You are on page 1of 306

IDENTIFICATION OF GENE REGULATORY NETWORKS CONTROLLING CELL

DIFFERENTIATION IN MAIZE ENDOSPERM

by

Junpeng Zhan

__________________________
Copyright © Junpeng Zhan 2018

A Dissertation Submitted to the Faculty of the

SCHOOL OF PLANT SCIENCES

In Partial Fulfillment of the Requirements

For the Degree of

DOCTOR OF PHILOSOPHY

In the Graduate College

THE UNIVERSITY OF ARIZONA

2018
3

STATEMENT BY AUTHOR

This dissertation has been submitted in partial fulfillment of the requirements for an
advanced degree at the University of Arizona and is deposited in the University Library to be
made available to borrowers under rules of the Library.

Brief quotations from this dissertation are allowable without special permission, provided
that an accurate acknowledgement of the source is made. Requests for permission for extended
quotation from or reproduction of this manuscript in whole or in part may be granted by the head
of the major department or the Dean of the Graduate College when in his or her judgment the
proposed use of the material is in the interests of scholarship. In all other instances, however,
permission must be obtained from the author.

SIGNED: Junpeng Zhan


4

ACKNOWLEDGMENTS

I thank my advisor Ramin Yadegari and my previous advisor Xiangfeng Wang for offering me
the opportunities to work in their laboratories, and their support in the past several years. Ramin
helped me to identify research projects of interest and guided me in almost every aspect of my
graduate studies. I appreciate the time and effort he spent with me to fill the gaps in my
knowledge about plant development and gene regulation. He taught me from the very beginning
how to think conceptually and critically as well as the importance of starting with questions in
doing research and communicating about science with people in talks and writing. I thank
Xiangfeng for teaching me how to handle high-throughput sequencing data, and for encouraging
me to read as many research articles as possible in the first few semesters. I also thank Ramin
and Xiangfeng for their suggestions about my future career.

I thank my dissertation committee members Mark Beilstein, Rebecca Mosher, Ravi Palanivelu,
and Karen Schumaker for their invaluable advice and suggestions. I also thank Eliot Herman for
serving on my comprehensive exam committee. I am grateful to Karen Schumaker and her
postdoc Shea Monihan for teaching me how to approach biological questions.

This entire dissertation is a product of collaboration. I thank all members of the Yadegari Lab,
particularly Guosheng Li, Dhiraj Thakare, Choong-Hwan Ryu, Shanshan Zhang, Dongfang
Wang, Brenda Hunter, and Megan Skaggs, and all members of the Wang Lab, particularly
Chuang Ma, for their training, advice, and the enlightening discussions with them either
individually or at group meetings. Many of them, as well as our collaborators Gary Drews (and
his lab members Alan Lloyd, Neesha Nixon, Angela Arakaki, William Burnett, and Kyle Logan),
Joanne Dannenhoffer (and her lab members Carla Spears and Devon Leroux), Philip Becraft
(and his lab members Hao Wu and Bryan Gontarek), and Brian Larkins, directly contributed
data/materials to my dissertation projects (additional information on my contributions and the
contributions of my co-authors/collaborators to each chapter of this dissertation is provided in
Appendix A). I express my heartfelt thanks to all these colleagues and collaborators for allowing
me to use their biological materials or to include their data in my dissertation. I also thank
undergraduate students of the Yadegari Lab, Ali Shareef, Tricia Ramsay, and Hailey Buell for
their help with various aspects of my research work.
5

I acknowledge the financial support from the China Scholarship Council, the U.S. National
Science Foundation, the Univ. Arizona Graduate College Fellowship, and the Boynton Graduate
Fellowship in Plant Molecular Biology, Univ. Arizona.

Finally, I am indebted to my parents for their unconditional support and encouragement all the
time.
6

TABLE OF CONTENTS

LIST OF FIGURES ........................................................................................................................ 8


LIST OF TABLES ........................................................................................................................ 12
ABSTRACT.................................................................................................................................. 15
CHAPTER 1: ENDOSPERM DEVELOPMENT AND CELL SPECIALIZATION .................. 17
1.1 INTRODUCTION ............................................................................................................................ 17
1.2 COENOCYTE FORMATION AND CELLULARIZATION .......................................................... 18
1.3 PATTERN FORMATION ................................................................................................................ 21
1.4 CELL FATE SPECIFICATION AND DIFFERENTIATION ......................................................... 22
1.5 KEY QUESTIONS AND FUTURE DIRECTIONS ........................................................................ 28
CHAPTER 2: RNA SEQUENCING OF LASER-CAPTURE MICRODISSECTED
COMPARTMENTS OF THE MAIZE KERNEL IDENTIFIES REGULATORY MODULES
ASSOCIATED WITH ENDOSPERM CELL DIFFERENTIATION .......................................... 32
2.1 ABSTRACT ...................................................................................................................................... 32
2.2 INTRODUCTION ............................................................................................................................ 32
2.3 RESULTS ......................................................................................................................................... 35
2.4 DISCUSSION ................................................................................................................................... 49
2.5 METHODS ....................................................................................................................................... 54
2.6 SUPPLEMENTAL DATA ............................................................................................................... 71
CHAPTER 3: OPAQUE-2 REGULATES A COMPLEX GENE NETWORK ASSOCIATED
WITH CELL DIFFERENTIATION AND STORAGE FUNCTION OF MAIZE ENDOSPERM
..................................................................................................................................................... 110
3.1 ABSTRACT .................................................................................................................................... 110
3.2 INTRODUCTION .......................................................................................................................... 111
3.3 RESULTS ....................................................................................................................................... 113
3.4 DISCUSSION ................................................................................................................................. 127
3.5 METHODS ..................................................................................................................................... 132
3.6 SUPPLEMENTAL DATA ............................................................................................................. 152
CHAPTER 4: NAKED ENDOSPERM TRANSCRIPTION FACTORS REGULATE
PARTIALLY OVERLAPPING GENE NETWORKS DURING MAIZE ENDOSPERM
DEVELOPMENT ....................................................................................................................... 193
4.1 ABSTRACT .................................................................................................................................... 193
4.2 INTRODUCTION .......................................................................................................................... 194
4.3 RESULTS ....................................................................................................................................... 196
4.4 DISCUSSION ................................................................................................................................. 205
7

4.5 METHODS ..................................................................................................................................... 208


4.6 SUPPLEMENTAL DATA ............................................................................................................. 220
CHAPTER 5: FUTURE QUESTIONS AND EXPERIMENTS ................................................ 245
APPENDIX A: DISSERTATION AUTHOR’S MAIN CONTRIBUTIONS TO EACH
CHAPTER .................................................................................................................................. 248
APPENDIX B: REPRINT OF “ENDOSPERM DEVELOPMENT AND CELL
SPECIALIZATION” .................................................................................................................. 252
APPENDIX C: REPRINT OF “RNA SEQUENCING OF LASER-CAPTURE
MICRODISSECTED COMPARTMENTS OF THE MAIZE KERNEL IDENTIFIES
REGULATORY MODULES ASSOCIATED WITH ENDOSPERM CELL
DIFFERENTIATION” ............................................................................................................... 268
REFERENCES ........................................................................................................................... 287
8

LIST OF FIGURES

FIGURE 1.1. Early proliferation of maize endosperm….…………………………………….....30

FIGURE 1.2. Cell differentiation in maize endosperm………………………………………31

FIGURE 2.1. Profiles of sequenced RNAs from the captured filial and maternal compartments of
8-DAP maize kernel……………………………………………………………………………...62
FIGURE 2.2 Relationship between the RNA populations obtained from the filial and maternal
compartments of 8-DAP kernel………………………………………………………………….63
FIGURE 2.3. Compartment-specific gene sets identified using the CS scoring method………..64
FIGURE 2.4. Gene coexpression modules detected using WGCNA……………………………65
FIGURE 2.5. De novo GATA-rich sequence motifs over-represented in the upstream sequences
of the coexpression module M18 genes identified by the MEME program……………………..67
FIGURE 2.6. A regulatory module of the MRP-1-regulated GRN based on the analysis of MRP-
1-bound sub-motifs.……………………………………………………………………………...68
SUPPLEMENTAL FIGURE 2.1. Structure of an 8-DAP B73 maize kernel.…………………...71
SUPPLEMENTAL FIGURE 2.2. Representative images of kernel compartments that were marked
and collected for the LCM-RNA-Seq analyses reported in this study……..……………72
SUPPLEMENTAL FIGURE 2.3. Representative quality assessments of LCM-derived RNAs used
in sequencing……………………………………………………………………………….73
SUPPLEMENTAL FIGURE 2.4. Reproducibility of RNA-seq reads for each triplicate of the filial
compartments………………………………………………………………………………74
SUPPLEMENTAL FIGURE 2.5. Distribution profile of RNA-Seq reads along the length of the
gene models……………………………………………………………………………………...75
SUPPLEMENTAL FIGURE 2.6. Profiles of sequenced RNAs from the captured kernel
compartments…………………………………………………………………………………….76
SUPPLEMENTAL FIGURE 2.7. Validation of the compartment-specific expression patterns
using in situ hybridization localization of the mRNAs…………………………………………..77
SUPPLEMENTAL FIGURE 2.8. Z-score plots showing expression profiles of all genes in the
WGCNA-generated coexpression modules M1-M18……………………………………………78
SUPPLEMENTAL FIGURE 2.9. Expression pattern of genes in the endosperm-associated
modules identified by WGCNA based on previously reported RNA-Seq data………………….79
SUPPLEMENTAL FIGURE 2.10. Expression pattern of genes in the endosperm-associated
modules identified by WGCNA based on RNA-Seq data from Chen et al. (2014a)…………….80
9

SUPPLEMENTAL FIGURE 2.11. Expression pattern of genes in the WGCNA coexpression


modules (M1, M2, and M9) associated with both AL and EMB (in comparison to the other two
EMB-associated modules M7 and M8) based on RNA-Seq data from Chen et al. (2014a)…….81
SUPPLEMENTAL FIGURE 2.12. Relationships of new sets of allele-biased genes with the
coexpression modules obtained using WGCNA…………………………………………………82
SUPPLEMENTAL FIGURE 2.13. Enrichment of biological processes in the filial compartment-
correlated modules……………………………………………………………………………….83
SUPPLEMENTAL FIGURE 2.14. Visualization of the five endosperm compartment-correlated
coexpression modules using the VisANT program……………………………………………...84
SUPPLEMENTAL FIGURE 2.15. Yeast one-hybrid assays for binding of MRP-1 to the sequence
motifs listed in Table 2.1………………………………………………………………85
SUPPLEMENTAL FIGURE 2.16. Yeast one-hybrid assays for binding of MRP-1 to the
promoters of the genes listed in Supplemental Table 2.12………………………………………87
SUPPLEMENTAL FIGURE 2.17. Expression pattern of the 93 genes containing MRP-1-binding
sub-motifs based on RNA-Seq data from Chen et al. (2014a)…………………………………..90
FIGURE 3.1. Identification of direct and indirect O2 target genes using RNA-Seq and ChIP-
Seq………………………………………………………………………………………………144
FIGURE 3.2. GO term enrichments of the O2-modulated and/or bound gene sets……………146
FIGURE 3.3. Temporal expression patterns of O2-modulated and/or bound genes during
endosperm development………………………………………………………………………..147
FIGURE 3.4. Full-length O2 and bZIP17 proteins interact and co-activate target gene
expression………………………………………………………………………………………148
FIGURE 3.5. O2 and NKD2 co-regulate a downstream gene network……………………...…149
FIGURE 3.6. Transmission electron microscopy of 20-DAP AL cells in W22o2 and W22…..150
FIGURE 3.7. Models for O2-mediated regulation of endosperm gene expression and
development…………………………………………………………………………………….151
SUPPLEMENTAL FIGURE 3.1. Genetic confirmation of the B72o2 mutant………………...155
SUPPLEMENTAL FIGURE 3.2. Quality assessment of the RNA-Seq data…………………..157
SUPPLEMENTAL FIGURE 3.3. Comparison of O2-modulated genes identified by us and prior
studies…………………………………………………………………………………………..158
SUPPLEMENTAL FIGURE 3.4. Visualization of ChIP-Seq reads and peaks around 5 previously
reported direct targets of O2 using the Integrative Genomics Viewer……………..159
10

SUPPLEMENTAL FIGURE 3.5. Association of the O2-modulated and/or bound gene sets with
previously reported O2-binding sites…………………………………………………………...160
SUPPLEMENTAL FIGURE 3.6. Expression levels of the direct O2 targets in seed tissues in
comparison to vegetative and reproductive tissues based on the available maize expression atlas
(Chen et al., 2014a)……………………………………………………………………………..161
SUPPLEMENTAL FIGURE 3.7. Distribution of mapped reads around the gene models of direct
O2 targets……………………………………………………………………………………….162
SUPPLEMENTAL FIGURE 3.8. Results of EMSAs for binding of O2 to the O2 peaks or promoter
regions associated with direct O2 targets…………………………………………….163
SUPPLEMENTAL FIGURE 3.9. DLR assays for binding of O2 to the O2 peaks associated with
direct O2 targets………………………………………………………………………………...164
SUPPLEMENTAL FIGURE 3.10. Gene numbers and log2FC (mean ± SEM) of zein gene
families/sub-families modulated and/or bound by O2………………………………………….165
SUPPLEMENTAL FIGURE 3.11. Expression patterns of the TF genes directly regulated by O2
based on the available maize expression atlas (Chen et al., 2014a)……………………………166
SUPPLEMENTAL FIGURE 3.12. Enrichments of TF gene families among the O2-modulated
and/or bound gene sets………………………………………………………………………….167
SUPPLEMENTAL FIGURE 3.13. Phylogenetic relationship among all annotated bZIP-family
proteins in maize………………………………………………………………………………..168
SUPPLEMENTAL FIGURE 3.14. Domain mapping auto-activator assay of O2……………..169
SUPPLEMENTAL FIGURE 3.15. Results of directed Y2H assay of protein-protein interaction
between O2 and bZIP17………………………………………………………………………...171
SUPPLEMENTAL FIGURE 3.16. In vitro interaction between O2 and bZIP17……………...172
SUPPLEMENTAL FIGURE 3.17. Schematic of constructs used in DLR assays for gene co-
activation by two TF proteins…………………………………………………………………..173
SUPPLEMENTAL FIGURE 3.18. Transmission electron micrographs of 12- and 16-DAP AL
cells of W22 and W22o2………………………………………………………………………..174
SUPPLEMENTAL FIGURE 3.19. Venn diagram of numbers of O2-activated direct targets
identified by Li et al. (2015) and this study…………………………………………………….175
FIGURE 4.1. Relationship between endosperm mRNA populations of all analyzed mutants and
the WT………………………………………………………………………………………….211
FIGURE 4.2. Differential gene expression analysis……………………………………………213
FIGURE 4.3. Analysis of gene sets cooperatively co-regulated by NKD1 and NKD2………..214
11

FIGURE 4.4. Comparison of the NKD-regulated gene sets with the previously identified O2-
modulated and/or bound gene sets……………………………………………………………...215
FIGURE 4.5. Putative direct target genes of NKD1 and NKD2……………………………….216
FIGURE 4.6. Coexpression networks identified using WGCNA………………………………217
SUPPLEMENTAL FIGURE 4.1. SCC heat map of biological samples based on log2-transformed
RPKM values of all the 34,348 genes tested for differential expression…………220
SUPPLEMENTAL FIGURE 4.2. Scatter plots of log2FC versus average log2CPM of all the genes
tested for differential expression…………………………………………………………221
SUPPLEMENTAL FIGURE 4.3. RT-qPCR validation of DEGs……………………………...222
SUPPLEMENTAL FIGURE 4.4. Analysis of gene sets antagonistically co-regulated by NKD1
and NKD2………………………………………………………………………………………223
SUPPLEMENTAL FIGURE 4.5. Log2FC heat map of NKD-regulated cell cycle control-related
genes……………………………………………………………………………………………224
SUPPLEMENTAL FIGURE 4.6. Log2FC heat map of NKD-regulated storage-protein genes and
Opaque/Floury genes…………………………………………………………………………...225
SUPPLEMENTAL FIGURE 4.7. Log2FC heat map of NKD-regulated carbohydrate biosynthesis
or degradation-related genes……………………………………………………………………226
SUPPLEMENTAL FIGURE 4.8. Log2FC heat map of NKD-regulated carotenoid biosynthesis-
related genes…………………………………………………………………………………….227
SUPPLEMENTAL FIGURE 4.9. Log2FC heat map of NKD-regulated anthocyanin biosynthesis-
related genes…………………………………………………………………………………….228
S U P P LE M E N T A L F IG U R E 4 . 1 0 . Lo g 2 F C h e a t m a p o f N K D - r e g u l a t e d A B A
biosynthesis/degradation/signaling-related genes………………………………………………229
SUPPLEMENTAL FIGURE 4.11. Log2FC heat map of NKD-regulated pathogen defense-related
genes. Asterisks indicate significant differential expression…………………………...230
SUPPLEMENTAL FIGURE 4.12. Relationships of the WGCNA-identified coexpression modules
with the DEG sets……………………………………………………………………..231
SUPPLEMENTAL FIGURE 4.13. Relationships of the coexpression modules with the O2-
modulated and/or bound gene sets……………………………………………………………...232
SUPPLEMENTAL FIGURE 4.14. Enrichments of TF families among the coexpression
modules…………………………………………………………………………………………233
12

LIST OF TABLES

TABLE 2.1. Results of Y1H assays for binding of MRP-1 to the MEME-identified sequence
motifs of M18……………………………………………………………………………………70
SUPPLEMENTAL TABLE 2.1. The origins, quality and quantities of RNAs isolated using laser-
capture microdissection…………………………………………………………………….91
SUPPLEMENTAL TABLE 2.2. Summary statistics of RNA-Seq reads and mapping…………92
SUPPLEMENTAL TABLE 2.3. Number of genes expressed in each of the ten compartments….93
SUPPLEMENTAL TABLE 2.4. Number of genes expressed at different FPKM levels in the ten
compartments…………………………………………………………………………………….94
SUPPLEMENTAL TABLE 2.5. Summary of principal component analysis (PCA) of the ten
compartments…………………………………………………………………………………….95
SUPPLEMENTAL TABLE 2.6. Summary of the available in situ hybridization and promoter
analysis data for endosperm cell-specific genes……………………..…………………………..96
SUPPLEMENTAL TABLE 2.7. Robustness of WGCNA-generated coexpression modules based
on a permutation test of average topological overlap (TO)……………………………………...97
SUPPLEMENTAL TABLE 2.8. Allele-biased genes enriched in the endosperm-associated
coexpression modules M10, M15, M17, and M18………………………………………………98
SUPPLEMENTAL TABLE 2.9. Motif analysis of the endosperm compartment-correlated
modules M10, M12, M15, M17, and M18………………………………………………………100
SUPPLEMENTAL TABLE 2.10. Occurrences of sub-motifs of each motif enriched in upstream
sequences of M18 genes as identified by MEME……………………………………………….101
SUPPLEMENTAL TABLE 2.11. Putative functions of the 93 M18 genes that contain at least one
MRP-1-binding sub-motifs (Motifs 3a/8a, 6a/8b, 8c, and 8f)……………………………...102
SUPPLEMENTAL TABLE 2.12. Results of Y1H assays for binding of MRP-1 to the promoters
of genes within M18……………………………………………………………………………104
SUPPLEMENTAL TABLE 2.13. Sequences of the primers used to generate clones for the in situ
hybridization probes…………………………………………………………………………….105
SUPPLEMENTAL TABLE 2.14. Sequences of the primers used to generate the constructs to test
the sub-motifs in Y1H assays………………………………………………………………106
SUPPLEMENTAL TABLE 2.15. Sequences of the primers used to generate the constructs to test
the promoters in Y1H assays……………………………………………………………….107
SUPPLEMENTAL TABLE 3.1. Opacity of mature kernels from reciprocal crosses between
B73o2, W22o2, and W64Ao2…………………………………………………………………..176
13

SUPPLEMENTAL TABLE 3.2. Summary statistics of RNA-Seq reads and mapping………..177


SUPPLEMENTAL TABLE 3.3. Summary statistics of ChIP-Seq reads and mapping………..178
SUPPLEMENTAL TABLE 3.4. Summary of the previously reported O2-binding sites……...179
SUPPLEMENTAL TABLE 3.5. Positions of EMSA probes as compared to O2 peaks……….180
SUPPLEMENTAL TABLE 3.6. Results of DLR assays of O2-target gene (peak) interaction
shown as fold induction values (mean ± SEM)………………………………………………...181
SUPPLEMENTAL TABLE 3.7. Results of MEME-ChIP analysis of the 401-bp sequences
flanking the summits of peaks associated with the putative direct O2 targets…………………182
SUPPLEMENTAL TABLE 3.8. Results of DLR assays of target-gene (peak) co-activation by O2
and bZIP17 shown as fold induction values (mean ± SEM)………………………………..183
SUPPLEMENTAL TABLE 3.9. Putative direct O2 targets that have previously been predicted as
direct NKD2 targets genes by Gontarek et al. (2016)…………………………………………..184
SUPPLEMENTAL TABLE 3.10. Results of DLR assays of target-gene (peak) co-activation by
O2 and NKD2 shown as fold induction values (mean ± SEM)………………………………...185
SUPPLEMENTAL TABLE 3.11. Sequences of the primers used to generate the EMSA
probes................................................................................................................................186
SUPPLEMENTAL TABLE 3.12. Sequences of the primers used to generate the competitor
sequences in the EMSAs………………………………………………………………………..187
SUPPLEMENTAL TABLE 3.13. Sequences of primers used for RT-qPCR……………….....188
SUPPLEMENTAL TABLE 3.14. Cloning strategies and sequences of primers used for directed
Y2H assays……………………………………………………………………………………...189
SUPPLEMENTAL TABLE 3.15. Cloning strategies and sequences of primers used for BiFC
assays…………………………………………………………………………………………...190
SUPPLEMENTAL TABLE 3.16. Cloning strategies and sequences of primers used for DLR
assays…………………………………………………………………………………………...191
Table 4.1. Abbreviations for genotypes and developmental stages…………………………….219
SUPPLEMENTAL TABLE 4.1. The origins of RNAs and experimental design for RNA-Seq...234
SUPPLEMENTAL TABLE 4.2. Summary statistics of RNA-Seq reads and mapping………..235
SUPPLEMENTAL TABLE 4.3. Normalized Ct values for genes analyzed using RT-qPCR…237
SUPPLEMENTAL TABLE 4.4. Robustness of coexpression modules based on a permutation test
of average topological overlap……………………………………………………………..238
14

SUPPLEMENTAL TABLE 4.5. Functional annotation of genes associated with GO terms


"response to auxin" and/or "auxin-activated signaling pathway" enriched for M6…………….239
SUPPLEMENTAL TABLE 4.6. Bonferroni-corrected P values for enrichment analysis of reported
NKD-binding motifs in 0.5 kb promoter regions determined using AME……………240
SUPPLEMENTAL TABLE 4.7. Results of de novo motif enrichment analysis of the coexpression
modules…………………………………………………………………………..................241
SUPPLEMENTAL TABLE 4.8. Sequences of primers used for RT-qPCR…………………...243
15

ABSTRACT

The endosperm of cereal grains accumulates large amounts of storage compounds such as
carbohydrates and proteins that are essential for seed germination. These compounds also
constitute a principal source of human food, animal feed, and industrial raw material globally.
However, the molecular mechanisms controlling the cell differentiation of cereal endosperm
remain largely unknown. This dissertation has been focused on identification of the gene
regulatory networks associated with the differentiation of basal endosperm transfer layer
(BETL), the starchy endosperm (SE), and the aleurone (AL) of the maize (Zea mays) endosperm,
which are the main cell types involved in nutrient uptake from the maternal tissue and the
synthesis and deposition of storage compounds.

As a first step toward characterizing the gene networks associated with cell differentiation of the
maize endosperm, the mRNAs in five major cell types of the differentiating endosperm and in
the embryo and four maternal compartments of the maize kernel were profiled using RNA
sequencing (RNA-Seq). Comparisons of these mRNA populations revealed the diverged gene
expression programs between filial and maternal compartments. Gene coexpression network
analysis identified coexpression modules associated with single or multiple kernel compartments
including modules for the endosperm cell types. Detailed analyses of a coexpression module
strongly associated with BETL identified a regulatory module activated by the MYB-Related
Protein-1 (MRP-1), a previously reported regulator of BETL differentiation. 1

To understand the interplay between the endosperm cell differentiation programs and the storage
program which is activated immediately after the main cell types become distinguishable, direct
and indirect target genes of Opaque-2 (O2), which is a bZIP family transcription factor (TF) that
has been known as a major regulator of the endosperm storage program, were identified using a
chromatin immunoprecipitation followed by high-throughput sequencing (ChIP-Seq) coupled
with RNA-Seq. The analysis identified 1,863 O2-modulated genes, including 186 putative direct
targets and 1,677 indirect targets, whose known/putative gene functions suggest a broad role for
O2 in the regulation of cell differentiation, storage product deposition, maturation, and stress
response of maize endosperm. Temporal examination of the expression patterns of O2 targets in

1
This paragraph has been published as part of the abstract of the paper in Appendix C.
16

the developing endosperm of an o2 mutant vs. wild type revealed at least two distinct modes of
O2-mediated gene activation. Target-gene co-activation and protein-protein interaction studies
identified two transcription factors—bZIP17 and NAKED ENDOSPERM2 (NKD2)—that are
encoded by O2-activated genes, which can in turn co-activate other O2-network genes with O2.
An ultrastructural analysis of AL cells of an o2 mutant showed that the mutant differs from the
wild type in cell wall thickness and in the abundance and size distribution of lipid bodies and
protein storage vacuoles, suggesting that O2 is required for proper differentiation of AL.

The NKD1 and NKD2 genes encode paralogous INDETERMINATE DOMAIN (IDD) family
TFs that have previously been shown to play critical roles in regulating many endosperm
developmental processes, including the differentiation of AL and SE. To understand the
individual functions of each NKD gene and to gain insights to the gene networks regulating
endosperm development, a time-course transcriptome analysis of endosperm isolated from the
single mutants and a double mutant of NKD1 and NKD2 was performed with comparison to a
wild type using RNA-Seq. The analysis identified 11,554 genes differentially expressed between
the mutants and the wild type. Comparison of genes dysregulated in the single and double
mutants revealed that NKD1 and NKD2 regulate partially overlapping downstream gene sets. A
detailed analysis of the NKD-regulated genes associated with the nkd mutant phenotype
suggested that the double mutant phenotype is primarily associated with the loss of NKD1
function. Nonetheless, the analysis also uncovered a unique role for NKD2 in regulation of a
subset of the mutant-phenotype-associated genes. An analysis of gene coexpression networks
identified gene sets that are regulated by the NKDs in a highly coordinated manner, possibly in
combination with or through regulation of other TFs.

Together, the analyses of the MRP-1, O2, and NKD-regulated gene networks provided
significant insights into the gene regulatory mechanisms underlying differentiation of the maize
endosperm cell types. Because some components of the gene networks identified here are
conserved across certain cereal species as well as eudicots, the resulting data sets have the
potential to guide future improvement of grain yield and quality in maize and other economically
important seed crops.
17

CHAPTER 1

Endosperm development and cell specialization 2

1.1 INTRODUCTION

The endosperm of angiosperms is a seed structure that provides nutrients and signals for embryo
development and seedling germination (Li and Berger, 2012; Olsen and Becraft, 2013). In cereal
crops, it occupies the largest portion of the mature grain, contains large amounts of storage
compounds including primarily carbohydrates and storage proteins, and is an important source of
biofuel (Lopes and Larkins, 1993; Sabelli and Larkins, 2009; FAO, 2015). Because of its value
and relatively large size, maize endosperm has become a model system for studies of endosperm
development.

Angiosperm seed development is initiated by a double fertilization during which one of two
sperm cells fuses with the egg cell within the female gametophyte (embryo sac) to produce the
diploid embryo (1 maternal:1 paternal) and the other fertilizes the central cell to form the triploid
endosperm (2 maternal:1 paternal) (Friedman et al., 2008; Hamamura et al., 2012).
Subsequently, in maize endosperm the nuclei undergo proliferation, creating a coenocyte that
becomes cellularized and then differentiates into at least seven recognizable cell types: the basal
endosperm transfer layer (BETL); aleurone (AL); embryo-surrounding region (ESR); central
starchy endosperm (CSE); subaleurone (SA); conducting zone (CZ); and basal intermediate zone
(BIZ). Concurrent with cell differentiation, the endosperm undergoes two major phases of
mitotic proliferation, an early period that lasts until 8 to 12 days after pollination (DAP) in the
central region, and a late period that continues until 20 to 25 DAP in the outer cell layers (AL
and SA). Starting around 8 to 10 DAP, cells in the central portion of the endosperm gradually
switch from a mitotic cell cycle to endoreduplication and become filled with starch and storage
proteins. They eventually undergo maturation and desiccation (Sabelli and Larkins, 2009;
Becraft and Gutierrez-Marcos, 2012). These developmental events correspond to three important
physiological periods: 1) a lag period (approximately the first two weeks after pollination); 2) a
grain-filling period (from approximately two weeks after pollination until six-seven weeks after

2
This chapter has been published as a chapter in Maize Kernel Development [2017, B.A. Larkins, ed (CABI), pp.
28-43] and is included as a reprint in Appendix B.
18

pollination); and 3) a final period during which grain filling ceases and the kernel matures
(Johnson and Tanner, 1972; Jones et al., 1985). Although most of the kernel dry weight is gained
during the grain-filling period, the preceding lag period is a critical formative phase during
which kernel sink strength and storage capacity are established through cell division and plastid
proliferation (Reddy and Daynard, 1983; Jones et al., 1985; Jones et al., 1996). Moreover, the
length of the lag period has recently been shown to correlate positively with kernel size (Sekhon
et al., 2014). This chapter provides an overview of the early period of maize endosperm
development (i.e., the lag period), with an emphasis on our current knowledge of molecular
mechanisms that regulate endosperm cell differentiation.

1.2 COENOCYTE FORMATION AND CELLULARIZATION

Upon fertilization, the central cell (primary endosperm cell) nucleus undergoes multiple rounds
of division without cytokinesis, forming a multinucleate coenocyte. The coenocyte consists of a
thin layer of cytoplasm surrounding a large central vacuole, and it fills the majority of the
volume of the embryo sac. As coenocytic divisions proceed synchronously, the nuclei spread
along the periphery of central vacuole from the micropylar end (near the embryo) toward the
antipodal end of the embryo sac (Randolph, 1936; Monjardino et al., 2007; Leroux et al., 2014).
In maize endosperm, coenocytic nuclear proliferation takes place within ~3 DAP and results in a
cell containing 128 to 512 nuclei (Figure 1.1A), without a significant increase in the size of the
unfertilized embryo sac (Randolph, 1936; Kiesselbach, 1999; Leroux et al., 2014).
Immunohistochemical staining of microtubules in barley and Arabidopsis show the nuclei are
evenly spaced within the coenocyte by internuclear radial microtubule systems (RMSs) that
emerge around nuclear-cytoplasmic domains (NCD) (Brown et al., 1994; Brown et al., 1999).
The uncoupling of mitosis and cytokinesis contrasts with what occurs in somatic cells, where
mitotic divisions involve formation of a phragmoplast that directs cell plate formation between
the daughter cells (Jurgens, 2005). Likely due to this difference, nuclear divisions in the
coenocytic endosperm proceed faster than in the embryo; by the time the zygote undergoes its
first cell division, endosperm nuclei have already divided two to three times (Randolph, 1936).
Therefore, coenocyte formation could be an evolutionary strategy to rapidly populate the central
cell with nuclei and support rapid mitotic proliferation afterward (Sabelli and Larkins, 2009). In
19

wheat and barley, a short-lived phragmoplast forms between dividing sister nuclei of the
coenocytic endosperm, and cell plates are transiently deposited in wheat, suggesting
phragmoplast function is suppressed after it is initiated in the coenocyte (Brown et al., 1994;
Tian et al., 1998). Therefore, the absence of cell membrane and cell wall synthesis during
coenocytic development could be due to suppression of phragmoplast formation.

Starting ~2.5 to 4 DAP, the coenocyte cellularizes within approximately one day in two
consecutive but distinguishable phases: alveolation and partitioning (Monjardino et al., 2007;
Leroux et al., 2014). Alveolation begins with formation of vesicles that deliver wall materials to
the space between NCDs to form anticlinal cell walls separating peripheral nuclei. This results in
tube-like alveoli that are closed at the end facing the coenocyte wall and open toward the central
vacuole. Subsequently, nuclei in the alveoli divide periclinally to generate a second layer of
nuclei that are displaced inward and separated by deposition of a periclinal wall thus forming an
outer layer of cells and an inner second layer of alveoli (Figure 1.1B) (Brown et al., 1994;
Brown et al., 1999; Monjardino et al., 2007; Leroux et al., 2014). The process of alveolation in
maize proceeds repeatedly until up to 4 layers of nuclei are formed, and is followed by random
cellular partitioning of the central vacuole (Leroux et al., 2014). Cellularization through this later
partitioning appears to be a mechanism unique to maize, as in other cereals cellularization is
completed by alveolation (Brown et al., 1994; Brown et al., 1996a; Brown et al., 1996b; Leroux
et al., 2014).

Evidence from many angiosperm model systems suggests timing of the transition from
coenocytic to cellular proliferation is a key decision in endosperm development, as it has a high
correlation with endosperm/seed size (Li and Berger, 2012; Orozco-Arroyo et al., 2015; Gehring
and Satyaki, 2017). For example, in maize and Arabidopsis, seeds with an excess of the maternal
genome generally show precocious cellularization and reduced endosperm size, while seeds with
paternal genomic excess generally exhibit delayed cellularization and increased endosperm size
(Cooper, 1951; Scott et al., 1998; von Wangenheim and Peterson, 2004; Pennington et al., 2008).
Similar correlation between cellularization timing and endosperm size was also observed in
interspecific crosses within multiple genera (Bushell et al., 2003; Ishikawa et al., 2011; Rebernig
et al., 2015; Garner et al., 2016; Oneal et al., 2016). In nearly all these crosses, with the Oryza
interspecific crosses being the only exception, failure in endosperm cellularization causes seed
20

lethality, and hence establishes interploidy or interspecific hybridization barriers. These


phenomena have been interpreted using the parental conflict theory (also known as the kinship
theory) of genomic imprinting (Haig and Westoby, 1989; Bushell et al., 2003; Haig, 2014).

Genomic imprinting is the allele-biased expression of genes in a parent-of-origin-dependent


manner (Abramowitz and Bartolomei, 2012; Peters, 2014). In plants, imprinting is observed
predominantly in the endosperm (Gehring, 2013; Rodrigues and Zilberman, 2015). The parental
conflict theory hypothesizes that in situations where a mother contributes nutritional resources to
offspring of multiple different fathers, the paternally inherited genes foster uptake of as many
nutrients as possible to increase fitness of the fathers-derived progeny, whereas the maternal
genes tend to evenly distribute nutrition to all offspring (Haig and Westoby, 1989; Haig, 2014).
According to this theory, maternally-expressed imprinted genes (MEGs) tend to limit endosperm
growth (size), while paternally-expressed genes (PEGs) tend to promote it. An extensive set of
studies in Arabidopsis have implicated chromatin-regulatory mechanisms, including many
histone modification- and DNA methylation-related processes in regulation of cellularization and
seed size (Li and Berger, 2012; Gehring and Satyaki, 2017). Mutations in components of the
Fertilization-Independent Seed (FIS)-Polycomb Repressive Complex 2 (PRC2), a putative
H3K27 methyltransferase, has been shown to result in endosperm cellularization failure,
increased coenocytic proliferation, and ultimately seed abortion (Li and Berger, 2012; Gehring
and Satyaki, 2017). The gene networks regulated by the FIS-PRC2 complex during early
endosperm development are beginning to be deciphered functionally. Among its direct targets is
the type I MADS-box family transcription-factor (TF) gene AGAMOUS-LIKE62 (AGL62);
mutations in AGL62 result in precocious endosperm cellularization and reduced seed size (Kang
et al., 2008; Hehenberger et al., 2012). Arabidopsis DNA METHYLTRANSFERASE1 (MET1)
is responsible for maintaining DNA (CpG) methylation (Finnegan and Dennis, 1993). Global
hypomethylation of the paternal genome in anti-sense or loss-of-function met1 mutants delayed
endosperm cellularization and increased seed size, while hypomethylation of the maternal
genome has opposite effects (Adams et al., 2000; Xiao et al., 2006). These observations support
a model in which genomic imprinting contributes to regulation of endosperm cellularization, and
establishment of interploidy/interspecific hybridization barriers. This concept has recently been
reinforced by the observations that paternal inheritance of mutations in some PEGs (ADMETOS,
21

SUVH7, PEG2 and PEG9) can lead to partial normalization of the timing of endosperm
cellularization in interploidy hybridizations (Kradolfer et al., 2013; Wolff et al., 2015).

Genetic studies in Arabidopsis have also uncovered a maternal or sporophytic contribution to the
regulation of endosperm cellularization and seed size based on the analysis of the HAIKU (IKU)
pathway genes (IKU1, IKU2, MINISEED3, and SHORT HYPOCOTYL UNDER BLUE1), and the
AP2 family TF gene APETALA2 (AP2) (Li and Berger, 2012; Orozco-Arroyo et al., 2015). Loss
of function of individual IKU pathway genes results in precocious endosperm cellularization and
reduced seed size, while loss of function of AP2 results in delayed endosperm cellularization and
increased seed size (Orozco-Arroyo et al., 2015). How these sporophytic genes regulate
endosperm development remains unclear; it is likely some signaling mechanisms mediate
interactions between the three major seed components, the seed coat, the embryo, and the
endosperm, with the latter perhaps acting as a central hub for regulatory activities that coordinate
interactions with the other two components (Berger et al., 2006; Orozco-Arroyo et al., 2015).
Available data on the genetic and epigenetic players in Arabidopsis support a key role for
cellularization in proper endosperm development—and by extension in proper development of
the whole seed. How this key step occurs and the nature of the gene networks that regulate the
transition from coenocytic to cellular proliferation remain to be determined. In maize, the nature
of such regulatory processes is even less well understood. Genes encoding a number of
chromatin-modifying enzymes, including MET1 and two FIS-PRC2-component-related proteins,
have been identified (Steward et al., 2000; Danilevskaya et al., 2003; Haun et al., 2007);
however, their roles, if any, in regulation of endosperm cellularization are elusive.

1.3 PATTERN FORMATION

Endosperm patterning proceeds in two main phases. The first establishes the micropylar and
chalazal domains of the endosperm along the micropylar-chalazal (MC) axis, while the second
results in differentiated cell types that are arranged according to radial symmetry (Costa et al.,
2004; Li and Berger, 2012). The first phase is reflected by the reported micropylar/embryo-
surrounding-region-specific gene expression patterns in the endosperm of Arabidopsis and maize
(Opsahl-Ferstad et al., 1997; Bonello et al., 2000; Tanaka et al., 2001; Baud et al., 2005; Ingouff
et al., 2005; Yang et al., 2008), and the distinct mitotic domains within the coenocytic endosperm
22

of Arabidopsis (Boisnard-Lorig et al., 2001; Brown et al., 2003). Mutations in genes encoding
components of Arabidopsis FIS-PRC2 display female-gametophytic defects and lead to ectopic
chalazal endosperm development (Sorensen et al., 2001; Guitton et al., 2004), suggesting female-
gametophytic or maternally-encoded functions are required for establishing the early MC
endosperm axis. In maize, the maternal gametophytic mutation baseless1 (bsl1) that disrupts
central cell polarity also alters the spatial expression patterns of BETL marker genes in both
coenocytic and cellularized endosperm (Gutierrez-Marcos et al., 2006), suggesting the regulatory
program controlling basal endosperm patterning is already active, at least partially, in the
unfertilized central cell.

Two cytokinin biosynthetic genes [ISOPENTENYL TRANSFERASE-4 (IPT-4) and IPT-8] were
identified as predominantly expressed in the chalazal region of the coenocytic endosperm of
Arabidopsis, indicating a potential role of polarized cytokinin localization in MC axis patterning
(Li et al., 2013). Interestingly, cytokinin has also been shown to be critical for the establishment
of the MC axis that pre-exists in the female gametophyte (Tekleyohans et al., 2017). Whether
this axis influences formation of the corresponding MC axis in the endosperm is unknown;
however, the extensive similarities between the female gametophyte and endosperm
developmental patterns, including the sequence of coenocytic profileration and cellularization,
suggest both female gametophyte development and endosperm development are regulated by an
ancient genetic program (Olsen et al., 1999). Therefore, elucidation of the role of cytokinin is
critical to understanding endosperm pattern formation and could provide insight into its
evolutionary history.

1.4 CELL FATE SPECIFICATION AND DIFFERENTIATION

The second phase of endosperm patterning involves cell fate specification of epidermal and inner
cells, which is described below in conjunction with other cell differentiation events. Starting
around 4 DAP, maize endosperm differentiates into four specialized compartments or cell types
that become cytologically distinguishable by around 6 DAP, namely the BETL, AL, SE, and
ESR (Figure 1.1C and Figure 1.2A). These main cell types then become further differentiated
(by ~8 DAP), leading to the emergence of three additional recognizable cell types: the BIZ, CZ,
23

and SA (Figure 1.2B). Although some of these individual cell types have not been characterized
fully or associated with any specialized biological function, each of the resulting seven cell types
occupies a specific territory within the early endosperm and possesses a unique set of cytological
characteristics and gene expression programs (Leroux et al., 2014; Zhang et al., 2015). The
BETL and AL are discussed in detail elsewhere (Chourey and Hueros, 2017; Gontarek and
Becraft, 2017). Here we focus primarily on cell differentiation events associated with the main
cell types in the context of their individual functions (if known) and their overall contribution to
the development of the endosperm as a storage compartment.

The BETL contains a single layer of cells that is located at the base of the endosperm in direct
contact with the maternal placento-chalazal zone (PC) (Figure 1.2). The main function of the
BETL is transport of nutrients from maternal tissue into the SE (Gunning and Pate, 1969; Pate
and Gunning, 1972; Shannon et al., 1986). BETL cell fate is likely specified during coenocytic
proliferation of the endosperm. In barley, mRNA of the BETL marker gene, END1, is localized
at the basal region of the coenocytic endosperm (Doan et al., 1996). Consistent with this, the
maize END1 ortholog, named BETL-9, also exhibits a basal endosperm-specific mRNA
accumulation pattern in the coenocytic and cellularized endosperm (Gutierrez-Marcos et al.,
2006; Royo et al., 2014). The phenotype of the maize globby-1 (glo-1) mutant supports the
hypothesis that BETL cell fate is already specified during coenocytic proliferation (Costa et al.,
2003). In this mutant, BETL differentiation is disrupted to variable extents and expression of
BETL marker genes is reduced at the coenocytic phase. Interestingly, the glo-1 mutant also
shows abnormal endosperm cellularization in the basal region, suggesting the GLO-1 gene plays
a broad role in both early cell proliferation and differentiation of the endosperm (Costa et al.,
2003).

Extensive evidence indicates BETL differentiation is regulated genetically and epigenetically. A


detailed set of studies recently sought to uncover the gene regulatory networks (GRNs) of the
BETL, with a particular focus on MYB-Related Protein-1 (MRP-1), a MYB-related family
member that plays a key role in BETL differentiation (Yuan et al., 2016; Chourey and Hueros,
2017; Doll et al., 2017). MRP-1 directly regulates numerous BETL-expressed genes, including
the MATERNALLY EXPRESSED GENE-1 (MEG-1) gene that encodes a small cysteine-rich
peptide (Zhan et al., 2015). The MEG-1 gene itself was shown to be necessary and sufficient for
24

regulation of BETL differentiation (Costa et al., 2012). MEG-1 is an imprinted gene that shows a
maternally-biased expression pattern in endosperm around 4 to 6 DAP (Gutierrez-Marcos et al.,
2004), strongly suggesting BETL differentiation is also epigenetically regulated. In support of
this notion, the 2 maternal:1 paternal genomic ratio was shown to be critical for BETL
differentiation, as both maternal and paternal genomic excess disrupt the process to varying
extents (Charlton et al., 1995; Pennington et al., 2008). Nonetheless, the nature of MRP-1- and
MEG-1-mediated regulation of BETL differentiation remains to be determined.

