinusart, N., Lim-1104

sirichaikul, S., Tanaka, M., Utsumi, Y., Sakurai, T., Seki, M., and1105

Narangajavana, J. (2015). Genome-wide analysis reveals phytohor-1106

mone action during cassava storage root initiation. Plant molecular1107

biology, 88(6):531–543.1108

Tanabe, M. and Kanehisa, M. (2012). Using the KEGG database resource.1109

Current protocols in bioinformatics, Chapter 1(1):Unit1.12–1.12.43.1110

Tang, J. R., Lu, Y. C., Gao, Z. J., Song, W. L., Wei, K. H., Zhao, Y., Tang,1111

Q. Y., Li, X. J., Chen, J. W., Zhang, G. H., Long, G. Q., Fan, W., and1112

Yang, S. C. (2019). Comparative transcriptome analysis reveals a1113

gene expression profile that contributes to rhizome swelling in Panax1114

japonicus var. major. Plant Biosystems.1115

Tanimoto, E. (2012). Tall or short? Slender or thick? A plant strategy1116

for regulating elongation growth of roots by low concentrations of1117

gibberellin. Annals of Botany, 110(2):373–381.1118

Tsaftaris, A., Pasentsis, K., Kalivas, A., Michailidou, S., Madesis, P., and1119

Argiriou, A. (2012). Isolation of a CENTRORADIALIS/TERMINAL1120

FLOWER1 homolog in saffron (Crocus sativus L.): characterization1121

and expression analysis. Molecular Biology Reports, 39(8):7899–7910.1122

Vaten, A., Dettmer, J., Wu, S., Stierhof, Y.-D., Miyashima, S., Yadav, S. R.,1123

Roberts, C. J., Campilho, A., Bulone, V., Lichtenberger, R., Lehes-1124

ranta, S., Mahonen, A. P., Kim, J.-Y., Jokitalo, E., Sauer, N., Scheres,1125

B., Nakajima, K., Carlsbecker, A., Gallagher, K. L., and Helariutta, Y.1126

(2011). Callose biosynthesis regulates symplastic trafficking during1127

root development. Developmental Cell, 21(6):1144–1155.1128

Vesely, P., Bures, P., Smarda, P., and Pavlıcek, T. (2011). Genome size1129

and DNA base composition of geophytes: the mirror of phenology1130

and ecology? Annals of Botany, 109(1):65–75.1131

Wang, Y., Liu, L., Song, S., Li, Y., Shen, L., and Yu, H. (2017). DOFT and1132

DOFTIP1 affect reproductive development in the orchid Dendrobium1133

Chao Praya Smile. Journal of experimental botany, 68(21-22):5759–5772.1134

Wilson, C. A. (2006). Patterns in evolution in characters that define Iris1135

subgenera and sections. Aliso, 22(1):425–433.1136

Xu, L., Wang, J., Lei, M., Li, L., Fu, Y., Wang, Z., Ao, M., and Li, Z. (2016).1137

Transcriptome analysis of storage roots and fibrous roots of the tradi-1138

tional medicinal herb Callerya speciosa (Champ.) ScHot. PLOS ONE,1139

11(8):e0160338–20.1140

Xu, X., van Lammeren AA, Vermeer, E., and Vreugdenhil, D. (1998).1141

The role of gibberellin, abscisic acid, and sucrose in the regulation of1142

potato tuber formation in vitro. Plant Physiology, 117(2):575–584.1143

Yang, M., Zhu, L., Pan, C., Xu, L., Liu, Y., Ke, W., and Yang, P. (2015).1144

Transcriptomic analysis of the regulation of rhizome formation in1145

temperate and tropical lotus (Nelumbo nucifera). Nature Publishing1146

Group, 5(1):13059.1147

Yockteng, R., Almeida, A. M. R., Yee, S., Andre, T., Hill, C., and Specht,1148

C. D. (2013). A method for extracting high-quality rna from diverse1149

plants for next-generation sequencing and gene expression analyses.1150

Applications in Plant Sciences, 1(12):1300070–6.1151

16

not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this preprint (which wasthis version posted September 20, 2020. ; https://doi.org/10.1101/845602doi: bioRxiv preprint