BIZ and CZ are two endosperm cell types located immediately central to the BETL and
differentiate later than BETL (Figure 1.2B). The CZ contains highly elongated cells (~3.5 times
longer than wide) with tapering end walls, sparse cytoplasm, and large nuclei (Leroux et al.,
2014). The CZ cells are believed to transport nutrient solutes throughout the endosperm (Becraft,
2001), but different from the BETL, the CZ cells lack cell wall ingrowths and distinct vacuoles
(Leroux et al., 2014). The BIZ contains 2 to 4 layers of cells that line between the BETL and the
CZ (Leroux et al., 2014). These cells are often grouped as part of the BETL (Sabelli and Larkins,
2009; Olsen and Becraft, 2013; Yuan et al., 2016), but increasing evidence suggests that the BIZ
cells constitute a unique cell type in the endosperm. The BIZ cells show intermediate, but largely
different, characteristics from the BETL and CZ in terms of cell elongation, the extent of cell
wall ingrowth, cytoplasmic density, and nuclear size, supporting their distinction as a unique cell
type (Monjardino et al., 2013; Leroux et al., 2014). Accordingly, recent mRNA in situ
hybridization assays detected mRNAs localized exclusively in the BETL or BIZ, as well as
mRNAs that are preferentially localized in both BETL and BIZ at 6 to 8 DAP (Li et al., 2014).
Because the differentiation of CZ and BIZ appears to begin later than in the four main cell types,
and they are related to the main cell types either clonally or functionally, the CZ and BIZ can be
viewed as specialized sub-regions of the main cell types. The delayed timing of their
differentiation relative to the BETL could be related to a need for the ever-increasing rate of
movement of photo-assimilates through the developing BETL for utilization and storage in the
inner endosperm cells, including the SE.

The AL is the peripheral layer of cells that covers the entire surface of the endosperm, except the
BETL and ESR regions (Figure 1.2). During endosperm development, the AL stores proteins,
lipids, non-starch carbohydrates, and mineral nutrients. During seed germination, the primary
25

role of the AL is production of hydrolytic enzymes to utilize storage proteins, nucleic acids, and
carbohydrates stored in the SE (Becraft, 2007; Becraft and Yi, 2011). As such, AL is the only
endosperm cell type that remains alive when the other compartments of the endosperm mature
and undergo programmed cell death (Young et al., 1997; Young and Gallie, 2000). The AL and
BETL constitute a continuous cell layer, but have distinct morphological and cytological
characteristics (Chourey and Hueros, 2017; Gontarek and Becraft, 2017). The cell fate of AL
also seems to be controlled by a regulatory program independent from that for the BETL,
because endosperm in the maize defective kernel1 (dek1) mutant lacks an AL but has a normal
BETL (Becraft et al., 2002; Lid et al., 2002).

A series of genetic experiments showed the AL and SE share a common cell lineage, and they
exhibit interchangeable cell fates regulated in response to positional cues (Becraft and Asuncion-
Crabb, 2000; Becraft et al., 2002; Lid et al., 2002). These observations indicate that AL and SE
are relatively plastic and maintain their differentiated states by sensing and responding to
environmental cues throughout development. A number of genes are known to be involved in
regulation of AL cell fate and/or differentiation (Gontarek and Becraft, 2017). These genes and
the corresponding mutants are beginning to provide a clearer picture of AL differentiation and its
relation to the underlying SA and CSE cells.

The SE is the major endosperm cell type that accumulates starch, DNA and storage proteins. It
occupies the largest, central portion of the mature grain and represents the bulk of the endosperm
mass. The SE cells are derived by differentiation of the centrally-localized cells formed after
cellularization and periclinal divisions of the AL/SA cells (Morrison et al., 1975; Lending and
Larkins, 1989; Becraft and Asuncion-Crabb, 2000). Late in differentiation, as the BIZ, CZ, and
SA cell types (here considered sub-regions of SE) become visible, the central region of the SE
can be further delineated and is termed CSE (Figure 1.2B). This region contains variably sized
cells and nuclei that increase in size from near the SA toward the center (Leroux et al., 2014).
The adjacent SA cells are filled by large vacuoles, mitochondria, and proplastids, but the
cytoplasm is slightly less dense than that of the AL cells, and only a small number of protein
bodies and starch grains are present (Khoo and Wolf, 1970; Lending and Larkins, 1989; Leroux
et al., 2014).
26

Due to a lack of mutants that specifically alter SE cell fate, little is known about how it is
controlled. Given the reversibility in cell fates of SE and AL, as discussed above, mutants with
interchangeable SE to AL phenotypes would be valuable for understanding the molecular
mechanisms controlling SE cell fate. Such mutants possibly exist in uncharacterized dek or
empty pericarp mutant collections (Neuffer and Sheridan, 1980; Sheridan and Neuffer, 1980;
Scanlon et al., 1994). By contrast, more is known about regulation of the storage programs of the
SE and AL, and in recent years this knowledge is beginning to shed light on SE/AL
differentiation. Expression of storage-protein genes has been detected in both SE and AL at the
mRNA and protein levels (Reyes et al., 2011). The bZIP family TF protein, Opaque-2 (O2), and
the DOF family TF, Prolamin Box-binding Factor (PBF), are two of the major regulators of the
storage program (Kawakatsu and Takaiwa, 2010; Thompson and Verdier, 2012). The naked
endosperm (nkd) mutant and DOF3 RNA interference (RNAi) knockdown lines that exhibit
defects in AL cell fate and differentiation also show altered storage product accumulation, and
both NKD1 and NKD2 directly regulate storage-protein gene expression (Yi et al., 2015;
Gontarek et al., 2016; Qi et al., 2016). Furthermore, both O2 and PBF are down-regulated in AL
of the nkd mutant (Gontarek et al., 2016). These findings indicate extensive interplay between
the gene regulatory programs that control SE/AL cell fate and differentiation, and the programs
controlling the storage function of SE and AL. Therefore, characterization of the GRNs regulated
by the NKDs, O2, and PBF is expected to provide valuable insight into the regulation of AL/SE
cell fate and differentiation.

The ESR consists of multiple layers of small, densely cytoplasmic and thin-walled cells that
surround the endosperm cavity where the embryo develops (Schel et al., 1984; Opsahl-Ferstad et
al., 1997; Leroux et al., 2014). Early in its differentiation, the ESR surrounds the entire embryo.
As differentiation proceeds, ESR cells are pushed away and crushed by the developing embryo
and remain exclusively in regions around the embryo suspensor (Figure 1.2) (Leroux et al.,
2014). This morphological change of the ESR is supported by mRNA localization patterns and
promoter activities of the ESR marker genes, ESR-1, 2, and 3 (Opsahl-Ferstad et al., 1997;
Bonello et al., 2000). In both assays, marker gene expression detected in the endosperm region
surrounding the entire embryo at 5 DAP becomes restricted to a small region surrounding the
embryo suspensor by 7 to 9 DAP, and to only the base of the suspensor by 12 to 15 DAP.
27

Increasing evidence in maize and Arabidopsis suggests the ESR (the analogous region in
Arabidopsis is referred to as the micropylar endosperm) is involved in nurturing and defense of
the embryo, and also mediates signaling between the endosperm and embryo. The nutritive
function is in part suggested by the finding that an invertase inhibitor gene (INVINH-1) is
expressed in the maize ESR and likely functions in modulating invertase activity to regulate
sugar transport into the embryo (Bate et al., 2004). The ESR-specific expression patterns of two
maize genes, the ANDROGENIC EMBRYO-3 (AE-3) gene encoding a small hydrophilic protein
with structural similarity to the BETL-expressed BAPs mentioned above (Magnard et al., 2000;
Sevilla-Lecoq et al., 2003), and ESR-6, encoding a defensin-like protein (Balandin et al., 2005),
suggest these gene products function within the ESR in defense of the embryo. The role of the
ESR in signaling between the endosperm and the embryo is supported by multiple lines of
evidence in both Arabidopsis and maize. The Arabidopsis ZHOUPI (ZOU) gene is expressed
predominantly in the ESR, and loss of function of ZOU results in retarded endosperm breakdown
and impaired epidermal development of the embryo. The latter indicates a critical role of the
ESR region in control of embryo epidermal development through a signaling pathway (Yang et
al., 2008; Xing et al., 2013). RNAi knockdown of ZOU in maize endosperm also results in
retarded breakdown of the ESR and the adjacent suspensor (Grimault et al., 2015). In addition,
the Arabidopsis EMBRYO SURROUNDING FACTOR 1 (ESF1) accumulates in the central cell
before fertilization and in the ESR after fertilization, and is required for early embryo patterning,
likely through a non-cell-autonomous pathway (Costa et al., 2014). These observations support a
conserved linkage between embryonic and endosperm cell fates in both eudicots and monocots.

The mechanisms controlling ESR cell fate specification are unknown, but the ESR-1, 2, and 3
genes are possibly involved in the process. These genes encode small secreted proteins that show
partial homology to Arabidopsis CLAVATA3 (CLV3) (Cock and McCormick, 2001; Bonello et
al., 2002), a signaling peptide that functions in the maintenance of the stem cell population in
shoot apical meristems (Fletcher et al., 1999; Ogawa et al., 2008). This suggests ESR proteins
could be involved in an equivalent fashion in signaling cell fate specification and/or
differentiation of the ESR region itself, or they may mediate signaling events required for
epidermal and/or suspensor development in the embryo. It is also possible differentiation of the
ESR region could require signals from the embryo. In support of embryo-to-endosperm
signaling, mutant embryo-less kernels form an embryo cavity within the endosperm, yet no
28

mRNAs of the ESR-1, 2, and 3 genes are detectable in cells surrounding the cavity and,
furthermore, these cells lack any ESR cell morphology (Opsahl-Ferstad et al., 1997). Therefore,
published data support a role for an extensive set of signaling processes between the endosperm
and the embryo that may underlie proper seed development (Widiez et al., 2017).

1.5 KEY QUESTIONS AND FUTURE DIRECTIONS

An understanding of the nature of regulatory programs that dictate endosperm development will
ultimately contribute to the improvement of yield and quality of the maize kernel. As in other
multi-cellular eukaryotic systems, cell fate specification and differentiation in the maize
endosperm is likely regulated by a combination of endogenous cues and positional information,
including signals within the developing endosperm as well as from surrounding kernel
compartments. These cell-differentiation regulatory programs exhibit extensive interplay with
related programs associated with early endosperm cell proliferation, development of polarity,
and later programs associated with storage product accumulation and endosperm maturation.
Therefore, a deep mechanistic understanding of key processes regulating endosperm cell
differentiation will require a holistic understanding of the temporal and spatial gene expression
programs occurring during endosperm development.

In recent years, numerous endosperm transcriptome profiling studies have been carried out
(Sekhon et al., 2011; Lu et al., 2013; Sekhon et al., 2013; Chen et al., 2014a; Li et al., 2014;
Zhan et al., 2015; Qu et al., 2016). In addition, many genome-wide analyses of histone
modifications, DNA methylation, gene imprinting, and profiles of proteins and phosphoproteins
have been published (Waters et al., 2011; Zhang et al., 2011; Makarevitch et al., 2013; Walley et
al., 2013; Xin et al., 2013; Zhang et al., 2014; Dong et al., 2016; Walley et al., 2016). Although
most of these studies were performed on the whole endosperm or kernel, integration of the
limited amount of spatial data within the endosperm and the available temporal data of whole
endosperm/kernel can provide insight into gene regulation dynamics, as illustrated by use of the
endosperm transcriptome (Li et al., 2014; Zhan et al., 2015). To generate a high-resolution
spatio-temporal atlas of gene expression in the endosperm, laser-capture-microdissection (LCM)-
based transcriptome profiling has recently been proven to be a powerful approach (Emmert-Buck
et al., 1996; Kerk et al., 2003; Thakare et al., 2014; Xiong et al., 2014; Yi et al., 2015; Zhan et
29

al., 2015). Other cell-type-specific genome-wide studies, such as mass spectrometry analysis of
protein profiles and chromatin immunoprecipitation assay of chromatin modifications, require
relatively larger amounts of tissues/cells. Therefore, alternative ‘single-cell omics’ approaches,
such as fluorescence-activated sorting of tagged cells or nuclei (Deal and Henikoff, 2010; Wang
and Bodovitz, 2010; Macaulay and Voet, 2014; Handley et al., 2015; Wang and Deal, 2015;
Clark et al., 2016), as have been carried out in other systems (Brady et al., 2007; Evrard et al.,
2012; Slane et al., 2014), can be employed to overcome such limitations. Recently, fluorescent
markers of AL, BETL, and SE proved useful for tracking cell fate and differentiation of these
cell types (Gruis et al., 2006). However, fluorescent markers of the other cell types remain to be
developed. Such markers, in turn, will likely have to be identified through LCM-assisted
transcriptome profiling.

Using these approaches, the resulting spatio-temporal transcriptomic and proteomic data can be
used to construct coexpression networks and infer GRNs (Kang et al., 2011; Downs et al., 2013;
Xue et al., 2013; Zhan et al., 2015; Walley et al., 2016). The hubs of the networks, particularly
TFs, are likely to play important roles in the regulation of the associated biological processes and
functions related to differentiated states. The function of the hubs can be further studied by
generating loss-of-function mutants using RNAi or genome-editing tools (Hannon, 2002;
Townsend et al., 2009; Zhang et al., 2013; Bortesi and Fischer, 2015), which are expected to
uncover novel functional information for endosperm-expressed genes.
30

Figure 1.1. Early proliferation of maize endosperm.

Confocal micrographs of kernels at stages of coenocytic proliferation (A), cellularization


(alveolation phase) (B), and beginning of cell differentiation upon cellularization (C). The arrow
in B indicates a nucleus undergoing periclinal division. Abbreviations: AL, aleurone; BETL,
basal endosperm transfer layer; CV, central vacuole; EMB, embryo; ESR, embryo-surrounding
region; SE, starchy endosperm. Scale bars = 100 μm.
31

Figure 1.2. Cell differentiation in maize endosperm.

Schematic diagrams of the maize kernel showing the relative position of endosperm cell types
and other kernel compartments at early (A) and late (B) differentiation phases. Abbreviations:
AL, aleurone; BETL, basal endosperm transfer layer; CSE, central starchy endosperm; CZ,
conducting zone; EMB, embryo; ESR, embryo-surrounding region; NU, nucellus; PC, placento-
chalaza; PE, pericarp; PED, pedicel; SA, subaleurone; SE, starchy endosperm.
32

CHAPTER 2

RNA sequencing of laser-capture microdissected compartments of the maize kernel


identifies regulatory modules associated with endosperm cell differentiation 3

2.1 ABSTRACT

Endosperm is an absorptive structure that supports embryo development or seedling germination


in angiosperms. The endosperm of cereals is a main source of food, feed, and industrial raw
materials worldwide. However, the genetic networks that regulate endosperm cell differentiation
remain largely unclear. As a first step toward characterizing these networks, we profiled the
mRNAs in five major cell types of the differentiating endosperm and in the embryo and four
maternal compartments of the maize (Zea mays) kernel. Comparisons of these mRNA
populations revealed the diverged gene expression programs between filial and maternal
compartments, and an unexpected close correlation between embryo and the aleurone layer of
endosperm. Gene coexpression network analysis identified coexpression modules associated
with single or multiple kernel compartments including modules for the endosperm cell types,
some of which showed enrichment of previously identified temporally activated and/or imprinted
genes. Detailed analyses of a coexpression module highly correlated with the basal endosperm
transfer layer (BETL) identified a regulatory module activated by MRP-1, a regulator of BETL
differentiation and function. These results provide a high-resolution atlas of gene activity in the
compartments of the maize kernel and help to uncover the regulatory modules associated with
the differentiation of the major endosperm cell types.

2.2 INTRODUCTION

Seed development is initiated by double fertilization of the haploid egg cell and the dikaryotic
central cell to produce two filial structures, a diploid embryo and a triploid endosperm,
respectively (Faure, 2001; Hamamura et al., 2012). Endosperm functions as an absorptive
structure that supports embryo development or seedling germination in angiosperms (Lopes and

3
This chapter has been published in The Plant Cell (2015, vol. 27, 513-531) and is included as a reprint in Appendix
C.
33

Larkins, 1993). Recent evidence also indicates that endosperm plays a critical role in regulation
of seed development through interaction with the embryo and the seed coat (Berger et al., 2006;
Lafon-Placette and Kohler, 2014). The endosperm of cereal grains occupies a large portion of the
mature seed, holds large amounts of proteins and carbohydrates required for seedling
development, and is an important source of food, feed, and renewable industrial raw materials
(Lopes and Larkins, 1993; Olsen, 2001, 2004; Sabelli and Larkins, 2009; FAO, 2012).

In most flowering plants, endosperm development begins with the formation of a coenocyte, as
the fertilized central cell undergoes multiple rounds of nuclear divisions without cytokinesis. The
multinucleated coenocyte then undergoes cellularization and cell differentiation (Olsen, 2004;
Sabelli and Larkins, 2009). In dicots, the endosperm is mostly absorbed by the developing
embryo shortly after cellularization. By contrast, in monocots, and particularly in cereals, the
endosperm enlarges significantly after cellularization through many rounds of cell division
accompanied by cell enlargement and organelle proliferation. Consequently, the cereal
endosperm acquires a high storage capacity of carbohydrates and proteins prepared for
mobilization upon seedling germination (Lopes and Larkins, 1993; Sreenivasulu and Wobus,
2013). The acquisition of endosperm storage capacity is enabled in part through the activity of
specialized cell types or compartments that mediate uptake of nutrients from the maternal
structures and their storage in the inner compartments of the endosperm. Therefore, elucidating
how cell differentiation is regulated during endosperm development is central to understanding
endosperm structure and function.

Because of its relatively large size and economic importance, the maize (Zea mays) endosperm
represents an excellent model system to study early regulatory processes that regulate regional
and cellular differentiation events. The initial coenocytic phase of endosperm growth in maize
occurs during the first 2 days after pollination (DAP) and this is followed by a period of
cellularization during 3-4 DAP. Following cellularization, the endosperm cells undergo two
major phases of mitotic proliferation, an early phase that lasts until 8-12 DAP in the central
region, and a late phase that continues until 20-25 DAP in the outer endosperm layers. Starting at
~8-10 DAP, the central portion of endosperm cells gradually switches from mitosis to
endoreduplication and becomes filled with starch and storage proteins (Brink and Cooper, 1947;
34

Olsen, 2001, 2004; Sabelli and Larkins, 2009; Becraft and Gutierrez-Marcos, 2012; Olsen and
Becraft, 2013; Leroux et al., 2014).

Differentiation of maize endosperm cells occurs primarily at 4-6 DAP (following endosperm
cellularization and before the initiation of mitotic proliferation), resulting in four main cell types
including the starchy endosperm (SE), the aleurone (AL), the embryo-surrounding region (ESR),
and the basal endosperm transfer layer (BETL)(Olsen, 2001; Becraft and Gutierrez-Marcos,
2012; Leroux et al., 2014). The SE is the cell type that accumulates starch and storage proteins.
The SE itself contains at least three sub-regions, including the central starchy endosperm (CSE),
the conducting zone (CZ) and the subaleurone (SA) (Becraft, 2001; Olsen, 2001, 2004; Sabelli
and Larkins, 2009). The AL is a single peripheral layer of cells that produces hydrolytic enzymes
to mobilize the storage products in the SE when activated during seed germination. The ESR is
believed to act as a physical barrier and messenger between endosperm and embryo (Olsen,
2004). The BETL is a transfer cell layer that transports nutrients from the maternal tissue into the
inner endosperm cells, including the developing SE, in order to enable starch and protein
synthesis (Sabelli and Larkins, 2009; Becraft and Gutierrez-Marcos, 2012). Recent studies have
identified many genes expressed specifically in the BETL including multiple genes encoding
cysteine-rich proteins that are thought to act as antimicrobial or intercellular signal molecules
(Tailor et al., 1997; Marshall et al., 2011), and MRP-1 (Myb-Related Protein-1), a MYB-related
transcription factor previously shown to activate a number of these genes in the BETL (Gomez et
al., 2002; Gutierrez-Marcos et al., 2004; Gomez et al., 2009). Moreover, ectopic expression of
MRP-1 in the AL has been shown to produce a transient BETL-like structure (Gomez et al.,
2009). Additional recent efforts have enabled genome-wide identification of gene expression
during nearly all stages of endosperm development (Sekhon et al., 2013; Chen et al., 2014a;
Mathelier et al., 2014). However, little is known about the gene regulatory networks (GRNs) that
regulate the differentiation and determine the function of the individual cell types or
compartments of the endosperm in maize.

Here, we used a coupled laser-capture microdissection (LCM) and RNA sequencing (RNA-Seq)
strategy to comprehensively profile the mRNA populations present in each of the main cell types
of the maize endosperm, as well as the embryo and four maternal compartments of the kernel at
8 DAP. We identified mRNAs that specifically accumulate in each of the captured
35

compartments. Also, using an unbiased network analysis tool, we detected modules of


coexpressed genes that are either predominantly expressed in a single compartment or expressed
in multiple compartments, including several endosperm-correlated modules that are enriched for
temporally up-regulated genes and/or imprinted genes that we had previously identified. By
focusing on the analysis of genes in a BETL-correlated coexpression module, we identified and
experimentally validated a regulatory module of the BETL GRN that is activated by MRP-1.

2.3 RESULTS

Capture and analysis of mRNA populations of filial and maternal compartments of 8-DAP
kernel

To identify the genes active in each of the endosperm cell types, we used laser-capture
microdissection to isolate and profile mRNA populations of five endosperm compartments (cell
types) of maize inbred line B73 at 8 DAP. The compartments analyzed included aleurone (AL),
basal endosperm transfer layer (BETL), embryo-surrounding region (ESR), and two sub-regions
of starchy endosperm (SE), the central starchy endosperm (CSE) and the conducting zone (CZ).
To compare endosperm gene expression programs with embryonic and maternal programs, we
also captured the embryo (EMB), nucellus (NU), placento-chalazal region (PC), pericarp (PE),
and the vascular region of the pedicel (PED) at 8 DAP (Figure 2.1A; Supplemental Figures 2.1,
2.2, and 2.3; Supplemental Table 2.1). We selected this time point because it follows
differentiation of the main cell types of the endosperm, which occurs at 4-6 DAP, and precedes
developmental programs associated with endosperm function, including the activation of storage
product synthesis and deposition program, which initiates at ~8-10 DAP (Becraft, 2001; Olsen,
2001, 2004; Sabelli and Larkins, 2009; Becraft and Gutierrez-Marcos, 2012; Olsen and Becraft,
2013; Leroux et al., 2014). Total RNA extracted from biological triplicates for the endosperm
compartments and embryo, and single replicates of maternal compartments (22 samples) were
reverse-transcribed to cDNA using Oligo(dT) and random primers, amplified, and paired-end
sequenced using an Illumina HiSeq 2000 platform.

The resulting reads were quality checked and mapped to the maize reference genome (B73
RefGen_v3). Of the resulting mapped reads (8.9 to 34.7 million, 52.3% to 89.7% of total reads),
36

3.2 to 11.3 million (22.1% to 39.5%) were mapped to exonic sequences (Supplemental Table
2.2). The exonic reads were normalized using Cufflinks (Trapnell et al., 2012) and reported as
fragments per kilobase of transcript per million mapped reads (FPKM). A gene was considered
expressed in a given sample if the lower boundary of its FPKM 95% confidence interval
(FPKM_conf_lo) was greater than zero (Hansey et al., 2012). Based on this criterion, 29,369
genes were identified as expressed in at least one of the 22 samples (Supplemental Data Set 2.1).
Using pairwise Spearman Correlation Coefficient (SCC) analysis, the triplicate FPKM values
from each of the endosperm compartments and the embryo were shown to be highly correlated (ρ
= 0.87 to 0.91, Supplemental Figure 2.4). Accordingly, we pooled each triplicate set of exonic
reads and re-normalized the data using Cufflinks. Using the same cut-off as indicated above, we
detected 30,665 genes expressed in at least one of the ten compartments (Supplemental Data Set
2.2 and Supplemental Figure 2.5), 10,725 genes expressed in all ten compartments
(Supplemental Figure 2.6A), and between 15,910 and 23,853 (NU and ESR, respectively)
expressed in individual compartments (Figure 2.1B; Supplemental Table 2.3). In all cases, the
proportion of transcription factor (TF) genes detected as expressed tracked closely with the total
number of expressed genes (between ~5.3% and 6.0% for CZ and PED, respectively; Figure
2.1B, Supplemental Table 2.3). The proportions of high-expressing (FPKM ≥ 10), medium-
expressing (2 ≤ FPKM < 10) and low-expressing (FPKM < 2) genes were relatively similar in all
compartments (Figure 2.1C, Supplemental Table 2.4). Collectively, 22,703 genes were detected
as expressed in EMB, and 28,078 and 22,989 genes expressed in at least one captured endosperm
and maternal compartment, respectively. The three sources of captured tissues shared 19,009
expressed genes in total (Supplemental Figure 2.6B). Taken together, our analysis of LCM-
derived RNA-Seq data indicates that we have obtained sufficient coverage of the transcriptome
of the filial and maternal compartments of the kernel for subsequent analysis of gene networks.

Filial and maternal compartments of the kernel exhibit distinct mRNA populations

To understand the relationships between the mRNA populations isolated from the individual
compartments, we performed a principal component analysis (PCA) (Figure 2.2A; Supplemental
Table 2.5) and a hierarchical clustering of data from the SCC analysis (Figure 2.2B) of
normalized expression levels for the 30,665 genes expressed in at least one compartment. Those
37

compartments with highest overlap in mRNA populations are expected to be more closely
associated in such analyses and are likely to share functions. Both analyses showed high
correlation among endosperm cell types and a distinct clustering of these cell types in
comparison to the maternal compartments (Figures 2.2A and 2.2B). As expected, the filial EMB
showed closer correlation with endosperm cell types as compared to the maternal compartments.
However, the endosperm AL showed a closer correlation with EMB (ρ = 0.80) than with any
other endosperm cell type (Figure 2.2B). An analysis of our data for expression of two AL
marker genes, namely VPP1 (Vacuolar H+-translocating inorganic pyrophosphatase 1)
(Wisniewski and Rogowsky, 2004) and AL-9 (Gomez et al., 2009), indicated that the captured
EMB RNA sample was not contaminated by the AL RNAs (Supplemental Data Set 2.2).
Therefore, our observation suggests that the AL is distinct in some zygotic functions typically
not found in the other endosperm cell types. Together, our data indicate that maternal and filial
gene expression programs are divergent as they arise from distinct genetic origins, and that the
captured compartments show sufficient diversity at the mRNA level to allow identification of
unique gene sets for each compartment.

Identification of gene sets specifically expressed in each of the 8-DAP kernel compartments

To discover the gene expression programs that characterize each kernel compartment, we
identified mRNAs that specifically accumulate in each compartment at 8 DAP by applying a
compartment specificity (CS) scoring algorithm to the genes with FPKM ≥ 2 in at least one
compartment. In this analysis, we defined the corresponding genes with CS score > 0.3 as being
expressed in a compartment-specific pattern. Using this cut-off, 13,009 compartment-specific
genes were identified in total for all captured compartments (Supplemental Data Set 2.3, Figure
2.3). In contrast to the similarity of the overall mRNA profiles detected among the compartments
as described above (Figure 2.1C), the numbers of detected compartment-specific genes showed
dramatic differences among the ten compartments (Figure 2.3B). The endosperm cell types
showed the lowest number of compartment-specific genes, ranging from a low of 331 in the CSE
(1.6% of all CSE-expressed genes) to a high of 912 in the BETL (4.4% of all BETL-expressed
genes), as compared to the maternal compartments that ranged from 1,390 in the PC (8.6% of all
PC-expressed genes) to 2,432 in the PE (13.2% of all PE-expressed genes) (Figure 2.3B). For the
38

EMB, 2,235 genes were identified as compartment specific, which corresponds to 9.8% of all
EMB-expressed genes (Figure 2.3B). The proportion of TF genes among the compartment-
specific genes did not track uniformly across all compartments, varying from a low of 3.9% (CZ)
to a high of 9.9% (EMB) (Figure 2.3B). The variable number and proportion of compartment-
specific genes and the associated variation in the proportion of TF genes suggest that most of the
endosperm cell types captured express less complex gene sets as compared to the maternal
compartments or the embryo. Alternatively, the complexity of expression may simply reflect the
complexity of the captured compartments, as the captured EMB and maternal compartments
likely contain more than one cell type.

We used three sets of expression localization data to validate the cell type-specific patterns of
mRNA accumulation in the endosperm. First, we carried out a series of mRNA in situ
hybridizations for genes that were shown to be highly specific to a single compartment based on
CS scores ranging from 0.74 to 0.99, including 20 genes expressed in the CSE (2), ESR (1), AL
(3), and BETL (14) (Supplemental Figure 2.7). Second, we previously carried out in situ
hybridization for 10 genes specifically expressed in CSE (3), CZ (1), ESR (3), AL (1), and BETL
(2) (Mathelier et al., 2014). These genes showed CS scores in the given compartments ranging
from 0.47 to 0.99. Third, we summarized previously reported in situ hybridization or promoter
activity data for cell-specific genes from the literature (Hueros et al., 1995; Hueros et al., 1999;
Magnard et al., 2000; Serna et al., 2001; Woo et al., 2001; Gomez et al., 2002; Magnard et al.,
2003; Gutierrez-Marcos et al., 2004; Wisniewski and Rogowsky, 2004; Balandin et al., 2005;
Massonneau et al., 2005; Muniz et al., 2006; Gomez et al., 2009; Muniz et al., 2010; Royo et al.,
2014) and also found these genes to have a relatively high range of CS scores (0.51 to 0.99).
Altogether, these comprise 44 genes showing cell-specific expression in the endosperm
(Supplemental Table 2.6). All of these genes showed highly specific mRNA localization patterns
within the cell types with high CS scores, indicating that our RNA-Seq data accurately reflect the
accumulation of endogenous mRNAs in the endosperm.

Identification of gene coexpression modules of 8-DAP maize kernel

To begin to understand the nature of the GRNs in each of the captured compartments or cell
types of the 8-DAP kernel, we identified coexpressed gene sets by applying weighted gene
39

coexpression network analysis (WGCNA) (Zhang and Horvath, 2005; Langfelder and Horvath,
2008) to the expressed genes after excluding the ones with low FPKM (average FPKM < 1)
and/or low coefficient of variance (CV < 1) across all ten compartments. The 9,361 genes that
fulfilled these stringent criteria fell into 18 coexpression modules (M1-M18), containing from 83
(M1) to 1,209 (M4) genes, including 3 (M1) to 111 (M8) coexpressed TF genes (Figure 2.4B;
Supplemental Data Set 2.4). Trend-plot analysis of Z-scores of genes in each module showed
that these gene sets were expressed in a highly coordinated manner (Supplemental Figure 2.8).
Significantly, a permutation test showed that the average topological overlap of the 18 observed
modules was greater than randomly sampled modules of the same size (P < 10-5; Supplemental
Table 2.7), suggesting that the assignment of gene sets to each of the modules was highly robust.
Association of each coexpression module with each compartment or cell type was quantified by
Pearson Correlation Coefficient (PCC) analysis and visualized using a hierarchically clustered
heat map (Figure 2.4A). Interestingly, the 18 coexpression modules fell into two distinct
categories showing a relatively high correlation (r ≥ 0.35) with either the filial (i.e., M1, M2,
M8, M9, M10, M12, M15, M17, and M18) or the maternal (i.e., M3-M6, M11, M13, M14, and
M16) compartments, except for only one module (M7) that was highly correlated with both
EMB (r = 0.68) and PE (r = 0.64). Furthermore, 10 of the 18 modules, including M3, M4, M6,
M8, M10, M11, M12, M15, M17, and M18, specifically correlated with individual
compartments (r ≥ 0.85 for one compartment and r < 0.35 for other compartments), indicating
that the expression of genes in these modules are highly compartment- or cell type-specific. The
other 8 modules showed relatively high correlation to at least two compartments (r ≥ 0.35).
Consistent with the high correlation of mRNAs detected between EMB and AL described above
(Figure 2.2B), three modules including M1, M2, and M9 showed high correlation with both
EMB and AL (r ≥ 0.35).

To confirm that the assignment of genes to these coexpression modules using WGCNA reflected
valid compartment-based patterns of mRNA accumulation, we examined the number of
overlapping genes between these modules and the compartment-specific gene sets identified
above (Supplemental Data Set 2.3). This analysis showed that from ~52% to 94% of genes in the
compartment-correlated coexpression modules were in fact detected within the corresponding
compartment-specific gene sets, and that the overlap between each compartment-correlated
module and the corresponding compartment-specific gene set was greater than expected by
40

chance (hypergeometric test, P < 10-5) (Figure 2.4C). These data indicate that a large proportion
of coexpressed genes in each module detected through WGCNA is related to compartment- or
cell type-specific functions. Taken together, these results indicate that each of the filial and
maternal compartments of the maize kernel is associated with one or more coexpression modules
that reflect the gene-regulatory processes specific to each compartment and are indicators of the
differentiation programs functioning within each compartment.

The endosperm-associated coexpression modules are associated with distinct temporal


programs of expression

We had previously described a set of temporal programs of gene activity during early kernel and
endosperm development in maize and had suggested that some of these programs correlated with
cell differentiation (Mathelier et al., 2014). The identified temporal programs included gene sets
exhibiting temporal up-regulation at (“up@”) specific stages (e.g., up@6DAP refers to a gene set
exhibiting low expression at 0-4 DAP and high expression at 6-12 DAP). Comparison of the
spatial coexpression modules described here with the temporally up-regulated gene sets showed
that all of the five endosperm compartment-correlated modules significantly overlapped with the
up@6DAP gene set (P < 10-5), while four of them significantly overlapped with the up@8DAP
gene set (Figure 2.4D). Consistent with this, an analysis of the overall expression levels for each
coexpression module using the available developmental RNA-Seq data generated by us and
others (Chen et al., 2014a; Mathelier et al., 2014) from whole kernel and whole endosperm
material showed that nearly all endosperm-correlated modules (M10, M12, M15, M17 and M18)
showed up-regulation at 6-8 DAP (Supplemental Figures 2.9 and 2.10). However, the extent of
the up-regulated patterns varied among the endosperm-correlated modules with the expression of
genes in the AL-, ESR-, and BETL-correlated modules (M12, M15, and M18, respectively)
showing a more rapid decline by 10 DAP whereas the expression of CSE-, and CZ-correlated
modules (M10 and M17, respectively) exhibited a more gradual decline beyond 22 DAP
(Supplemental Figures 2.9 and 2.10). Interestingly, genes in modules M1, M2, and M9 with high
correlations with both EMB and AL also showed a similar temporal expression pattern as those
of the endosperm-correlated modules (Supplemental Figure 2.11). Together, these data indicate
41

that the coexpression modules associated with the major endosperm compartments at 8-DAP are
regulated in a highly coordinated manner in both space and time.

The endosperm-associated modules are enriched for endosperm-imprinted genes

Gene imprinting has been suggested to be involved in the regulation of nutrient allocation from
the maternal tissues to endosperm (Haig and Westoby, 1989; Moore and Haig, 1991; Costa et al.,
2012). To investigate the relationship between the spatial programs of gene expression and gene
imprinting, we examined the overlap between the WGCNA-generated coexpression modules
with the imprinted genes that we previously identified in developing endosperm (Xin et al.,
2013). This analysis showed that a subset of genes in the CZ- and BETL-correlated coexpression
modules (M17 and M18, respectively) significantly overlapped with a subset of the previously
described paternally-expressed gene sets (PEGs); the CSE-correlated module M10 and the
modules associated with the maternal compartments (M3, M5, M6, M11, M13, M14, and M16)
showed significant overlap with the maternally expressed gene sets (MEGs); and a subset of the
ESR-correlated coexpression module exhibited significant overlap with both the previously
described PEGs and MEGs (hypergeometric P < 10-5; Figure 2.4E). In support of these data, a
similar pattern of overlaps was observed when we applied less stringent criteria to identify genes
with allele-biased expression patterns using the same set of normalized RNA-Seq data
(Supplemental Figure 2.12). Interestingly, many of the imprinted/allele-biased genes assigned by
WGCNA to the endosperm-associated coexpression modules were TF genes from multiple
families (Supplemental Table 2.8). These results indicate an extensive interplay between
epigenetic programs that regulate allelic expression and the transcriptional regulatory programs
involved in the cellular differentiation and function of endosperm.

Biological processes enriched in coexpression modules of the filial compartments of the


kernel

The identified coexpression modules (Figure 2.4A) are likely associated with specific biological
processes or pathways involved in the development or function of each compartment. To identify
the major biological processes associated with the filial coexpression modules, we used
42

Blast2GO (Conesa et al., 2005; Conesa and Gotz, 2008; Gotz et al., 2008) to identify the
processes that were significantly enriched (FDR < 0.05) in the modules that showed high
correlation with endosperm compartments and/or the EMB. These included modules M1, M2,
M8, M9, M10, M17, M12, M15, and M18. As expected, the CSE-correlated M10 was shown to
be enriched for “starch biosynthetic process” and “glycogen biosynthetic process” (Supplemental
Figure 2.13), with the former GO category including the Shrunken-2 (Sh2), Brittle-2 (Bt2),
Starch-Branching Enzyme 1 (SBE1), and Waxy1 genes, which all have well characterized
functions in starch biosynthesis (Shure et al., 1983; Giroux et al., 1994; Blauth et al., 2002).
Additionally, close inspection of genes in M10 revealed that this module also contained other
starch-synthesis-related genes without any current GO annotation including the Shrunke-1 (Sh1),
Sugary1 (Su1), and Starch Synthase 1 (SS1) genes (Chourey and Nelson, 1979; James et al.,
1995; Commuri and Keeling, 2001).

In the case of zein-related genes, the mRNAs for only four genes encoding the 15-kD beta-zein,
the 16-kD gamma-zein, the 27-kD gamma-zein, and the 18-kD delta-zein were detected in our
analysis (FPKM_conf_lo > 0) in at least one cell type (Supplemental Data Set 2.2). Interestingly,
all four zein genes, as well as the Floury-1 gene, which encodes an ER membrane protein
involved in the targeted localization of an 22-kD alpha-zein in protein body formation (Holding
et al., 2007), were contained within the M10 module (Supplemental Data Set 2.4). This
observation correlates well with previous reports that the formation of zein-containing protein
bodies start as small accretions consisting primarily of beta- and gamma-zeins (Woo et al.,
2001), suggesting that the M10 module contains the key early genes necessary for storage
protein body biogenesis. The M10 module also included the TF genes Opaque-2 (O2) and PBF
(Prolamin-box Binding Factor), with the relatively high M10 module membership (MM) scores
of 0.93 and 0.97, respectively (Supplemental Data Set 2.4). Furthermore, visualization of M10
using VisANT (Hu et al., 2004) showed that these two TFs are among the most highly connected
intramodular hubs of this module (Supplemental Figure 2.14A). O2 and PBF have previously
been shown to regulate storage-program gene expression in maize endosperm (Schmidt et al.,
1990; Schmidt et al., 1992; Marzabal et al., 2008). For example, the 15-kD beta-zein and the 27-
kD gamma-zein have been shown to be regulated by O2 and PBF, respectively (Cord Neto et al.,
1995; Marzabal et al., 2008). Therefore, the M10 coexpression module likely includes a number
of direct gene targets of both TFs.
43

For the CZ-correlated M17 module, key functional over-representations included “glycolysis”,
“response to hydrogen peroxide”, “response to cadmium ion”, and “response to heat”
(Supplemental Figure 2.13). This module showed a modest correlation with the CSE (r = 0.27;
Figure 2.4A), suggesting that the CZ may express an overlapping set of genes that function
similarly to those detected in the CSE. Conversely, the captured CSE cells are expected to have
contained a portion of the CZ cells (Supplemental Figure 2.1) as the latter likely extend basally
and centrally into the captured CSE region (Cooper, 1951; Charlton et al., 1995; Becraft, 2001).
This may explain the modest level of correlation of M10 with CZ (r = 0.29; Figure 2.4A).

The AL-correlated module M12 was enriched for “single-organism process” (Supplemental
Figure 2.13), and the two previously described markers of the aleurone, VPP1 (Wisniewski and
Rogowsky, 2004) and AL-9 (Gomez et al., 2009), were assigned within this module with high
MM values (0.91 and 0.85, respectively).

Likely due to the greater structural complexity of the EMB as compared to the captured
endosperm compartments, the EMB-specific module M8 was enriched for more diverse GO
categories in comparison to the endosperm-specific modules. The four biological processes that
were most significantly enriched for this module included “regulation of transcription, DNA-
templated”, “floral organ development”, “phyllome development”, and “response to hormone”.
Modules M1, M2, and M9 showed relatively high correlation with both the AL and the EMB (r
≥ 0.35). Among these, M9 also showed a slightly lower correlation with CSE (r = 0.33). Among
these modules, M2 contained the largest number (477) of genes. Similar to M8, this module was
also enriched for many GO categories, most of which were biological processes related to DNA
replication and mitotic cell division (e.g., “DNA replication”, “DNA repair”, “mitotic spindle
assembly checkpoint”, “mitotic spindle assembly checkpoint”, and “microtubule-based
movement”, etc.). Similarly, M9 (containing 194 genes) was enriched for “cell cycle process”,
“chromosome segregation”, “regulation of DNA replication”, and “single−organism organelle
organization”, etc. Furthermore, M1, M2, and M9 shared enrichment for “nucleosome assembly”
and “DNA duplex unwinding”, which are also biological processes involved in DNA replication
and cell division (Supplemental Figure 2.13). These GO enrichments suggest that the EMB and
AL, and to some degree the CSE, are programmed to undergo extensive mitotic cell proliferation
at 8 DAP via the coordinated expression of a relatively extensive gene network.
44

Consistent with the presumptive function of the ESR in mediating endosperm-embryo interaction
and expressing anti-microbial products (Sabelli and Larkins, 2009), the ESR-specific module
M15 was enriched for “cell-cell signaling involved in cell fate commitment” (Supplemental
Figure 2.13). This module also included many genes isolated previously by virtue of their highly
specific ESR expression pattern, including ESR-1, ESR-2, ESR-3, ESR-6, ESR-6B, and AE-3
(Schel et al., 1984; Opsahl-Ferstad et al., 1997; Bonello et al., 2000; Balandin et al., 2005; Sosso
et al., 2010). In our analysis, many of the same genes, including ESR-1, ESR-2, ESR-3, and ESR-
6 were positioned within the intramodular hubs with high MM (e.g., 0.99) to M15 (Supplemental
Figure 2.14C and Supplemental Data Set 2.4).

As expected from the BETL’s reported role as mediator of sugar and metabolite uptake into the
endosperm through its interaction with the underlying maternal placenta-chalazal region (Sabelli
and Larkins, 2009), the BETL-correlated module M18 was found to be enriched for
“transmembrane transport”, “ion transport”, and “sucrose transport” functions (Supplemental
Figure 2.13). The M18 module contained nearly all of the previously identified BETL-expressed
genes (Hueros et al., 1995; Cheng et al., 1996; Doan et al., 1996; Hueros et al., 1999; Serna et
al., 2001; Magnard et al., 2003; Gutierrez-Marcos et al., 2004; Massonneau et al., 2005; Gruis et
al., 2006; Muniz et al., 2006; Brugiere et al., 2008; Gomez et al., 2009; Muniz et al., 2010)
including BETL-1, 3, 4, 9, and 10; BAP-1A, 1B, 2, 3A, and 3B; TCRR-1, and 2; and INCW2 (Cell
wall invertase 2), EBE-2 (Embryo sac/Basal endosperm-layer/Embryo-surrounding region-2),
IPT-2 (Isopentenyl transferase-2), and CC-8 (Corn Cystatin-8) (Supplemental Data Set 2.4). In
addition, this module contained 11 of the 13 MEG genes (with MEG-4 and MEG-14 being the
two exceptions) that have been identified in the B73 genome (Supplemental Data Set 2.4),
including MEG-1, which is a BETL-specific gene that has been shown to be important for the
development and differentiation of the BETL (Costa et al., 2012). The BETL-, BAP-, and MEG-
type genes encode small, secreted, cysteine-rich proteins (CRPs) that have been suggested to
protect the embryo from maternally-transmitted pathogens (Tailor et al., 1997) and to serve as
signaling molecules that coordinate the supply of nutrients to the embryo during kernel
development (Marshall et al., 2011), while the TCRR genes encode type-A response regulators
(Muniz et al., 2006; Muniz et al., 2010). Significantly, all of these genes showed high MM
(>0.90) to module M18, and many of them were among the top-scoring hubs (Supplemental
Figure 2.14E and Supplemental Data Set 2.4). Furthermore, the BETL-specific TF gene MRP-1,
45

previously shown to be involved in regulation of BETL differentiation (Gomez et al., 2002;


Gomez et al., 2009), was also detected in M18 with a high MM score (1.00, Supplemental Data
Set 2.4). This is consistent with its role as an activator of many BETL-specific genes, including
BETL-1, BETL-9, BETL-10, BAP-2, MEG-1, TCRR-1 and TCRR-2 (Gomez et al., 2002;
Gutierrez-Marcos et al., 2004; Muniz et al., 2006; Gomez et al., 2009; Muniz et al., 2010).
Therefore, these data suggest that MRP-1 acts as a major regulator of a subset of genes in the
M18 coexpression module. Together, our results suggest that the WGCNA-identified
coexpression modules can be used as starting points for identification of GRNs functioning in
each endosperm compartment.

De novo identification of cis-motifs associated with the endosperm coexpression modules

As a first step toward identification of the endosperm GRNs, we used MEME software (Bailey
and Elkan, 1994) to detect putative cis-regulatory elements in upstream gene sequences from
each of the five endosperm coexpression modules identified using WGCNA including M10,
M12, M15, M17, and M18 (Figure 2.4A). We searched for 10- to 12-bp sequence motifs over-
represented within -1 kb to +0.5 kb (relative to transcription start site, TSS) of genes in each
module, and further identified motifs that were significantly similar (q-value < 0.05) to the
known plant cis-motifs available in the JASPAR CORE database using TOMTOM program
(Gupta et al., 2007). We found that most of the detected motifs were shared among the majority
of the modules as exemplified by the motifs that contained exclusively CG- or AT-rich
sequences (Supplemental Table 2.9) with significant similarity to the reported binding sites of
ABI4 (a maize AP2-EREBP protein) and SOC1 (an Arabidopsis thaliana MIKC-type MADS-
box protein), respectively (Riechmann et al., 1996; Niu et al., 2002). Coincidently, four of the
five endosperm modules (M10, M12, M15, and M17) included at least one MIKC-type MADS-
box gene while all five modules included at least one AP2-BREBP gene based on the current
maize genome annotation (Supplemental Data Set 2.4), indicating that the MIKC-type MADS-
box and AP2-EREBP families may play broad regulatory roles in maize endosperm
development.

In contrast, a small number of motifs were detected specifically in single modules. In one
instance, the Motif 10 that was enriched in the CSE-correlated module M10 showed significant
46

similarity to a cis-motif that has been shown to bind a bZIP TF in snapdragon (Antirrhinum
majus) (Supplemental Table 2.9)(Martinez-Garcia et al., 1998). Accordingly, two bZIP TF
genes, bZIP46 and the storage program regulator O2 were detected within M10 (Supplemental
Data Set 2.4). These results suggest that a subset of genes in module M10 may be regulated by
the bZIP genes in the same module. In a second case, Motifs 3, 6, and 8 that were detected in
upstream sequences of M18 genes each contained a repeated GATA sequence (Figure 2.5A)
similar to a sequence previously shown to be involved in binding and activation of target genes
by MRP-1, a MYB-related (MYBR) TF (Baranowskij et al., 1994; Barrero et al., 2006), whereas
no GATA-containing motifs were detected in any of the other endosperm coexpression modules.
In support of a transcriptional regulatory role for the motifs, our analysis indicated a biased
distribution of Motifs 3, 6, and 8 upstream to the TSS (Figure 2.5B). A close inspection of the
sub-motifs, namely the variant sequences of each motif, of Motifs 3, 6, and 8 revealed that at
least 148 genes in the M18 module contained one or more sub-motifs within their flanking
sequences. Collectively, our data suggest that a large subset of genes in the M18 coexpression
module is likely regulated by MRP-1 through binding at GATA-rich sequences.

Identification of a gene regulatory module associated with BETL cell differentiation

We focused on the BETL-associated coexpression module M18 to decipher a portion of the


BETL GRN. The BETL transports nutrients from the maternal tissue into the endosperm and is
important for the proper development of the endosperm and the endosperm’s capacity as a
storage organ (Thompson et al., 2001; Costa et al., 2012). Based on the available data
(Baranowskij et al., 1994; Gomez et al., 2002; Gutierrez-Marcos et al., 2004; Barrero et al.,
2006; Muniz et al., 2006; Gomez et al., 2009; Muniz et al., 2010) including our identification of
the coexpression module M18 and its association with GATA-rich sequence motifs (discussed
above), we hypothesized that a GRN for BETL differentiation is minimally composed of MRP-1
and a large set of target genes that are activated upon binding of MRP-1 to Motifs 3, 6, and 8. A
close examination of all the sub-motifs of the 10 motifs enriched for M18 showed that each of
the motifs is represented by two to 21 sub-motifs that appeared one to 90 times within the
promoters of the M18 genes (Supplemental Table 2.10).
47

We tested for binding of MRP-1 to Motifs 3, 6, and 8 using directed yeast one-hybrid (Y1H)
assays. We introduced two constructs into yeast cells, one that resulted in expression of MRP-1
and a second that comprised the test sequence fused upstream of the yeast AUR1-C gene. AUR1-
C confers resistance to aureobasidin A (AbA) (Heidler and Radding, 1995). In this assay, MRP-1
binding to the test sequence results in AUR1-C activation and, thus, growth of cells on plates
containing AbA, whereas absence of MRP-1 binding results in absence of growth on plates
containing AbA. Testing for MRP-1 binding to all of the 24 sub-motifs comprising Motifs 3, 6,
and 8 indicated strong binding to four of the sub-motifs. These sub-motifs included Motifs 3a/8a,
6a/8b, 8c, and 8f, with Motifs 3a and 6a identical to 8a and 8b, respectively (Table 2.1;
Supplemental Figure 2.15). Comparison of these sub-motifs with the previously reported MRP-1
binding sites showed that Motifs 6a/8b and 8c were identical to the reported MRP-1 binding site
in the promoters of BETL-1 (Motif IV) and BETL-2, respectively (Barrero et al., 2006), while
Motifs 3a/8a and 8f represent newly identified MRP-1 binding sites.

We identified 93 genes that contain at least one of the four sub-motifs that were shown to be
bound by MRP-1 in the Y1H assays (Supplemental Table 2.11). All the 93 genes showed
relatively high MM values to M18, ranging from 0.62 to 1.00, with 78 of them higher than 0.95
(Supplemental Table 2.11). These results suggest that these 93 genes constitute a regulatory
module of the MRP-1-regulated GRN. Significantly, 7 of the 14 BETL-expressed genes
described previously, including MEG-1, BETL-1, BETL-10, BAP-1A, BAP-2, BAP-3A, and BAP-
3B, were present in this regulatory module (Supplemental Table 2.11). Based on the available
functional annotation of maize genes, this regulatory module contained at least 17 CRP-coding
genes (including 3 BETL-, 4 BAP-, and 10 MEG-type genes; Supplemental Table 2.11), among
which BETL-1, BETL-2, BETL-10, and MEG-1 have previously been shown to be regulated by
MRP-1. Notably, 6 TF genes were detected within this regulatory module, including three
MYBR-, two C2C2-GATA-, one AP2-EREBP-, and one DBB-family genes, suggesting strongly
that MRP-1 indirectly regulates a subset of the BETL-expressed genes by regulating these TFs.
In addition, this regulatory module also included a gene (GRMZM2G406552) that encodes a
putative nonspecific lipid-transfer protein (Supplemental Table 2.11). Furthermore, comparison
of the genes in the regulatory module to the GenBank non-redundant (nr) and Swissprot protein
databases (BLASTX, E-value < 10-6) revealed that a number of genes encoding putative
transporters of peptide (GRMZM2G156794), calcium (GRMZM5G836886), phosphate
48

(GRMZM2G466545), and magnesium (GRMZM2G054632) were also included in this


regulatory module (Supplemental Data Sets 2.5 and 2.6).

To validate this putative network further, we used Y1H assays to test for binding of MRP-1 to 22
gene promoters from the M18 module containing motifs positive for MRP-1 binding. We also
tested 19 gene promoters lacking these motifs. Of the 41 promoters tested, 31 were shown to
bind MRP-1 and 10 did not. Significantly, all promoters containing the MRP-1-binding sub-
motifs were shown to bind MRP-1 (Supplemental Figure 2.16 and Supplemental Table 2.12).
Interestingly, 7 of the 19 promoters lacking the MRP-1-binding sub-motifs were found to bind
MRP-1 in our assays, indicating that MRP-1 binds to additional sequences not identified in the
MEME analysis.

These results suggest strongly that the 93 genes containing the MRP-1-binding sub-motifs, as
well as many additional genes (at least seven), are directly regulated by MRP-1 and that this
gene set constitutes a regulatory module within the BETL GRN (Figure 2.6A). If so, these genes
should exhibit a similar temporal pattern of expression to that of MRP-1. An analysis of the
expression of the 93 putative MRP-1 target genes throughout development using the RNA-Seq
data generated and/or processed by Chen et al. (2014a) showed that a large subset of these genes
(including most of the cysteine-rich protein-coding genes) were primarily expressed in the
endosperm peaking at 6-8 DAP, mirroring the pattern of MRP-1 expression (Figure 2.6B).
Interestingly, this set included genes that displayed restricted patterns of mRNA accumulation
throughout development, with most showing low mRNA prevalence in the vegetative organs, but
showing a wide range of mRNA levels in the 6-8DAP-endosperm/kernel (Supplemental Figure
2.17). In comparison to the rest of the M18 genes, the MRP-1 regulatory module genes showed a
significantly higher relative level of expression in the 8-DAP BETL (Figure 2.6C; P = 6.4e-15
based on unpaired t-test). As the M18 genes showed significant overlap with the genes in the two
temporal programs up@6DAP and up@8DAP (Figure 2.4D), we examined the overlap between
the 93 genes containing the MRP-1-binding sub-motifs and the two temporal clusters. The result
showed that 49 of the 93 genes lie within the up@6DAP gene set (Figure 2.6D, hypergeometric
P = 8.1e-89), while only 3 genes were found to overlap with the up@8DAP gene set, indicating
that a significant portion of the MRP-1 regulatory module is coordinately up-regulated by MRP-
1 at 6 DAP in BETL. Taken together, these results suggest that a large subset of M18 genes
49

constitutes a portion of the MRP-1-regulated gene network, and that these genes are activated by
MRP-1 through binding to specific upstream cis-regulatory sequences.

2.4 DISCUSSION

We used an LCM RNA-Seq profiling approach to comprehensively detect mRNA populations


for ten filial and maternal compartments of an 8-DAP maize kernel, and subsequently identified
highly correlated gene expression programs associated with each compartment using WGCNA.
The endosperm coexpression modules are expected to reflect the state of cellular differentiation
within individual endosperm compartments or cell types. Our data indicate that the timing and
extent of these differentiation processes are unique to each compartment as suggested previously
(Olsen, 2001, 2004; Sabelli and Larkins, 2009; Becraft and Gutierrez-Marcos, 2012; Leroux et
al., 2014; Mathelier et al., 2014). As a test case, we deciphered an MRP-1 regulatory module
containing 93 genes that are likely involved in BETL cellular differentiation.

The high-quality RNAs isolated from the laser-captured cells (Supplemental Figures 2.2 and 2.3,
and Supplemental Table 2.1) and the resulting highly reproducible RNA-Seq data (Supplemental
Figure 4) enabled us to detect mRNAs of 30,666 genes accumulated in at least one captured
compartment (Supplemental Data Set 2.2) and 28,078 genes expressed in at least one endosperm
compartment (Supplemental Figure 2.6B). The latter is similar to the 33,084 genes that we
previously detected as expressed in the 8-DAP whole endosperm (Mathelier et al., 2014). The
difference is likely due to the use of different cut-off criteria for defining genes as expressed in
the two studies, and to the fact that we likely did not collect every portion of the 8-DAP
endosperm in our LCM analysis.

Application of a CS scoring method to the 30,666 expressed genes identified 13,009 genes that
were predominantly expressed in single compartments (Supplemental Data Set 2.3). These cell-
specific patterns were validated with 20 genes using in situ hybridization and by comparisons
with the expression patterns of previously reported endosperm-expressed genes (Supplemental
Figure 2.7 and Supplemental Table 2.6). In all 44 cases tested, the CS patterns closely matched
50

the experimentally observed patterns, indicating that our RNA-Seq data accurately reflect the
accumulation of endogenous mRNAs in the endosperm.

As expected, the PCA and SCC analysis showed that each captured compartment exhibited a
distinct mRNA population, with the maternal and filial compartments forming two separate
groups. Consistent with this, WGCNA identified both coexpression modules that were correlated
with multiple compartments (r ≥ 0.35) and modules that were specifically correlated with each of
the captured compartments (r ≥ 0.85 for one compartment and r < 0.35 for other compartments),
with the filial and maternal compartment-correlated coexpression modules falling into two nearly
distinct groups (Figure 2.4A). Notably, the SCC analysis also showed that the AL was more
closely related to the EMB than to any of the other endosperm compartments (ρ = 0.80; Figure
2.2B). Accordingly, WGCNA identified three coexpression modules (M1, M2, and M9) that
showed relatively high correlation (r ≥ 0.35) to both AL and EMB, with M9 also exhibiting a
modest correlation to CSE (r = 0.27) (Figure 2.4A). GO analysis indicated that these modules
were enriched for genes involved in mitotic cell proliferation (Supplemental Figure 2.13). This
suggests that partially overlapping sets of cell-proliferative programs distinguish the AL and
EMB from the rest of the filial compartments of an 8-DAP kernel.

GO enrichment of the coexpression modules correlated with single endosperm compartments


(M10, M12, M15, M17, and M18) were generally consistent with the presumptive functions of
each compartment as described previously (Olsen, 2001, 2004; Sabelli and Larkins, 2009;
Becraft and Gutierrez-Marcos, 2012). This indicates that these coexpression modules can be used
to decipher the key regulatory programs associated with cellular differentiation and related
functions of each compartment. In support of this notion, these modules were shown to contain
many cell type-specific genes that had been described previously and also identified by us using
the CS scoring method (Figure 2.4C; Supplemental Data Set 2.4). The latter, in effect, constitutes
a high-resolution atlas of spatially specific gene expression programs in the differentiating maize
endosperm.

We previously performed RNA-Seq with whole kernels or isolated endosperm at 8 temporal


stages and identified gene sets exhibiting temporal patterns of gene expression (Mathelier et al.,
51

2014). Among these were temporally up-regulated genes exhibiting relatively low expression at
early stages and relatively high expression at later stages. Analysis of the temporally up-
regulated genes among the WGCNA coexpression modules detected significant enrichment of
the up@6DAP and/or up@8DAP temporal programs in the endosperm compartment-specific
modules including the BETL-correlated M18 (Figure 2.4D). Similarly, a broad survey of the
endosperm-correlated modules for expression during development indicated that these modules
follow distinct dynamics of expression with an onset of activation at about 6 DAP, yet the ESR,
BETL, and AL (M15, M18, and M12, respectively) show a more rapid down-regulation pattern
as compared to the CZ and CSE (M17 and M10, respectively) (Supplemental Figures 2.9, 2.10,
and 2.11). These data suggest that although the endosperm compartments captured for this study
already exhibit characteristics of the differentiated state, the associated coexpression modules
nonetheless reflect active regulatory processes that may underlie continuous differentiation and
specialization of the relevant cell types. However, the fact that the previously identified temporal
programs lack spatial resolution within the endosperm limits our ability to accurately correlate
the spatial programs with the developmental dynamics of individual cell types at this time.
Therefore, further capture and analysis of endosperm compartments from multiple early stages of
endosperm will enable a more comprehensive understanding of these regulatory processes.

Analysis of endosperm-imprinted genes (Xin et al., 2013) in the WGCNA-identified


coexpression modules (Figure 2.4E) revealed that the BETL- and CZ-associated modules (M17
and M18, respectively) were enriched for PEGs, the CSE-correlated module M10 were enriched
for MEGs, while the ESR-associated module M15 was enriched for both PEGs and MEGs.
These observations indicate that the five endosperm compartment-correlated modules identified
in this study can be differentiated in part by memberships of imprinted genes that are likely a
reflection of compartment’s function. For example, BETL and CZ presumably function in the
nutrient transport from the maternal tissue to the inner endosperm cells including the developing
SE (Becraft, 2001; Sabelli and Larkins, 2009). The association of these compartments with the
expression of PEGs is consistent with the parental conflict model, which predicts opposite roles
for MEGs and PEGs in regulating nutrient allocation from the mother to offspring (Haig and
Westoby, 1989; Moore and Haig, 1991). On the other hand, the association of CSE with some
52

MEGs, and the dual association of ESR with PEGs and MEGs, suggest a more complex
relationship between gene imprinting and endosperm cell function.

The CSE, as a major sub-region within the SE, is responsible for the storage of most starch and
storage proteins in the endosperm (Olsen, 2001, 2004; Sabelli and Larkins, 2009).
Correspondingly, as revealed by the GO enrichment analysis, the CSE-correlated module M10
included many starch biosynthetic genes (Supplemental Figure 2.13). In addition, a number of
early zein genes, as well as the storage-program regulators O2 and PBF were also detected in
this module (Supplemental Data Set 2.4). Because the expression of a large subset of the storage-
protein genes, considered to be regulated by O2 and/or PBF, is not fully activated by 8 DAP,
only a few known targets of O2 and PBF were detected in our dataset. These included the 15-kD
beta-zein gene and the cyPPDK1 gene (encoding a cytoplasmic pyruvate orthophosphate
dikinase) known to be regulated by O2, and the 27-kD gamma-zein gene regulated by PBF
(Cord Neto et al., 1995; Gallusci et al., 1996; Maddaloni et al., 1996; Marzabal et al., 2008).
Therefore, although our data may not allow us to fully decipher the GRNs regulating the storage
function of the SE, they provide an insight into the early phase of the storage program activation.

The BETL-correlated module M18 contained numerous previously described BETL-specific


genes (Supplemental Data Set 2.4), including MRP-1 and seven genes regulated by MRP-1
(Gomez et al., 2002; Gutierrez-Marcos et al., 2004; Muniz et al., 2006; Gomez et al., 2009;
Muniz et al., 2010). The identification of GATA-containing motifs among a subset of M18 genes
(Figure 2.5) and confirmation of binding of MRP-1 to these sequences using Y1H assays
(Supplemental Table 2.11) allowed us to propose a 93-gene, direct-target regulatory module for
MRP-1 (Figure 2.6). Functional annotation of the 93-gene regulatory module indicates a diverse
array of gene functions activated by MRP-1 including putative signaling and nutrient transport
(Supplemental Table 2.11). The larger M18 gene set also exhibits a wide range of putative
functions including those expected for a transfer cell layer (Supplemental Figure 2.13). These
gene functions are likely sufficient to support BETL differentiation, as a recent study showed
that the ectopic expression of MRP-1 via an aleurone-specific gene promoter was capable of
inducing differentiation of an ectopic BETL in the aleurone albeit in a transient manner (Gomez
et al., 2009). On the other hand, ectopic activation of these genes may not produce viable cells
53

outside the endosperm context, as attempts with overexpressing MRP-1 in maize using
ubiquitously expressing promoters produced no transformants (Gomez et al., 2009). Further
understanding of the MRP-1-regulated network and the associated gene functions will likely
require characterization of complete or partial loss-of-function mutants.

The gene set activated by MRP-1 is likely significantly larger than the 93-gene set discussed
above for two reasons. First, our Y1H assays showed activation of 7 genes that lack motifs
identified in our MEME analysis (Supplemental Figure 2.16 and Supplemental Table 2.12),
indicating that MRP-1 binds to additional sequences not identified in our MEME analysis and
that M18 contains additional MRP-1-regulated genes. Second, MRP-1 likely directly regulates
six TFs, including MYBR24, MYBR33, GATA7, GATA33, and EREB137, and a DBB-family
TF (Supplemental Table 2.11). Each of these TFs may activate a gene set within M18.
Additional protein-DNA interaction studies such as electrophoretic mobility shift assays, and
transient directed-expression assays would be necessary to further characterize the interaction
between MRP-1 and the full spectrum of its direct targets. Thus, the entire regulatory module
activated by MRP-1 (i.e., both directly and indirectly) probably encompasses a much larger
proportion of the M18 module than what we report here. Studies devoted to identifying the full
spectrum of MRP-1 binding sites and to the identification of the target genes activated by the
TFs activated by MRP-1 are in progress to fully characterize this module.

Furthermore, it is notable that in addition to MRP-1 itself, M18 contains at least 48 coexpressed
TF genes including 6 MYBR genes (Supplemental Data Set 2.4). The latter may regulate
overlapping gene sets with MRP-1, possibly through binding to the identified GATA-containing
motifs. Limitations of the de novo cis-motif detection approaches utilized here and an absence of
an extensive cis-motif-TF database for plants have precluded identification of additional TF
targets in M18. Therefore, further approaches including directed Y1H assays for specific TF-
target interactions in combination with transient expression studies of the TFs, and analysis of
any available TF gene mutants will be required to determine the full extent of the BETL GRN.

In summary, our dataset provides a high-resolution atlas of gene expression in differentiating


endosperm compartments and maternal compartments of an early maize kernel. This dataset
54

provides insights into the functions of the endosperm cell types and into the coexpressed gene
sets that establish the differentiated states and functions of these cell types. Furthermore, as
exemplified by our initial analysis of the MRP-1 regulatory module, this dataset can be used as a
starting point to dissect the modules regulating endosperm cell differentiation. The analyses
provided here constitutes a significant step toward the identification of GRNs that regulate maize
endosperm cell differentiation and determine its function.

2.5 METHODS

Plant materials and growth

Plants of the reference maize (Zea mays) genotype, B73, were grown under greenhouse
conditions (16-h day) at the University of Arizona during April-July 2012 and self-pollinated to
obtain 8-DAP kernels (Mathelier et al., 2014) for LCM. The kernels for morphological analysis
shown in Supplemental Figure 2.1 were collected during October-November 2013. Kernel
compartments were delineated for capture using tissue sections obtained from Farmer’s fixed
(see below), or from paraformaldehyde-fixed material that were further stained with Toluidine
Blue as described previously (Drews, 1998). The tissue sections for each biological replicate of a
given compartment were captured from multiple kernels of a single ear (Supplemental Table
2.1).

Laser-capture microdissection, RNA isolation, and cDNA synthesis and amplification

Kernels were harvested from plants mid-day and cut at the pedicel, punctured through the
pericarp, vacuum infiltrated with cold Farmer’s fixative (ethanol:glacial acetic acid, 3:1) (Kerk et
al., 2003) and stored in cold fixative overnight. Fixed kernels were then dehydrated in a graded
ethanol series, cleared in n-Butanol, embedded in Paraplast X-tra (McCormick Scientific Leica)
using microwave (Takahashi et al., 2010), cut to 10-μm sections, and mounted on PEN-coated
slides. Shortly before capture, sections were deparaffinized in Xylenes and air dried (Takahashi
et al., 2010). Individual cell types or kernel compartments were captured directly into an aliquot
of the ARCTURUS PicoPure RNA extraction buffer (Applied Biosystems) using a Leica
LMD6500 Laser Microdissection System (Leica Microsystems). RNA was extracted following
manufacturer’s suggested protocol (ARCTURUS PicoPure kit, Applied Biosystems), checked
for quality, DNase treated using TURBO DNase (Life Technologies), and further purified using
55

the ARCTURUS PicoPure columns (Applied Biosystems). For each of the 22 samples
(Supplemental Table 2.1), 10 ng purified RNA was used for cDNA synthesis and amplification.
cDNA synthesis with oligo(dT) and random primers, and cDNA amplifications were carried out
using an Ovation RNA-Seq System V2 kit (Nugen Technologies Inc.) following manufacturer’s
protocols with minor modifications. The quality and profile of the RNA and amplified cDNA
samples were checked on an Agilent 2100 Bioanalyzer (Agilent Technologies) using an Agilent
RNA 6000 Pico Kit (Agilent Technologies) and an Agilent High Sensitivity DNA Kit (Agilent
Technologies), respectively.

RNA-Seq library construction and sequencing

Construction and sequencing of RNA-Seq libraries were performed at the University of Arizona
Genetics Core. Using an Illumina TruSeq DNA Sample Preparation Kit v2 (Part # 15026486
Rev. A, Illumina, Inc., San Diego, California), nearly 1 µg amplified cDNA for each of the 22
samples was used to generate multiplexed RNA-Seq libraries (mean size 350-380 bp, including
120-bp adapters) by following the manufacturer’s suggested procedures. The samples were
sequenced in batches on four flow cell lanes of an Illumina HiSeq 2000 platform using a TruSeq
SBS Kit v3 (Cat # FC-401-3001) to produce 2×100-nt paired-end reads. Two of the four lanes
each contained libraries for a single replicate of the captured endosperm compartments (AL,
BETL, ESR, CSE, and CZ); the third lane contained libraries for a single replicate of the
endosperm compartments plus the library for a single replicate of the maternal compartment PC;
and the fourth lane contained libraries for EMB (three replicates) and the other three maternal
compartments (NU, PE, and PED, one replicate each). After the raw reads were generated,
adapter sequences were trimmed using the Trimmomatic program (Bolger et al., 2014). Quality
of the trimmed reads were checked using the FastQC program (Andrews, 2010).

Reads mapping and analysis

RNA-Seq reads were aligned to the maize reference genome version 3 (B73 RefGen_v3)
(Hubbard et al., 2002) using TopHat v2.0.9 (Trapnell et al., 2009). Intron length was set to 30 to
8000 nt while maximum number of mismatches per read was set to 3. To eliminate the effect of
reads mapping in intergenic and/or repeated genomic regions on the estimation of effective
library size that may be caused by the cDNA amplification method, reads mapped to exonic
56

regions were extracted using the intersect function of BEDTools v2.17.0 (Quinlan and Hall,
2010), and were provided as input to Cufflinks v2.1.1 (Trapnell et al., 2010) for normalization
and estimation of gene expression level. The multi-mapped reads correction and fragment bias
correction options of Cufflinks were used. Gene expression levels were reported as fragments per
kilobase of transcript per million mapped reads (FPKM) (Mortazavi et al., 2008). The upper and
lower bound FPKM values (FPKM_conf_hi and FPKM_conf_lo, respectively) for the 95%
confidence interval of each gene were also provided by Cufflinks. A gene was defined as
expressed in a sample if the FPKM_conf_lo was greater than zero. Information regarding maize
genome annotation used in these analyses was obtained from Ensembl Plants
(plants.ensembl.org, Release 19).

Spearman Correlation Coefficient (SCC) analysis (Zar, 1972; Hollander et al., 2013) was used to
quantify the reproducibility of data between the triplicates of endosperm compartments and
EMB. SCCs was calculated from log2-transformed FPKM values [i.e., log2 (FPKM+1)] of the
expressed genes. Based on the high correlation of gene expression profiles among the replicated
samples, exonic reads were merged to create a union of each triplicate and FPKM were
recalculated for the 6 merged samples. With the resulting FPKM data (log2-transformed) of the
expressed genes (FPKM_conf_lo > 0 after re-normalization), principal component analysis
(PCA) and SCC analysis were used to compare gene expression profiles among all ten
compartments. The prcomp and cor.test functions in R were used for PCA and SCC analysis,
respectively.

Identification of compartment-specific gene sets

The genes specifically expressed in each kernel compartment were identified using a
compartment specificity (CS) scoring algorithm that compares the expression level of a gene in a
given compartment with its maximal expression level in the other nine compartments (Ma and
Wang, 2012; Ma et al., 2014). For a given gene i, its expression values in ten compartments are

denoted as EVi = ( E1 , E2 , E3 ,..., E10 ) , the CS score of this gene in compartment j is defined as:
i i i i

max Eki
CS (i, j ) = 1 − , where 1 ≤ k ≤10, k ≠ j. Thus, CS scores range from 0 to 1, and the higher
E ij
57

the CS score of a gene for a compartment, the more likely the gene is specifically expressed in
that compartment.

In situ hybridization localization of mRNAs

Staged kernels were obtained from B73 plants grown at the University of Utah. Kernels were
harvested, fixed, processed, and hybridized to probes as previously described (Mathelier et al.,
2014). Primers used to generate the probe clones are listed in Supplemental Table 2.13.

Identification of coexpression modules

The R package Weighted Gene Coexpression Network Analysis (WGCNA) (Zhang and Horvath,
2005; Langfelder and Horvath, 2008) was used to identify modules of highly correlated genes
based on the FPKM data. Genes with low FPKM (mean FPKM < 1 for ten compartments) or low
coefficient of variation of FPKM (CV < 1 among ten compartments) were filtered out. Using the
FPKM values of the remaining 9,361 genes, a matrix of pairwise SCCs between all pairs of
genes was created, and transformed into a matrix of connection strengths (an adjacency matrix)
1+ correlation β
by raising the correlation matrix to the power β =12 (connection strength = ( )
2
). The power β was interpreted as a soft-threshold of the correlation matrix. The resulting
adjacency matrix was then converted to a topological overlap (TO) matrix by the TOMsimilarity
algorithm. Genes were hierarchically clustered based on TO similarity. The Dynamic Tree Cut
algorithm was used to cut the hierarchal clustering tree, and modules were defined as branches
from the tree cutting. Modules with fewer than 30 genes were merged into their closest larger
neighbor module. Each module was summarized by the first principal component of the scaled
module expression profiles (referred to as module eigengene, ME). Module membership (MM,
also known as module eigengene-based connectivity kME) of a gene to a given module was
calculated as Pearson Correlation Coefficient (PCC) between the expression levels (FPKMs, in
ten compartments) of the gene and the ME of the module using the signedKME algorithm.
Finally, genes were reassigned using the moduleMergeUsingKME algorithm to ensure each gene
possesses the highest MM in its own assigned module. Module-compartment associations were
quantified by PCC analysis where each module was represented by its ME, each compartment
58

was represented with a numeric vector with “1” for the compartment of interest, and “0” for all
other compartments.

Computational validation of module robustness

Average TO for each identified coexpression module were calculated and compared to the
average TO of modules of the same size generated by randomly assigning the 9,361 tested genes
to 18 modules; 100,000 permutations of randomly sampled modules were tested. The observed
modules were considered robust if the average TOs were significantly higher than the randomly
generated modules (P < 10-5).

Visualization of hub genes

Genes with highest degree of connectivity within a module are referred to as intramodular hub
genes (Langfelder and Horvath, 2008). The top 200 connections (based on topological overlap)
among the top 100 genes in each module ranked by kME was visualized by VisANT (Hu et al.,
2004).

Gene annotation and functional enrichment analysis

Locus names and functional annotation of maize genes were obtained from Ensembl Plants. The
recently annotated maternally expressed gene (MEG) family members (Xiong et al., 2014) were
also incorporated. Annotation of TF family members were based on information from Plant
Transcription Factor Database v3.0 (Jin et al., 2014) and GrassTFDB of GRASSIUS (Gray et al.,
2009; Yilmaz et al., 2009). Annotation of zein genes in the B73 genome were based on the
information as summarized by Chen et al. (2014a). Putative functions of genes of interest were
identified and inspected manually. cDNA sequences of the longest isoform of each gene were
obtained from Ensembl Plants and used for homology searches against the NCBI non-redundant
(nr) and Swissprot protein databases, respectively, using the BLASTX program. Only the top
five hits for each gene (E-value < 10-6) were considered in our analysis of putative gene
functions.

Gene Ontology (GO) term enrichment analyses of the WGCNA-identified coexpression modules
were performed using a modified Fisher's exact test in Blast2GO software (FDR < 0.05). GO
59

annotations for maize genes were obtained from Gramene (gramene.org, Release 40). For each
module, only protein-coding genes (i.e., transposable elements, miRNA genes, and pseudogenes
excluded) were subject to GO enrichment analysis, with the most specific biological processes
reported.

Identification of cis-motifs

MEME program was used to identify de novo motifs in the promoter regions of genes in each
coexpression module. We defined the promoter regions as 1 kb upstream and 500 bp
downstream of transcription start sites, and obtained the genomic sequences using a customized
Perl script. For each module, 10 motifs (Motifs 1 through 10) were reported by MEME, and the
ones with E-value higher than 10-6 were excluded manually. Using the TOMTOM motif
comparison tool, the resulting motifs were aligned with motifs in the JASPAR CORE Plantae
database (Mathelier et al., 2014) to identify significantly similar known cis-motifs (q-value <
0.05). A customized Perl script was used to identify the genes that contain at least one of the four
sub-motifs shown to be bound by MRP-1 in the Y1H assays.

Yeast one-hybrid assays

The Matchmaker Gold kit from Clontech (clontech.com) for yeast one-hybrid assays was used to
validate MRP-1-target sequence interactions following manufacturer’s procedures with
modifications, including the use of the yeast CUP1 promoter (Etcheverry, 1990) in place of the
yeast ADH promoter to drive MRP-1 expression in yeast cells. The plasmid containing the MRP-
1 coding region driven by the CUP1 promoter was generated as follows. The CUP1 promoter
was provided in plasmid pMB465 by Marcus Babst. A Not I/Sac I fragment containing the
CUP1 promoter fragment was subcloned into yeast vector pRS425 to make pCUP425. The
MRP-1 open reading frame was generated by PCR using primers MRP1-FEco
(GGCCGAATTCAATCCCAACTTCAACAGTGTG) and MRP1-RBam
(GGCCGGATCCTCGGTTATATATCTGGCTCTCC). The resulting PCR product was cloned
into pGADT7 (Clontech) using the Eco RI and Bam HI restriction sites introduced during PCR.
The resulting plasmid was called MRP1/pGADT7. Using primers NoAD-FNot
(GATCGCGGCCGCATGGAGTACCCATACGACG) and PGAD-R2Sal
(GATCGTCGACGGAATATGTTCATAGGGTAG), a fragments containing an HA tag, the
MRP-1 open reading frame, and the ADH1 terminator was amplified. The resulting PCR product
60

was cloned into pCUP425 through the use of the Not I and Sal I restriction sites introduced
during PCR. The final plasmid was called MRP1/pCUP425. The identical plasmid lacking the
MRP-1 coding region was called pCUP425.

The motif:AUR1-C constructs included 147 bp of sequence upstream of the translational start
codon of TCRR-1 (GRMZM2G016145). The motif-promoter fragments were generated by PCR
amplification using the primers listed in Supplemental Table 2.14. The promoter:AUR1-C
constructs included 257-1187 bp of sequence upstream of the translational start codon. The
promoter fragments were generated by PCR amplification using the primers listed in
Supplemental Table 2.15. The resulting PCR products were cloned into pAbAi (Clontech)
through the use of unique 5’ and 3’ restriction sites introduced during PCR (Supplemental Tables
2.14 and 2.15). The resulting plasmids then were then integrated into the yeast genome of strain
Y1HGold following the manufacturer’s recommended protocol (Clontech).

We introduced the MRP1/pCUP425 plasmid into each of the promoter:AUR1-C and


motif:AUR1-C yeast strains; these were called +MRP-1 strains in the Y1H figures. To generate
control strains, we also introduced pCUP425 into each of the promoter:AUR1-C yeast strains and
motif:AUR1-C; these were called -MRP-1 (or empty vector) strains in the Y1H figures. Yeast
transformations were performed using the lithium acetate procedure.

Yeast growth assays were performed as follows. Equal numbers of cells from single colonies
were spotted in a 1:5 dilution series (~625 cells/spot, ~125 cells/spot, ~25 cells/spot, and ~5
cells/spot) onto plates. The plate media was deficient in leucine, and contained 750-2000 ng/ml
AbA (variation in the concentration of AbA used was due to manufacturer batch variability) and
0 or 0.1 mM CuSO4. As a growth control, each cell suspension was also spotted onto plates
lacking AbA. This procedure was followed for both the +MRP1 and -MRP strains. The plates
containing the spotted yeast were incubated at 30°C for ~48 hours and then images of the yeast
cells were captured. This procedure was carried out with two colonies per strain for each
experiment, and each experiment was performed twice on separate days with independently
transformed strains and independently prepared plates (i.e., four independent colonies tested per
strain). In all cases, all four colonies exhibited very similar growth patterns.

Yeast growth was scored as follows. Cell growth fell into three general categories. In the first
category, the +MRP-1 and -MRP-1 strains both grew well on plates lacking AbA but failed to
61

grow on plates containing AbA. This growth pattern indicated no binding of MRP-1 to the test
sequence and these strains were scored as negative. In the second category, the -MRP-1 strains
grew well on plates lacking AbA but failed to grow on plates containing AbA, and the +MRP-1
strains grew equally well (or nearly so) on plates lacking and containing AbA. This growth
pattern indicated strong binding of MRP-1 to the test sequence and these strains were scored as
positive. In the third category, growth of the +MRP-1 strains was reduced on plates containing
AbA relative to plates lacking AbA but this growth was significantly greater than that of the –
MRP-1 strains on plates containing AbA. This growth pattern suggested weak binding of MRP-1
to the test sequence and these strains were scored as weak positives.
62

Figure 2.1. Profiles of sequenced RNAs from the captured filial and maternal
compartments of 8-DAP maize kernel.

(A) Graphic representation of an 8-DAP maize kernel showing the relative position of the 10
captured filial and maternal compartments used for RNA sequencing. (B) Numbers of TF genes,
non-TF protein-coding genes, miRNA genes, transposable elements, and pseudogenes expressed
in the 10 captured compartments. (C) Proportions of genes expressed at different levels (based
on FPKM) in the 10 kernel compartments.
63

Figure 2.2. Relationship between the RNA populations obtained from the filial and
maternal compartments of 8-DAP kernel.

(A) Principal component analysis of genes expressed in the captured kernel compartments.
Principal components one through three (PC1, PC2, and PC3) collectively explained 61.9% of
the variance in the mRNAs obtained from the ten compartments. (B) Spearman correlation
coefficient (SCC) analysis of the mRNA data for the ten kernel compartments using log2-
transformed FPKM values of the 30,665 expressed genes. The hierarchical clustering
dendrogram was inferred by applying (1 - SCC) as distance function.
64

Figure 2.3. Compartment-specific gene sets identified using the CS scoring method.

(A) Heat map of scaled FPKM values of the 13,009 compartment-specific genes identified in all
10 kernel compartments. (B) Numbers of TF and non-TF genes in each compartment-specific
gene set.
65

Figure 2.4. Gene coexpression modules detected using WGCNA.

(A) Heat map of the correlations between detected modules (M1-M18) and kernel compartments
hierarchically clustered based on Euclidean distance. The PCC values are quantitative indicators
of relative expression levels of all genes in each module. (B) Numbers of TF and non-TF genes
in each coexpression module. (C) Relationships of the compartment-specific gene sets with the
corresponding compartment-correlated coexpression modules obtained using WGCNA. The heat
map indicates P values (-log10) of hypergeometric tests of over-representation of genes in a given
tested pair of gene sets. Compartment-specific gene sets are noted on the x axis and the
66

corresponding WGCNA coexpression modules (those with a specific pattern for a single
compartment) on the y axis. Boxes contain the numbers of overlapping genes and proportions (in
parentheses) of these genes in the WGCNA-identified modules. (D) Relationships of the
temporal gene sets (Mathelier et al., 2014) with the coexpression modules obtained using
WGCNA. The heat map indicates P values (-log10) of hypergeometric tests of over-
representation of genes in a given tested pair of gene sets. Temporal gene sets are noted on the x
axis and all WGCNA coexpression modules on the y axis. Boxes contain the numbers of
overlapping genes. Numbers of genes in each temporal gene set: up@2DAP, 54; up@3DAP, 68;
up@4DAP, 92; up@6DAP, 523; up@8DAP, 1,402; up@10DAP, 552; up@12DAP, 241. (E)
Relationships of the imprinted gene sets (Xin et al., 2013) with the coexpression modules
obtained using WGCNA. The heat map indicates P values (-log10) of hypergeometric tests of
over-representation of genes in a given tested pair of gene sets. Imprinted gene sets (Xin et al.,
2013) are noted on the x axis and all WGCNA coexpression modules on the y axis with the total
number of genes in each imprinted gene set indicated in parentheses. Boxes contain the numbers
of overlapping genes. Numbers of genes in each imprinted gene set: 7-DAP MEG, 37; 10-DAP
MEG, 185; 15-DAP MEG, 15; 7-DAP PEG, 80; 10-DAP PEG, 50; 15-DAP PEG, 48.
67

Figure 2.5. De novo GATA-rich sequence motifs over-represented in the upstream


sequences of the coexpression module M18 genes identified by the MEME program.

(A) Motifs 3, 6 and 8 enriched among the M18 genes. (B) Position frequencies of the same
motifs determined as number of motifs per 50-bp bins in the upstream gene regions.
68

Figure 2.6. A regulatory module of the MRP-1-regulated GRN based on the analysis of
MRP-1-bound sub-motifs.
69

(A) A network of the 93 M18 genes associated with at least one MRP-1-bound sub-motifs
visualized using Cytoscape (Shannon et al., 2003). Thickness of the arrows indicates the number
of motifs associated with each gene. (B) Expression pattern of the MRP-1 target genes in seed
tissues as compared to vegetative and reproductive tissues based on the log2-transformed RPKM
data from Chen et al. (2014a). The fraction of MRP-1 target genes that have available expression
data is indicated in parentheses. Blue line indicates the expression of MRP-1 itself. The selected
data include vegetative (Veg) tissues shoots (Sh), roots (R), Leaf (L), shoot apical meristem
(SAM, Replicate 1); reproductive (Rep) tissues ear (E), tassel (T, Replicate 1), pre-emergence
cob (C), silk (Si), Anther (A), ovule (O), pollen (P); and whole kernels, endosperm, and embryos
of different developmental stages (in days after pollination, DAP). (C) Expression level of genes
within the MRP-1-regulated regulatory module in BETL as compared to all other genes in M18.
Asterisk indicates P < 10-5 (unpaired t-test). (D) Venn diagram showing overlaps between the
genes that are up-regulated at 6 DAP (Mathelier et al., 2014) with those of the M18 coexpression
module, and those with detected MRP-1-binding sites. Asterisk indicates hypergeometric P < 10-
5
. The boxes in (B) and (C) represent the interquantile range, green lines the median, red dots the
mean, and whiskers 1.5 times the interquantile range.
70

Table 2.1. Results of Y1H assays for binding of MRP-1 to the MEME-identified sequence
motifs of M18.
Sub-motif Sequencea Y1H Resultsb
3a/8a TAGATAGATAGA +
3b TAGATATATAGA –
3c TAGATAAATAGA –
6a/8b TAGATATAGATA +
6b TAGATATATATA –
6c TAAATATAAATA –
6d TAGATATAGAAA –
6e TAGATATAAATA –
6f TAAATATAAAAA –
6g TAAATATATAAA –
6h TATATATATATA –
6i TAAATATATATA –
6j TAGATAAAAAAA –
6k TAAATATAGAAA –
6l TAGATATATAAA –
6m TAGATATAAAAA –
8c TAGATAGAGATA +
8d TAGAGAGAGAGA –
8e TAGATATAGAGA –
8f TAGATAGATATA +
8g AAGATAGAGAGA –
8h TAGATAGAGAGA –
aFlanking sequence information is provided in Supplemental Table 2.14.
bScoring: + indicates strong positive, – indicates negative. Images of the Y1H growth assays are shown in

Supplemental Figure 2.15.


71

2.6 SUPPLEMENTAL DATA

Supplemental Figure 2.1. Structure of an 8-DAP B73 maize kernel.

(A) Image of a Toluidine Blue-stained section of 8-DAP kernel showing different compartments.
Multiple images of a single kernel section obtained using a low-magnification objective lens
were merged using Adobe Photoshop CS5, and background was subsequently removed in Adobe
Illustrator CS6 to generate the image. (B) through (H) Details of the captured kernel
compartments. Colored dashed lines indicate the margins of excised compartments in a typical
laser-capture experiment. Bars: (A) and (B), 400 µm; (C) and (D), 50 µm; (E), 400 µm; (F), 200
µm.
72

Supplemental Figure 2.2. Representative images of kernel compartments that were marked
and collected for the LCM RNA-Seq analyses reported in this study.

(A) Regions of the CSE and the captured regions of CZ; (B) CZ and the captured regions of
ESR; (C) ESR; (D) BETL, and the captured regions of ESR and PC; (E) Regions of AL; (F) The
EMB including suspensor and the captured regions of ESR; (G) Regions of NU; (I) Region of
PE; and (I) Regions of PED. Bars: (A) and (B), 400 µm; (C), 200 µm; (D), 400 µm; (E), 100 µm;
(F), 200 µm; (G), (H) and (I), 400 µm.
73

Supplemental Figure 2.3. Representative quality assessments of LCM-derived RNAs used


in sequencing.

Electropherograms of sample RNAs extracted from the captured kernel compartments prior to
DNase treatment. RIN, RNA Integrity Number, is a measure of RNA quality (Agilent
Technologies). Controls refer to known concentrations of total RNA obtained from 2-DAP
kernels.
74

Supplemental Figure 2.4. Reproducibility of RNA-seq reads for each triplicate of the filial
compartments.

Reproducibility of RNA-seq reads for each triplicate of the filial compartments. Log2-
transformed FPKM values of the 29,369 genes expressed in at least one of the 22 sequenced
samples are shown as scatter plots and were used as input for the Spearman correlation
coefficient (SCC) analysis. The red diagonal line in each scatter plot denotes equal FPKMs
between two samples.
75

Supplemental Figure 2.5. Distribution profile of RNA-Seq reads along the length of the
gene models.

Reads of all 22 sequenced samples were pooled and exonic reads were analyzed for depth of
coverage using the available reference gene models to show the combination of Oligo(dT) and
random primers provided sufficient coverage of the sequenced cDNAs.
76

Supplemental Figure 2.6. Profiles of sequenced RNAs from the captured kernel
compartments.

(A) Number of expressed genes detected in one or more kernel compartments. Among the total
30,665 genes that were expressed in at least one compartment, 10,725 were detected in all
compartments while expression of 4,293 genes was restricted to single compartments. (B) Venn
diagram of the numbers of genes expressed in EMB, expressed in at least one captured
compartment of endosperm (END), or expressed in at least one maternal compartment (MAT).
77

Supplemental Figure 2.7. Validation of the compartment-specific expression patterns using


in situ hybridization localization of the mRNAs.

For each panel, the gene ID/locus name is indicated. (A) through (N) BETL-specific genes. (O)
An ESR-specific gene. (P) and (Q) CSE-specific genes. (R) through (T) AL-specific genes.
Stage of kernels: (A), 6 DAP; (B) through (I), 7 DAP; (J) through (N), 8 DAP; (O) 7 DAP; (P)
through (T), 10 DAP. Bars: 2 mm.
78

Supplemental Figure 2.8. Z-score plots showing expression profiles of all genes in the
WGCNA-generated coexpression modules M1-M18.
79

Supplemental Figure 2.9. Expression pattern of genes in the endosperm-associated modules


identified by WGCNA based on previously reported RNA-Seq data.

(A) Log2-transformed normalized read counts data from Li et al. (2014). (B) Log2-transformed
Reads per kilobase of transcript per million mapped reads (RPKM) data from Chen et al.
(2014a). The selected data included kernels and endosperm of different developmental stages (in
days after pollination, DAP). The associated kernel compartment and fractions of genes in each
module with available data are indicated in parentheses.
80

Supplemental Figure 2.10. Expression pattern of genes in the endosperm-associated


modules identified by WGCNA based on RNA-Seq data from Chen et al. (2014a).

The selected data included vegetative (Veg) tissues shoots (Sh), roots (R), Leaf (L), shoot apical
meristem (SAM, Replicate 1); reproductive (Rep) tissues ear (E), tassel (T, Replicate 1), pre-
emergence cob (C), silk (Si), Anther (A), ovule (O), pollen (P); and whole kernels, endosperm,
and embryos of different developmental stages (in DAP). The associated kernel compartment
and fractions of genes in each module that have available data are indicated in parentheses.
81

Supplemental Figure 2.11. Expression pattern of genes in the WGCNA coexpression


modules (M1, M2, and M9) associated with both AL and EMB (in comparison to the other
two EMB-associated modules M7 and M8) based on RNA-Seq data from Chen et al.
(2014a).

The selected data included the same tissues as Supplemental Figure 2.10. The associated kernel
compartment and fractions of genes in each module with available data are indicated in
parentheses.
82

Supplemental Figure 2.12. Relationships of new sets of allele-biased genes with the
coexpression modules obtained using WGCNA.

The new allele-biased gene sets were identified using the normalized read data from Xin et al.
(2013) by applying less stringent criteria, namely, 75% of the maternal reads and 55% of the
paternal reads of the SNP associated reads (with a minimum 20 reads) were defined as
maternally and paternally biased genes, respectively. The heat map indicates P values (-log10) of
hypergeometric tests of over-representation of genes in a given tested pair of gene sets. Allele-
biased gene sets are noted on the x axis and all WGCNA coexpression modules on the y axis
with the total number of genes in each set indicated in parentheses. Boxes contain the numbers of
overlapping genes.
83

Supplemental Figure 2.13. Enrichment of biological processes in the filial compartment-


correlated modules.

The heat map indicates enriched GO terms detected for modules M1, M2, M8, M9, M10, M12,
M15, M17, and M18 based on FDR (-log10). The grey boxes represent GO terms that are not
significantly enriched.
84

Supplemental Figure 2.14. Visualization of the five endosperm compartment-correlated


coexpression modules using the VisANT program.

The top 200 connections among top 100 intramodular hub genes (based on kME) are shown for
modules M10, M12, M15, M17, and M18 corresponding to CSE, AL, ESR, CZ, and BETL,
respectively. Either a locus name (if available) or a gene ID is used to identify each gene. The 10
to 11 genes with the highest number of connections are shown in a bigger node size with red
names/gene IDs.
85
86

Supplemental Figure 2.15. Yeast one-hybrid assays for binding of MRP-1 to the sequence
motifs listed in Table 2.1.

For each motif, two yeast strains were tested, one lacking MRP-1 (-MRP-1) and one expressing
MRP-1 (+MRP-1). Each strain was spotted onto plates containing (+AbA) and lacking (-AbA)
Aureobasidin A. Each strain was spotted in a 1:5 dilutions series.
87
88
89

Supplemental Figure 2.16. Yeast one-hybrid assays for binding of MRP-1 to the promoters
of the genes listed in Supplemental Table 2.18.

For each promoter, two yeast strains were tested, one lacking MRP-1 (-MRP-1) and one
expressing MRP-1 (+MRP-1). Each strain was spotted onto plates containing (+AbA) and
lacking (-AbA) Aureobasidin A. Each strain was spotted in a 1:5 dilutions series.
90

Supplemental Figure 2.17. Expression pattern of the 93 genes containing MRP-1-binding


sub-motifs based on RNA-Seq data from Chen et al. (2014a).

Genes were hierarchically clustered based on Euclidean distance. The selected data included the
same tissues as Supplemental Figure 2.10.
91

Supplemental Table 2.1. The origins, quality and quantities of RNAs isolated using laser-capture microdissection.
RNA quality as RNA
# kernels # sections used Total area (mm2) Total amount of RNA (ng)
Integrity Numbers (RINs)
Rep. 1 Rep. 2 Rep. 3 Rep. 1 Rep. 2 Rep. 3 Rep. 1 Rep. 2 Rep. 3 Rep. 1 Rep. 2 Rep. 3 Rep. 1 Rep. 2 Rep. 3
CSE 9 8 7 26 23 23 36.64 34.69 33.67 7.10 6.80 6.10 236.98 183.70 203.13
CZ 9 7 7 95 137 95 22.04 33.53 35.58 6.00 6.20 5.10 143.65 156.20 195.80
ESR 20 9 9 150 246 208 6.86 10.69 7.25 4.90 6.50 4.20 343.18 191.40 314.60
BETL 6 6 7 150 149 150 15.75 14.77 19.34 4.10 4.90 3.50 556.92 200.20 422.40
AL 8 6 6 57 68 63 10.77 6.43 8.73 5.30 6.30 5.00 242.01 244.20 334.40
EMB 11 3 3 89 61 87 8.86 11.15 9.52 6.00 5.50 4.00 101.20 235.40 458.04
PC - - 7 - - 118 - - 6.09 - - 2.90 - - 25.62
NU - 14 - - 110 - - 27.15 - - 4.20 - - 171.60 -
PE - 11 - - 10 - - 19.70 - - 4.80 - - 193.60 -
PED - 12 - - 41 - - 19.39 - - 5.00 - - 411.40 -
Control (10 ng) - - - - - - - - - 8.40 9.00 9.00 - - -
Control (5 ng) - - - - - - - - - 8.40 9.30 9.10 - - -
Control (2.50 ng) - - - - - - - - - 8.80 9.40 9.20 - - -
Control (1.25 ng) - - - - - - - - - 8.10 8.10 9.30 - - -
92

Supplemental Table 2.2. Summary statistics of RNA-Seq reads and mapping.


All mapped reads Mapped to exons
Samplesa Total # reads %
# % #
(of mapped)
CSE-1 10,222,204 8,861,426 86.7 3,299,695 37.2
CSE-2 31,450,130 25,808,488 82.1 8,616,333 33.4
CSE-3 33,686,796 27,230,237 80.8 8,924,753 32.8
CZ-1 29,475,616 24,887,580 84.4 9,755,319 39.2
CZ-2 36,099,924 28,111,041 77.9 10,407,325 37.0
CZ-3 31,555,056 24,171,723 76.6 7,456,103 30.8
ESR-1 40,439,606 34,729,514 85.9 11,344,960 32.7
ESR-2 33,883,182 26,160,199 77.2 8,323,640 31.8
ESR-3 32,478,306 29,137,023 89.7 8,238,762 28.3
BETL-1 17,109,048 14,341,367 83.8 4,599,768 32.1
BETL-2 27,823,360 21,678,667 77.9 6,237,856 28.8
BETL-3 29,420,758 20,380,515 69.3 6,319,413 31.0
AL-1 11,496,690 9,247,655 80.4 3,581,977 38.7
AL-2 30,289,840 23,899,741 78.9 7,937,483 33.2
AL-3 21,498,076 16,487,384 76.7 5,114,948 31.0
EMB-1 23,091,244 12,142,139 52.6 4,792,036 39.5
EMB-2 21,651,262 13,972,219 64.5 4,075,913 29.2
EMB-3 22,895,586 11,973,058 52.3 3,284,420 27.4
PC 28,140,418 21,025,363 74.7 5,095,299 24.2
NU 24,447,238 15,850,406 64.8 4,363,845 27.5
PE 23,581,668 16,044,397 68.0 3,549,164 22.1
PED 21,838,282 13,136,224 60.2 3,363,514 25.6
a
Biological replicates are indicated as -1, -2 and -3.
93

Supplemental Table 2.3. Number of genes expressed in each of the ten compartments.

Protein- Transposable TFs


Compartment Pseudogenes miRNAs Total #
coding genes elements # %
CSE 19,928 664 544 0 1,137 5.4 21,136
CZ 20,775 819 646 0 1,172 5.3 22,240
ESR 22,228 925 699 1 1,305 5.5 23,853
BETL 19,281 704 550 0 1,105 5.4 20,535
AL 20,646 680 578 0 1,173 5.4 21,904
EMB 21,392 685 623 3 1,343 5.9 22,703
PC 15,605 354 305 0 964 5.9 16,264
NU 15,173 397 340 0 899 5.7 15,910
PE 17,615 417 369 1 1,057 5.7 18,402
PED 16,740 366 338 0 1,040 6.0 17,444
94

Supplemental Table 2.4. Number of genes expressed at different FPKM levels in the ten
compartments.

Compartment FPKM < 2 2 ≤ FPKM < 10 10 ≤ FPKM < 200 FPKM ≥ 200
# % # % # % # %
CSE 14,501 47.29 7185 23.43 8,509 27.75 470 1.53
CZ 14,438 47.08 7570 24.69 8,237 26.86 420 1.37
ESR 13,859 45.19 7244 23.62 9,315 30.38 247 0.81
BETL 15,585 50.82 6104 19.91 8,685 28.32 291 0.95
AL 12,839 41.87 8204 26.75 9,182 29.94 440 1.43
EMB 11,112 36.24 9395 30.64 9,765 31.84 393 1.28
PC 15,613 50.91 7098 23.15 7,647 24.94 307 1.00
NU 14,905 48.61 7699 25.11 7,748 25.27 313 1.02
PE 11,176 36.45 10197 33.25 9,001 29.35 291 0.95
PED 12,354 40.29 9273 30.24 8,712 28.41 326 1.06
95

Supplemental Table 2.5. Summary of principle component analysis (PCA) of the ten
compartments. Ten principle components (PCs) of each of the ten kernel compartments, the
standard deviation of each PC, and the percentage of variance explained by each PC are shown.
PC1 PC2 PC3 PC4 PC5 PC6 PC7 PC8 PC9 PC10

AL -91.9 66.0 -1.3 57.3 0.6 17.4 34.3 93.6 -20.3 1.7E-12
BETL -72.2 -125.1 26.8 -64.2 -18.2 -70.6 -58.0 42.6 12.9 -2.1E-13
CSE -99.2 -6.0 -12.1 32.0 41.2 81.2 -15.1 -15.5 55.4 6.8E-13
CZ -102.3 -51.3 -0.9 -14.6 34.3 51.8 -29.3 -46.6 -54.7 -3.8E-13
EMB -55.3 152.4 3.7 58.4 -31.7 -82.5 -33.9 -44.8 2.3 9.1E-13
ESR -71.3 -44.3 8.3 -47.5 -32.0 -31.7 105.5 -35.6 7.2 4.3E-13
NU 112.3 -30.5 -180.6 5.5 27.0 -22.9 1.5 3.3 -1.6 -8.6E-13
PC 146.2 -106.4 58.8 99.4 -73.0 22.4 -1.4 -8.7 -2.6 -1.2E-12
PE 97.0 111.6 7.6 -107.5 -71.9 62.0 -16.1 9.1 0.9 -8.3E-13
PED 136.7 33.8 89.8 -18.8 123.7 -27.0 12.5 2.7 0.5 -2.2E-13
Standard deviation 107.6 90.5 70.8 63.4 59.4 55.8 45.1 42.7 27.3 2.5E-13
% of variance 28.9 20.5 12.5 10.0 8.8 7.8 5.1 4.6 1.9 0.0
Cumulative % of variance 28.9 49.4 61.9 71.9 80.7 88.5 93.6 98.1 100.0 100.0
96

Supplemental Table 2.6. Summary of the available in situ hybridization and promoter analysis
data for endosperm cell-specific genes.
Cell CS
Gene ID Locus Name Assays References
Type Score
GRMZM2G170400 PEAMT AL 0.94 In situ hybridization This study
GRMZM2G437435 - AL 0.90 In situ hybridization This study
GRMZM2G091774 - AL 0.87 In situ hybridization This study
GRMZM2G069095 VPP1 AL 0.70 In situ hybridization Wisniewski et al. (2004)
In situ hybridization; Li et al. (2014);
GRMZM2G091054 AL-9 AL 0.51
promoter fusion Gómez et al. (2009)
AC199820.4_FG006 - BETL 0.99 In situ hybridization Li et al. (2014)
GRMZM2G016145 TCRR-1 BETL 0.99 In situ hybridization Muñiz et al. (2006)
GRMZM2G426158 - BETL 0.99 In situ hybridization This study
GRMZM2G111306 MRP1 BETL 0.99 In situ hybridization Gómez et al. (2002)
GRMZM2G133382 BAP-3A BETL 0.99 In situ hybridization Serna et al. (2001)
GRMZM2G132162 - BETL 0.99 In situ hybridization This study
GRMZM2G072219 - BETL 0.99 In situ hybridization Li et al. (2014)
GRMZM2G073290 BETL-4 BETL 0.97 In situ hybridization Hueros et al. (1999)
GRMZM2G158400 - BETL 0.97 In situ hybridization This study
GRMZM2G332483 - BETL 0.97 In situ hybridization This study
GRMZM2G121137 - BETL 0.96 In situ hybridization This study
In situ hybridization; Gutiérrez-Marcos et al.
GRMZM2G354335 MEG-1 BETL 0.96
promoter fusion (2004)
This study;
GRMZM2G008271 BAP-1A BETL 0.95 In situ hybridization
Serna et al. (2001)
GRMZM2G107302 - BETL 0.94 In situ hybridization This study
GRMZM2G123411 - BETL 0.93 In situ hybridization This study
GRMZM2G175912 MEG-13 BETL 0.93 In situ hybridization This study
GRMZM2G058703 - BETL 0.92 In situ hybridization This study
GRMZM2G158407 - BETL 0.92 In situ hybridization This study
GRMZM2G152655 BAP-2 BETL 0.89 In situ hybridization Hueros et al. (1999)
GRMZM5G812796 - BETL 0.89 In situ hybridization This study
In situ hybridization;
GRMZM2G167733 EBE-2 BETL 0.86 Magnard et al. (2003)
promoter fusion
GRMZM2G082785 BETL-1 BETL 0.79 In situ hybridization Hueros et al. (1995)
GRMZM2G401374 CC8 BETL 0.76 In situ hybridization Massonneau et al. (2005)
In situ hybridization; This study;
GRMZM2G087413 BETL-9 BETL 0.74
promoter fusion Royo et al. (2014)
GRMZM2G090264 TCRR-2 BETL 0.57 In situ hybridization Muñiz et al. (2010)
GRMZM2G138727 27-kD γ-zein CSE 0.96 In situ hybridization Woo et al. (2001)
GRMZM2G060429 16-kD γ-zein CSE 0.95 In situ hybridization Woo et al. (2001)
GRMZM2G089713 Sh1 CSE 0.89 In situ hybridization This study
GRMZM2G369799 - CSE 0.77 In situ hybridization Li et al. (2014)
GRMZM2G429899 Sh2 CSE 0.68 In situ hybridization This study
GRMZM2G023872 SCL1 CSE 0.60 In situ hybridization Li et al. (2014)
GRMZM2G154182 ASN1 CSE 0.47 In situ hybridization Li et al. (2014)
GRMZM2G006585 - CZ 0.70 In situ hybridization Li et al. (2014)
GRMZM2G315601 ESR-2 ESR 0.99 In situ hybridization Li et al. (2014)
GRMZM2G048353 ESR-6 ESR 0.99 In situ hybridization Balandín et al. (2005)
GRMZM2G149869 - ESR 0.99 In situ hybridization This study
GRMZM2G120008 - ESR 0.99 In situ hybridization Li et al. (2014)
GRMZM2G046086 ESR-1 ESR 0.99 In situ hybridization Li et al. (2014)
GRMZM2G372553 AE3 ESR 0.54 In situ hybridization Magnard et al. (2000)
97

Supplemental Table 2.7. Robustness of WGCNA-generated coexpression modules based on a


permutation test of average topological overlap (TO).
Module Mean TO observeda Mean TO randomb Maximal TO randomc P-valued
1 0.20032682 0.04206503 0.05808373 < 10-5
2 0.13829484 0.03242548 0.03794889 < 10-5
3 0.05589406 0.03135009 0.03492353 < 10-5
4 0.05180793 0.03119400 0.03506029 < 10-5
5 0.11182849 0.03251804 0.03856629 < 10-5
6 0.05307235 0.03117166 0.03413336 < 10-5
7 0.10177977 0.03284855 0.03858159 < 10-5
8 0.05829481 0.03152054 0.03555416 < 10-5
9 0.26306783 0.03539109 0.04476182 < 10-5
10 0.11657907 0.03440147 0.04239688 < 10-5
11 0.04351963 0.03179076 0.03609683 < 10-5
12 0.08102435 0.03258770 0.03830303 < 10-5
13 0.11766000 0.03433284 0.04294189 < 10-5
14 0.14466003 0.03906574 0.05298123 < 10-5
15 0.05718379 0.03269729 0.03842810 < 10-5
16 0.10289539 0.03315041 0.04010339 < 10-5
17 0.11984522 0.03364567 0.04121327 < 10-5
18 0.07067248 0.03197153 0.03666547 < 10-5
a
the average TO for the genes in an observed module.
b
the average TO of 100,000 iterations.
c
the maximum of the average TO for the 100,000 iterations.
d
the empirical probability of finding an average TO greater than or equal to the observed TO in 100, 000
collections of modules comprised of randomly selected genes.
98

Supplemental Table 2.8. Allele-biased genes enriched in the endosperm-associated


coexpression modules M10, M15, M17, and M18.

Allelic
Gene ID Module Expression Stringencya TF Family Functional Annotation
Pattern
GRMZM2G014119 10 MEG High - Uncharacterized protein
GRMZM2G170099 10 MEG High - Uncharacterized protein
GRMZM2G169695 10 MEG High - AT-rich element binding factor 3
GRMZM2G124708 10 MEG High - -
GRMZM2G121546 10 MEG Low - Uncharacterized protein
GRMZM2G072174 10 MEG Low - Uncharacterized protein
Putative NAC domain
GRMZM2G154182 10 MEG Low NAC transcription factor superfamily
protein; Uncharacterized protein
GRMZM2G500698 10 MEG Low - -
Paired amphipathic helix repeat
GRMZM2G170201 10 PEG High Orphans
family protein
GRMZM2G043417 10 PEG High - -
MADS-box transcription factor 47;
GRMZM2G059102 10 PEG Low MADS
Uncharacterized protein
GRMZM2G359589 15 MEG High C2H2 Uncharacterized protein
GRMZM2G370991 15 MEG High - Uncharacterized protein
GRMZM2G343972 15 MEG High - Uncharacterized protein
GRMZM2G130580 15 MEG High Orphans Uncharacterized protein
GRMZM2G024468 15 MEG High MYB Uncharacterized protein
Putative MYB DNA-binding
GRMZM2G150680 15 MEG High MYB domain superfamily protein;
Uncharacterized protein
GRMZM2G069840 15 MEG High - -
GRMZM2G476175 15 MEG High - -
GRMZM2G120085 15 MEG Low - Uncharacterized protein
GRMZM2G474602 15 MEG Low - -
GRMZM2G092101 15 PEG High - Uncharacterized protein
Putative leucine-rich repeat
GRMZM2G103164 15 PEG High - receptor-like protein kinase family
protein
C2C2- Putative GATA transcription factor
GRMZM2G324131 15 PEG High
GATA family protein
Putative AP2/EREBP
GRMZM5G830365 15 PEG High - transcription factor superfamily
protein
DNA-binding protein;
GRMZM2G089562 15 PEG High -
Uncharacterized protein
AC209624.2_FG001 15 PEG High - -
AC209624.2_FG003 15 PEG High - -
GRMZM2G028366 15 PEG High - -
GRMZM2G093947 15 PEG High - -
GRMZM2G374169 15 PEG High - -
GRMZM2G494808 15 PEG High - -
GRMZM2G099353 15 PEG Low - Uncharacterized protein
GRMZM2G110531 15 PEG Low Orphans Speckle-type POZ protein
Putative homeobox/lipid-binding
GRMZM2G004334 15 PEG Low Homeobox
domain family protein
GRMZM2G450822 15 PEG Low - -
GRMZM2G062650 17 MEG High NAC NAM-related protein 1
99

Supplemental Table 2.8 Continued


GRMZM2G063498 17 MEG High - -
Pyrophosphate-energized
GRMZM2G041065 17 MEG High -
vacuolar membrane proton pump
GRMZM2G179777 17 MEG Low - Triacylglycerol lipase
GRMZM2G085078 17 MEG Low - -
Putative homeodomain-like
GRMZM2G047104 17 PEG High - transcription factor superfamily
protein
AC191534.3_FG003 17 PEG High - -
GRMZM2G006732 17 PEG High - -
GRMZM2G339663 17 PEG High - -
GRMZM5G897856 17 PEG High - -
GRMZM2G112925 18 MEG High Orphans -
GRMZM2G103247 18 MEG Low Orphans Uncharacterized protein
GRMZM2G147226 18 MEG Low - Uncharacterized protein
GRMZM2G012071 18 MEG Low - -
GRMZM2G152764 18 MEG Low - -
GRMZM2G459363 18 MEG Low - -
GRMZM2G447406 18 PEG High - Uncharacterized protein
GRMZM2G449489 18 PEG High - Uncharacterized protein
GRMZM2G007736 18 PEG High - Uncharacterized protein
GRMZM2G139406 18 PEG High - Uncharacterized protein
C2C2-
GRMZM2G048850 18 PEG High Uncharacterized protein
GATA
Putative RING/U-box superfamily
GRMZM2G145123 18 PEG High -
protein; Uncharacterized protein
Putative AP2/EREBP
AP2-
GRMZM2G104866 18 PEG High transcription factor superfamily
EREBP
protein; Uncharacterized protein
GRMZM2G121570 18 PEG High MYB MYB-type transcription factor
GRMZM2G084462 18 PEG High - Isopentenyl transferase IPT2
Galactosylgalactosylxylosylprotein
GRMZM2G037469 18 PEG High -
3-beta-glucuronosyltransferase 1
GRMZM2G136465 18 PEG High - -
GRMZM2G164314 18 PEG High - -
GRMZM2G366435 18 PEG High - -
GRMZM2G479318 18 PEG High - -
GRMZM5G871454 18 PEG High - -
GRMZM2G156794 18 PEG Low - Uncharacterized protein
Hexokinase-1; Uncharacterized
GRMZM2G046686 18 PEG Low -
protein
Anthranilate N-benzoyltransferase
GRMZM2G131165 18 PEG Low -
protein 1
AP2-
GRMZM2G480434 18 PEG Low -
EREBP
GRMZM2G311401 18 PEG Low - -
GRMZM2G345189 18 PEG Low - -
GRMZM2G474620 18 PEG Low - -
a
Stringency of criteria used for defining imprinted/allele-biased gene expression patterns. High indicates genes
identified by applying the relatively stringent criteria used by Xin et al. (2013); Low indicates additional genes
identified by applying the less stringent criteria used in this study.
100

Supplemental Table 2.9. Motif analysis of the endosperm compartment-correlated modules


M10, M12, M15, M17, and M18.
Module Motif E-value Motif Sequence Matching TFa
10 1 1.30E-136 [CG]CGCCGCCGCCG abi4
2 4.40E-62 AAA[AT]AAA[AT]A[AT]A[AT] SOC1
3 2.50E-56 [GA][GA][AG]GAG[AG][AG][GA]GAG -
4 2.20E-54 CAGC[AG]GCAGC -
5 1.50E-41 AA[AG]AAAAAAAAA SOC1
6 4.60E-39 C[GT]CC[GAT]CC[GT]CC -
7 2.70E-39 [GC]C[GC][GC]CG[CG][CG][CG]GC[GC] -
8 2.40E-33 [AT][AT]A[TA][TA]TAT[AT]T[AT][TA] -
9 5.10E-28 GG[AG]GG[AG][GA]G[AG][AG]G[GA] -
10 3.00E-25 C[GC]TCG[TG]CG[CTG]C bZIP911
12 1 7.70E-196 CGCCGCCGCCGC abi4
2 4.10E-92 CG[CG]CGCCGCC abi4
3 1.20E-76 [GA][GA]AGAG[AG][GA]AGAG -
4 9.30E-83 AAAAAAAA[AT]A SOC1
5 2.10E-78 C[TG]CCTCC[TG]CC[TG]C -
6 4.30E-65 [TA]A[TA]ATATA[TA][AT]TA -
7 4.40E-55 C[GC][GC]CG[GC]CGGC ERF1
8 3.10E-53 CG[TC]CGTCG[CT]C -
9 8.70E-51 C[TC]GC[TC]GC[TG]GC -
10 1.30E-37 [AT]TTTTT[AT]TTT -
15 1 7.00E-178 CACTCGGCAAAG -
2 7.10E-175 GCCGCCGCCGCC abi4
3 4.30E-86 GAGAG[AG][GA]AGA[GA][AG] -
4 1.00E-68 C[GC]GCG[GC]CG[GC]C ERF1
5 3.00E-64 AAAAA[AT]AAAA SOC1
6 2.70E-43 G[CG][GC]G[CA]GG[CA][GC]G[CG]G abi4
7 8.30E-48 C[AG]GCAGC[AG]GC -
8 3.70E-46 C[CT][CT]C[CT][CT]C[CT]CC[CT]C -
9 9.60E-43 [CT]C[GC][TCG]CG[CT]CG[GCT]CG -
10 5.10E-30 A[AT]A[TA]A[AT]A[TA][AT]AA[AT] -
17 1 1.20E-154 [GC][CG]CGCCGCCGCC abi4
2 3.70E-88 ATATA[TA]A[TA]A[TA]AT -
3 1.50E-83 A[AG]AAAAAAAA SOC1
4 7.40E-97 [AG]GAGA[GA][AG]GAGAG -
5 7.30E-48 CCGCCGCCGC ERF1
6 3.40E-50 CG[CT]CG[TCG]CGCC abi4
7 3.10E-44 AAA[AT][AT][AT]AA[TA]AAA SOC1
8 6.40E-34 CTCC[TG]CCTCC -
9 7.40E-34 C[AG]GC[AG]GC[AG]GC
10 1.90E-37 CC[GC]C[GC]C[GC]C[CG][GC][CG]C
18 1 1.80E-203 [CG]CGCCGCCGCCG abi4
2 2.40E-182 ATATATATATAT
3 1.00E-174 TAGATAGATAGA --
4 1.90E-101 CACTCGGCAAAG --
5 2.00E-99 CG[CG]CGCCGCC abi4
6 9.20E-81 TA[GA]ATATA[GTA]A[TA]A
7 1.90E-81 [AT]AAA[TA]AAA[TA]AAA SOC1
8 1.80E-104 TAGATA[GT]A[GT]A[GT]A --
9 2.80E-115 [GA]GAGAG[AG]GAGAG --
10 1.40E-74 C[GC]GCG[GC]CGGC ERF1
a
Transcription factors of which the known binding motif most significantly matches the identified motif (q-value <
0.05).
101

Supplemental Table 2.10. Occurrences of sub-motifs of each motif enriched in upstream


sequences of M18 genes as identified by MEME.
Motif Sub-motifs Sites Motif Sub-motifs Sites Motif Sub-motifs Sites
1 CCGCCGCCGCCG 63 5 CGCCGCCGCC 65 8 TAGATAGATAGA 36
GCGCCGCCGCCG 18 CGGCGCCGCC 32 TAGATATAGATA 31
TCGCCGCCGCCG 13 CGCCGCCGGC 3 TAGATAGAGATA 10
CCGCCGCCGCCC 3 6 TAGATATAGATA 24 TAGAGAGAGAGA 8
GCGCCGCCGCCC 3 TAGATATATATA 13 TAGATATAGAGA 7
2 ATATATATATAT 90 TAAATATAAATA 11 TAGATAGATATA 6
ATATATATAGAT 10 TAGATATAGAAA 10 AAGATAGAGAGA 1
3 TAGATAGATAGA 72 TAGATATAAATA 8 TAGATAGAGAGA 1
TAGATATATAGA 18 TAAATATAAAAA 7 9 AGAGAGAGAGAG 40
TAGATAAATAGA 10 TAAATATATAAA 6 GGAGAGAGAGAG 21
4 CACTCGGCAAAG 45 TATATATATATA 6 AGAGAGGGAGAG 6
CACTCGACAAAG 4 TAAATATATATA 4 GGAGAGGGAGAG 6
CACTCGGCACAG 3 TAGATAAAAAAA 4 GGAGGAGGAGAG 6
CACTCGGTAAAG 3 TAAATATAGAAA 3 GGAGAAGGAGAA 4
CTCTCGGCAAAG 3 TAGATATATAAA 3 GGAGAAGGAGAG 4
CTCTCGGCACAG 3 TAGATATAAAAA 1 GGAGAGAGAGAA 3
CACCCGGCAAAG 2 7 AAAATAAATAAA 23 GGAGAAAGAGAG 2
CACTCAGCAAAG 2 AAAATAAAAAAA 12 GGAGGGAGAGAG 2
CACTCGGCAAAA 2 AAAAAAAAAAAA 11 GGAGGGAGAGGG 2
CACTCGGCAAAC 2 TAAAAAAATAAA 10 GGAGAGAGAGGG 1
CACTCGGCAAAT 2 AAAAAAAATAAA 8 GGAGAGGGAGGG 1
CACTTGGCAAAG 2 TAAATAAATAAA 8 GGAGGGGGAGAG 1
CGCTCGGCAAAG 2 TAGAAAAATAAA 7 GGAGGGGGAGGG 1
CTCTCGACAAAG 2 AAGAAAAATAAA 5 10 CGGCGGCGGC 40
TACTCGGCAAAG 2 TAAAAAAAAAAA 4 CGGCGCCGGC 31
AACTCGGCAAAG 1 TAAATAAAAAAA 4 CCGCGGCGGC 26
CACTCGGCGAAG 1 AATATAAATAAA 3 CGGCGGCGCC 3
CACTCGGTACAG 1 AAGATAAATAAA 2
CATTCGGCAAAG 1 TATATAAATAAA 2
CCCTCGGCAAAG 1 AAAATATATAAA 1
CTCTCAGCAAAG 1
102

Supplemental Table 2.11. Putative functions of the 93 M18 genes that contain at least one
MRP-1-binding sub-motifs (Motifs 3a/8a, 6a/8b, 8c, and 8f).
Locus/
Gene ID MM TF Family Functional Annotation
TF Name
AC196417.3_FG008 0.997666935 - - -
AC204876.3_FG006 0.995274098 - - -
AC233883.1_FG006 0.98545352 - - Uncharacterized protein
GRMZM2G002903 0.980405318 - - Uncharacterized protein
GRMZM2G003493 0.709293417 - - Uncharacterized protein
GRMZM2G005710 0.925096716 - - Uncharacterized protein
GRMZM2G007784 0.97271802 - - -
GRMZM2G008271 0.998357469 - BAP-1A Basal layer antifungal peptide
GRMZM2G008403 0.997103295 - BAP-1B Basal layer antifungal peptide
GRMZM2G009427 0.995625792 - - -
GRMZM2G010762 0.994304527 - - Early nodulin-like protein 3
Putative leucine-rich repeat receptor-
GRMZM2G011896 0.988628322 - - like protein kinase family protein;
Uncharacterized protein
GRMZM2G017869 0.995815117 - - Uncharacterized protein
GRMZM2G024350 0.995946537 - - Uncharacterized protein
GRMZM2G027472 0.995775449 - - Uncharacterized protein
GRMZM2G028386 0.995641682 AP2-EREBP EREB137 -
GRMZM2G034862 0.995837595 - - Uncharacterized protein
GRMZM2G045082 0.995397205 - - Putative uncharacterized protein
GRMZM2G045638 0.978759996 - - Uncharacterized protein
GRMZM2G047833 0.995274098 - - -
GRMZM2G048850 0.697116411 C2C2-GATA GATA33 Uncharacterized protein
Putative MYB DNA-binding domain
GRMZM2G049695 0.930118941 MYB-related MYBR24
superfamily protein
GRMZM2G049912 0.950990813 - - Uncharacterized protein
GRMZM2G050407 0.966973351 - - Uncharacterized protein
GRMZM2G054359 0.996226267 - - Uncharacterized protein
GRMZM2G054632 0.977265787 - - Uncharacterized protein
GRMZM2G058703 0.998757072 - - Uncharacterized protein
GRMZM2G060991 0.984451944 - - -
GRMZM2G066546 0.982044708 - - Uncharacterized protein
GRMZM2G076370 0.997249693 - - Uncharacterized protein
GRMZM2G079547 0.994976509 - - Uncharacterized protein
GRMZM2G082785 0.991592203 - BETL-1 BET1 protein
GRMZM2G082799 0.978839534 - - BET1 protein
GRMZM2G086827 0.998053357 - MEG-10 -
GRMZM2G086835 0.923876737 - - Zinc finger protein 7
GRMZM2G088896 0.998693431 - MEG-9 -
GRMZM2G091445 0.999221086 - BETL-10 Putative defensin
GRMZM2G094054 0.997047242 - MEG-6 Putative uncharacterized protein
GRMZM2G095337 0.885645943 - - Uncharacterized protein
Acid phosphatase 1; Uncharacterized
GRMZM2G103526 0.7672607 - -
protein
GRMZM2G106655 0.99900411 - - -
GRMZM2G107302 0.999258566 - - Uncharacterized protein
GRMZM2G107754 0.99443088 - - Uncharacterized protein
GRMZM2G110395 0.833343634 - - -
GRMZM2G110816 0.987045749 - - Uncharacterized protein
GRMZM2G115574 0.993363479 - - Uncharacterized protein
GRMZM2G128531 0.995574221 - - Blue copper protein
GRMZM2G131982 0.99682921 DBB - -
103

Supplemental Table 2.11 Continued


Putative GATA transcription factor
GRMZM2G118214 0.789535171 C2C2-GATA GATA7 family protein; Uncharacterized
protein
GRMZM2G122437 0.970709069 - - Uncharacterized protein
Putative cytochrome P450
GRMZM2G126055 0.972985678 - - superfamily protein; Uncharacterized
protein
GRMZM2G133370 0.998120923 - BAP-3B Basal layer antifungal peptide
GRMZM2G133382 0.996008427 - BAP-3A Basal layer antifungal peptide
GRMZM2G133407 0.966870998 - - Uncharacterized protein
GRMZM2G136771 0.995682732 - - Uncharacterized protein
GRMZM2G137959 0.998934848 - MEG-4 MEG4
GRMZM2G141574 0.950302131 - - -
GRMZM2G147226 0.968661535 - - Uncharacterized protein
GRMZM2G148270 0.984627461 - - Putative uncharacterized protein
Basal layer antifungal protein2;
GRMZM2G152655 0.999157709 - BAP-2
Uncharacterized protein
GRMZM2G156794 0.995986927 - - Uncharacterized protein
GRMZM2G158400 0.997452659 - - Uncharacterized protein
GRMZM2G158407 0.999145715 - - -
GRMZM2G165535 0.986253278 - - Uncharacterized protein
GRMZM2G170181 0.695195223 - - -
GRMZM2G172596 0.998045775 - - GAST1 protein
Putative cytochrome P450
GRMZM2G175250 0.994699115 - -
superfamily protein
GRMZM2G175896 0.997778567 - MEG-12 -
GRMZM2G175912 0.999396321 - MEG-13 -
GRMZM2G181051 0.998899063 - MEG-11 -
GRMZM2G332483 0.99739216 - - Uncharacterized protein
GRMZM2G332703 0.983249958 - - -
GRMZM2G341010 0.998049519 - - Uncharacterized protein
GRMZM2G342246 0.99535502 - - Beta-expansin 7
Plastid-specific 30S ribosomal protein
GRMZM2G347956 0.843165264 - -
1
GRMZM2G354335 0.998651579 - MEG-1 Putative uncharacterized protein
GRMZM2G383937 0.995865784 - - -
GRMZM2G406552 0.791376554 - - Nonspecific lipid-transfer protein
GRMZM2G409309 0.996479121 - - -
Putative MYB DNA-binding domain
GRMZM2G422083 0.995972492 MYB-related MYBR33
superfamily protein
GRMZM2G426158 0.995470482 - - -
GRMZM2G466545 0.964270672 - - Uncharacterized protein
GRMZM2G487702 0.993630476 - - -
MEG2; Maternally expressed gene 2;
GRMZM2G502035 0.998543642 - MEG-2
Uncharacterized protein
GRMZM5G800473 0.996024857 - - -
GRMZM5G802585 0.808982972 - - Uncharacterized protein
GRMZM5G803611 0.922753269 - - Uncharacterized protein
GRMZM5G812796 0.999152 - - Uncharacterized protein
GRMZM5G836886 0.93249873 - - -
GRMZM5G864689 0.623012716 - - Uncharacterized protein
GRMZM5G888860 0.957073782 - - -
GRMZM5G890815 0.954454282 - - -
GRMZM5G897740 0.998634244 - - Uncharacterized protein
104

Supplemental Table 2.12. Results of Y1H assays for binding of MRP-1 to the promoters of
genes within M18.
Gene ID Locus/TF Name Sub-motifs Contained in Promotera Y1H Resultsb
AC184794.2_FG003 - None –
GRMZM2G009854 - None +
GRMZM2G016145 TCRR-1 None (+)
GRMZM2G027472 - 3a/8a (3), 6a/8b (2), 8c (1), 8f (1) +
GRMZM2G030403 MYBR9 None –
GRMZM2G034862 - 3a/8a (1), 8f (1) +
GRMZM2G054359 - 3a/8a (5), 6a/8b (1), 8f (3) +
GRMZM2G055621 - None –
GRMZM2G058703 - 3a/8a (2) +
GRMZM2G062565 - None +
GRMZM2G063042 - None +
GRMZM2G070399 - None –
GRMZM2G072219 - None –
GRMZM2G082799 - 6a/8b (1), 8c (1) +
GRMZM2G084462 - None –
GRMZM2G087413 - None +
GRMZM2G091445 - 6a/8b (1) +
GRMZM2G107302 - 3a/8a (6) +
GRMZM2G111306 MRP-1 None –
GRMZM2G112925 - None –
GRMZM2G121111 MYBR81 None –
GRMZM2G123411 - None +
GRMZM2G125386 - None –
GRMZM2G133370 - 3a/8a (1), 6a/8b (1), 8c (1) +
GRMZM2G133382 - 3a/8a (5), 8c (1) +
GRMZM2G136771 - 3a/8a (5), 6a/8b (1) +
GRMZM2G138392 - None (+)
GRMZM2G141574 - 6a/8b (2), 8f (1) +
GRMZM2G152655 BETL-2 3a/8a (4), 8c +
GRMZM2G158407 - 3a/8a (2) +
GRMZM2G175912 - 6a/8b (3) (+)
GRMZM2G175976 BETL-3 None –
GRMZM2G181051 - 8f (1) +
GRMZM2G332483 - 3a/8a (1), 8c (1), 8f (1) +
GRMZM2G354335 MEG-1 6a/8b (2), 8f (1) +
GRMZM2G386276 - None –
GRMZM2G402156 MYBR19 None –
GRMZM2G422083 MYBR33 6a/8b (2) +
GRMZM2G426158 - 3a/8a (2) +
GRMZM5G800473 - 6a/8b (1) +
GRMZM5G812796 - 6a/8b (1) +
GRMZM5G897740 - 8f (1) +
aThe number in parenthesis indicates the number of copies of that sub-motif.
bScoring: + indicates strong positive, (+) indicates weak positive, – indicates negative. Images of the Y1H

growth assays are shown in Supplemental Figure 2.16.


105

Supplemental Table 2.13. Sequences of the primers used to generate the clones for the in situ
hybridization probes.
Gene ID Forward Primer Reverse Primer
GRMZM2G008271 TCTAGTTCATCACCCATGGC GATCATGCATTAAGAGGGTC
GRMZM2G058703 CACATTGTTAGTATTAGACC TAAAGCAGGAAGGGGAGAGG
GRMZM2G087413 TAAAGATCTCAGTGTCACAC AGCCAGGTTGAAACGGCCTC
GRMZM2G089713 TACATTCTGGATCAGGTCCG CCCTTCAACTTGTACTCGTC
GRMZM2G091774 GGAGGCCTCCTCAAGATTGG AGCATAATAGCATCAGTTGG
GRMZM2G107302 AGTACACGACTCAGCCATTG GACACCAAAATACTATATGC
GRMZM2G121137 TATCAGTATCTTCCACAAGC CAAAACACCAACACAAATGC
GRMZM2G123411 CTCTAAACTCAAACAAATAC CGAGTGCTTCAGACACTCGG
GRMZM2G132162 AGCATACTATTAGACATACC GAGGCCACGTATGGAGAACC
GRMZM2G133370 GATCTTATCCATCACCTATG CTAATATCATTGAAGACACAC
GRMZM2G149869 CGCCGACGCAAGACTACCAG ACCTAACACCTGATGAATCC
GRMZM2G158400 GGAGAAGAAATTACAGATCC TAGTAAAGATTTATGAGCAC
GRMZM2G158407 GAACATCGAAAGTATCATGC AACAGTTTTGTTCATATCTC
GRMZM2G170400 AAAGCTTGTGGGGAAAATGG GTGAAGGATGGTGTCACGGC
GRMZM2G175912 CATGTCACAAATGATATCGG TGACTAAATTTCGCGAATCG
GRMZM2G332483 CAAGATCAGCCCCAAGCACG ACTTATCTTATTGAGTTCAC
GRMZM2G426158 ATAGGCAATGTAAGAGTAGC TTGAGACTGTTTCGAGATAG
GRMZM2G429899 GGCTACACAAATGCCTGAAG ACATCCAGAGCTGACACGTG
GRMZM2G437435 TGGATGCTACACTCCTATCG ATGGCTCCTCGTTGCTCAGC
GRMZM5G812796 CGATCGTCGTCGCAGAAAGG GAAATACTCCACATGCATTG
106

Supplemental Table 2.14. Sequences of the primers used to generate the constructs to test the
sub-motifs in Y1H assays. The sub-motif sequences are in bold red letters and are underlined. In
all constructs, the sequence of the reverse primer was
GATCCTCGAGGGACTAGCTAGACAAGCTCTC.
Sub-motif Forward Primer
3a/8a GATCAAGCTTATAGATAGATAGATTAAGCGTATCGCTAG
3b GATCAAGCTTTAGATATATAGATTAAGCGTATCGCTAG
3c GATCAAGCTTTAGATAAATAGATTAAGCGTATCGCTAG
6a/8b GATCAAGCTTATAGATATAGATATTAAGCGTATCGCTAG
6b GATCAAGCTTATAGATATATATATTAAGCGTATCGCTAG
6c GATCAAGCTTTAAATATAAATATTAAGCGTATCGCTAG
6d GATCAAGCTTTAGATATAGAAATTAAGCGTATCGCTAG
6e GATCAAGCTTTAGATATAAATATTAAGCGTATCGCTAG
6f GATCAAGCTTTAAATATAAAAATTAAGCGTATCGCTAG
6g GATCAAGCTTTAAATATATAAATTAAGCGTATCGCTAG
6h GATCAAGCTTTATATATATATATTAAGCGTATCGCTAG
6i GATCAAGCTTTAAATATATATATTAAGCGTATCGCTAG
6j GATCAAGCTTTAGATAAAAAAATTAAGCGTATCGCTAG
6k GATCAAGCTTTAAATATAGAAATTAAGCGTATCGCTAG
6l GATCAAGCTTTAGATATATAAATTAAGCGTATCGCTAG
6m GATCAAGCTTTAGATATAAAAATTAAGCGTATCGCTAG
8c GATCAAGCTTTAGATAGAGATATTAAGCGTATCGCTAG
8d GATCAAGCTTTAGAGAGAGAGATTAAGCGTATCGCTAG
8e GATCAAGCTTTAGATATAGAGATTAAGCGTATCGCTAG
8f GATCAAGCTTTAGATAGATATATTAAGCGTATCGCTAG
8g GATCAAGCTTAAGATAGAGAGATTAAGCGTATCGCTAG
8h GATCAAGCTTTAGATAGAGAGATTAAGCGTATCGCTAG
107

Supplemental Table 2.15. Sequences of the primers used to generate the constructs to test the
promoters in Y1H assays.
Gene ID Forward Primer Reverse Primer
GATCAAGCTTGTCTAGAATGGTTTA GATCCTCGAGGCACTTTCCGCAGATCC
AC184794.2_FG003
GTGTAG CATATG
GATCAAGCTTCAGCTACACCTGAA GATCCTCGAGCTTTTTGATGATTTAGCA
GRMZM2G009854
AACCTAAAC GGTGG
GATCAAGCTTCGTCTACTTCTTGCT GATCCTCGAGGGACTAGCTAGACAAGC
GRMZM2G016145
CTAGATC TCTC
GATCGGTACCATGGTGTGGAACGA GATCCTCGAGATGGATGACCGATAAGA
GRMZM2G027472
CCGTTG GC
GATCAAGCTTGAATAGTGGTATGAA GATCCTCGAGGAGAGTTGAGGATTGAG
GRMZM2G030403
CTATGC TAC
GATCAAGCTTGGCCTCTAACGTAAT GATCCTCGAGGGGTGATTAGAAAATACT
GRMZM2G034862
GGTTAAGG AATGCC
GATCAAGCTTGGCATGTACATAGG GATCCTCGAGTATTCTATCCTTTGAGAT
GRMZM2G046532
GCTCTG GGAATG
GATCAAGCTTGGTGTTGGTTAGTC GATCCTCGAGGGGTGCTAGATGAGAGC
GRMZM2G054359
ACAAGTC TAATACC
GATCAAGCTTGATGGGGAGAGGCA GATCCTCGAGGTCGTATGAACAGGAGT
GRMZM2G055621
TGTAAC GC
GATCAAGCTTATCAGGATACAATCG GATCCTCGAGGGGTGATGAACTAGAAA
GRMZM2G058703
TTCTG GTG
GATCGGTACCGGTTTGTGCTGAAC GATCCTCGAGGGACAGTGAGTGACCTA
GRMZM2G062565
CTTTGG ACTATG
GATCGGTACCGACCGAAGGGGAAA GATCCTCGAGGGTTTGTTTATATTTGCT
GRMZM2G063042
AAGAC AG
GATCAAGCTTCATGGGTTGCACGT GATCCTCGAGCGACGATCTACACTGCT
GRMZM2G070399
GAAC GTC
GATCAAGCTTCTATCTTCTACATTC GATCCTCGAGTTTGTAGTGCTATATGTG
GRMZM2G072219
TTTGCC GATTGG
GATCAAGCTTCCTTTTCTTTATACC GATCCTCGAGAAGGAAAGATGATCAAG
GRMZM2G082799
ACATGCC CTTC
GATCAAGCTTGGATCGATGACGTG GATCCTCGAGGTCTTGATGATCTTGATT
GRMZM2G084462
ATCAAG ATTGTAGCC
GATCAAGCTTGGGATTTTGTAGGA GATCCTCGAGGGGTATAACTTCAACTGT
GRMZM2G087413
ATCATC TG
GATCAAGCTTACCACTGCTTGCTA GATCCTCGAGTGAGGAGAACAGCATAC
GRMZM2G091445
GTGATTTC ACATG
GATCAAGCTTGTAGTTCAGCAGAA GATCCTCGAGCGAGCGTGTACTGCACG
GRMZM2G107302
AACAGCACGC CACAT
GATCAAGCTTATATCACAAGGAAAG GATCCTCGAGGAGGTGCGAGGGATTAA
GRMZM2G111306
ATATG GTAC
GATCAAGCTTGCCTTTACGGGCTTT
GRMZM2G112925 GATCCTCGAGGTTTCCTTGGGCTTTGG
TTAGG
GATCGGTACCAAAAGAGGTGGGAT GATCCTCGAGCCATCAGCAACTAACGTT
GRMZM2G115340
CACCC TC
GATCAAGCTTAGAGGAGCGGTTGA GATCCTCGAGCGGCCGGGGTTAGCTAA
GRMZM2G121111
ACACTG GCTAG
GATCAAGCTTGCCCGTCCTGTCCT GATCCTCGAGGGGTGAAAAGGGTAAGA
GRMZM2G123411
ACAACG GC
GATCAAGCTTCTTGGTTTTGGTGAA GATCCTCGAGAGAGGAAACAATTGTTCT
GRMZM2G125386
TTGTGCC TGC
GATCGGTACCGCGGTTGGTAGACA GATCCTCGAGAGGTGATGGATAAGATC
GRMZM2G133370
GGTAGG TAATA
108

Supplemental Table 2.15 Continued


GATCAAGCTTCTCGAATAATCTAAT GATCCTCGAGTGTGATGGATAAGAGCT
GRMZM2G133382
GTATTC AATAATA
GATCAAGCTTAAAGCAATGCTCATC GATCCTCGAGGTGAAATGAAGAGGAGT
GRMZM2G136771
TTTAACC GG
GATCCTCGAGGTCAGACACGATGAGAG
GRMZM2G138392 GATCAAGCTTCGGTTTTGAATGTGC
AG
GATCAAGCTTATGCACTTTGGATAT GATCCTCGAGAAGAAGAGGATATATAGA
GRMZM2G141574
ACC TC
GATCGGTACCCGTGGGTTTCGTAC GATCCTCGAGATCAGCACAAGATCCAA
GRMZM2G150680
CCAC GG
GGCCGGTACCAGTTGATATAACTA GATCCTCGAGGGGTGACAGATGATATG
GRMZM2G152655
GATAGG AGC
GATCAAGCTT GATCCTCGAGGGTTAATTATTTGGGTGA
GRMZM2G158407
CTGTCTCGGATGGAAAAAC GGA
GATCGGTACCACGGTTAATAGTAG GCCCTCGAGGACGCAAGAAAATCTAAA
GRMZM2G175912
AGCCAG GAAC
GATCAAGCTTCAAATTACCCGCAG GATCCTCGAGTGTAGTTTGCTATCACCC
GRMZM2G175976
GGGTATG TT
GATCAAGCTTAACTTCTGCAGAGT GATCCTCGAGGTCGCAAGAAAATTTAAG
GRMZM2G181051
GTTTTG GAAC
GATCAAGCTTCGAGAGCCTAACTT GATCCTCGAGCACAAATTAAATGCATAG
GRMZM2G314094
CACCCAAC AAG
GATCGGTACCATACATAGATGGAC GATCCTCGAGGAGATGAGAGCACAATA
GRMZM2G332483
AAACTG TTAC
GGCCGGTACCTCTCGACACAGGTA GATCCTCGAGGTCGCAAGAAATGTTAA
GRMZM2G354335
GGTAG GGAAC
GATCCCCGGGGTAGGAACTCACAT GATCCTCGAGCTTTAAGTTAAATATGGT
GRMZM2G386276
AAG ACC
GATCAAGCTTGCCCTGGACGGTTC GATCCTCGAGGGGGTTGGGTACTACGA
GRMZM2G402156
GCGATTAG TATATG
GATCAAGCTTCCATACAGATATAGG GATCCTCGAGACATGGATGGAGAAAAG
GRMZM2G422083
TATATG GGC
GATCAAGCTTCGAGCCTTCGCATC GATCCTCGAGGGTCGTATGAACAGGAG
GRMZM2G426158
GTTGATC TGCTAG
GATCAAGCTTCGGAAAGGACAGGA GATCCTCGAGGATGCAACAAATACCTCC
GRMZM2G846314
CGTGTG TAC
GATCAAGCTTCGAGCGAGCCGGAA GATCCTCGAGATTGGAATGTATGTGAGT
GRMZM5G800473
GTTTG GAG
GATCAAGCTTCTTCTGACCATTGGC GATCCTCGAGGAGGAAACGAGTATTCTT
GRMZM5G812796
TCTG GC
GATCAAGCTTCAAGTTTCATACCAA GATCCTCGAGCGCCACCACTTCTTCTCT
GRMZM5G897740
TATACG G
109

Supplemental Data Set 2.1. Normalized expression levels in FPKM for the 29,369 expressed
genes expressed in at least one of the 22 sequenced samples.

Supplemental Data Set 2.2. Normalized expression levels in FPKM for the 30,665 genes
expressed in at least one of the ten compartments after merge of replicates.

Supplemental Data Set 2.3. Compartment specificity (CS) scores for the 13,009 compartment-
specific genes.

Supplemental Data Set 2.4. Module assignment and module memberships (MMs) for the 9,361
genes selected for WGCNA.

Supplemental Data Set 2.5. BLASTX search of the 93 M18 genes that contain at least one
MRP-1-binding sub-motifs against the NCBI nr database (E-value < 10-6).

Supplemental Data Set 2.6. BLASTX search of the 93 M18 genes that contain at least one
MRP-1-binding sub-motifs against the Swissprot database (E-value < 10-6).
110

CHAPTER 3

Opaque-2 regulates a complex gene network associated with cell differentiation and
storage function of maize endosperm

3.1 ABSTRACT

Development of the cereal grain endosperm involves extensive cell differentiation processes to
enable uptake of sugar from the maternal plant and accumulation of large quantities of storage
products, such as carbohydrates and proteins, and their subsequent utilization during
germination. However, little is known about the molecular mechanisms that link the gene
regulatory programs underlying cell differentiation with the programs controlling storage product
synthesis and deposition, including the well-described activation of the zein gene family
members by the maize (Zea mays) bZIP transcription factor Opaque-2 (O2). To better understand
these processes, we mapped the in vivo binding sites of O2 in the inbred B73 endosperm early in
its expression using a chromatin immunoprecipitation coupled with high-throughput sequencing
approach in comparison with that of an o2 mutant (B73o2), and characterized the genes
differentially expressed in the two genotypes using RNA sequencing. The analysis identified
1,863 O2-modulated genes, including 186 putative direct targets and 1,677 indirect targets,
whose annotated/putative gene functions suggested a broad role for O2 in the regulation of cell
differentiation, storage product accumulation, maturation, and stress response of maize
endosperm. Temporal examination of the expression patterns of O2 targets in developing
endosperm of B73 versus B73o2 revealed at least two distinct modes of O2-mediated gene
activation. Target-gene co-activation and protein-protein interaction studies identified two
transcription factors—bZIP17 and NAKED ENDOSPERM2 (NKD2)—that are encoded by O2-
activated genes, which can in turn co-activate other O2-network genes with O2. NKD2 has
previously been shown to be involved in regulation of aleurone cell differentiation along with its
paralog NKD1. Accordingly, our ultrastructural analysis of aleurone cells of an o2 mutant
showed that the mutant differs from the wild type in anticlinal cell wall thickness and in the
abundance and size distribution of lipid bodies and protein storage vacuoles, suggesting that O2
is required for proper differentiation of aleurone. Collectively, our results provide new insights
into the complexity of the O2 regulatory network and the role of O2 in the regulation of cell
differentiation and function of maize endosperm.
111

3.2 INTRODUCTION

Endosperm is a filial seed structure that provides nutrients and signals essential for
embryogenesis and seedling germination (Li and Berger, 2012; Yan et al., 2014). In contrast to
dicotyledonous plants such as Arabidopsis (Arabidopsis thaliana) in which the endosperm is
eventually absorbed in part by the developing embryo, the endosperm of cereals persists
throughout seed development, constitutes a large proportion of the mature grain, and
accumulates a high level of storage compounds, including primarily starch and storage proteins
(Sabelli and Larkins, 2009; Li and Berger, 2012). As such, cereal grains are the foremost source
of human calories as well as industrial raw materials (FAO, 2015). However, our understanding
of how the gene regulatory programs controlling the storage function of cereal endosperm are
activated early on during endosperm development and the interplay between the storage program
and the programs associated with the earlier developmental processes such as cell differentiation
is limited.

Development of angiosperm seeds is initiated by double fertilization, a process during which one
of the two sperm cells fuses with the egg cell to produce the embryo and the other fertilizes the
central cell to produce the endosperm (Friedman et al., 2008; Hamamura et al., 2012). Upon
fertilization, in maize (Zea mays) seed, the nucleus of the fertilized central cell (primary
endosperm cell) undergoes free nuclear division, resulting in a coenocyte that becomes
cellularized by approximately 4 days after pollination (DAP) and then differentiates into four
main cell types that become distinguishable by about 6 DAP. These cell types are termed the
aleurone (AL), basal endosperm transfer layer (BETL), embryo-surrounding region (ESR), and
starchy endosperm (SE). By about 8 DAP, further differentiation of the SE results in at least four
distinct subregions, namely the basal intermediate zone (BIZ), central starchy endosperm (CSE),
conducting zone (CZ), and subaleurone (SA). Concurrent with cell differentiation, mitotic
proliferation of the maize endosperm also begins immediately after cellularization, and lasts until
about 8 to 12 DAP in the CSE, and 20 to 25 DAP in the SA and AL. Starting at about 8 to 10
DAP, the SE cells gradually switch from mitosis to endoreduplication, and accumulate storage
compounds. Eventually, the SE cells undergo maturation that begins at about 16 DAP and
involves programmed cell death and desiccation (Becraft, 2001; Olsen, 2001, 2004; Sabelli and
Larkins, 2009; Becraft and Gutierrez-Marcos, 2012; Leroux et al., 2014; Zhan et al., 2017).
112

One of the key characteristics of the highly differentiated SE cells is the expression of storage-
protein genes (Sabelli and Larkins, 2009; Becraft and Gutierrez-Marcos, 2012). Cereal
endosperm accumulates four main classes of storage proteins, including albumins, globulins,
prolamins, and glutelins (Shewry et al., 1995; Shewry and Halford, 2002). Prolamins, which
constitute the major storage proteins in maize endosperm, are termed zeins. Zein proteins are
further classified as α- (19- and 22-kD), β- (15-kD), γ- (16-, 27- and 50-kD), and δ- (10- and 18-
kD) zein and are encoded by distinct gene families (Coleman and Larkins, 1999; Woo et al.,
2001). Opaque-2 (O2) has been shown to regulate genes in nearly all these families through
binding to a GENERAL CONTROL OF NITROGEN4 (GCN4)-like motif (a.k.a. O2 box)
(Schmidt et al., 1990; Schmidt et al., 1992; Muth et al., 1996; Li et al., 2015). A mutation in the
maize O2 locus results in opaque mature kernels that show reduced levels of 22-kD α-zeins and
increased lysine content (Mertz et al., 1964; Schmidt et al., 1987). Previous transcriptome
analyses of o2 mutant endosperm indicated that, in addition to storage-protein gene expression,
O2 regulates genes involved in a diverse array of biological functions or processes, such as
synthesis and metabolism of carbohydrates and lipids (Hunter et al., 2002; Jia et al., 2007; Frizzi
et al., 2010; Hartings et al., 2011; Jia et al., 2013; Li et al., 2015). Nonetheless, whether these
processes are directly or indirectly regulated by O2 remains largely unknown. Protein-DNA
binding assays enabled identification of a number of direct target genes of O2; the best
characterized, canonical targets of O2 include multiple zein genes of various families, the b-32
gene that encodes a 32-kD type I ribosome-inactivating protein which has been implicated in
pathogen defense, the cyPPDK1 gene that encodes the endosperm-specific cytosolic isoform of
pyruvate orthophosphate dikinase putatively involved in carbon partitioning, the LKR/SDH gene
that encodes the bifunctional enzyme lysine-ketoglutarate reductase/saccharopine dehydrogenase
involved in lysine catabolism, and the O2 gene itself (Lohmer et al., 1991; Schmidt et al., 1992;
Cord Neto et al., 1995; Gallusci et al., 1996; Maddaloni et al., 1996; Kemper et al., 1999; Li et
al., 2015).

O2 likely functions in a complex with other regulatory proteins. First, O2 as a bZIP family
protein can homodimerize or heterodimierze in vitro with the O2-heterodimerizing proteins
(OHPs)—OHP1 and OHP2—that are paralogs of O2 (Pysh et al., 1993; Unger et al., 1993; Pysh
and Schmidt, 1996; Xu and Messing, 2008). The OHPs are also capable of binding the O2 box in
zein gene promoters in vitro (Pysh et al., 1993; Yang et al., 2016). Second, O2 has been shown to
113

physically interact with the DOF family TF PROLAMIN-BOX BINDING FACTOR (PBF) in
vitro (Vicente-Carbajosa et al., 1997). Third, PBF and OHP1/OHP2 have been shown to interact
in vitro and in vivo (Zhang et al., 2015). Genetic analyses have shown that O2, the OHPs, and
PBF co-activate expression of zein genes in a synergistic or additive manner, with O2 playing
the major role in regulation of 19- and 22-kD α-zein genes, and the OHPs playing the major role
in regulation of the 27-kD γ-zein gene (Zhang et al., 2015; Yang et al., 2016). In addition, O2 has
also been shown to interact with several other nuclear proteins, including the MADS-box family
TF protein MADS47, the transcriptional-adaptor protein ALTERATION/DEFICIENCY IN
ACTIVATION2 (ADA2) and the histone acetyltransferase GENERAL CONTROL OF
NITROGEN5 (GCN5) to co-regulate O2 target gene expression (Bhat et al., 2003; Bhat et al.,
2004; Qiao et al., 2016). The partnership of O2 with multiple nuclear factors is likely associated
with the complexity of the O2 gene regulatory network (GRN).

Here, we used a coupled RNA sequencing (RNA-Seq) and chromatin immunoprecipitation


followed by high-throughput sequencing (ChIP-Seq) approach to identify genes directly or
indirectly regulated by O2 in the maize endosperm. Using RNA-Seq, we identified 1,863 genes
differentially expressed between the wild-type (WT) B73 inbred line and an o2 mutant in the
B73 background (B73o2). Among these, 186 genes were detected as putative direct targets of O2
by ChIP-Seq. Analyses of the direct and indirect O2 targets revealed a broad role of O2 in the
regulation of the gene expression programs associated with multiple key aspects of endosperm
development and function. An analysis of the temporal expression patterns of O2 targets in WT
versus mutant endosperm revealed two distinct modes of gene activation by O2. The
identification of several direct target gene-associated cis-motifs that are putatively bound by O2
cofactors, which we defined as other TFs that co-regulate direct target genes with O2, and the
observation that O2 directly activates two genes encoding O2 cofactors, provided further insights
into the key players in the O2 GRN. An ultrastructural analysis of the developing endosperm of
an o2 mutant revealed an important role for O2 in AL cell differentiation.

3.3 RESULTS

Identification of O2 target genes using chromatin immunoprecipitation coupled with


differential expression analysis
114

To identify O2-regulated genes, we first screened for genes differentially expressed in the
endosperm of the WT B73 versus a B73o2 mutant (Supplemental Text 3.1, Supplemental Figure
3.1, and Supplemental Table 3.1) using RNA-Seq of endosperm dissected at 15 DAP, a stage
shortly after the initiation of the storage program (Sabelli and Larkins, 2009; Li et al., 2014). For
each genotype, total RNAs extracted from biological triplicates were polyA-selected, reverse-
transcribed to cDNA, and sequenced using an Illumina HiSequation 2500 platform. Between
52.9 and 66.8 million paired-end reads were obtained for each sample, and 92.3% to 93.3% of
the reads were mapped to the maize reference genome (Supplemental Table 3.2). After excluding
the genes expressed at extremely low levels, the read counts for the remaining 28,580 genes were
normalized by edgeR, and statistically tested for differential expression. Among the tested genes,
1,863 genes were detected as differentially expressed genes (DEGs) [false discovery rate (FDR)
< 0.05], including 1,024 that were downregulated [log2-transformed fold change (FC) < 0] and
839 upregulated (log2FC > 0) in the mutant (Figure 3.1A; Supplemental Text 2, Supplemental
Figures 3.2 and 3.3, and Supplemental Data Set 3.1).

To determine the genome-wide direct target genes of O2, we performed a ChIP-Seq assay on two
biological replicates of the B73 endosperm at 15 DAP using a previously described anti-O2
antibody (Schmidt et al., 1992). As a negative control, we performed the same assay in
duplicates using endosperm from the B73o2 mutant. The ChIPed genomic fragments were
sequenced using an Illumina HiSequation 2500 platform to generate paired-end reads. Each
sample produced 25.6 to 38.3 million reads of which 68.5% to 79.8% were mapped to the maize
reference genome (Supplemental Table 3.3). The mapped reads of the biological replicates for
each genotype were subsequently merged and were used to identify O2-bound sequences using
the MACS software (Zhang et al., 2008; Feng et al., 2012). With a cutoff q-value of 0.05, MACS
detected 6,365 peaks corresponding to putative O2-bound regions in the maize genome. A major
fraction (70.0%) of the identified peaks were located >5 kb away from gene models (Figure
3.1B) as previously described for other maize TFs (Morohashi et al., 2012; Eveland et al., 2014).
However, this was a higher fraction than previously reported by Li et al. (2015), who showed
that 66% O2 peaks were located within 1 kb up- or downstream of gene models (Li et al., 2015).
We defined O2-bound genes as those with peaks detected in the genic regions or within 5 kb up-
or downstream of the annotated gene models. Based on these criteria, 3,282 putative O2-bound
genes were identified (Supplemental Data Set 3.2). Of the peaks associated with these genes,
115

most were located within either 5 kb upstream (18.5% of all peaks) or 5 kb downstream (8.5%)
to the gene models, while much smaller fractions were located within intronic (2.7%), 5’-
untranslated region (UTR; 0.1%), coding DNA sequence (CDS; 0.1%), or 3’-UTR (0.1%)
sequences (Figure 3.1B).

Overlay of the identified O2-bound genes with the O2-modulated genes identified 186 bound
modulated genes (here referred to as direct target genes of O2), including 147 shown to be O2
activated and 39 as O2 repressed (Figure 3.1C; Supplemental Data Set 3.3). By exclusion, the
unbound modulated genes (here referred to as the indirect targets of O2) included 877 O2-
activated genes and 800 O2-repressed genes, while the group that were detected as bound but
unmodulated by O2 included 3,096 genes (Figure 3.1C).

To gain insights into the association of the O2-modulated and/or bound gene sets with O2-
binding sites, we summarized the previously reported O2-binding sequences from the literature
(Supplemental Table 3.4) and determined the occurrences of these sequences within the
promoter regions of the O2-modulated and/or bound gene sets. The result showed that a
relatively larger fraction of the direct targets was associated with known O2-binding sites as
compared to the indirect targets or the bound unmodulated genes (Supplemental Figure 3.5).
Correspondingly, the enrichments of the O2-binding-site-associated genes among the activated
direct targets and the bound unmodulated genes were statistically significant (P < 10-5,
hypergeometric test), suggesting that our ChIP-Seq analysis successfully distinguished the direct
O2 targets from the indirect targets with respect to the association with previously identified O2-
binding sites. Notably, only a small fraction of the O2-bound genes (630/3,282) were associated
with the previously identified O2-binding sites, suggesting that O2 likely binds additional cis-
regulatory sequences that are yet to be characterized (see below).

We next evaluated the extent to which the absence of O2 modified the endosperm transcriptome
either directly or indirectly through a boxplot analysis of the log2-transformed RPKM data. The
analysis indicated that the O2-activated direct targets showed the largest decrease in overall gene
expression in the mutant as compared to the WT endosperm (average log2FC = -3.75) (Figure
3.1D). Conversely, the O2-repressed genes (both direct and indirect) showed a more modest
level of increase in mRNA in the mutant (average log2FC = 1.98 and 1.90, respectively). These
data support the primary role of O2 as a transcriptional activator (Schmidt et al., 1992).
116

An analysis of the expression patterns of the O2-activated direct targets using the available maize
gene expression atlas (Chen et al., 2014a) showed that the expression of these genes are highly
endosperm specific, mirroring the pattern of O2 itself (Supplemental Figure 3.6). Interestingly,
the expression of the O2-repressed direct targets also showed a slightly higher expression in the
endosperm/kernel than in other tissues. However, their median expression levels were
consistently lower than the O2-activated direct targets throughout endosperm/kernel
development, and the differences were highly significant (P < 0.001, Student’s t test) for all the
examined developmental stages after 8 DAP (Supplemental Figure 3.6).

Based on our criteria for defining O2-bound genes, we detected 166 O2 peaks within the range
of -5 kb to +5 kb of the gene models of the 186 direct O2 target genes (Supplemental Data Set
3.3). To gain insights into the O2-binding sites among these genes, we profiled the distribution of
the 166 peaks by counting the number of peaks per 100-bp bins and found an enrichment within
the -400 to +100 bp genomic regions of the genes (Figure 3.1E). A profiling of read distribution
over the same annotated gene models using a similar approach confirmed this pattern
(Supplemental Figure 3.7). These results were consistent with the reported localization of several
known O2-binding sites, which were detected within the 300-bp region upstream of the
translation start sites (Schmidt et al., 1992; Wu and Messing, 2012). We used two independent
approaches to verify binding of O2 to the O2 peaks. First, using electrophoretic mobility shift
assays (EMSAs), we tested 9 peaks (or promoter regions that partially overlap with an O2 peak)
associated with O2-activated genes, including the 18-kD δ-zein gene, bZIP17, G-BOX BINDING
FACTOR1 (GBF1), bZIP104, GRMZM2G084296, NAC78, NAKED ENDOSPERM2 (NKD2),
and TRYPTOPHAN AMINOTRANSFERASE RELATED3 (TAR3); and one peak that was
associated with the O2-repressed gene GRMZM2G443668 (Supplemental Table 3.5). The results
showed that all the tested sequences were bound by O2, and that this binding activity could be
partially or completely abolished by addition of competitor probes containing O2-binding sites
(Supplemental Figure 3.8). Second, in dual luciferase reporter (DLR) assays performed in
Nicotiana benthamiana with the O2 protein fused with a strong transcription activation domain
(Supplemental Figure 3.9A), 7 out of 13 peaks were confirmed to be bound by O2 (P < 0.05,
Student’s t test) (Supplemental Figure 3.9B and Supplemental Table 3.6). Among these 7 peaks,
six were associated with O2-activated genes, including the 15-kD β-zein gene, the 18-kD δ-zein
117

gene, Floury-2 (FL2), bZIP17, NKD2, and TAR3; and one was associated with the O2-repressed
gene LIP15 (encoding a 15-kD low temperature-induced protein).

To identify additional cis-regulatory sequences that mediate O2 binding and to gain further
insights into the regulatory protein complex involved in the regulation of O2 target genes, we
performed an analysis of DNA sequence motifs enriched within a 401-bp region centered at the
summits of each of the 166 peaks for the direct O2 target genes using the MEME-ChIP software
(Machanick and Bailey, 2011). This analysis identified 19 motifs (E-value < 0.05; Supplemental
Table 3.7), two of which (Motifs 5 and 15) contained a common sequence CCACGTCA that is
significantly similar to several known plant bZIP motifs, all characterized by an ACGT core
sequence (P < 0.05). Both motifs were shown to be centrally enriched in the analyzed genomic
regions (P < 0.05) (Figure 3.1F; Supplemental Table 3.7). These results indicate that Motifs 5
and 15 likely contain the cis-regulatory sequences bound by O2.

In addition, MEME-ChIP also identified several significantly enriched cis-motifs that were not
centrally enriched (Supplemental Table 3.7). The enrichment of these motifs within the detected
O2-bound genomic regions suggest that they may contain binding sites of O2 cofactors. For
example, the most prevalent sequence variants of Motif 2 all contained the common sequence
TGTAAAG (Supplemental Table 3.7), which was indeed identical to the core sequence of the
known PBF-binding site (i.e., the P box) (Vicente-Carbajosa et al., 1997). Consistent with
previous reports showing the co-existence of the O2 box and P box in the promoter regions of
seed storage-protein genes (Vicente-Carbajosa et al., 1997; Hwang et al., 2004), the enrichment
of Motif 2 suggests that many of the direct O2 targets are co-regulated by PBF or, alternatively,
other endosperm-expressed DOF family TFs. Moreover, Motif 10 [G(A/G)CATGCATG]
(Supplemental Table 3.7) was shown to be significantly similar to the known binding site of
ABI3 (P < 0.05), a seed-specific ABI3-VP1 family TF in Arabidopsis. Close inspection of the
sequence of Motif 10 showed that this motif in fact contained the previously reported RY motif
(CATGCATG), which is a conserved cis-regulatory sequence shown to be involved in the
activation of many seed-expressed genes by ABI3-VP1 family TFs in Arabidopsis, legumes and
cereals (Santos-Mendoza et al., 2008). This suggests that an endosperm-expressed ABI3-VP1
family TF in maize may bind to the RY-like Motif 10 to regulate a subset of the direct O2 target
genes. In addition, the identification of the motifs similar to known binding sites of other TFs
118

(e.g., SOC1, AtMYB15, FLC, etc.; Supplemental Table 3.7) indicates the involvement of TFs
from other families in the regulation of O2 targets. Together, these results suggest that many
additional TFs, which are yet to be identified, are involved in the regulation of O2 target genes in
combination with O2.

O2 regulates a diverse set of genes involved in endosperm storage product synthesis and
accumulation

To gain a comprehensive understanding of functions of the O2-regulated gene network, we


identified the GO terms significantly enriched (FDR < 0.05) among the O2 target genes (directly
activated/repressed and indirectly activated/repressed) using the Blast2GO software (Conesa and
Gotz, 2008). The most significantly enriched GO term for the O2-activated direct targets was the
molecular function “nutrient reservoir activity” (Figure 3.2A), with nearly all the associated
genes encoding zein proteins (21 of 22, with GRMZM2G325920 annotated as having an
uncharacterized function). A less significant molecular function enriched for the activated direct
target genes was “proline dehydrogenase activity” that was associated with two proline oxidase
genes (GRMZM2G053720 and GRMZM2G117956). The most significant biological process
enriched for this target gene set was “carboxylic acid metabolic process,” which was associated
with 16 genes, including the LKR/SDH, cyPPDK1, and cyPPDK2 genes, while the other two
associated enriched biological processes were “glutamine family amino acid biosynthetic
process” and “proline catabolic process”. Consistent with the GO term enrichments, an analysis
of metabolic pathways associated with the O2-activated direct targets using the CornCyc tool
revealed genes involved in glycine biosynthesis, L-glutamine biosynthesis, proline metabolism
(as part of the L-Nδ-acetylornithine biosynthesis pathway) and L-glutamate degradation
(Supplemental Data Set 3.4). Together, these GO term enrichments and pathway associations
indicate that O2 directly activates genes encoding storage proteins and enzymes involved in
amino acid synthesis and metabolism, as well as regulating genes associated with partitioning of
carbon between carbohydrates and proteins in maize endosperm.

We also closely examined all the known zein genes, as canonical direct targets of O2, for
expression and O2 binding. Of the 40 zein genes annotated in the B73 genome (Chen et al.,
2014a), three genes including one 22-kD α-zein (GRMZM2G535393) and two 19-kD z1A sub-
119

family α-zein genes (GRMZM2G489581 and GRMZM2G514379) were excluded from our
differential gene expression analysis due to low levels of expression; and of the remaining 37
zein genes, 34 were shown to be modulated by O2 and three (all γ-zein genes) did not display
significant differential expression levels at 15 DAP although found to be bound by O2
(Supplemental Figure 3.8). All the differentially expressed zein genes were downregulated in the
o2 mutant (log2FC < -4.0), indicating that O2 acts as an activator of all the modulated zein genes.
The combination of the differential gene expression analysis and O2-binding data suggested that
different zein families are regulated by O2 in different manners at 15 DAP. The 15-kD β-zein
and the 18-kD δ-zein genes were directly activated by O2, whereas the α-zein gene family
members (31 genes) were identified as either direct or indirect target genes of O2 (Supplemental
Figure 3.10). Among the latter, the 19-kD α-zein gene sub-families z1A and z1B, and the 22-kD
α-zein gene family each included both direct and indirect, while the 19-kD subfamily z1D
included only indirect gene targets (Supplemental Figure 3.10). Furthermore, different families
of zein genes were shown to be downregulated in the mutant to different extents. The most
dramatically downregulated gene family was the 22-kD α-zein family, with log2FC ranging from
-10.5 to -9.8, while the 19-kD α-zein sub-families z1A, z1B, and z1D showed a relatively lower
extent of downregulation, with log2FC ranging from -8.7 to -7.6, -7.5 to -5.3, and -9.36 to -4.79,
respectively (Supplemental Figure 3.10 and Supplemental Data Set 3.1). These data suggest that
the 22-kD α-zein family is most crucially dependent on activation by O2, which is consistent
with a previous report showing that the 22-kD zein proteins are almost absent in the o2 mutant
endosperm while the other zein proteins were reduced to a lesser extent (Kodrzycki et al., 1989).
As noted above, the γ-zein genes (including the 16-kD, 27-kD, and 50-kD zein genes) were
associated with O2 peaks but were not modulated by O2 based on our data (Supplemental Figure
3.10); this observation suggests that although O2 binds the upstream regions of some target
genes, their expression is not significantly dependent on this binding, at least at the assayed stage
of 15 DAP.

The O2-activated indirect targets were enriched for a more diverse set of GO terms likely due to
the larger number of associated genes as compared to the direct targets (Figure 3.2B). Notably,
the most significantly enriched GO term for this gene set was also “nutrient reservoir activity”
associated with 16 genes, including at least 10 known zein genes (see above). Accordingly, many
other GO terms enriched for this gene set, including the molecular functions “3-deoxy-7-
120

phosphoheptulonate synthase activity,” “2-oxoisovalerate dehydrogenase (acylating) activity,”


and “methylmalonate-semialdehyde dehydrogenase (acylating) activity,” and the biological
processes “aromatic amino acid family biosynthetic process,” and “valine biosynthetic process,”
were associated with amino acid or protein synthesis/metabolism. Likewise, metabolic pathway
analysis also detected genes involved in synthesis of aromatic compounds (e.g., chorismate, 4-
hydroxybenzoate, and benzoate) or branched-chain amino acids among the indirect target genes
(Supplemental Data Set 3.4).

Interestingly, the O2-activated indirect targets were also enriched for many GO terms related to
the synthesis/metabolism of carbohydrates and lipids. For example, the most significantly
enriched biological process was “single-organism carbohydrate metabolic process,” which was
associated with 41 genes, including the genes encoding a sucrose synthase gene [SUCROSE
SYNTHASE2 (SUS2)], a malate synthase [MALATE SYNTHASE1 (MAS1)], a glucose-1-
phosphate adenyltransferase [ADP GLUCOSE PYROPHOSPHORYLASE2 (AGP2)], and two
fructokinases [FRUCTOKINASE1 (FRK1) and FRK2], that are involved in carbohydrate
synthesis/degradation pathways (Supplemental Data Set 3.4). On the other hand, this gene set
was enriched for many lipid storage/metabolism-associated GO terms, including “acyl-CoA
oxidase activity,” “lipid metabolic process,” “lipid storage,” and “monolayer-surrounded lipid
storage body,” and the metabolic pathway analysis confirmed the putative functions of many of
these genes in the biosynthesis/metabolism of lipid/fatty acids (Supplemental Data Set 3.4). In
addition, the O2-activated indirect targets were also enriched for “response to abiotic stimulus,”
“response to herbicide,” “response to absence of light,” “response to desiccation,” etc.,
suggesting that O2 also indirectly regulates abiotic stress responses of the endosperm.

The subset of genes identified as directly repressed by O2 did not show significant enrichment of
any GO terms, which was likely due to the limited functional annotations for most of these
genes. The genes with available annotation included only three TF genes (LIP15, MYBR10, and
Orphan154) and one diacylglycerol O-acyltransferase gene (GRMZM2G042356) that functions
in triacylglycerol biosynthesis (Supplemental Data Set 3.3). Manual inspection of putative
functions of the other genes in this group identified a putative sugar transporter gene
(GRMZM2G159559) and a putative glutathione γ-glutamylcysteinyltransferase gene
(GRMZM2G038170) (Supplemental Data Sets 3.5 and 3.6) that are likely associated with
121

storage product synthesis/metabolism. On the other hand, the gene subset identified as O2-
repressed indirect targets was most significantly enriched for the biological process “translational
termination” and the molecular function “translation release factor activity, codon specific”
(Figure 3.2C), indicating that O2 likely facilitates storage product synthesis and deposition in the
endosperm through indirect repression of translational termination. In addition, this group of
genes was also enriched for “protein serine/threonine kinase activity” and “protein
phosphorylation,” suggesting that O2 indirectly represses some protein kinase activity in the
endosperm. Taken together, O2 directly or indirectly regulates a large number of downstream
genes involved in the synthesis/metabolism of a wide range of storage products deposited in the
endosperm, including storage proteins, carbohydrates, and lipids.

Two modes of O2-mediated gene activation can be distinguished from the temporal
patterns of target gene expression

To understand the dynamics of O2 network gene expression during early endosperm


development, we interrogated the previously identified time-course programs of endosperm
mRNA accumulation (Li et al., 2014) for timing of O2 activation versus its targets. Based on the
temporal expression data, O2 exhibited an “up@8DAP” pattern; that is, it showed a relatively
low expression in the 0- through 6-DAP kernels but high expression in the 8-, 10-, and 12-DAP
endosperm. Comparison of the O2-modulated and/or bound gene sets with the temporally
upregulated gene sets showed that the O2-activated direct targets significantly overlapped with
the up@8DAP and up@10DAP gene sets (P < 10-5, hypergeometric test) (Figure 3.3A).
Similarly, the O2-activated indirect target gene set also significantly overlapped with up@8DAP
and up@10DAP gene sets (Figure 3.3A), indicating a similar timing of activation of both the
direct and indirect targets of O2. Interestingly, a significant portion of the O2-repressed indirect
target gene set also showed an upregulated pattern at 8 and 10 DAP in WT endosperm (i.e.,
up@8DAP and up@10DAP) (Figure 3.3A). This suggests a role for O2 in dampening or tightly
modulating the expression of genes activated through O2-dependent or O2-independent
activation during early endosperm development. These results nonetheless suggest that the
temporal upregulation of a substantial portion of the direct and indirect O2 targets correlates
strongly with O2’s upregulation.
122

From our previous temporal data, we also detected a significant number of O2-activated indirect
target genes that showed a temporally upregulated pattern in the WT kernel/endosperm at 4 or 6
DAP (i.e., up@4DAP or up@6DAP) (Figure 3.3A), preceding the detected upregulation of O2 at
8 DAP (Li et al., 2014). Coincidently, we also detected a handful of direct targets as showing the
same upregulated pattern at 4 and 6 DAP (Figure 3.3A). Therefore, we hypothesized that a
portion of the O2-regulated network would be activated by an O2-independent process and then
subsequently maintained or further activated by O2 after 8 DAP. To test this, we used reverse
transcription quantitative PCR (RT-qPCR) to measure the mRNA levels of representative genes
that were detected to be directly or indirectly activated by O2, including O2 itself, in a
developmental time series of endosperm from 6 to 30 DAP in both B73 and B73o2 (Figures 3.3B
to 3.3M; Supplemental Data Set 3.8). Among the 18 O2-activated direct targets tested, 11 genes
were significantly downregulated in the o2 mutant at 12, 15, and/or 18 DAP (Figures 3.3C to
3.3M; Supplemental Data Set 3.8). Surprisingly, these genes exhibited at least two distinct
temporal patterns of activation in the WT. Group 1 genes showed either an undetectable [b-32
and GRMZM2G025763 (a 19-kD z1B sub-family α-zein gene)] or nearly undetectable (azs22-4
and the 15-kD β-zein gene) level of expression in the WT before O2 mRNA was detected (6 to 8
DAP; Figure 3.3B), and showed an increase in expression in subsequent stages (Figures 3.3C to
3.3F). On the other hand, Group 2 genes, including the 18-kD δ-zein gene, bZIP17, GBF1,
GRMZM2G117956, LKR/SDH, NAC122, and NKD2, were already highly expressed even before
the detectable accumulation of O2 mRNA [Ct ≤ 31.37 ± 2.31 (mean ± SEM) at 6 DAP]
(Figures 3.3G to 3.3M). These results suggest that O2 is required for the early activation of
Group 1 genes whereas activation of Group 2 genes is O2 independent. In addition, for all Group
2 genes, the most dramatic downregulation in the mutant endosperm were observed at 15 or 18
DAP, indicating that O2 is essential for maintenance of their expression. Therefore, not all O2
direct target genes are necessarily activated by O2 during early stages (6 to 8 DAP) of
endosperm development, some being activated via O2-independent regulatory processes.
Moreover, it is noteworthy that all genes in Groups 1 and 2, as well as O2 itself, exhibited a
dramatic temporal upregulation detected during the period from 15 or 18 DAP to 22 DAP in the
o2 background, with some genes showing mRNA levels as high as those of the WT by 22 DAP
(Figures 3.3B to 3.3M). These results highlight a late, O2-independent gene regulatory program
that is also responsible for maintaining a high expression level of the O2-activated direct targets
123

later in endosperm development. Taken together, O2 regulates a highly dynamic network of


genes whose regulation highlights three periods of endosperm development.

O2 directly activates a novel O2-heterodimerizing-protein gene

In order to begin to identify key regulatory components of the O2 network, we explored the
nature of TF genes acting downstream of O2. Our RNA-Seq analysis detected at least 93 TF
genes modulated by O2 (Supplemental Data Set 3.1). Overlay of the O2-modulated genes with
the O2-bound genes identified 15 of the 93 TF genes as putative direct O2 target genes,
including 12 O2-activated and three O2-repressed TF genes; the former included 4 bZIP genes
(including O2 itself), 3 NAC genes, 2 Homeobox genes, and 3 genes belonging to the C2H2,
C3H, and CCAAT-HAP5 families, respectively, while the latter included a bZIP gene, a MYB-
related gene, and a putative TF gene without a recognized gene family (Orphan154)
(Supplemental Data Set 3.3). An analysis of the expression patterns of these O2-regulated TF
genes throughout development showed that, similar to O2, most of these genes are preferentially
expressed in the endosperm albeit at relatively lower levels than O2 (Supplemental Figure 3.11).
Notably, two of the five bZIP genes, namely LIP15 (repressed by O2) and GBF1 (activated by
O2), have been shown to be induced by low temperature and hypoxia, respectively (de Vetten
and Ferl, 1995; Kusano et al., 1995). The recently characterized C2H2 zinc finger gene NKD2
encoding one of the two paralogous INDETERMINATE domain TFs (NKD1 and NKD2), which
are central regulators of many endosperm developmental processes, including proper AL and SE
cell differentiation and storage product deposition (Yi et al., 2015; Gontarek et al., 2016), was
also detected as an O2-activated direct target gene (Supplemental Data Set 3.3). Together, the
detection of these 15 TF genes as direct targets of O2 suggests a broad, direct role for O2 in
endosperm development that encompasses regulation of cell differentiation, storage function, and
response to abiotic stresses.

Nearly 78 TF genes were detected among the indirect O2 target genes (58 O2-activated and 20
O2-repressed genes), corresponding to 29 different TF families. Many of these TF genes have
been shown to be involved in the regulation of various aspects of endosperm development. For
instance, the O2-repressed genes included PBF, which is another key regulator of the endosperm
storage program as discussed above. The PBF mRNA level was increased by about two fold in
124

the B73o2 mutant (log2FC = 0.98; Supplemental Data Set 3.1), suggesting that PBF expression is
indirectly repressed by O2 in the WT. In addition, among the O2-activated indirect targets is the
recently reported Floury-3 gene encoding a PLATZ family TF that functions in grain filling (Li
et al., 2017), and the Viviparous-1 (VP1) gene encoding an ABI3-VP1 family TF that plays an
essential role in regulation of maize seed maturation and germination (Hoecker et al., 1995).
Therefore, as each O2-regulated TF likely regulates a set of target genes in the endosperm, the
direct or indirect regulation of the TF genes by O2 in part can explain the large number (1,677)
of indirect O2 targets and their associated diversity of biological functions.

To further understand the nature of the TF genes under O2 regulation, we tested for enrichment
of the annotated TF gene families among the O2-modulated and/or bound gene sets and detected
a significant enrichment of bZIP family TF genes among the O2-activated direct targets (P =
3.1×10-5, hypergeometric test) (Supplemental Figure 3.12). To understand the relationship
between O2 and the (other) O2-regulated bZIP genes, we performed a phylogenetic analysis of
all the annotated maize bZIP proteins. The phylogeny showed that the proteins encoded by two
of the O2-regulated bZIP genes—bZIP17 and bZIP104—are in fact in the same well-supported
clade as O2 and the previously reported OHP proteins (Supplemental Figure 3.13). Interestingly,
in a yeast two-hybrid (Y2H) screen of a 15-DAP maize endosperm cDNA library using a
fragment of the O2 protein (termed O2F4) as the bait, we detected bZIP17 as a putative O2-
interacting protein with moderate confidence of interaction (Supplemental Text 3 and
Supplemental Figure 3.14). Therefore, we hypothesized that bZIP17 can form heterodimer with
O2 and co-regulate O2 target genes with O2, and tested this hypothesis using three approaches.
Using a directed Y2H assay, we were unable to detect interaction between O2F4 that was fused
with the DNA-binding domain of the yeast transcriptional activator GAL4, and the full-length
bZIP17 protein fused with the GAL4 transcriptional activation domain (Supplemental Figure
3.15). Nonetheless, a reciprocal set of in vitro pull-down assays of HA-/Myc-tagged O2 with
Myc-/HA-tagged bZIP17 proteins showed that the two proteins interact in vitro (Supplemental
Figure 3.16). This interaction was further confirmed in vivo using a bimolecular fluorescence
complementation (BiFC) assay performed in leaf epidermal cells of N. benthamiana, in which
O2 and bZIP17 were fused to the N- and C-terminal portions, respectively, of a yellow
fluorescent protein (YFP) variant, Venus, and strong fluorescence was observed in the nuclei of
the infiltrated N. benthamiana leaf cells (Figure 3.4A). To understand if O2 and bZIP17 can co-
125

activate target gene expression, we tested the six peaks that were associated with O2-activated
targets and were confirmed to be bound by O2 in N. benthamiana (Supplemental Figure 3.9B)
for gene co-activation by O2 and bZIP17 using DLR assay with the TFs each expressed as an
intact protein (Supplemental Figure 3.17). For one of the six peaks, which was upstream to the
FL2 gene (encoding a 22-kD α-zein protein), the reporter gene expression was significantly
elevated when O2 and bZIP17 were both expressed compared to when O2 alone was expressed
[P < 0.05, one-way ANOVA with post-hoc Tukey’s honestly significant difference (HSD) test]
(Figure 3.4B; Supplemental Table 3.8). This indicates that bZIP17 can enhance the trans-
activation of FL2 by O2 by binding to the associated O2 peak, whereas bZIP17 alone was unable
to trans-activate it. Together, these data suggest that O2 interacts with bZIP17 and they can co-
activate at least a subset of target genes in a synergistic manner likely through binding to the
same cis-regulatory sequence in the form of a bZIP heterodimer.

O2 and NKD2 co-regulate a gene network associated with AL cell differentiation

Our identification of NKD2 as a direct target of O2 suggested that O2 and NKD2 co-regulate a
downstream gene network with direct mutual regulation. The NKD TFs have previously been
shown to directly trans-activate expression of O2 and a 22-kD α-zein gene in isolated AL
protoplasts (Gontarek et al., 2016). A reported transcriptome analysis of an nkd1;2 double
mutant coupled with analysis of NKD2-binding sites led to the identification of 1,059 and 1,050
putative direct targets of NKD2 in the SE and AL, respectively (Gontarek et al., 2016). We
identified a total of 34 of these genes as direct O2 targets in our analysis, and the overlap was
highly significant (P = 1.7×10-26, hypergeometric test) (Figure 3.5A; Supplemental Table 3.9).
To understand if NKD2 and O2 can co-regulate target gene expression through O2-bound
sequences, we performed DLR assays on the six peaks that were associated with O2-activated
targets and were confirmed to be bound by O2 in N. benthamiana (Supplemental Figure 3.9B),
with O2 and NKD2 each expressed as intact proteins (Supplemental Figure 3.17 and
Supplemental Table 3.10). In fact, five of these peaks were determined to be associated with
putative common direct targets of both O2 and NKD2 , and included the 15-kD β-zein gene, the
18-kD δ-zein gene, FL2, TAR3, and NKD2 itself (Supplemental Table 3.6). The results showed
that O2 and NKD2 co-activated three of the six peaks in a synergistic manner (Figure 3.5B),
126

suggesting that the two TFs act collaboratively to activate a portion of the O2- and NKD2-
regulated gene networks. As a TF protein that also functions immediately downstream of O2, we
hypothesized that NKD2 could directly regulate a subset of the indirect O2 target genes. To test
this hypothesis, we examined the indirect O2 target genes for presence of putative NKD2 direct
target genes (Gontarek et al., 2016). The result showed that 120 of the 1,677 indirect O2 gene
targets are likely regulated directly by NKD2 in SE and/or AL based on a highly significant
overlap (P = 1.3×10-44, hypergeometric test) (Figure 3.5C; Supplemental Data Set 3.7). These
results indicate that NKD2 can act both as a co-activator and as a downstream effector of O2 in
regulation of both direct and indirect target genes.

The above data suggested that O2 may perform a role in AL development by regulating NKD2
and its downstream genes. To understand if O2 is required for proper AL differentiation, we
performed an ultrastructural analysis of the AL of an o2 mutant (in W22 background) in
comparison to a WT W22. A preliminary analysis of 12- and 16-DAP AL showed that the
mutant differs from the WT in cell wall thickness as well as in the abundance and size of lipid
bodies and protein storage vacuoles (PSVs), albeit the effects were rather subtle (Supplemental
Figure 3.18). Previously it has been shown that development of AL cells (and in particular the
PSV and their accumulation of proteinaceous contents) begins around 15 DAP (Kyle and Styles,
1977). Therefore, we used 20-DAP samples for a full ultrastructural and quantitative analysis of
AL. The results showed that the o2 AL cells had significantly thinner anticlinal cell walls than
the WT (P < 0.05, Student’s t test), and displayed cytoplasmic contents with fewer lipid bodies
but more PSVs than the WT (based on the area of cytoplasm occupied; Figures 3.6A to 3.6D).
An examination of lipid body sizes revealed that the mean size of the lipid bodies in the o2 AL
was significantly smaller than in the WT (Figure 3.6E). In fact, the mutant significantly differed
from the WT in size class distribution (P < 0.05, Pearson’s Chi-squared test), having more lipid
bodies of the smallest sizes (≤ 0.7 µm2) (Figure 3.6F). In contrast, the mean size of the PSV was
significantly larger in o2 than in the WT (Figure 3.6G); the size class distribution of PSVs
differed between the two genotypes with o2 having more large-size PSVs (> 1.5 µm2) (Figure
3.6H). Taken together, our data indicate that O2 plays an important role in regulation of AL cell
differentiation and its storage function, likely through a direct mutual regulatory interaction with
NKD2.
127

3.4 DISCUSSION

We have begun dissecting the O2 GRN in maize endosperm by initially identifying endosperm-
expressed genes directly or indirectly regulated by O2. Our mRNA profiling of 15-DAP
endosperm from B73o2 versus B73 identified 1,863 O2-modulated genes (Figure 3.1A and
Supplemental Data Set 3.1). The number resembles those of the DEGs previously identified in a
W22o2 mutant versus W22 (Li et al., 2015) and in W64Ao2 versus W64A (Zhang et al., 2016)
using RNA-Seq. However, comparison of the DEGs that we identified versus four previously
published DEG sets (Hunter et al., 2002; Frizzi et al., 2010; Jia et al., 2013; Zhang et al., 2016)
showed poor overlap, with only 6 commonly detected DEGs identified from all studies
(Supplemental Text 3.2 and Supplemental Figure 3.3). Such discrepancy could be attributed to at
least two reasons. First, as has been shown previously, O2 may modulate partially overlapping
sets of genes depending on the genetic background (Bernard et al., 1994), which may also be
affected by the specific type of o2 mutant allele (Jia et al., 2007). Second, as suggested by our
temporal expression assays, the extent of dysregulation of a given O2-modulated gene in the o2
mutant versus WT depends on the developmental staging; therefore, each of the previous studies
may have detected a distinct set of O2-modulated genes due to the specific stages selected for
analysis (ranging from 15 to 25 DAP).

The ChIP-Seq analysis detected 6,365 putative O2-bound genomic regions associated with 3,282
genes (Supplemental Data Set 3.2). Among these putative O2-bound genes, 186 were detected as
direct target genes of O2 based on their differential expression in WT versus o2 backgrounds
(Supplemental Data Set 3.3). Of these, we detected 12 of 35 O2-activated direct target genes
(Supplemental Figure 3.19 and Supplemental Data Set 3.9) identified by Li et al. (2015).
Compared to this previous study, our ChIP-Seq and RNA-Seq analysis also identified 134
additional O2-activated direct targets, including six 22-kD α-zein genes, the LKR/SDH gene, and
the O2 gene itself (Supplemental Data Set 3.9), which were previously identified as canonical
targets of O2 based on EMSAs, DNase I footprinting, ChIP-quantitative PCR, in vivo trans-
activation assays, and/or differential mRNA/protein accumulation in an o2 mutant versus WT
(Schmidt et al., 1990; Lohmer et al., 1991; Schmidt et al., 1992; Kemper et al., 1999; Locatelli et
al., 2009).
128

The GO term enrichment and metabolic pathway analyses of O2-modulated and/or bound gene
sets revealed that O2 likely regulates storage-protein gene expression and the associated amino
acid synthesis/metabolism processes both directly and indirectly (Figure 3.2; Supplemental
Figure 3.10 and Supplemental Data Set 3.4). The varied extent of downregulation of the different
zein gene families/sub-families in o2 is consistent with the previous studies of zein mRNA
and/or protein levels in the mutant (Kodrzycki et al., 1989; Zhang et al., 2015; Yang et al., 2016),
supporting the notion that distinct zein gene families are regulated by diverse gene regulatory
mechanisms, likely involving combinatorial activities between O2 and other endosperm-
expressed TFs. In addition, our analyses provide evidence that the regulation of carbohydrate and
lipid synthesis/metabolism by O2 indirectly is primarily an indirect effect.

The analysis of the overlaps between the O2-modulated and/or bound gene sets and the
temporally upregulated gene sets that we identified previously (Li et al., 2014) revealed a tightly
linked timing of activation between O2 and the O2 targets, and also detected a number of targets
that are likely activated before the activation of O2 itself (Figure 3.3A). Accordingly, our
temporal real-time expression assays of o2 versus WT endosperm identified at least two distinct
modes of gene activation by O2, with three distinguishable periods of gene regulatory functions
associated with endosperm development (Figure 3.7A). The two modes differ most dramatically
in the early (by ~12 DAP) and mid (~12 to 18 DAP) periods. In the first mode, the activation of
the O2 targets at the early period is likely dependent on the expression of O2 (Figures 3.3C to
3.3F), whereas in the second mode, O2 is not required for the activation of the O2 targets at the
early period, but is likely required for the maintenance of their expression at the mid period
(Figures 3.3G to 3.3M). In both cases, there are likely additional regulators that can activate the
O2 target-gene expression in an O2-independent manner during the late period, which begins at
about 18 to 22 DAP (Figures 3.3C to 3.3M). The nature of the TFs regulating the early activation
of O2 target genes remains unclear but this program is likely involved in establishing the initial
active state of the target genes with or without O2. Once activated, another set of regulators
seems to be required for the maintenance of the high level of expression of the O2 network genes
in the subsequent periods. These regulators likely include, in addition to O2 itself, the TFs that
co-regulate target genes with O2 (e.g., bZIP17) and/or the TFs that are encoded by O2 target
genes and acting downstream of O2 (e.g., NKD2). Interestingly, the expression of such
regulators may be initiated early in endosperm development in an O2-independent manner but
129

their continued high expression may be dependent on O2 (Figures 3.3H and 3.3M). Moreover,
these factors may not be necessarily expressed in an endosperm-specific manner during the plant
life cycle (Supplemental Figure 3.11). The distinct behavior of the O2-network genes during the
mid and late periods—that is, the significant downregulation of O2 target genes in the mid
period, and a near but not complete recovery of expression in the late period—is likely dictated
by the requirement of O2 in the mid period and a partial requirement of O2 in the late period.
The latter may be due to a partial compensation or redundancy displayed by otherwise non-
endosperm-specific bZIP proteins in absence of the endosperm-specific O2. Together, although
our data do not allow us to determine whether O2 binds the direct target genes (detected at 15
DAP) during all the analyzed stages of endosperm development, the temporal expression
patterns suggest that the O2 GRN is modulated finely based on the period of endosperm
development. Therefore, further understanding of the O2 network will require comprehensive
temporal analyses of gene expression programs and the genome-wide DNA-binding patterns of
O2 and its target genes throughout endosperm development.

O2 has long been suggested as a transcriptional activator, and there was little prior evidence to
suggest that O2 also functions as a direct transcriptional repressor. In our analysis, we detected
839 O2-repressed genes, including 39 putative direct targets. The O2-repressed direct target gene
set showed temporally upregulated patterns throughout endosperm development (Supplemental
Figure 3.6), suggesting that O2 represses these genes quantitatively to a certain degree but does
not seem to affect their otherwise global and temporally upregulated expression pattern in the
WT. These RNA-Seq-based observations were further supported by our temporal RT-qPCR
analysis of individual genes. For instance, two of the O2-repressed direct targets examined in our
temporal analysis (GRMZM2G443668 and GRMZM2G034623) showed upregulated expression
patterns in both B73 and B73o2 endosperm (Supplemental Data Set 3.8). As discussed above, it
is yet to be determined if these genes are bound by O2 at the later developmental stages, and thus
it is unclear if the upregulation of mRNA levels in the o2 mutant is due to the loss of direct
transcriptional repression by O2. Nevertheless, these data strongly suggest that O2 functions as
both an activator and a repressor.

Our cis-motif analysis of the O2-bound genomic regions revealed centrally enriched bZIP-like
binding motifs, which likely act as cis-regulatory sequences for O2 and related bZIP TFs (Figure
130

3.1F). In fact, the conserved sequence of the two bZIP-like motifs (CCACGTCA) was essentially
identical to the O2-binding motif that was identified by Li et al. (2015) and exhibited strong
binding to the O2 protein in an EMSA. On the other hand, the identification of other cis-motifs
sheds new light on additional storage program regulators in maize endosperm. For example, the
MEME-ChIP-identified Motif 10 was significantly similar to the RY motif in Arabidopsis,
which has been shown to be involved in the activation of seed storage-protein gene expression
by several ABI3-VP1 family TFs, including FUSCA3 (FUS3) and LEAFY COTYLEDON2
(LEC2) (Reidt et al., 2000; Braybrook et al., 2006). In fact, the RY-like motifs are likely
functionally conserved between monocots and eudicots, because a RY-like motif in the cereal
species barley (Hordeum vulgare) is bound by an ortholog (HvFUS3) of the Arabidopsis FUS3
(Moreno-Risueno et al., 2008). Moreover, the Arabidopsis ABSCISIC ACID INSENSITIVE3
(ABI3) protein physically interact with bZIP10 and bZIP25, which are orthologs of the maize O2
in Arabidopsis, to synergistically regulate storage-protein gene expression in the Arabidopsis
embryo (Lara et al., 2003). In addition, heterologous expression in yeast indicates that HvFUS3
physically interacts with the BARLEY LEUCINE ZIPPER2 (BLZ2) protein, which is an O2-like
TF (Moreno-Risueno et al., 2008). These reported data on the conserved binding between ABI3-
VP1 family TFs and RY motifs, and protein-protein interactions between ABI3-VP1 TFs with
bZIP TFs in Arabidopsis and cereal plants, together with our identification of the RY motif
within the O2-bound genomic regions, strongly suggest that O2 may interact with an ABI3-VP1
TF to regulate a subset of its target genes in maize endosperm. Moreover, these interactions
suggest that this regulatory module evolved prior to the divergence of monocots and eudicots.

The identification of many TF genes, including several bZIP genes, as direct targets of O2
provides further insights into the complexity of the O2 GRN. The bZIP17 gene, which is a
paralog of O2, was identified as a direct target of O2 based on our ChIP-Seq and RNA-Seq
analysis, and the bZIP17 protein can interact with O2 to co-activate zein-gene expression in DLR
assays. This suggests that O2 and bZIP17 may form a coherent feed-forward loop—that is, the
output (activation or repression) of the direct regulation (e.g., O2 directly activates FL2) is the
same as the overall output of the indirect regulation (e.g., O2 indirectly activates FL2 through
activation of bZIP17) (Mangan and Alon, 2003)—in regulation of some of O2-regulated genes
(Figure 3.4; Supplemental Data Set 3.3). Furthermore, these data indicate that bZIP17 is a
putative novel regulator of the endosperm storage program that may play redundant roles with
131

the O2/OHP proteins. Interestingly, our DLR assays showed that O2 and NKD2 co-activate
bZIP17 (Figure 3.5B), indicating a potential role for bZIP17, in combination with O2 and the
NKDs, in regulation of endosperm cell differentiation. In addition, the detection of the two
stress-responsive bZIP genes—LIP15 and GBF1—as direct O2 targets suggests that O2 regulates
responses to low temperature and hypoxia through direct regulation of these TF genes
(Supplemental Data Set 3.3); coincidently, the GO term enrichment analyses indicate that the
O2-activated indirect targets were enriched for biological processes related to abiotic stress
responses (Figure 3.2B). Because each bZIP TF has the potential to dimerize with certain other
bZIPs and to co-regulate target genes, the O2-regulated bZIP genes remain to be further
characterized for interaction at the protein level and for target-gene co-activation activities, in
order to further understand the O2/bZIP-regulated gene network and the roles of the bZIP
proteins in the regulation of endosperm development.

O2 is directly regulated by NKD1 and NKD2 based on the downregulation of O2 mRNAs in the
15-DAP AL of an nkd1;2 double mutant and the trans-activation of an O2 promoter by NKD1
and NKD2 individually in DLR assays performed with AL protoplasts (Gontarek et al., 2016).
Here, our data indicate that NKD2 is directly activated by O2 in the maize endosperm and the
NKD2 protein can co-activate target genes with O2, forming potentially another set of coherent
feed-forward loops within the O2 network (Figure 3.5; Supplemental Data Set 3.3). Furthermore,
our microscopic analysis of anticlinal cell wall, lipid body and PSVs in the AL showed that O2
plays an essential role in the regulation of cell differentiation and storage function of AL (Figure
3.6). NKD1 and NKD2 have been shown to directly activate the VP1 gene, which was detected
as an O2-activated indirect target in our analysis (log2FC = -2.10; Supplemental Data Set 3.1).

In sum, we hypothesize that O2, bZIP17, NKD1, NKD2, and VP1 co-regulate a complex gene
network associated with the endosperm cell differentiation, storage, maturation and seed
germination programs, and endosperm’s responses to abiotic stresses (Figure 3.7B). In support of
this model, our recent profiling of mRNAs in maize kernel compartments detected spatially
overlapping expression patterns of all these TF genes, with all of them showing moderate levels
of expression in the AL [fragments per kilobase of transcript per million mapped reads (FPKM)
≥ 2.12], while the O2, bZIP17, NKD1, and NKD2 genes were also expressed in CSE and CZ
(FPKM ≥ 0.80) (Zhan et al., 2015). However, the downregulation of O2 mRNAs was detected
132

only in AL, but not in SE, of the nkd1;2 double mutant (Gontarek et al., 2016), indicating distinct
roles of O2 and NKDs in AL versus SE. Therefore, a spatial (and temporal) analysis of gene
activities in endosperm cell types of single mutants and higher order mutant combinations for
O2, bZIP17, NKDs, and/or VP1 would be necessary to fully understand the GRNs controlled by
them.

3.5 METHODS

Plant Materials and Growth

Maize plants used for RNA-Seq, ChIP-Seq, RT-qPCR, immunoblot, and O2 gene/cDNA
sequencing experiments were grown in a greenhouse at University of Arizona with 30oC/25oC
(day/night) temperature cycles under natural light supplemented with high pressure sodium and
metal halide lamps for 16 h per day. The endosperm tissues for the RNA-Seq and ChIP-Seq
experiments were manually dissected from 15-DAP kernels obtained from self-pollinated B73
and B73o2 plants grown during June to October 2012. The tissues for the RT-qPCR assays (and
sequencing of the O2 cDNA in B73o2) were manually dissected from self-pollinated B73 and
B73o2 plants grown during November 2013 to February 2014. For these experiments, each
biological replicate of a given sample included multiple endosperms from a single ear (plant).
Biological replicates 1 and 2 of each ChIP-Seq sample was derived from the same ear as
biological replicates 1 and 2 of each RNA-Seq sample, respectively. Leaf tissues were collected
from B73o2 and W22o2 mutants at about one week after germination. The W22 and W22o2
plants used for the ultrastructural analysis of kernel phenotype were grown in the greenhouse at
University of Arizona (for 12- and 16-DAP kernels, which were collected during April to May
2017) or in the field at Iowa State University (for 20-DAP kernels, which were collected during
August 2017). The N. benthamina plants were grown, after the seeds were stratified at 4oC for 2
days, in a growth chamber at 22oC with 16-h light per day.

Immunoblot assays

Nearly 100 mg endosperm tissue was ground to fine powder in liquid nitrogen and homogenized
using 200 µl of protein extraction buffer [1% SDS, 100 mM Tris-Cl (pH 7.5), 100 mM NaCl, 5
mM EDTA] supplemented with 2% β-mercaptoethanol. The homogenate was centrifuged at
133

12,000× g at 4oC for 15 min. About 10 µl supernatant was mixed with 6× SDS-PAGE loading
buffer and heated at 95℃ for 5 min prior to loading on a SDS-PAGE gel. After separation, the
protein was blotted onto a nitrocellulose membrane (Sigma-Aldrich) using the Owl HEP-1 Semi
Dry Electroblotting System (Thermo Scientific). The nitrocellulose membrane was then probed
with the anti-O2 antibody using the AmershamTM ECLTM Select Western Blotting Detection
Reagent (GE Healthcare Life Sciences) by following the manufacturer’s instruction. The probed
membrane was then exposed to X-ray film (BioExpress) to detect the signals.

RNA-Seq

Nearly 4 µg of total RNA were isolated from each of the six endosperm samples (2 genotypes ×
3 replicates) using an SDS-phenol method as described previously (Li et al., 2014), and quality-
checked using an Agilent 2100 Bioanalyzer (Agilent Technologies). Using a TruSeq DNA
Sample Preparation Kit v2 (Illumina), poly(A)-containing mRNAs were purified, fragmented,
and reverse-transcribed by the University of Arizona Genetics Core facility to construct
multiplexed RNA-Seq libraries. The resulting libraries were sequenced on a single flow cell lane
of an Illumina HiSequation 2500 platform at the same facility using a TruSeq SBS kit v3
(Illumina) to produce 2× 100-nucleotide paired-end reads.

ChIP-Seq

Chromatin purification was performed by following a previously published protocol (Saleh et al.,
2008) with minor modifications. Dissected endosperms were cross-linked (fixed) using 1%
formaldehyde in phosphate buffer (10 mM Sodium Phosphate, 50 mM NaCl, 0.1 mM Sucrose,
pH 7.0) with vacuum infiltration for 30 min. Fixation was quenched with 125 mM Glycine, and
the fixed materials were washed three times with distilled water, dried with paper towels, frozen
in liquid nitrogen, and stored at -80°C. About 2 g of the fixed tissue was ground to fine powder
in liquid nitrogen, and homogenized in 25 ml of extraction buffer 1 [0.4 M Sucrose, 10 mM Tris-
Cl (pH 8.0), 10 mM MgCl2, 5 mM β-mercaptoethanal]. The homogenate was then filtered
through two layers of Miracloth (Calbiochem-Novabiochem Corporation). The filtered
homogenate was centrifuged at 12,000× g for 10 min at 4oC. After removing the supernatant, the
pellet was resuspended in 5 ml extraction buffer 2 [0.25 M Sucrose, 10 mM Tris-Cl (pH 8.0),10
mM MgCl2, 1% Triton X-100, 5 mM β-mercaptoethanal] and centrifuged as described above.
The pellet was then resuspended in 0.3 ml extraction buffer 3 [1.7 M Sucrose, 10 mM Tris-Cl
134

(pH 8.0), 2 mM MgCl2, 0.15% Triton X-100, 5 mM β-mercaptoethanal]. The suspension was
carefully layered onto 0.3 ml extraction buffer 3, and centrifuged at 16,000× g at 4oC for 1 h. The
resulting pellet (chromatin) was used sonicated as described below. Solutions for chromatin
purification contained fresh 1 mM PMSF and 1× Protease inhibitor Cocktail (Thermo Scientific
Company LLC).

ChIP was performed using Chromatin Immunoprecipitation (ChIP) Assay Kit (Millipore
Corporation, Billerica, MA) by following the manufacturer’s instruction with modifications. The
chromatin pellet from 1 g tissue was resuspended in 1 ml immunoprecipitation buffer [0.1%
SDS, 1% Triton X-100, 2 mM EDTA, 20 mM Tris-Cl (pH 8.0), 150 mM NaCl], with 1×
Protease inhibitor cocktail and 1 mM PMSF freshly added), and sonicated in 0.25-ml aliquots
with a probe sonicator (Fisher Scientific Model 120 Sonic Dismembrator) for six cycles of 20
sec at 25% amplitude and cooled in ice water for 40 sec between pulses. The sonicated
chromatin was centrifuged at 14,000 rpm for 10 min at 4°C and the supernatant was used for
immunoprecipitation. A 10 μl aliquot of the 1 ml of supernatant (1%) was saved as “input”
DNA. The chromatin solution was precleared using 60 μl of protein A agarose/Salmon Sperm
DNA (50% Slurry) (Upstate) for 1 h at 4°C on a rotating Labquake tube shaker. The mixture was
then centrifuged at 13,000 rpm at 4°C for 1 min. The supernatant was mixed with 5 μl anti-O2
antibody and incubated overnight at 4°C on a rotating tube shaker. 60 μl protein A
agarose/Salmon Sperm DNA (50% Slurry) was added to the mixture and the mixture was
incubated for 2 h at 4°C on a rotating tube shaker. The agarose beads were pelleted at 1000 rpm
for 1 min at 4°C and the supernatant was removed carefully. The beads were washed and eluted
by following the manufacturer’s instruction The elute and the “input” were reverse-cross-linked
by heating at 65°C for overnight and the DNA was purified by using Zymo ChIP DNA Clean &
Concentrator (Zymo Research). The ChIPed DNA was eluted with 10 μl Elution Buffer, and
quality-checked using an Agilent 2100 Bioanalyzer. Using an Illumina Pico DNA Seq Library
Preparation Kit (Gnomegen Inc.), nearly 10 ng of purified ChIPed DNA from each endosperm
sample (2 genotypes × 2 replicates) was used by the University of Arizona Genetics Core facility
to generate a set of multiplexed ChIP-Seq libraries by following the manufacturer’s instruction.
The resulting libraries were sequenced on a single flow cell lane of an Illumina HiSequation
2500 platform at the same facility to produce 2× 100-nucleotide paired-end reads.
135

Analysis of RNA-Seq and ChIP-Seq data

Raw reads from RNA-Seq and ChIP-Seq were trimmed to remove adapter sequences and low-
quality bases using the Trimmomatic program (Bolger et al., 2014) with the parameter
TRAILING set to three. The resulting RNA-Seq reads were mapped to the maize reference
genome (B73 RefGen_v3) using Tophat v2.0.9 (Trapnell et al., 2009) with parameters as
previously reported (Zhan et al., 2015). Reads mapped to each gene were counted using the
multicov function of BEDTools v2.17.0 (Quinlan and Hall, 2010) and normalized using the R
package edgeR v3.2.4 (Robinson et al., 2010) with the trimmed mean of M-values method. The
genes with raw counts per million (CPM) > 1 in at least 3 samples were tested for differential
expression analysis using edgeR with an exact test analogous to Fisher’s exact test. FDR < 0.05
was used as the cutoff to call DEGs. The trimmed ChIP-Seq reads were mapped to the reference
genome using Bowtie v2.1.0 (Langmead and Salzberg, 2012) with the following parameters: --
score-Min L, -0.6, -0.18 --end-to-end --very-sensitive --no-discordant. Mapped reads for
duplicates of the B73 samples and the B73o2 samples (control) were merged, respectively, and
were used for peak-calling using MACS v2.1.0.20140616 (Zhang et al., 2008; Feng et al., 2012)
with the following parameters: -g 2.06e9 -q 0.01 -mfold 10 30. ChIP-Seq read coverage across
the genome were visualized using the Integrative Genomics Viewer (Thorvaldsdottir et al.,
2013). The “slop” and “intersect” functions of BEDTools were used to determine gene-peak
association to identify O2-bound genes. Information about maize genome annotation and gene
functional description were obtained from Ensembl Plants (plants.ensembl.org, release 19) and
MaizeGDB (maizegdb.org) (Andorf et al., 2016). Annotation of TFs was obtained from Plant
Transcription Factor Database v3.0 (Jin et al., 2014) and GrassTFDB of GRASSIUS (Gray et al.,
2009; Yilmaz et al., 2009). Annotation of zein genes was curated based on the information
summarized by Chen et al. (2014a). The GO term annotations for maize genes were obtained
from Gramene (gramene.org, release 40) and the agriGO database v1.2 (Du et al., 2010). GO
term enrichment analysis was performed using Blast2GO v3.0.11 (Gotz et al., 2008). Metabolic
pathway association analysis was performed using CornCyc v8.0 on MaizeGDB (Walsh et al.,
2016). Putative functions for genes of interest but lack functional annotations were inspected by
BLASTX search of the NCBI nr and Swissprot databases (accessed September 2017) as
described previously (Zhan et al., 2015). A customized Perl script was used to detect the
presence of previously reported O2-binding sites within the promoter regions of O2-modulated
136

and/or bound genes, which we defined as 1 kb upstream and 0.5 kb downstream of the annotated
transcription start sites. MEME-ChIP v4.9.1 was used for cis-motif enrichment analysis, and the
JASPAR 2014 Plantae motif database (Mathelier et al., 2014) was used for motif comparison by
TOMTOM.

BLAT

The available sequences of microarray probes or cDNAs were aligned to the B73 cDNA
sequences (obtained from Ensembl Plants) using BLAT v34x10 (Kent, 2002) to identify the
corresponding genes in the B73 genome. The best alignment (based on matches) for each
probe/cDNA was selected using a customized Perl script. The alignments with identity < 80%
were manually excluded.

Expression of GST fusions in Escherichia coli and purification

The open reading frame (ORF) of O2 (GRMZM2G015534_T02) was amplified with primers
O2T2PGEX-FXba (5’-GCCTCTAGACATGGAGCACGTCATCTCAATG-3’) and
O2T2PGEX-RHind3 (5’-GCCAAGCTTCTAATACATGTCCATGTG-3’), digested with XbaI
and HindIII, and cloned into the XbaI and HindIII sites of pGEX-KG-Kan vector. This clone,
called O2-pGEX-KG-Kan, was introduced into E. coli strain ArcticExpress Blue (Agelent
Technologies). For protein expression, cells were grown in Superboth containing 50 µg/ml
kanamycin at 37°C to an OD600 of 0.6 and then cooled to 30°C. Protein expression was then
induced with 0.05 mM IPTG and the cells were grown at 30°C for 3 hours following induction.
For protein purification, induced cells were centrifuged, and the pellet was frozen and thawed,
resuspended in lysis buffer [10 mM Tris (pH 8.0), 10 mM EDTA, 10 mM NaCl] containing 500
µg/ml lysozyme, incubated at room temperature for 30 minutes, and then sonicated to reduce
viscosity. The sonicated cells were then centrifuged, and the supernatant was collected and then
passed through a 0.45 µm filter. The O2 protein was then purified using a 1 ml glutathione
sepharose column (GSTrap FF, GE Healthcare) following the manufacturer’s instruction.
Purification of GST protein was similar except that protein expression was carried out at 37°C
for 1 hour.

EMSA
137

All EMSA probes were labeled with Cy5 and were generated by PCR amplification from pCRII-
TOPO clones using Cy5-labeled M13 forward and reverse primers. For the bZIP17 clone, two
80-nt oligonucleotides were annealed as described below. For all other clones, the sequences of
interest were amplified by PCR using the primers listed in Supplemental Table 3.11. All inserts
were inserted into the pCRII-TOPO vector using the TOPO TA Cloning kit (Fisher Scientific).
Unlabeled competitors were prepared by annealing oligonucleotides of 25 to 50 nt in length. The
sequences of the oligonucleotides are listed in Supplemental Table 3.12. Annealing of
oligonucleotides was performed as follows: 1 nmol of each oligonucleotide was mixed in
binding buffer, heated to 95°C in a thermal cycler and cooled slowly to room temperature.
Binding buffer consisted of 10 mM Tris (pH 7.5), 50 mM KCl, 1 mM DTT, and 0.1% Tween 20.
All binding reactions were carried out at room temperature for 20 minutes in a total volume of 20
µl containing binding buffer; 2.5% glycerol; 25 nM probe; and 13 nM, 38 nM, or 115 nM O2
protein. GST protein alone at 100 nM was included as a control. Binding reactions containing
competitor DNAs contained 115 nM O2 protein and 500 nM or 5 µM competitor DNA.
Reaction products were separated by gel electrophoresis using 1% agarose in 0.5× TBE and
scanned on a GE Typhoon FLA9500.

RT-qPCR

RT-qPCR assays were performed on a time series of endosperm material collected at 6, 8, 10, 12,
15, 18, 22, and 30 DAP from both B73 and B73o2 with three biological replicates. RNA
isolation, cDNA synthesis, and qPCR experiments were performed essentially as described
previously (Wang et al., 2010; Li et al., 2014) using the primers listed in Supplemental Table
3.13. The raw threshold cycle (Ct) values were normalized against the Ct values of the
Thioredoxin (TXN) gene, which showed a similar and stable expression in B73 versus B73o2 at
all the analyzed stages (Supplemental Data Set 3.8). The normalized Ct values were manually cut
off at 36. The relative difference (i.e., -log2FC) in mRNA levels of a gene between B73 and
B73o2 at a given stage was determined as: ΔΔCt = CtB73o2 - CtB73, where both CtB73o2 and CtB73
are normalized Ct values.

Amplification and sequencing of O2 genomic DNA and cDNA

Maize genomic DNA (gDNA) was isolated from leaf tissues using a urea-based method as
described previously (Holding et al., 2008). Each biological replicate of a given leaf sample
138

included about 20 mg leaf tissue from a single plant. Using the same pair of primers, 5’-
TTAACTCATGGGTGCATAGAGA-3’ and 5’-GAAACCCGCAGTGCCTAATA-3’, the O2
gene sequence was amplified from the leaf gDNA of B73o2 and W22o2 (3 and 2 biological
replicates, respectively), while the O2 cDNA was amplified from a 22-DAP endosperm cDNA
sample from B73o2 (2 biological replicates) by PCR using the Phusion Hot Start II DNA
Polymerase (Fisher Scientific). The PCR products from gDNA amplification were cloned into
the pGEM-T vector (Promega) and were Sanger-sequenced using the vector-specific M13F(-21)
and M13R Reverse primers and the gene-specific primers 5’-ATCTTTATTATTCCCTCGCT-3’
and 5’-GCTCTCAGCACCCTGTTGTC-3’; the PCR products from cDNA amplification were
Sanger-sequenced directly using the same primers as used for PCR and primer 5’-
TCAAGAGCCTGCATCCAT-3’.

Phylogenetic analysis

Protein sequences of bZIP family members were obtained from Ensembl Plants, and the longest
annotated protein sequence encoded by each gene was selected for phylogenetic analysis using a
customized Perl script. Multiple sequence alignment was performed using Clustal Omega
(v1.2.4; https://www.ebi.ac.uk/Tools/msa/clustalo/)(Sievers et al., 2011). A phylogenetic tree
based on the alignment was inferred using RAxML v8.0.23 (Stamatakis, 2014) with the
maximum likelihood method using a Whelan and Goldman (WAG) model of amino acid
evolution (Whelan and Goldman, 2001) and a Gamma model of rate heterogeneity. The
confidence level of nodes in the tree was determined by inferring phylogeny in 100 bootstrap
replicates. The tree was visualized using FigTree v1.4.3 (tree.bio.ed.ac.uk/software/figtree/).

Domain mapping auto-activator assay and Y2H screen

The domain mapping auto-activator assay and Y2H screen were performed by Hybrigenics
Services SAS (Paris, France). The full-length or truncated O2 ORF was PCR-amplified from the
cDNA of 15-DAP B73 endosperm using primers containing a 50-nt region homologous to the
LexA DNA-binding domain (BD) vector pB27 and a 20-nt region homologous to the
corresponding bait fragment. Each PCR product and the linearized pB27 vector were co-
transformed with the empty activation domain (AD) vector pP7 into yeast, allowing homologous
recombination between the PCR product and pB27 to occur. Positive transformants were
selected on synthetic dropout (SD) medium lacking Trp and Leu (SD/-Trp/-Leu). Auto-activation
139

activity of the full-length or truncated O2 was assessed by detecting the growth rate of the
transformed yeast on SD/-Trp/-Leu/-His plates in presence of a concentration series of 3-
aminotriazole (3-AT; 0 mM, 1 mM, 5 mM, 10 mM, and 50 mM).

Using total RNA isolated from 15-DAP B73 endosperm, a random-primed prey library for the
Y2H screen was constructed by cloning the cDNA fragments into the pP6 vector as a C-terminal
fusion to the GAL4 AD. The CDS of O2F4 was PCR amplified and cloned into pB27 as a C-
terminal fusion to the LexA BD (N-LexA-O2F4-C) and used as the bait for the screen. The pP6
and pB27 vectors were derived from the original pBTM116 (Vojtek and Hollenberg, 1995) and
pGADGH (Bartel, 1993) vectors, respectively. The screen was performed using a mating method
with YHGX13 (Y187 ade2-101::loxP-kanMX-loxP, matα) and L40ΔGal4 (mata) yeast strains on
SD/-Trp/-Leu/-His plates, as previously described (FromontRacine et al., 1997). Each prey
fragment from the positive clones was PCR-amplified and sequenced at the 5’ and 3’ junctions.
A confidence score (predicted biological score) was assigned to each positive interaction,
ranging from “A” to “E” corresponding to decreasing confidence of interaction, as previously
described (Formstecher et al., 2005).

Directed Y2H assay

The ORF of each protein was PCR-amplified from a 15-DAP B73 endosperm cDNA (O2 and
O2F4) or a clone from the GRASSIUS maize TFome collection (bZIP17; clone pUT6435)
(Burdo et al., 2014) and cloned into pGBKT7 and/or pGADT7 vectors, allowing the protein to be
expressed as fusion protein with the GAL4 BD domain or the GAL4 AD domain. The primers
and restriction enzymes used to generate the clones are listed in Supplemental Table 3.14. All
clones were Sanger-sequenced to confirm the inserts. Each pair of BD and AD constructs were
co-transformed into the AH109 yeast strain. Positive clones were selected on a SD/-Trp/-Leu
plate, and spotted in serial dilutions of 104, 103, 102, and 101 cells on a SD/-Trp/-Leu plate as
well as a SD/-Trp/-Leu/-His/-Ade plate, and cultured for 4 d at 30oC.

In vitro pull-down assay

Tagged full-length proteins were synthesized in vitro with the cDNA clones in pGADT7 and
pGBKT7 vectors as templates using the TnT® Quick Coupled Transcription/Translation
Systems (Promega) by following the manufacturer’s instruction. The 35S-labeled proteins were
140

synthesized in presence of [35S]methionine (PerkinElmer). The in vitro pull-down assays were


performed using a previously reported method (Cifuentes-Rojas et al., 2011) with minor
modifications. Briefly, for a given pair of assayed proteins, the TnT-produced proteins (10 µl
each) were mixed, incubated at 30⁰C for 20 min, and then at 4°C on a rotating VWR tube rotator
for 20 min. Subsequently, 300 µl blocking buffer (buffer W-100 [20 mM TrisOAc (pH 7.5), 10%
glycerol, 1 mM EDTA, 5 mM MgCl2, 0.2 M NaCl, 1% NP-40, 0.5 mM sodium deoxycholate,
and 100 mM potassium glutamate] containing 0.5 mg/ml BSA, 0.5 mg/ml lysozyme, 0.05 mg/ml
glycogen, and 1 mM DTT) was added and the mixture was incubated at 4°C for 1 hour while
rotating. In the meantime, anti-Myc agarose (Sigma Aldrich) was blocked 3 times (15 min each)
in 1 ml blocking buffer at 4⁰C while rotating. The blocked protein mixture was mixed with the
anti-Myc agarose and incubated at 4°C for 2 hours while rotating. After centrifugation at 2,500×
g for 1 min at 4oC, the supernatant was saved as “input” protein. The precipitated agarose beads
were then washed 5 times with 1 ml W-300 buffer (W-100 containing 300 mM potassium
glutamate) and proteins were eluted by adding 25 µl 5× SDS-PAGE sample buffer. After elution,
20 µl pulled-down proteins, along with approximately 6% of the “input” proteins were resolved
by SDS-PAGE electrophoresis with 10% acrylamide and visualized using a storage phosphor
screen GP (Kodak) and a storm 860 molecular imager (Molecular dynamics).

BiFC assay

The wild-type O2 ORF was PCR-amplified from a 15-DAP B73 endosperm cDNA; the mutated
O2 ORF (named O2LA), which encodes an O2 protein with two leucine residues in the leucine-
zipper domain substituted by alanine residues (L265A and L272A), was generated using a PCR-
based site-directed mutagenesis method. The attB adapter PCR method (Gateway Technology,
Invitrogen) was used to generate full attB PCR products. The attB PCR products were cloned
into the pDONR207 vector by BP reaction to generate entry clones for O2 and O2LA, while the
pUT6435 clone was directly used as the entry clone for bZIP17. The expression clones were
generated by LR reaction using pDEST-gwVYCE and pDEST-gwVYNE (Gehl et al., 2009) as
the destination vectors, allowing the O2/O2LA/bZIP17 protein to be expressed as fusion protein
with the C terminus (VenusC) or N terminus (VenusN) of the Venus YFP. The primers used to
generate clones for O2 and O2LA are listed in Supplemental Table 3.15. All clones were Sanger-
sequenced to confirm the inserts. For each assayed pair of fusion proteins, leaves of 5- to 6-
141

week-old N. benthamiana plants were co-infiltrated with two individually transformed


Agrobacterium tumefaciens strain GV3101 (pMP90)—each harboring one of the two fusion-
protein-coding constructs—and the p19 helper strain (culture OD600 all adjusted to 0.5). Leaf
tissue was collected about 48 h after infiltration and fluorescence was examined using an
Axiophot compound epifluorescence microscope (Zeiss, Jena, Germany) as described previously
(Wang et al., 2006).

DLR assay

DLR assays for O2-peak binding was performed as previously described (Taylor-Teeples et al.,
2015) with minor modifications. The O2 ORF was PCR-amplified from the 15-DAP B73
endosperm cDNA and was cloned into the pDONR221-P1-P4 vector by BP reaction to generate
the pLAH-O2 construct; each O2 peak region was PCR-amplified from B73 leaf gDNA and was
cloned into the pDONR221-P3-P2 vector by BP reaction to generate a pLAH-peak construct. A
multi-site LR cloning was performed using these two entry clones, a third entry clone pLAH-R4-
R3-VP64Ter, and the destination vector pLAH-LARm, to generate the expression construct
pLAH-LARm-O2-VP64-peak. For each peak, an expression construct containing the ORF of the
Citrine YFP instead of O2, named pLAH-LARm-Citrine-VP64-peak, was also generated as a
negative control. All clones were Sanger-sequenced to confirm the inserts. Leaves of 4- to 5-
week-old N. benthamiana plants were co-infiltrated with A. tumefaciens strain GV3101 (pMP90)
harboring an expression construct (culture OD600 adjusted to 0.12) and the p19 helper strain
(culture OD600 adjusted to 0.075). Leaf tissue was collected about 42 h after infiltration and the
activities (luminescence) of the firefly luciferase (LUC) and the Renilla luciferase (REN) were
measured with a GloMax 20/20 Luminometer (Promega) using the Dual-Luciferase Reporter
Assay System (Promega) by following the manufacturer’s instruction. Briefly, about 2.5 mg leaf
tissue from a single plant was frozen in liquid nitrogen, ground using a bead-beater, and lysed
using 200 ul Passive Lysis Buffer. 50 ul Luciferase Assay Reagent II was added to a 10-ul
aliquot of the lysate, and firefly luciferase activity was measured using a GloMax 20/20
Luminometer with 10 s integration time. Firefly luciferase activity was subsequently quenched
by adding 50 ul of Stop & Glo Reagent, which contains Renilla luciferin substrate, and Renilla
luciferase activity was measured with a 10 s integration time. For each peak, a ratio between
LUC and REN activities (LUC/REN) was determined from the activity of the pLAH-LARm-O2-
142

VP64-peak construct. The ratio for each biological replicate was then divided by the ratio
determined using the pLAH-LARm-Citrine-VP64-peak construct (mean of biological replicates)
to calculate a fold induction; the difference between the LUC/REN ratios obtained using the two
expression constructs were statistically tested using Student’s t test. For the DLR assays of
target-gene co-activation by two TFs, the ORF of each TF was cloned into the pBN-d35Stev
vector (linearized with BamHI/KpnI) using In-Fusion cloning to generate a pBN-d35Stev-TF
construct to enable the expression of the TF as an intact protein. For each tested peak, the
corresponding Citrine-containing expression construct used in the DLR assay for O2 binding was
also used as the negative control, and was co-infiltrated with one or two pBN-d35Stev-TF
constructs in the experimental assays. The LUC/REN ratios determined from each biological
replicate of the experimental assays were divided by the mean ratio determined from the
corresponding negative control assay to calculate a mean fold induction. Difference between the
LUC/REN ratios among all the control and experimental assays were statistically tested using
one-way ANOVA with post-hoc HSD test. For the peaks that were tested in both assays shown
in Figure 3.4 (Supplemental Table 3.8) and Figure 3.5 (Supplemental Table 3.10), all assays for a
given peak were performed at the same time, and thus the fold-induction data (mean ± SEM) for
“YFP” and “YFP+O2”, respectively, are the same in the two figures (tables). All DLR assays
were performed with 5 biological replicates of N. benthamiana leaf sample. The primers and
cloning methods used to generate the clones for DLR assays are listed in Supplemental Table
3.16.

Transmission electron microscopy

Kernels were fixed in 4% formaldehyde and 2% glutaraldehyde in 50mM sodium phosphate


buffer. Tissue pieces were washed 3 times in cold 50 mM sodium phosphate buffer (pH 7.0) and
then post-fixed in 2% osmium tetroxide in 50 mM sodium phosphate buffer overnight at 4°C.
After fixation, pieces were washed 3 times in cold 50 mM sodium phosphate buffer, then
dehydrated in a graded ethanol series of 30%, 50%, 70%, 80%, 95%, and then 3 times in 100%
for 15 minutes each. Pieces were then transferred to a 2:1, then 1:1, then 1:2 ratio of 100%
ethanol to Spurr’s resin (Electron Microscopy Sciences) for 4-16 h at room temperature with
rotation. Pieces were transferred to fresh 100% resin 3 times for a minimum of 7 h, maximum of
21 h before embedding samples in 100% Spurr’s resin and hardening at 60°C for at least 25 h.
143

Blocks were trimmed and thin sections cut with a diamond knife on a Powertome
Ultramicrotome X (Boekeler Instruments). Sections were collected on 150-mesh copper grids
and stained with 2% uranyl acetate for 20 minutes followed by lead citrate for 3 minutes.
Observations and images were made with a Hitachi HT7700 Scanning/Transmission Electron
Microscope (Hitachi Instruments, Tokyo, Japan).

Measurement of cell features

Measurements of AL cells were made from TEM images using Image Pro Plus (Media
Cybernetics, Inc.). When needed, a one-time adjustment of image contrast was made to an entire
image using the “best-fit” function. Micrographs for each genotype (W22 n = 20; W22o2 n = 12)
were obtained from two plants (1 or 2 kernels per plant). The cytoplasmic area occupied by
protein sorting vacuoles (PSVs) and by lipid bodies was measured. For each cell, all PSVs and
lipid bodies were traced using the polygon tool within a 600 × 600 pixel square (about 80 µm2)
anchored randomly in 1 of 5 sectors of the cytoplasm of the cell (excluding areas with cell wall
and nuclei). PSVs were identified based on their low electron density, polymorphic appearance,
bounding membrane, and typical electron dense inclusions. Lipid bodies were identified based
on their high electron density, circular shape and lack of delineating unit membrane. Mean area
of PSVs (W22 n = 235; W22o2 n = 173) and lipid bodies (W22 n = 862; W22o2 n = 573) was
determined within the samples. Cell wall thickness (n = 17) was measured at a random location
along a single anticlinal wall per cell.
144

Figure 3.1. Identification of direct and indirect O2 target genes using RNA-Seq and ChIP-
Seq.

(A) Scatter plot of log2FC versus log2RPKM (average in all 6 RNA-Seq samples) of the 28,580
genes tested for differential expression. DEGs are represented by red dots, and the other genes
blue. Dashed lines indicate 2-fold changes (log2FC = ±1). (B) Distribution of O2 peaks in the
145

maize genome based on localization of peak summits. (C) Venn diagram of numbers of O2-
modulated genes and O2-bound genes detected by RNA-Seq and ChIP-Seq. (D) Expression
levels of the O2-modulated and/or bound gene sets in WT versus o2 mutant. Colored boxes
represent the interquartile range, horizontal lines within the boxes the median, whiskers 1.5 times
the interquartile range, and black dots the outliers. (E) Distribution of O2 peaks around the gene
models of direct O2 targets. The 5 kb genomic regions up- or downstream gene models were
divided into 50 bins, while the representative gene model showing the average length of maize
genes (about 2.2 kb based on the B73 gene model annotation) were divided into 22 bins. (F) Two
bZIP-like motifs and their positional distributions around the summits of peaks associated with
direct O2 targets. The P values were calculated using the CentriMo program.
146

Figure 3.2. GO term enrichments of the O2-modulated and/or bound gene sets.

Treemap representations of GO term enrichments for direct target (A), O2-activated indirect
target (B) and O2-repressed indirect target (C) gene sets. GO terms are grouped based on the
main categories, including molecular function, biological process, and cellular component. Size
and color of the squares/rectangles indicates significance of enrichment (-log10FDR).
147

Figure 3.3. Temporal expression patterns of O2-modulated and/or bound genes during
endosperm development.

(A) Relationships of the O2-modulated and/or bound gene sets with the previously published
temporally upregulated gene sets (Li et al., 2014). The heat map indicates P values (-log10) of
hypergeometric tests of overrepresentation of genes in a given pair of gene sets. Boxes contain
the numbers of overlapping genes and the associated P values (in parentheses). The O2-
modulated and/or bound gene sets are noted on the x axis and the temporal gene sets are noted on
the y axis. Numbers of genes in each temporal gene set: up@2DAP, 54; up@3DAP, 68;
up@4DAP, 92; up@6DAP, 523; up@8DAP, 1,402; up@10DAP, 552; up@12DAP, 241. (B)
through (M) The mRNA levels of 12 O2-activated direct targets in the developing endosperm of
WT versus o2 quantified using RT-qPCR. As a canonical non-zein target of O2, the b-32 gene,
although not detected as an O2-bound gene in our ChIP-Seq analysis, was also included. Data
are normalized Ct values (mean ± SEM; n = 3). Asterisks indicate significant difference between
WT and o2 (|ΔΔCt| > 1; P < 0.05, Student’s t test). Blue dashed lines indicate undetectable
mRNA level (Ct = 36).
148

Figure 3.4. Full-length O2 and bZIP17 proteins interact and co-activate target gene
expression.

(A) O2 interacts with bZIP17 in planta. Fluorescent and corresponding bright field images from
the BiFC assay of interaction between O2-VenusN and bZIP17-VenusC. O2-VenusN/O2-VenusC
was used as a positive control. O2LA-VenusN/O2-VenusC, O2LA-VenusN/bZIP17-VenusC, and
O2LA-VenusN/O2LA-VenusC were used as negative controls. Bars = 40 μm. Each assay was
repeated at least 4 times. (B) Co-activation of reporter gene (LUC) by O2 and bZIP17 through
binding to O2 peaks. The gene symbol of the O2 target gene associated with the peak is indicated
in parenthesis. Data are fold inductions (mean ± SEM; n = 5). Different letter indicates P < 0.05,
one-way ANOVA with post-hoc Tukey’s HSD test.
149

Figure 3.5. O2 and NKD2 co-regulate a downstream gene network.

(A) Venn diagram of numbers of putative direct targets of O2 and NKD2. (B) Co-activation of
LUC gene by O2 and NKD2 through binding to O2 peaks. The gene symbol of the O2 target
gene associated with each peak is indicated in parenthesis. Data are fold inductions (mean ±
SEM; n = 5). Different letter indicates P < 0.05, one-way ANOVA with post-hoc Tukey’s HSD
test. (C) Venn diagram of numbers of putative indirect O2 targets and direct NKD2 targets.
150

Figure 3.6. Transmission electron microscopy of 20-DAP AL cells in W22o2 and W22.

(A) and (B). Transmission electron micrographs of AL cells of WT (A) and o2 (B). Bar = 1 µm.
(C), (E), and (G). Distribution of anticlinal cell wall thickness, lipid body area, and PSV area.
Boxes represent the interquartile range, yellow lines the median, blue dots the mean, and
whiskers 1.5 times the interquartile range. (D) Percentages of cytoplasm that is occupied by lipid
bodies and PSVs. (F) and (H). Size class distribution of lipid bodies and PSVs. P values in (C),
(D), (E), and (G) were calculated using Student’s t test.
151

Figure 3.7. Models for O2-mediated regulation of endosperm gene expression and
development.

(A) Two distinct modes of O2-mediated gene activation. (B) Co-regulation of maize seed
development and germination by O2, bZIP17, and NKD2. With direct mutual regulation and/or
protein-protein interactions with bZIP17 or NKD2, O2 directly regulates genes involved in AL
versus SE differentiation, storage-protein synthesis, and abiotic stress responses of the
endosperm; O2 indirectly regulates genes involved synthesis/metabolism of storage proteins,
carbohydrates, and lipids; and O2 indirectly regulates VP1, which in turn regulates seed
maturation and germination.
152

3.6 SUPPLEMENTAL DATA

Supplemental Text 1: Genetic confirmation of the B73o2 mutant

To confirm the o2 mutation in the B73o2 mutant, we examined kernel phenotype of B73o2 in
comparison to a W22o2, a W64Ao2, and the reciprocal F1 seeds obtained from crosses among
B73o2, W22o2, and W64Ao2. Similar to the inbred mutants, 100% of the mature kernels
produced by all reciprocal crosses showed an opaque phenotype (Supplemental Figure 3.1A and
Supplemental Table 3.1). The failure of complementation among all three o2 alleles confirmed
the presence of an o2 mutation in the B73o2 mutant line used in our analyses.

To confirm the nature of the o2 mutant sequences in the B73o2 inbred used in our studies, we
cloned and sequenced the O2 gDNA in B73o2 in comparison to W22o2. A 2,711-bp sequence
spanning 154 bp upstream the start codon to 11 bp downstream the stop codon (based on gene
model annotation in B73) were obtained for B73o2; and a 2,714-bp sequence spanning from 154
bp upstream the start codon to 14 bp downstream the stop codon, were obtained for W22o2.
These two sequences were 100% identical within the 2,711 bp covered by the B73o2 sequence,
indicating that the o2 alleles in the B73o2 and W22o2 mutants are of the same origin. Alignment
of the 2,711-bp sequence with O2 genomic sequence in B73 using SnapGene v4.0.5 (from GSL
Biotech; available at snapgene.com) detected 39 mismatches and 23 gaps, whereas alignment of
the same sequence with the O2 genomic sequence in W22 detected 32 mismatches and 21 gaps.
In both alignments, many mismatches and gaps were found in both exonic and intronic
sequences, including many missense mutations and insertions/deletions in the CDS regions
(Supplemental Figure 3.1B). An online BLASTN search against the NCBI nr nucleotide database
(accessed July 2017) using the 2,711-bp sequence showed that the most similar WT gene is the
O2 gene in the maize accession Ac 1503 GM 1407 (X15544.1; with 34 mismatches and 23 gaps,
based on subsequent SnapGene alignment), whereas the most similar mutated gene is the O2
gene in the quality protein maize (QPM) inbred line CA339 (KF831426.1; with 3 mismatches)
(Maddaloni et al., 1989; Chen et al., 2014b).

Because a modest level of O2 mRNAs were detected in the B73o2 mutant endosperm at stages
beyond 10 DAP (Figure 3.3B), and to confirm that the O2 mRNAs detected by RT-qPCR in the
153

mutant were derived from the mutated O2 gene, we sequenced the O2 cDNA in 22-DAP mutant
endosperm, and obtained a 2,675-bp sequence spanning 105 bp upstream the start codon to 14 bp
downstream the stop codon. The putative coding sequence within this 2,675 bp was 100%
identical to the O2 gDNA in B73o2, with five gaps corresponding to the five introns in the
annotated gene model of O2 in the B73 genome (Supplemental Figure 3.1B). Immunoblot assays
using the anti-O2 antibody detected accumulation of the O2 protein at a relatively high level by
15 DAP in the WT but no accumulation in the B73o2 mutant (Supplemental Figures 3.1C and
3.1D).

Supplemental Text 2: Quality assessment of RNA-Seq and ChIP-Seq data

We evaluated the quality of our RNA-Seq data in two ways. First, we used pairwise Spearman
correlation coefficient (SCC) analysis of the normalized reads (in reads per kilobase per million
mapped reads, RPKM) of all the genes tested for differential expression. This analysis revealed
SCCs ranging from 0.92 to 0.95 for the WT samples and 0.96 to 0.97 for the B73o2 samples
(Supplemental Figure 3.2A) that indicated high reproducibility of the sampling. Second, we
compared our data to the previously published reports of the O2-modulated genes. An
examination of expression levels of several canonical direct O2 targets showed that almost all the
known targets, as exemplified by O2 itself, a 22-kD α-zein gene (azs22-4) and the LKR/SDH
gene, were dramatically downregulated in our mutant data (Supplemental Figure 3.2B). We also
compared our DEGs with the DEGs identified by prior transcriptomic studies of o2 mutants by
mapping the published transcript/probe sequences [available for four reports (Hunter et al., 2002;
Frizzi et al., 2010; Jia et al., 2013; Zhang et al., 2016)] to the B73 cDNAs using the BLAT
program (Kent, 2002). The result showed that the DEGs that we identified included 494 genes
that have been detected by at least one prior study (Supplemental Figure 3.3A). Surprisingly,
among these 494 genes, only 6 have been commonly identified by all four previous studies
(Supplemental Figure 3.3A). An analysis of the expression level of DEGs detected by previous
studies but not by us showed that nearly all of them exhibited low fold changes (|log2FC| < 1) in
our dataset (Supplemental Figure 3.3B).

To evaluate the accuracy of our identification of direct O2 targets, we examined the differential
expression and peak association of several canonical O2 targets including O2 itself, cyPPDK1,
154

azs22-4, LKR/SDH, b-32, and the 15-kD β-zein gene (Lohmer et al., 1991; Schmidt et al., 1992;
Cord Neto et al., 1995; Gallusci et al., 1996; Maddaloni et al., 1996; Kemper et al., 1999). The
results showed that nearly all of these genes were identified as direct O2 targets by our data
(Supplemental Figure 3.4 and Supplemental Data Set 3.3). The only exception was the b-32
gene, which exhibited a dramatic downregulated pattern in the B73o2 mutant (log2FC = -9.65;
Supplemental Data Set 3.1) but was not associated with an O2 peak.

Supplemental Text 3: Identification of O2-interacting proteins through Y2H screen

To identify O2 cofactors, an Y2H screen of a 15-DAP endosperm cDNA library was carried out.
Similar to the observations in a previous study (Zhang et al., 2012), a pilot screen using the full-
length O2 protein as bait resulted in strong auto-activation of the reporter gene. Therefore, an
independent domain mapping auto-activator assay was performed and identified a fragment of
the O2 protein (O2F4) that did not cause auto-activation (Supplemental Figure 3.14). In a
subsequent screen using O2F4 as bait, 53.3 million prey clones were analyzed and 111 positive
clones were detected. To identify the corresponding genes/proteins, we aligned sequences of the
111 positive clones to the B73 cDNAs using BLAT and detected 26 genes encoding putative O2-
interacting proteins, including bZIP17 (with a “D” score for the corresponding clone).
155

Supplemental Figure 3.1. Genetic confirmation of the B72o2 mutant.

(A) Opacity of mature kernels from reciprocal crosses between B73o2, W22o2, and W64Ao2
compared to those from inbred WT and mutants. Top, intact kernels on a light box; bottom,
156

latitudinal sections of kernels. Brightness of the top image was uniformly adjusted using Adobe
Photoshop CC 2015. Scale bars = 1 cm. (B) Sequence alignment between the gDNA and cDNA
of O2 in B73o2, and the O2 gDNA in W22o2, W22, and B73. The alignment was performed
using SnapGene 4.0.5 with the O2 gDNA in B73o2 as the reference sequence. The O2 gDNA
sequences of B73 and W22 were based on gene model annotation of the reference genomes B73
RefGen_v3 and Zm-W22-REFERENCE-NRGENE-2.0, respectively. The O2 ORF in the B73o2
mutant was predicted according to the annotation of O2 gene model in B73 RefGen_v3. Red
triangles/rectangles, insertion; white boxes, deletion or substitution. (C) Detection of O2 protein
in a developmental time series of B73 endosperm using immunoblot assay. (D) Immunoblot
detection of O2 protein in B73 and B73o2 at 15 DAP.
157

Supplemental Figure 3.2. Quality assessment of the RNA-Seq data.

(A) Pairwise scatter plots and SCCs between biological replicates based on log2RPKM values of
all the genes tested for differential expression. (B) Expression levels of the tested genes in each
RNA-Seq sample. Boxes represent the interquartile range, horizontal lines within the boxes the
median, whiskers 1.5 times the interquartile range, and black dots the outliers. The expression
levels of three known direct targets of O2, including the O2 gene, the LKR/SDH gene, and a 22-
kD α-zein gene (azs22-4), as well as the housekeeping Thioredoxin (TXN) gene, are indicated.
Biological replicates are indicated by -1, -2, and -3.
158

Supplemental Figure 3.3. Comparison of O2-modulated genes identified by us and prior


studies.

(A) Venn diagrams of the numbers of DEGs in o2 mutant versus WT identified by us, Zhang et
al. (2016), Jia et al. (2013), Frizzi et al. (2010), and Hunter et al. (2002). Total number of DEGs
identified by each study is indicated in parenthesis. (B) Scatter plot of log2FC versus log2RPKM
of DEGs identified by us and/or at least one prior studies (Hunter et al., 2002; Frizzi et al., 2010;
Jia et al., 2013; Zhang et al., 2016). Black dashed lines indicate 2-fold changes.
159

Supplemental Figure 3.4. Visualization of ChIP-Seq reads and peaks around 5 previously reported direct targets of O2 using
the Integrative Genomics Viewer.
160

Supplemental Figure 3.5. Association of the O2-modulated and/or bound gene sets with
previously reported O2-binding sites.

P values were calculated by enrichment analysis of O2-binding site-associated genes in each


gene set using hypergeometric test. Fraction of O2 binding site-associated genes in each gene
set: unbound repressed, 126/800; bound repressed, 10/39; bound unmodulated, 567/3096; bound
activated, 53/147; unbound activated, 141/877.
161

Supplemental Figure 3.6. Expression levels of the direct O2 targets in seed tissues in
comparison to vegetative and reproductive tissues based on the available maize expression
atlas (Chen et al., 2014a).

The fractions of O2 targets that have available data are indicated in parentheses. Colored boxes
represent the interquartile range, horizontal lines within the boxes the median, whiskers 1.5 times
the interquartile range, and black dots the outliers. The red line indicates the expression of O2
itself. Asterisks indicate highly significant differences between the activated and repressed
targets (P < 0.001, Student’s t test). The selected data include vegetative (Veg) tissues shoots
(Sh), roots (R), Leaf (L), shoot apical meristem (SAM, replicate 1); reproductive (Rep) tissues
ear (E), tassel (T, replicate 1), pre-emergence cob (C), silk (Si), Anther (A), ovule (O), pollen
(P); and whole kernels, endosperm, and embryos of different developmental stages (in DAP).
162

Supplemental Figure 3.7. Distribution of mapped reads around the gene models of direct
O2 targets.

The 5 kb genomic regions up- or downstream the annotated gene models were divided into 50
bins, while the representative gene model showing the average length of maize genes were
divided into 22 bins. Read numbers were scaled by dividing the read number in each bin by the
maximal number.
163

Supplemental Figure 3.8. Results of EMSAs for binding of O2 to the O2 peaks or promoter
regions associated with direct O2 targets.

Numbers indicate the amount of WT or mutant (mut) competitor (comp) probes. The WT probes
each contain one or more sequences similar to the known O2-binding sites, whereas the mutant
probes contain substitutions in those sequences.
164

Supplemental Figure 3.9. DLR assays for binding of O2 to the O2 peaks associated with
direct O2 targets.

(A) Schematic of constructs used for the DLR assays. For each tested peak, an experimental
construct containing the coding sequence of O2 fused with a strong transcriptional activation
domain VP64, and a control construct containing the coding sequence of Citrine (an YFP
variant) instead of O2 was used. Abbreviations: 35S Pro, Cauliflower mosaic virus (CaMV) 35S
promoter; 35S Ter, CaMV 35S terminator; VP64, transcriptional activation domain; 35Smini,
CaMV 35S minimal promoter; LUC, firefly luciferase; REN, Renilla luciferase. (B) Results of
DLR assays shown as fold inductions (mean ± SEM; n = 5). bZIP12, GRMZM2G034623, and
GRMZM2G443668 were identified as O2-repressed direct targets, while all the other tested
genes were identified as O2-activated direct targets. Asterisks indicate P < 0.05, Student’s t test.
165

Supplemental Figure 3.10. Gene numbers and log2FC (mean ± SEM) of zein gene
families/sub-families modulated and/or bound by O2.
166

Supplemental Figure 3.11. Expression patterns of the TF genes directly regulated by O2


based on the available maize expression atlas (Chen et al., 2014a).

Either a gene symbol (if available) or a GRASSIUS TF name is used to identify each gene. The
heat map was hierarchically clustered based on Euclidean distance. The selected data included
the same tissues as Supplemental Figure 3.6.
167

Supplemental Figure 3.12. Enrichments of TF gene families among the O2-modulated


and/or bound gene sets.

The heat map indicates P values (-log10) of hypergeometric tests of overrepresentation of genes
in a given pair of gene sets/families. Boxes contain the numbers of overlapping genes and the
associated P values (in parentheses). The heat map was hierarchically clustered based on
Euclidean distance. The number of genes in each gene set/family is shown in parenthesis.
168

Supplemental Figure 3.13. Phylogenetic relationship among all annotated bZIP-family


proteins in maize.

A gene ID and, if available, either a gene symbol or a GRASSIUS TF name, is used to identify
each gene. Proteins encoded by putative direct target genes of O2 are indicated by red
arrowheads. The phylogeny was inferred from protein sequences using a maximum likelihood
method. Values next to branches represent the percent of 100 bootstrap replicates that support
the topology. Bar, amino acid substitution per site.
169

Supplemental Figure 3.14. Domain mapping auto-activator assay of O2.

(A) The full-length O2 protein and its truncations (fragments) were tested for auto-activation
activity. The full-length protein (spanning amino acid residues 1 to 441) and the fragments are
170

shown as blue or grey bars, respectively, with numbers indicating amino acid positions within
the full-length protein. The position of the bZIP domain, predicted using InterPro 65.0
(www.ebi.ac.uk/interpro/), within the O2 protein sequence is indicated by a green bar. (B)
Results of the domain mapping auto-activator assay shown as yeast growth. The tested O2
fragments in the BD vector are indicated. The AD vector tested in each combination was empty
pP6. Eight transformants for each combination were spotted on a SD/-Trp/-Leu plate, and SD/-
Trp/-Leu/-His plates with a concentration series of 3-AT. The original bait construct used for the
pilot screen (Supplemental Text 3), with the full-length O2 ORF in pB27 vector, was used as
positive control; an empty pP6 vector, in combination with a circularized or linearized empty
pB27 (one transformant each), were tested as negative controls. Images in (B) were obtained
from Hybrigenics Services SAS.
171

Supplemental Figure 3.15. Results of directed Y2H assay of protein-protein interaction


between O2 and bZIP17.

Yeast AH109 cells co-transformed with the indicated plasmids were spotted in 10-fold serial
dilutions on a SD/-Trp/-Leu plate and a SD/-Trp/-Leu/-His/-Ade plate and incubated at 30oC for
4 days. BD-O2F4/AD-O2 was used as positive control, and BD-O2F4/Empty AD and Empty
BD/AD-O2 were used as negative controls.
172

Supplemental Figure 3.16. In vitro interaction between O2 and bZIP17.

(A) Pull-down was performed using an anti-Myc antibody. O2-Myc/O2-HA* was used as a
positive control, and O2-HA* alone was used as a negative control. (B) Pull-down was
performed using the indicated antibodies. O2-Myc* alone or bZIP17-HA* alone was used as a
negative control. Asterisks indicate [35S]methionine-labeled proteins. Brightness and contrast of
the image in (B) was uniformly adjusted using Adobe Photoshop CS3.
173

Supplemental Figure 3.17. Schematic of constructs used in DLR assays for gene co-
activation by two TF proteins.

For each tested peak, the same Citrine-expressing control construct as used in the binding assay
(Supplemental Figure 3.9) was used as negative control and was included in all experimental
assays. TFs were expressed individually from separate constructs as intact proteins.
Abbreviations: 35S Pro, CaMV 35S promoter; 35S Ter, CaMV 35S terminator; VP64,
transcriptional activation domain; NOS Ter, nopaline synthase terminator; 35Smini, CaMV 35S
minimal promoter; LUC, firefly luciferase; REN, Renilla luciferase.
174

Supplemental Figure 3.18. Transmission electron micrographs of 12- and 16-DAP AL cells
of W22 and W22o2. Bar = 1 µm.
175

Supplemental Figure 3.19. Venn diagram of numbers of O2-activated direct targets


identified by Li et al. (2015) and this study.
176

Supplemental Table 3.1. Opacity of mature kernels from reciprocal crosses between B73o2,
W22o2, and W64Ao2.
Cross Ear # # F1 Seeds F1 Seed Phenotypea
B73o2 × W22o2 1 406 100% opaque
2 433 100% opaque
W22o2 × B73o2 1 141 100% opaque
B73o2 × W64Ao2 1 285 100% opaque
2 392 100% opaque
W64Ao2 × B73o2 1 199 100% opaque
2 266 100% opaque
3 180 100% opaque
W22o2 × W64Ao2 1 103 100% opaque
2 255 100% opaque
W64Ao2 × W22o2 1 202 100% opaque
a
The opacity of the seeds was analyzed by placing the shelled kernels on a light box.
177

Supplemental Table 3.2. Summary statistics of RNA-Seq reads and mapping.


Samplea Total # reads Unique-map Multi-map Unmapped
34,164,299 18,470,737 4,241,156
B73-1 56,876,192 (60.07%) (32.48%) (7.46%)
37,199,016 15,282,531 4,385,853
B73-2 56,867,400 (65.41%) (26.87%) (7.71%)
28,614,800 20,440,662 3,851,022
B73-3 52,906,484 (54.09%) (38.64%) (7.28%)
48,169,603 14,044,210 4,581,301
B73o2-1 66,795,114 (72.12%) (21.03%) (6.86%)
42,437,810 12,861,375 3,958,251
B73o2-2 59,257,436 (71.62%) (21.70%) (6.68%)
44,522,204 12,892,046 4,498,326
B73o2-3 61,912,576 (71.91%) (20.82%) (7.27%)
a
Biological replicates are indicated as -1, -2 and -3.
178

Supplemental Table 3.3. Summary statistics of ChIP-Seq reads and mapping.


Samplea Total # reads Unique-map Multi-map Unmapped
14,444,794 16,123,111 7,731,529
B73-1 38,299,434 (37.72%) (42.10%) (20.19%)
10,611,258 10,773,945 6,781,075
B73-2 28,166,278 (37.67%) (38.25%) (24.08%)
11,910,627 12,786,781 8,471,960
B73o2-1 33,169,368 (35.91%) (38.55%) (25.54%)
8,462,481 9,045,172 8,073,197
B73o2-2 25,580,850 (33.08%) (35.36%) (31.56%)
a
Biological replicates are indicated as -1 and -2.
179

Supplemental Table 3.4. Summary of the previously reported O2-binding sites.


Binding Site Associated Gene Assay Reference
GATGACATGG b-32 DNase I footprinting assay Lohmer et al. (1991)
GATGATATGG
GATGATGTGG
GATGAGATGA
GTTGACGTGA
GTTGACGTTG O2 DNase I footprinting assay; Lohmer et al. (1991)
Electrophoretic mobility shift assay
(EMSA)
TCACATGTGT 22-kD α-zeins DNase I footprinting assay; Schmidt et al. (1992);
TCATGCATGT EMSA Muth et al. (1996)
TCCACGTAGAT
GCCACGTGGC EMSA Izawa et al. (1993)
GCCACGTCAC
GTGACGTCAC
ATGAGTCAT Pea lectin gene (psI), EMSA de Pater et al. (1994)
ATGACGTCAT etc.
GACACGTGTC
GACATGTC 25-kD α-coixin DNase I footprinting assay; Yunes et al. (1994)
EMSA
TCCACGTCAT 15-kD β-zein; DNase I footprinting assay Cord Neto et al. (1995)
17-kD β-coixin
CATGACGTGT cyPPDK1 DNase I footprinting assay; Maddaloni et al. (1996)
EMSA
TGACGTGG 15-kD β-zein, etc. EMSA Li et al. (2015)
TGGCGTGGCA MYBR13 EMSA Li et al. (2015)
TGACATGTAA 50-kD γ-zein EMSA Li et al. (2015)
TGTCGTGTCA cyPPDK2 EMSA Li et al. (2015)
180

Supplemental Table 3.5. Positions of EMSA probes as compared to O2 peaks.


Gene and
Pattern Peak starta Peak enda Probe starta
associated peak Probe enda
18-kD δ-zein bound activated -255 -14 -268 -14
(O2_peak_4457)
bZIP17 bound activated -113 151 -22 58
(O2_peak_6149)
GBF1 bound activated -1051 -892 -1053 -892
(O2_peak_4628)
bZIP104 bound activated -194 9 -195 10
(O2_peak_3672)
GRMZM2G084296 bound activated -403 -157 -404 -156
(O2_peak_3344)
GRMZM2G443668 bound repressed -461 -202 -460 -202
(O2_peak_4670)
NAC78 bound activated -730 -399 -731 -1
(O2_peak_13)
NKD2 bound activated -653 -307 -654 -1
(O2_peak_1306)
TAR3 bound activated -136 190 -137 190
(O2_peak_4476)
a
Positions relative to TSSs.
181

Supplemental Table 3.6. Results of DLR assays of O2-target gene (peak) interaction shown as
fold induction values (mean ± SEM).
Peaka Pattern Mean Fold Induction SEM P-valueb
Peak 4228 bound activated
(15-kD β-zein) 30.19870897 2.1270946 0.00016
Peak 4457 bound activated
(18-kD δ-zein) 13.7398991 0.9430607 0.00014
Peak 2807 bound activated
(FL2) 11.74464428 0.1693459 3.60E-10
Peak 4176 bound repressed
(bZIP12) 3.279100614 0.7133988 0.031
Peak 6149 bound activated
(bZIP17) 44.55504314 6.9191863 0.0032
Peak 3672 bound activated
(bZIP104) 0.872333458 0.0883905 0.38
Peak 5262 bound activated
(NAC122) 1.078265429 0.1469045 0.63
Peak 1306 bound activated
(NKD2) 14.79399112 2.3349731 0.0041
Peak 4688 bound activated
(O2) 1.209254634 0.3076512 0.56
Peak 6041 bound repressed
(GRMZM2G034623) 1.288271074 0.1604575 0.27
Peak 3344 bound activated
(GRMZM2G084296) 1.136903273 0.1502386 0.47
Peak 4476 bound activated
(TAR3) 5.286058847 1.082415 0.016
Peak 4670 bound repressed
(GRMZM2G443668) 0.571812759 0.1044283 0.012
a
The gene symbol (if available) or ID of the gene associated with each peak is indicated in
parenthesis.
b
Student's t test.
182

Supplemental Table 3.7. Results of MEME-ChIP analysis of the 401-bp sequences flanking the
summits of peaks associated with the putative direct O2 targets.
Discovery
Motif # Sequence E value Matching TFa
Program
1 AAA[AG]AAAA[GC]A MEME 1.1E-026 SOC1, PI, SEP3, AGL15,
FLC, SVP, AP3, PEND,
HMG-I/Y, id1
2 AT[GA]T[GA]TA[AT]A[GA] MEME 1.5E-023 -
3 CGGACCGTCC MEME 5.0E-014 AtMYB15
4 ACTAAATGTC MEME 1.6E-014 AtMYB77, HAT5
5 CC[AT][CT][GA]TCA[TC]C MEME 6.8E-012 bZIP911, TGA1, PIF5,
TGA1A, PIL5, EmBP-1,
bZIP910, PIF3, ABF1
6 AGATAAG[GC]TA MEME 1.3E-008 -
7 C[AC]AAAT[CA]CA[AT] MEME 5.3E-010 FLC
8 CTTTCCA[CG][CT]A MEME 1.8E-010 -
9 T[GA]T[GC]T[TG]GCT[TA] MEME 2.2E-007 RAV1
10 G[AG]CATGCATG MEME 2.5E-003 ABI3
11 CCA[CA][CG]TAGAT MEME 5.6E-005 -
12 [TC]TA[AG]CTACT[AG] MEME 1.2E-002 -
13 AGA[GT][AG]AG DREME 4.4E-009 -
14 [AG][AG]AAAAAA DREME 1.4E-007 SOC1, id1, PI
15 CCAC[AG]TCA DREME 2.5E-003 bZIP911, TGA1A, EmBP-1,
PIF5, TGA1, PIF3, bZIP910,
PIL5, PIF4
16 AAAGGTG[AT] DREME 4.0E-003 myb.Ph3
17 TTTATA[GTC] DREME 1.6E-002 -
18 ACATTTAG DREME 1.8E-002 -
19 C[AT]TGGCA DREME 3.0E-002 HY5, ABF1, PIF3, PIF4,
bZIP911, BES1
a
Transcription factors of which the known binding motif most significantly matches the identified motif ( P <
0.05).
183

Supplemental Table 3.8. Results of DLR assays of target-gene (peak) co-activation by O2 and bZIP17 shown as fold induction values
(mean ± SEM).

Effector(s) Peak 2807 Peak 4457 Peak 6149 Peak 1306 Peak 4228 Peak 4476
(FL2) (18-kD δ-zein) (bZIP17) (NKD2) (15-kD β-zein) (TAR3)
YFP 1 1 1 1 1 1
YFP+O2 5.0722821 9.97186481 6.321675053 6.4261866 15.8951153 2.2553364
Mean Fold YFP+bZIP17 1.154322 1.298354437 1.211848883 1.357575 0.873996971 1.0172567
Induction YFP+O2+bZIP17 6.2490692 5.552213495 6.876862651 4.8711119 13.34425015 2.491189
YFP 0.0964522 0.056416247 0.134519653 0.0449098 0.199182478 0.044755
YFP+O2 0.3799527 0.601335187 0.699890191 0.7005547 2.16573012 0.1632843
YFP+bZIP17 0.1734506 0.229449135 0.096803722 0.1894703 0.10984541 0.1064258
SEM YFP+O2+bZIP17 0.2162841 0.84042507 1.013333602 0.6031377 1.183713467 0.266953
YFP+O2 vs. YFP 1.16E-08 1.20E-08 8.92E-05 2.50E-06 1.38E-06 0.0003616
YFP+bZIP17 vs. YFP 0.9677708 0.977903737 0.994868475 0.9491393 0.999860704 0.9998532
YFP+O2+bZIP17 vs. YFP 2.68E-10 8.65E-05 2.82E-05 0.0001457 1.51E-05 5.46E-05
YFP+bZIP17 vs. YFP+O2 2.02E-08 1.94E-08 0.000140235 5.99E-06 1.24E-06 0.0004171
Adjusted YFP+O2+bZIP17 vs. YFP+O2 0.0151413 0.000120394 0.920245413 0.1329973 0.485319552 0.7515733
P-valuea YFP+O2+bZIP17 vs. YFP+bZIP17 4.25E-10 0.000183129 4.35E-05 0.0004087 1.33E-05 6.24E-05
a
one-way ANOVA with post-hoc HSD test.
184

Supplemental Table 3.9. Putative direct O2 targets that have previously been predicted as direct NKD2 targets
genes by Gontarek et al. (2016).
Gene
Gene ID NKD2 Target in Zein (Sub-)Family TF Family Symbol/ Functional Annotation
TF Name
GRMZM2G346897 AL only 22-kD α-zein (z1C) - AZ22Z4 22kD alpha zein 4; Zein
GRMZM2G088441 AL only 22-kD α-zein (z1C) - - Uncharacterized protein
GRMZM2G388461 AL only 22-kD α-zein (z1C) - - Uncharacterized protein
GRMZM2G160739 AL only 22-kD α-zein (z1C) - ZP3 Zein-alpha B49
GRMZM2G397687 AL only 22-kD α-zein (z1C) - FL2 Zein-alpha PZ22.1/22A1
GRMZM2G346895 AL only 22-kD α-zein (z1C) - - Uncharacterized protein
GRMZM2G044625 AL only 22-kD α-zein (z1C) - AZ22Z3 Zein-alpha PZ22.3
GRMZM2G044152 AL only 22-kD α-zein (z1C) - AZS22-8 Zein-alpha ZA1/M1
GRMZM2G086294 AL only 15-kD β-zein - ZP15 Zein-beta
GRMZM2G181362 AL only - - LKRSDH1 Lysine-ketoglutarate reductase/
saccharopine dehydrogenase1
GRMZM2G052266 AL only - - - Uncharacterized protein
GRMZM2G008607 AL only - - - Uncharacterized protein
GRMZM2G141810 AL only - - TAR3 Alliin lyase
GRMZM2G152764 AL only - - - -
GRMZM2G162690 AL only - - TRE1 Uncharacterized protein
GRMZM2G072240 AL only - - - Uncharacterized protein
GRMZM2G044498 AL only - - - -
GRMZM2G034623 AL only - - - Uncharacterized protein
GRMZM2G002630 SE and AL - - - Uncharacterized protein
GRMZM2G008913 SE and AL 19-kD α-zein (z1A) - ZP1 Zein-alpha PMS2
GRMZM2G059620 SE and AL 19-kD α-zein (z1A) - ZEIN Zein-alpha 19B1
GRMZM2G353268 SE and AL 19-kD α-zein (z1A) - - Zein-alpha A30
GRMZM2G008341 SE and AL 19-kD α-zein (z1A) - - Zein-alpha 19A2
GRMZM2G353272 SE and AL 19-kD α-zein (z1A) - FL4 Zein-alpha PMS1
GRMZM2G100018 SE and AL 18-kD δ-zein - DZS18 18kD delta zein
GRMZM2G053120 SE and AL 19-kD α-zein (z1A) - ZMPMS1 Zein-alpha PMS1
GRMZM5G817886 SE and AL - - AAA1 -
GRMZM2G038991 SE and AL - - - Uncharacterized protein
GRMZM5G884137 SE and AL - C2H2 NKD2 Uncharacterized protein
GRMZM2G141535 SE and AL - - AO1 Indole-3-acetaldehyde oxidase
GRMZM2G406204 SE only - NAC NACTF78 -
GRMZM2G024104 SE only - - GLN2 Glutamine synthetase root isozyme 2
GRMZM2G081060 SE only - - - Uncharacterized protein
GRMZM5G800764 SE only - - - Uncharacterized protein
185

Supplemental Table 3.10. Results of DLR assays of target-gene (peak) co-activation by O2 and NKD2 shown as fold induction values (mean
± SEM).

Peak 2807 Peak 4457 Peak 6149 Peak 1306 Peak 4228 Peak 4476
Effector(s)
(FL2) (18-kD δ-zein) (bZIP17) (NKD2) (15-kD β-zein) (TAR3)
YFP 1 1 1 1 1 1
YFP+O2 5.0722821 9.97186481 6.3216751 6.4261866 15.8951153 2.255336403
Mean Fold YFP+NKD2 1.3002807 1.039921507 1.0802283 14.149507 2.292685777 1.254241639
Induction YFP+O2+NKD2 7.325064 10.15087747 11.07454 15.953283 18.91220739 4.120227019
YFP 0.0964522 0.056416247 0.1345197 0.0449098 0.199182478 0.044754955
YFP+O2 0.3799527 0.601335187 0.6998902 0.7005547 2.16573012 0.163284341
YFP+NKD2 0.1381118 0.154100517 0.0468093 1.3205547 0.29430875 0.158760201
SEM YFP+O2+NKD2 0.1812176 1.007983766 0.426038 0.8629003 3.433991745 0.828667322
YFP+O2 vs. YFP 5.06E-09 5.87E-08 5.98E-07 0.0020767 0.000488537 2.07E-01
YFP+NKD2 vs. YFP 0.7859448 0.999959424 0.9990538 5.40E-08 0.968951862 9.75E-01
YFP+O2+NKD2 vs. YFP 6.99E-12 4.44E-08 5.43E-11 8.58E-09 6.64E-05 5.30E-04
YFP+NKD2 vs. YFP+O2 1.53E-08 6.25E-08 7.34E-07 5.42E-05 0.001191107 3.83E-01
Adjusted YFP+O2+NKD2 vs. YFP+O2 1.55E-05 0.996411261 2.67E-06 4.19E-06 0.725162271 3.37E-02
P-valuea YFP+O2+NKD2 vs. YFP+NKD2 1.37E-11 4.72E-08 6.14E-11 0.4729196 0.000153646 1.22E-03
a
one-way ANOVA with post-hoc HSD test.
186

Supplemental Table 3.11. Sequences of the primers used to generate the EMSA probes.
Gene and
Forward Primer Reverse Primer
associated peak
18-kD δ-zein GCCAAGCTTAGTCTCTTAATCCA GATCCTCGAGATGGCCATGTAG
(O2_peak_4457) CATCAG ATGAAAAAC
bZIP17 CGACGCGGATCCGAGTCATCCC GGAAAGGTTGGTTGGGGTGCTG
(O2_peak_6149) GCGTCACCATCCATTTGACGAG GGGGATAAACCGCCTCGTCAAA
GCGGTTTATCCCCCAGCACCCC TGGATGGTGACGCGGGATGACT
AACCAACCTTTCCA CGGATCCGCGTCGA
GBF1 GCCAAGCTTCACCCATCCACCTC GCCCTCGAGCCATGTGCGCCTG
(O2_peak_4628) CCAAGTC CGTATGAG
bZIP104 GATCAAGCTTAGCACGGACCGC GATCCTCGAGATCGGAATTCGG
(O2_peak_3672) ACCGCAGTG AGGCGAAAG
GRMZM2G084296 GATCAAGCTTAGTCGCCGAGAC GATCCTCGAGGAAAGGAGAAAG
(O2_peak_3344) AAAAAAAAGC GACAGTG
GRMZM2G443668 GATCAAGCTTCGAACACGAAGC GATCCTCGAGGGTGTGAGATCT
(O2_peak_4670) CTAATCAAAAC CTGGGTTG
NAC78 GATCAAGCTTAAGGGAAGTCGC GATCGTCGACTGCTCCCTCTGC
(O2_peak_13) GGGGCAAC CCTTCCTC
NKD2 GCCAAGCTTACACGCTTCTTCAT GCCCTCGAGATTCCCGAAATTCC
(O2_peak_1306) TAGTGTACGC ACCCCTG
TAR3 GATCAAGCTTATAGGTTTTAGAC GATCCTCGAGCGTCCCGCGGAT
(O2_peak_4476) TTTGGCTC TTAAAG
187

Supplemental Table 3.12. Sequences of the primers used to generate the competitor sequences in the
EMSAs.

Gene and Sequence


Primer 1 Primer 2
associated peak Name

18-kD δ-zein 100018comp GTCTCTTAATCCACATCAG GTATGACTAGGCTGATGTG


(O2_peak_4457) CCTAGTCATAC GATTAAGAGAC
100018mut GTCTCTTAATCCCCATGCA GTATGACTAGGCTGCTGGG
GCCTAGTCATAC GATTAAGAGAC
bZIP17 103647comp1 CGACGCGGATCCGAGTCAT ACGCGGGATGACTCGGATC
(O2_peak_6149) CCCGCGT CGCGTCG
103647comp2 GTCATCCCGCGTCACCATC TCAAATGGATGGTGACGCG
CATTTGA GGATGAC
103647comp3 TCACCATCCATTTGACGAG GGATAAACCGCCTCGTCAA
GCGGTTTATCC ATGGATGGTGA
GBF1 011932comp1 GCTCCCGTGGGCCCCCAC CCACGTAGCCCGCACCCC
(O2_peak_4628) GGCTATAGTACTGCGGGGT GCAGTACTATAGCCGTGGG
GCGGGCTACGTGG GGCCCACGGGAGC
011932comp2 ATAAGCTAGCTAGGTCCCT GACTAGGTGACAGCTCAGG
GAGCTGTCACCTAGTC GACCTAGCTAGCTTAT
bZIP104 098904comp1 CGCCGCGGAGTGGCAAAC TCTTCCGGGCCGCTTTCGC
(O2_peak_3672) TCGTAGATATCTCGCGAAA GAGATATCTACGAGTTTGC
GCGGCCCGGAAGA CACTCCGCGGCG
098904comp2 CGGAAGACACAGCGATCTG GGCCGGCTGCAACGGAGT
CTTTCCACGCCGGACTCCG CCGGCGTGGAAAGCAGAT
TTGCAGCCGGCC CGCTGTGTCTTCCG
098904comp3 GCCGGCCACCTTTTCCTCC AGGCGAAAGGAGACTAGG
CTCCTTCCCCTCTCCTAGT AGAGGGGAAGGAGGGAGG
CTCCTTTCGCCT AAAAGGTGGCCGGC
GRMZM2G08429 084296comp TGCCTCTGCTCCGCGTCAC AACGGGATGCGGTGACGC
6 (O2_peak_3344) CGCATCCCGTT GGAGCAGAGGCA
084296mut TGCCTCTGCTCCCCGGCAC AACGGGATGCGGTGCCGG
CGCATCCCGTT GGAGCAGAGGCA
GRMZM2G44366 443668comp CTTTTTTTTCGATGACATGC GTTATACACGGCATGTCAT
8 (O2_peak_4670) CGTGTATAAC CGAAAAAAAAG
443668mut CTTTTTTTTCGATGCCAGGC GTTATACACGGCCTGGCAT
CGTGTATAAC CGAAAAAAAAG
NAC78 406204comp TCCAACGGGAGCTGACTCG CAGAGCGATGCCGAGTCA
(O2_peak_13) GCATCGCTCTG GCTCCCGTTGGA
406204mut TCCAACGGGAGCTGCCTG CAGAGCGATGCCCAGGCA
GGCATCGCTCTG GCTCCCGTTGGA
NKD2 884137comp TAGAAGTAATCATGATGTG ATGTGATGACTCACATCAT
(O2_peak_1306) AGTCATCACAT GATTACTTCTA
884137mut TAGAAGTAATCATGCTGGG ATGTGATGACTCCCAGCAT
AGTCATCACAT GATTACTTCTA
TAR3 141810comp CGCCGCATGCGCACGTCA CGGCCGGGCCGGTGACGT
(O2_peak_4476) CCGGCCCGGCCG GCGCATGCGGCG
141810mut CGCCGCATGCGCCCGGCA CGGCCGGGCCGGTGCCGG
CCGGCCCGGCCG GCGCATGCGGCG
188

Supplemental Table 3.13. Sequences of primers used for RT-qPCR.


Gene ID Forward Primer Reverse Primer
GRMZM2G015534 GGATTATGAGCTGCTGGGTC CCAAACGGAACACTCATGGTTA
GRMZM2G044625 ACCAACTGACTATGTCGAAC ACAAGACAGGGTTTATCAAAGAGAA
GRMZM2G086294 TACGAGCCAGCTCTGATG CAGTAGTAGGGCGGAATG
GRMZM2G025763 AGAGCCGAGTTTTCACATAC AATAGAGCCACAATGGCG
GRMZM2G063536 TGCGCTCGTTAAGAATCAAACTA TTTAGACCATATAAGGAACACCG
GRMZM2G181362 GTGGAGTATATTGTAAGAGACGG CTCTGACGTCTCATCTTGTT
GRMZM2G306345 GGATGTGGTGATCAACAGTAT TGATGATGTATTGAAGAACAGAGT
GRMZM2G097457 GAGTATTAGTAGCAGCGCTC GGCCTCAATAGGTGGAATG
GRMZM2G103647 ACAAAGCTGCAAATGAGAAG GGTGCAAATCTACTAGATGACT
GRMZM2G100018 CCACAATGTTACTCTGATTCCA TGGGTACAACACAAATGCTTA
GRMZM2G011932 TTCCAGTGAAGTAACCGAGG TGACGGTCTCGAAAGTTTCAA
GRMZM5G884137 AGAAAGATGCGTCTTCTCTGA GATCCAACAAAGATTATCCAACTAC
GRMZM2G098904 TGCTGAGAGACTGAACTCG ATAGAGTCACAACACCTAACG
GRMZM2G014653 AAGAACTTGGCGATAATGTG CCATGGGCATCGAATACT
GRMZM2G430849 TGGATTGGAGGCCATTAATAAAG GTACCTGTACTCTACAACGGA
GRMZM2G053720 CGTGAGTGAAGTGCTTGT CGAACACTATTTCATTCATTCGAT
GRMZM2G117956 CGTGAGTGAAGTGAACCTG CTCGATAAGCCTCATTATTCCAT
GRMZM2G440949 CTTGAAGTTCTCCCACGC GCCCTCACATGGATGCATAATA
GRMZM2G084296 GGCAGTTACTTGAGCTCTTT TTCATCACTCTTAACATCACCG
GRMZM2G448607 CATAAATCATCCTTGTCATGGT ATAGGCATATTCGATTTCATAAGC
GRMZM2G443668 TCGAGGAGCAGGCTATGA TGTACAACGAGCCAGTGTC
GRMZM2G034623 GCCGTGATCTCGTCATATT GCGCGATCCAGCTATATACT
GRMZM2G036976 CAAAGTTCCAATTATGGCTGC ACGTCAACATCAACAATTGC
GRMZM2G088469 GTACTACTGGTGGACGATGT ACTAATTAAGACTTCCCTATGCC
GRMZM2G143008 TCCAGCAAGCAACTGATCTAA GTCGGAAGCTACATAGGATTAC
GRMZM2G018193 GCCCTCTTTTATATTTCTTATGGGT TCCCGTTCTTAGAGCAACA
GRMZM2G125130 CTGCCATTCTACCAGCTATTT TAGGAGGCCACATACATCAG
AF546187.1_FG007 GCTTCCATTCCATCTACAAGT GACATTGTTTGCCACATCCT
GRMZM2G048910 CAAGAACTGACAAATTCGGG GCATGAGAAGTCGGAAGC
GRMZM2G146283 GTAGTACCCAGTGAAATCAGG CCAATGATATATAGCTTACTGCAA
GRMZM2G154182 CCGTAATAACCAGTAGTTGGTAA ACGATAATATTGTACCAAACTCGAA
AC209080.3_FG004 GGTGCACGGATGCTTAAGA GGTTCGGCGTTATGATGAC
GRMZM2G071433 TTTCCTGTATGTCATGTGTTCG GAATAAGATAGAACTTACCAAACCG
GRMZM2G124207 AGCAAGCGTAGTATATATCATCC ACTATTCATCACACGCTCTAAC
GRMZM2G060429 GTTGGTTCTGAGTCTTGCTA ACTCTATGTCCATGAAACTGT
GRMZM2G014705 TAGCTGGACGCACTAGTTC GTAGATACAGTAAACTGGAGCAGA
GRMZM5G802403 TGATTACACTCTTGAGTTCATGTC TGAACATTCCAATCCGAGC
GRMZM2G138689 AGTGGTGTTTGCAATTGAAG ACACAAGGGTTTAGTAGCAT
GRMZM2G174883 CTGCCTGAATAAGTACTCGT GCTTACAACAATCTGCTCATTT
GRMZM2G089713 TGTCCTTCGATTAGTACGGG CTGCCCAAAAACGCTCAA
GRMZM2G028393 GGCTGATAATATATAAGCTAGCCG GGCGACATTTGAATACGTTT
GRMZM2G066612 TTCATGGGACATCCGTGC CCACGAGCTTGTCGATCT
189

Supplemental Table 3.14. Cloning strategies and sequences of primers used for directed Y2H assays.
Enzymes used Enzymes used to
CDS Purpose Forward Primer Reverse Primer
to cut insert linearize vector
O2 Cloning into NdeI/XmaI NdeI/XmaI GGCCATATGATGGAGCACG CCCCCCGGGCTAATACATGT
pGBKT7 vector TCATCTCAAT CCATGTGTAT
Cloning into NdeI/XmaI NdeI/XmaI
pGADT7 vector
O2F4 Cloning into NdeI/XmaI NdeI/XmaI GGCCATATGGTGGTGCCGA CCCCCCGGGCTAGCTCATC
pGBKT7 vector ACTCTTGTTG TCTATCACCCGCTT
bZIP17 Cloning into EcoRI/SalI EcoRI/SalI CCGGAATTCATGAAGAAGTG CGCGTCGACTCAAGGCCAA
pGBKT7 vector CGCGTCGGA ATGTCCGCCGCGCAA
Cloning into EcoRI/SalI EcoRI/XhoI
pGADT7 vector
190

Supplemental Table 3.15. Cloning strategies and sequences of primers used for BiFC assays.
Primer Name Primer Sequence Notes
attB1-adapter GGGGACAAGTTTGTACAAAAAAGCAGGCT Adapter primers for attB adapter PCR
attB2-adapter GGGGACCACTTTGTACAAGAAAGCTGGGT
attB1-O2ORF-F AAAAAAGCAGGCTTCATGGAGCACGTCATCTCAAT O2-ORF-specific primers
attB2-O2ORF-R CAAGAAAGCTGGGTTATACATGTCCATGTGTATGGC
O2LA-F CAGGTAGCACAGGCAAAAGCCGAGAATTCTTGCGCGCTGAGGCGC Primer with substitutions
O2LA-R GCGCCTCAGCGCGCAAGAATTCTCGGCTTTTGCCTGTGCTACCTG Primer with substitutions;
reverse complement of O2LA-F
191

Supplemental Table 3.16. Cloning strategies and sequences of primers used for DLR assays.
Cloning
cDNA/gDNA Purpose Forward Primer Reverse Primer
Strategy
Citrine ORF Cloning into Gateway GGGGACAAGTTTGTACAAAAAAGCAGGCTT GGGGACAACTTTGTATAGAAAAGTTGGGTG
pDONR221-P1-P4 Cloning AATGGTGAGCAAGGGCGAG CTTGTACAGCTCGTCCATGCC
O2 ORF Cloning into Gateway GGGGACAAGTTTGTACAAAAAAGCAGGCTT GGGGACAACTTTGTATAGAAAAGTTGGGTG
pDONR221-P1-P4 Cloning AATGGAGCACGTCATCTCAATGG ATACATGTCCATGTGTATGGCCC
O2 ORF Cloning into pBN- In- CGACTCTAGAGGATCCATGGAGCACGTCAT GGAAATTCGAGCTCGGTACCCTAATACATGT
35Stev Fusion CTCAAT CCATGTGTAT
Cloning
bZIP17 ORF Cloning into pBN- In- CGACTCTAGAGGATCCATGAAGAAGTGCGC GGAAATTCGAGCTCGGTACCTCAAGGCCAA
35Stev Fusion GTCGGA ATGTCCGCCGCGCAA
Cloning
NKD2 ORF Cloning into pBN- In- CGACTCTAGAGGATCCATGATGGCGTCGAA GGAAATTCGAGCTCGGTACCTCATGGCATC
35Stev Fusion TTCACC CTGCCTCCAT
Cloning
Peak 4228 Cloning into Gateway GGGGACAACTTTGTATAATAAAGTTGTATTG GGGGACCACTTTGTACAAGAAAGCTGGGTT
(15-kD β-zein) pDONR221-P3-P2 Cloning ATCATTCAGAGCAATGTATTGGTTGT GACGATCCTCTCTACGCACTGC
Peak 4457 Cloning into Gateway GGGGACAACTTTGTATAATAAAGTTGTACAC GGGGACCACTTTGTACAAGAAAGCTGGGTT
(18-kD δ-zein) pDONR221-P3-P2 Cloning ATCAGCTAGTCATACTTTAGCAAAAGC AATGGCCATGTAGATGAAAAACAATTTCTGG
Peak 2807 (FL2) Cloning into Gateway GGGGACAACTTTGTATAATAAAGTTGTAGTC GGGGACCACTTTGTACAAGAAAGCTGGGTT
pDONR221-P3-P2 Cloning ATTCCACATAAATGAAAAGAATTCCTATATAA TCATATATTTTTTTTTGGATTTTGTCTTTTTCC
AAATGACA AAACAACAGATAAG
Peak 4176 (bZIP12) Cloning into Gateway GGGGACAACTTTGTATAATAAAGTTGTAAAG GGGGACCACTTTGTACAAGAAAGCTGGGTT
pDONR221-P3-P2 Cloning ACGGCGCGAGG ATGATAGTTACCGGAGATGACACAATGAA
Peak 6149 (bZIP17) Cloning into Gateway GGGGACAACTTTGTATAATAAAGTTGTACAC GGGGACCACTTTGTACAAGAAAGCTGGGTT
pDONR221-P3-P2 Cloning CGTGACAGCCACGG CCCGTCCCGGCGTG
Peak 3672 Cloning into Gateway GGGGACAACTTTGTATAATAAAGTTGTAGCA GGGGACCACTTTGTACAAGAAAGCTGGGTT
(bZIP104) pDONR221-P3-P2 Cloning CGGACCGCACC ATCGGAATTCGGAGGCGAAAGG
Peak 5262 Cloning into Gateway GGGGACAACTTTGTATAATAAAGTTGTAGAT GGGGACCACTTTGTACAAGAAAGCTGGGTT
(NAC122) pDONR221-P3-P2 Cloning CCGGCCGCGGCCC AGATCGAGGACTGACTGACTGG
Peak 1306 (NKD2) Cloning into Gateway GGGGACAACTTTGTATAATAAAGTTGTAACA GGGGACCACTTTGTACAAGAAAGCTGGGTT
pDONR221-P3-P2 Cloning CGCTTCTTCATTAGTGTACGCT TATGGGTCGCAAGCATGCGTTTTCAC
Peak 4688 (O2) Cloning into Gateway GGGGACAACTTTGTATAATAAAGTTGTAGGC GGGGACCACTTTGTACAAGAAAGCTGGGTT
pDONR221-P3-P2 Cloning ATGCATGGGTACACG TGCTGGGCTCTCTCTCTTGC
Peak 6041 Cloning into Gateway GGGGACAACTTTGTATAATAAAGTTGTACTT GGGGACCACTTTGTACAAGAAAGCTGGGTT
(GRMZM2G034623) pDONR221-P3-P2 Cloning TTGTGAACGTACGGTACTATGACG GGTGCATGAGTTCGGCGTGCA
Peak 3344 Cloning into Gateway GGGGACAACTTTGTATAATAAAGTTGTAAGT GGGGACCACTTTGTACAAGAAAGCTGGGTT
(GRMZM2G084296) pDONR221-P3-P2 Cloning CGCCGAGACAAAAAAAAGC GAAAGGAGAAAGGACAGTGCGTG
Peak 4476 (TAR3) Cloning into Gateway GGGGACAACTTTGTATAATAAAGTTGTATAG GGGGACCACTTTGTACAAGAAAGCTGGGTT
pDONR221-P3-P2 Cloning GTTTTAGACTTTGGCTCTTCTTGT GCGTCCCGCGGATTTAAAG
Peak 4670 Cloning into Gateway GGGGACAACTTTGTATAATAAAGTTGTAACG GGGGACCACTTTGTACAAGAAAGCTGGGTT
(GRMZM2G443668) pDONR221-P3-P2 Cloning AACACGAAGCCTAATCAAAACGG GGGTGTGAGATCTCTGGGTTGTG
192

Supplemental Data Set 3.1. Normalized expression levels in RPKM, results of differential
expression analysis, and functional annotation for genes tested for differential gene expression.

Supplemental Data Set 3.2. Associated peaks and functional annotation of O2-bound genes.

Supplemental Data Set 3.3. Normalized expression levels in RPKM, associated peaks, and
functional annotation of direct O2 targets.

Supplemental Data Set 3.4. GO term enrichments and pathway associations of direct and
indirect target gene sets.

Supplemental Data Set 3.5. BLASTX search of the O2-repressed direct targets against the
NCBI nr database (E-value < 10-6).

Supplemental Data Set 3.6. BLASTX search of the O2-repressed direct targets against the
Swissprot database (E-value < 10-6).

Supplemental Data Set 3.7. Putative indirect O2 targets that have previously been predicted as
direct NKD2 targets by Gontarek et al. (2016).

Supplemental Data Set 3.8. Ct values for genes analyzed using RT-qPCR.

Supplemental Data Set 3.9. O2-bound and activated genes identified through ChIP-Seq and
RNA-Seq analysis by Li et al. (2015) and/or this study.
193

CHAPTER 4

NAKED ENDOSPERM transcription factors regulate partially overlapping gene


networks during maize endosperm development

4.1 ABSTRACT

Endosperm is a nutritive organ that supports the development of the embryo and seedling in
flowering plants. However, the molecular mechanisms underlying endosperm development
remain elusive. The recently duplicated NAKED ENDOSPERM (NKD) genes NKD1 and NKD2
encode INDETERMINATE DOMAIN (IDD) transcription factor (TF) proteins that have been
shown in combination to be required for endosperm development in maize (Zea mays). To
understand the individual functions of each NKD gene and to gain insights to the transcriptional
networks regulating endosperm development, we performed a transcriptome analysis of
endosperm from the single and double Dissociation transposon insertion mutants of NKD1 and
NKD2 in maize W22 background as compared to the wild type at 8, 12, and 16 days after
pollination. Analysis of these strong loss-of-function mutants identified a total of 11,554 genes
differentially expressed between the mutants and the wild type. Analysis of genes co-regulated
by NKD1 and NKD2 revealed several distinct patterns of downstream gene co-regulation by the
NKDs, indicating the two TFs co-regulate downstream genes predominantly in a cooperative
manner. A detailed analysis of the NKD-regulated genes supported the seed phenotypic
observations that NKD1 plays a more critical role than NKD2 in regulation of endosperm gene
expression. Nonetheless, our analysis also uncovered a unique role for NKD2 in regulation of a
subset of endosperm genes. Based on the reported binding sequences of NKD1 and NKD2, we
predicted 2,872 and 2,500 direct target genes of NKD1 and NKD2, respectively, many of which
can be linked to the nkd mutant phenotype. As a further step toward deciphering the downstream
gene networks regulated by the NKDs, we performed a gene coexpression network analysis on
all the NKD-regulated genes and identified gene modules that are regulated by the NKDs in a
highly coordinated manner. Through de novo cis-motif enrichment and TF family enrichment
analyses, we predicted families of TFs that are likely involved in the regulation of the
coexpressed gene sets under regulation of the NKDs. These results help to uncover the
regulatory functions of individual NKD proteins and provide further insights into the gene
regulatory networks controlling endosperm development in maize.
194

4.2 INTRODUCTION

Endosperm is a key evolutionary innovation in flowering plants and functions as a nutritive


organ in support of the developing embryo in eudicots or the germinating seedlings in cereal
grains (Li and Berger, 2012; Olsen and Becraft, 2013). The latter is a major source of human
calories (Sabelli and Larkins, 2009; FAO, 2015). Endosperm development is initiated by double
fertilization, during which one of the two sperm cells fertilizes the egg cell to produce the
embryo and the other fuses with the central cell to give rise to the endosperm (Friedman et al.,
2008; Hamamura et al., 2012). In maize (Zea mays), upon fertilization, the fertilized central cell
undergoes a period of coenocytic growth followed by cellularization at about 4 days after
pollination (DAP). Subsequently, the endosperm differentiates into four main cell types,
including the aleurone (AL), the basal endosperm transfer layer (BETL), the embryo-
surrounding region (ESR), and the starchy endosperm (SE). Mitotic proliferation of the
endosperm also begins immediately after cellularization, and lasts until about 8 to 12 DAP in the
center of the SE, and until 20 to 25 DAP in the AL and the periphery of the SE. Beginning at
about 8 to 10 DAP, the SE cells gradually switch from mitotic cell cycle to an endoreduplicative
cell cyle, and accumulate storage compounds, including starch and storage proteins. When the
maize kernel begins to mature by about 16 DAP, the SE cells undergo programmed cell death
and desiccation (Olsen, 2004; Sabelli and Larkins, 2009; Becraft and Gutierrez-Marcos, 2012;
Zhan et al., 2017). The NAKED ENDOSEPRM1 (NKD1) and NKD2 genes encode
INDETERMINATE DOMAIN (IDD) family transcription-factor (TF) proteins and have been
shown to regulate in combination multiple aspects of maize endosperm development, including
cell cycle control, cell differentiation, storage compound synthesis, and maturation (Yi et al.,
2015; Gontarek et al., 2016). However, it is still unclear how these distinct developmental
processes are coordinately regulated by the individual NKD proteins.

The NKD phenotype was initially based on the analysis of an nkd double-mutant (referred to as
the nkd-R mutant hereafter) containing a missense mutation in the coding region of NKD1 and a
retrotransposon insertion in the first exon of NKD2 (Yi et al., 2015). nkd-R was shown to be a
partial loss-of-function mutant of both NKD genes and it displays a pleiotropic phenotype in
endosperm development (collectively referred to as the nkd mutant phenotype hereafter). These
defects include an abnormal AL replaced with multi-layered peripheral cells that lack either AL
195

or SE identity, an opaque (floury) endosperm, reduced starch and storage-protein content,


decreased levels of carotenoid synthesis, sporadic anthocyanin accumulation, occasional vivipary
(i.e., precocious germination), and an increased susceptibility to fungal infection (Becraft and
Asuncion-Crabb, 2000; Yi et al., 2015). Correspondingly, a transcriptome analysis of 15-DAP
endosperm from the nkd-R mutant showed that many dysregulated genes are associated with cell
cycle control, storage-protein gene expression, starch biosynthesis, carotenoid biosynthesis,
anthocyanin biosynthesis, abscisic acid (ABA)-biosynthesis and signaling, and pathogen defense
(Gontarek et al., 2016). Differential gene expression analysis of the nkd-R mutant versus wild
type (WT) and an in vitro selection and amplification binding (SAAB) assay of NKD-binding
sites led to identification of many putative direct target genes of NKDs, including Opaque-2
(O2) and Vivipary-1 (VP1) that are well-characterized regulators of the endosperm storage
function and kernel maturation/germination programs, respectively (Schmidt et al., 1992;
Hoecker et al., 1995; Gontarek et al., 2016). The trans-activation of O2 and VP1 by NKDs was
further confirmed by TF-target interaction assays performed in isolated AL protoplast (Gontarek
et al., 2016). A new set of mutant alleles have been generated via insertion of Dissociation (Ds)
transposon in NKD1 (nkd1-Ds) and NKD2 (two alleles, nkd2-Ds0766 and nkd2-Ds0297) (Yi et
al., 2015). The nkd1-Ds mutant displayed a phenotype resembling that conferred by the nkd-R
mutant allele, whereas neither nkd2-Ds mutants displayed any discernible phenotype in the seed.
Therefore, as compared to NKD2, loss of NKD1 function was likely the primary contributor to
the original nkd-R mutant phenotype, and thus required for AL differentiation and overall
endosperm development (Yi et al., 2015).

Here, we addressed the nature of gene expression programs regulated by each individual NKD
protein by performing a transcriptome analysis of a time series of endosperm development from
the Ds insertion mutants of NKD1, NKD2 and a double mutant of NKD1 and NKD2 (referred to
hereafter as nkd1-Ds, nkd2-Ds and the double mutant, respectively) in comparison to a WT W22.
Differential gene expression analysis of the mutants versus WT identified genes that are
regulated by each individual NKD protein or by both. Identification of gene sets co-regulated by
NKD1 and NKD2 uncovered distinct patterns of co-regulation of downstream genes by the NKD
proteins. These data supported a more profound regulatory role for NKD1 than NKD2 in
endosperm development, but also uncovered a separate molecular function for NKD2. Taking
advantage of the previously reported binding sequences of NKD1 and NKD2 (Gontarek et al.,
196

2016), we predicted 2,872 and 2,500 direct target genes of NKD1 and NKD2, respectively.
Through a gene coexpression network analysis, we identified the NKD-regulated genes that are
coordinately regulated by the NKDs, possibly in combination with or through regulation of other
TFs.

4.3 RESULTS

Identification of NKD1 and NKD2 regulated gene sets

To identify genes regulated by the NKD proteins during maize endosperm development, we
performed an mRNA profiling of endosperm of the nkd1-Ds (nkd1-Ds/nkd1-Ds; +/+) and nkd2-
Ds (+/+; nkd2-Ds0766/nkd2-Ds0766) single mutants, and the double mutant (nkd1-Ds/nkd1-Ds;
nkd2-Ds0766/nkd2-Ds0766) at 8, 12, and 16 DAP in comparison to a wild-type W22 using RNA
sequencing (RNA-Seq) (Table 4.1). The three stages of endosperm development corresponded to
cell differentiation and mitotic proliferation, transition to endoreduplication cell cycle and onset
of storage product accumulation, and initiation of maturation, respectively (Sabelli and Larkins,
2009; Becraft and Gutierrez-Marcos, 2012; Zhan et al., 2017). RNA-Seq reads from two to four
biological replicates of each sample were mapped to the maize reference genome (Supplemental
Tables 4.1 and 4.2). Using edgeR, counts of the reads mapped to each gene were normalized as
reads per kilobase per million mapped reads (RPKM) (Supplemental Data Set 4.1). Principal
component analysis (PCA) and Spearman correlation coefficient (SCC) analysis showed that, at
a given developmental stage, the endosperm transcriptome of the double mutant was more
similar to the nkd1-Ds mutant than to the nkd2-Ds mutant (Figure 4.1), suggesting that the
perturbation of the double mutant transcriptome is primarily due to the loss of function of NKD1.
In addition, the transcriptome profiles of the nkd2-Ds endosperm were substantially different
from that of the corresponding WT, indicating that, NKD2 plays a smaller but nonetheless a
detectable role in regulation of gene expression in the developing endosperm (Figure 4.1;
Supplemental Figure 4.1).

By comparing the transcriptome profiles of each mutant endosperm with the WT of the same
stage, we detected 1,386 to 6,258 statistically significant differentially expressed genes (DEGs)
[|fold change (FC)| > 1.5, FDR < 0.05] (Figure 4.2A; Supplemental Figure 4.2). Union of these
197

gene sets yielded 11,554 DEGs (referred to hereafter as NKD-regulated genes) that were
significantly dysregulated in at least one mutant at one or more stages (Supplemental Data Set
4.2). Strikingly, the level of NKD2 mRNA in the nkd1-Ds mutant was as low as in the nkd2-Ds
mutant (RPKM < 1.2), indicating that the expression of NKD2 is largely dependent on activation
by NKD1, either directly or indirectly. To verify the DEGs, we performed reverse transcription
quantitative PCR (RT-qPCR) assays on four of them, including NKD1 (downregulated in the
nkd1-Ds mutant and double mutant at all stages), NKD2 (downregulated in all mutants at all
stages), O2 (downregulated in the nkd1-Ds mutant and the double mutant at 8 DAP), and VP1
(downregulated in the nkd1-Ds mutant at 12 DAP), using the same biological materials as were
used for RNA-Seq. The results confirmed the RNA-Seq-based differential expression patterns
(|FC| ≥ 1.87), and showed that the NKD1 and NKD2 mRNAs were undetectable in their
respective single mutants as well as in the double mutant (Supplemental Figure 4.3;
Supplemental Table 4.3). This supported the notion that the Ds insertion alleles of NKD1 and
NKD2 used here represent stronger alleles than the previously described nkd-R reference mutant
allele, which had shown detectable levels of NKD1 and NKD2 mRNAs (Yi et al., 2015).
Consistent with the PCA and SCC analyses showing extremely high similarity of transcriptome
profiles between the nkd1-Ds mutant and the double mutant at each stage, the number of DEGs
detected in the nkd1-Ds single mutant was close to the double mutant at each stage, whereas the
number of DEGs detected in the nkd2 single mutant was much smaller (Figure 4.2A).

An overall comparison of DEGs detected in all the 9 staged mutant samples versus the three WT
stages revealed extensive overlaps among all 9 DEG sets (Figure 4.2B). GO term enrichment
analysis showed that the NKD-regulated genes were involved in a diverse set of biological
processes (BPs), including those related to chromatin replication, transcription, translation, post-
translational modification of proteins, synthesis/metabolism/transport of primary metabolites and
secondary metabolites, development, and responses to hormone/stress/stimulation (Supplemental
Data Set 4.3). However, further comparisons of DEG sets parsed for mutant types and
developmental stages revealed dramatic temporal changes in the genes regulated by a single
NKD protein or both (Figures 4.2C to 4.2F). The genes that were dysregulated in nkd1 (i.e.,
detected as regulated by NKD1) at all three stages constituted only 36.1% (1,361/3,765), 31.3%
(1,361/4,347), and 22.5% (1,361/6,061) of the NKD1-regulated genes detected at each of the
individual developmental stages 8, 12 and 16 DAP, respectively (Figure 4.2C). Similarly for
198

NKD2, the DEGs that were detected as regulated by NKD2 at all stages were only 42.3%
(672/1,587), 26.8% (672/1,386), and 40.9% (672/1,642) of the NKD2-regulated genes detected
at each of the same stages (Figure 4.2D). These data suggested each NKD protein regulates an
extensive set of BPs through both a broad and stage-specific gene networks. For example, in the
NKD1-regulated gene sets, the only two BPs that were commonly enriched at all three stages
were “auxin-activated signaling pathway” and “response to salt stress,” whereas for the NKD2-
regulated gene sets, no enriched BP was commonly found in all three stages. Similarly, in the
double mutant, only “auxin-activated signaling pathway” and “response to karrikin” were
commonly enriched in DEGs of the three stages (Supplemental Data Set 4.3). GO term
enrichment analysis of stage-specific DEGs further underscored the dynamic nature of NKD-
regulated gene networks. The 8-DAP-specific DEG set (containing 1,756 genes; Figure 4.2F)
were significantly enriched for BPs “translation” and “ribosome biogenesis,” indicating a role for
NKDs in regulation of the translational machinery to control protein synthesis globally and/or to
control the expression of storage protein genes at 8 DAP (Supplemental Data Set 4.4). In
contrast, the 12- and 16-DAP-specific DEG sets (containing 1,204 and 3,057 genes, respectively;
Figure 4.2F) were most significantly enriched for “DNA-dependent DNA replication” and
“nucleosome assembly,” respectively, indicating that the NKDs are involved in regulation of
cellular proliferation and/or the endosperm endoreduplication processes at the later stages of
endosperm development (Supplemental Data Set 4.4).

Stage-specific comparisons of DEGs detected in each mutant versus WT showed that the DEGs
detected in nkd1-Ds versus WT were nearly inclusive of DEGs detected in nkd2-Ds versus WT
(Figures 4.2G to 4.2I). In total, the genes dysregulated in nkd1-Ds included 79.03%
(2,217/2,805) of the NKD2-regulated genes (Figure 4.2J). These data support the notion that
NKD1 plays a broader role than NKD2 in the regulation of endosperm gene expression.
However, NKD1 and NKD2 likely each have unique regulatory functions, as many genes were
detected to be specifically dysregulated in one of the two single mutants at each stage (Figures
4.2G to 4.2I). GO term enrichment analysis showed that the united nkd1-Ds-specific DEG set
(containing 2,185 genes; Figure 4.2J) were most significantly enriched for BPs related to rRNA
processing and translation, whereas the double-mutant-specific DEGs (containing 2,051 genes;
Figure 4.2J) were enriched for several BPs associated with DNA replication (Supplemental Data
Set 4.5). The latter, in effect, suggests that the associated genes are regulated by NKD1 and
199

NKD2 in a redundant manner. Together, these results indicate that NKD1 and NKD2 regulate
overlapping but substantially different downstream gene networks.

Analysis of gene co-regulation patterns conferred by NKD1 and NKD2

In all mutants analyzed here, the number of upregulated genes was larger than the number of
downregulated genes (Figure 4.2A), suggesting strongly that the NKDs can function both as
activators and repressors, perhaps with the latter being the primary role. In fact, many
downstream genes seem to be co-regulated by NKD1 and NKD2 (Figures 4.2G to 4.2J).
Therefore, we interrogated all the DEGs for genes that were likely co-regulated directly or
indirectly by NKD1 and NKD2. We defined the genes that were significantly downregulated in
both nkd1-Ds and nkd2-Ds single mutants as co-activated genes (Figure 4.3A), those upregulated
in both nkd1-Ds and nkd2-Ds as co-repressed genes (Figure 4.3B), those downregulated in nkd1-
Ds and upregulated in nkd2-Ds as NKD1-activated and NKD2-repressed genes (Supplemental
Figure 4.4A), and those upregulated in nkd1-Ds and downregulated in nkd2-Ds as NKD1-
repressed and NKD2-activated genes (Supplemental Figure 4.4B). The numbers of genes in the
resulting gene sets indicate that in most cases NKD1 and NKD2 function in a cooperative
manner, whereas only a small number of gene are regulated by them antagonistically. GO term
enrichment analysis showed that the genes co-activated by NKDs at all three stages (Figure
4.3C) were enriched for “transposition, DNA-mediated,” “lipid metabolic process,” and two BPs
related to spermine biosynthesis (Supplemental Data Set 4.6). The gene set co-repressed by
NKDs at all three stages (Figure 4.3C) did not show significant GO term enrichments likely due
to limited available annotation, but it contained the COLORED1 (R1) gene (Supplemental Data
Set 4.6), which is a known transcriptional regulator of the anthocyanin biosynthesis pathway
(Selinger and Chandler, 1999). Together, these results support a model in which NKDs co-
regulate different gene sets in different manners and ultimately to regulate a diverse set of
associated BPs that underlie the nkd mutant phenotype.

Differential roles of NKD1 and NKD2 in regulation of key biological processes and
functions of endosperm
200

To understand the roles of NKD1 and NKD2 in the regulation of biological processes/functions
previously associated with the nkd mutant phenotype (Yi et al., 2015), we systematically
examined the differential expression patterns of all the NKD-regulated genes found to be
associated with cell cycle, storage program, carotenoid biosynthesis/metabolism, anthocyanin
biosynthesis, ABA biosynthesis/signaling, and pathogen defense. Analysis of the differential
expression patterns of 74 genes associated with cell cycle control showed that most of the genes
were upregulated in the mutants at one or more stages, and significant upregulated patterns were
detected predominantly in the nkd1-Ds and/or the double mutant at 12 and 16 DAP, whereas the
number of genes that were significantly up- or down-regulated in the nkd2-Ds mutant was much
smaller (Supplemental Figure 4.5). Moreover, many of the genes were found to be dysregulated
only in the double mutant (primarily at 12 and/or 16 DAP); for instance, the Retinoblastoma-
related3 gene, which has been implicated in regulation of mitotic cell cycle in maize endosperm
(Sabelli et al., 2005), was upregulated only in the double mutant at 12 DAP. These data indicate
that the multi-layered peripheral endosperm cells of the nkd double mutant and the nkd1 single
mutant are likely produced as a result of loss of repression of the cell cycle control genes, and
that NKD1 and NKD2 regulate some of those genes in a redundant manner.

Among the known storage-protein genes and the related Opaque/Floury genes, we detected 30
(out of 55 total) NKD-regulated genes, of which most were dysregulated primarily or exclusively
in nkd1-Ds and the double mutant (Supplemental Figure 4.6). Interestingly, many of the genes
that were significantly downregulated in nkd1-Ds mutant at 8 DAP were upregulated moderately
in the 8-DAP nkd2-Ds endosperm (mean log2FC = 1.01), suggesting that they are activated by
NKD1 but repressed by NKD2 in the WT. However, a close examination of the carbohydrate
biosynthesis and degradation pathways for NKD-regulated genes revealed that at least 14 genes
involved in the superpathway of sucrose degradation or starch biosynthesis were dysregulated
exclusively in the nkd1-Ds and the double mutant at one or more stages (Supplemental Figure
4.7 and Supplemental Data Set 4.7), including 10 that were downregulated and four upregulated.
Therefore, NKD1 likely plays an important role in regulation of sucrose metabolism and starch
accumulation in the endosperm.

Similarly, most of the NKD-regulated genes involved in the carotenoid biosynthesis


superpathway were downregulated in the analyzed mutants, and the downregulation was detected
201

primarily in nkd1-Ds and the double mutant (Supplemental Figure 4.8 and Supplemental Data
Set 4.7), suggesting that NKD1 plays the primary role in the regulation of these genes. On the
other hand, three of the carotenoid-biosynthesis genes (GRMZM2G152135, GRMZM2G010221,
and GRMZM2G164318) were downregulated in all three mutants at one or more stages,
suggesting that, in addition to NKD1, NKD2 is also required for the proper regulation of a subset
of these genes. This was in contrast to most of the NKD-regulated, anthocyanin biosynthesis-
related genes, including the R1 gene, which were upregulated in both the single mutants and the
double mutant (Supplemental Figure 4.9 and Supplemental Data Set 4.7), indicating that both
NKD proteins are required for repressing these genes in the WT.

Interestingly, VP1, which is involved in regulation of both anthocyanin biosynthesis and ABA
signaling (McCarty et al., 1989; McCarty et al., 1991; Hoecker et al., 1995) was significantly
downregulated in the nkd1-Ds mutant at 12 DAP. This suggests that the NKD proteins regulate
R1 and VP1 in opposite manners to regulate coordinately the anthocyanin biosynthesis pathway.
Furthermore, the differential expression patterns of several of the anthocyanin-biosynthesis
pathway genes change temporally. For example, the GRMZM2G036409 gene was significantly
downregulated in all three mutants at 8 DAP, but was upregulated in in the nkd1-Ds and double
mutants at 12 and 16 DAP (Supplemental Figure 4.9). Similarly, the differential expression
patterns of ANTHOCYANINLESS2 and GRMZM2G155911 also changed temporally from
significant downregulation to upregulation in the double mutant (from 8 DAP to 12 and 16 DAP,
and from 12 DAP to 16 DAP, respectively). These data suggest that the regulatory role of NKDs
switch from activators to repressors in some instances as endosperm undergoes development.

In addition to VP1, we detected significant dysregulation of three genes involved in ABA


biosynthesis (1 downregulated and 2 upregulated), three genes involved in ABA degradation (all
upregulated), with two of them also dysregulated in the nkd2-Ds mutant (Supplemental Figure
4.10 and Supplemental Data Set 4.7). Therefore, NKDs may coordinately regulate the ABA
biosynthesis, degradation, and signaling processes to repress precocious germination in the WT.
At least 13 pathogen-defense-related genes, including two putative defensin genes, were also
significantly dysregulated in the three mutants (Supplemental Figure 4.11), suggesting that both
NKD proteins are critical for the regulation of pathogen defense in the endosperm.
202

An overall pattern that emerged from the analysis of DEGs was that most of the mutant
phenotype-associated genes were dysregulated more strongly at 12 and 16 DAP, but to a lesser
extent at 8 DAP (Supplemental Figures 4.5 to 4.11). Together, the data support the notion that
the kernel phenotypes of the double Ds mutants of NKD1 and NKD2 are primarily caused by the
loss of function of the NKD1 gene most crucially at 12 and 16 DAP, but also help distinguish the
unique role of NKD2 in proper regulation of endosperm development.

NKD-regulated genes are enriched for previously identified O2-regulated genes

Our previous analysis of the Opaque-2 gene regulatory network showed that O2 and NKD2
directly co-regulate target gene expression, and that many direct or indirect target genes of O2
are putative direct targets of NKD2 (Chapter 3). In line with these observations, comparisons
between the previously identified O2-modulated (i.e., O2-regulated) genes and the NKD-
regulated genes identified in this study showed that 765 of the 1,863 O2-regulated genes (41.1%)
were also regulated by NKD1 or NKD2 (Figure 4.4A). Furthermore, a comparison between the
O2-modulated and/or bound gene sets with the NKD-regulated gene sets showed that nearly all
the NKD-regulated gene sets were significantly enriched for indirect O2 target genes; however,
only the gene sets that were downregulated in the nkd1-Ds mutant and the double mutant were
significantly enriched for genes that were directly activated by O2 (P < 10-5, hypergeometric
test) (Figure 4.4B). A close inspection of all 186 direct O2 target genes showed that 75 of the
genes (40.3%) were also regulated by NKD1 and NKD2, including many zein genes that are
canonical target genes of O2 (Schmidt et al., 1990; Schmidt et al., 1992; Li et al., 2015)
(Supplemental Data Set 4.8). These data indicate that O2 likely co-regulates many downstream
genes with both NKD1 and NKD2, and that NKD1 plays the primary role in co-activation of the
direct O2 target genes along with O2.

Prediction of putative direct target genes of NKD1 and NKD2 and their relationship to the
O2 network genes

To begin to identify the genes directly regulated by the NKD TFs, we analyzed the promoter
sequences of all DEGs in the nkd1-Ds and nkd2-Ds single mutants for presence of previously
203

reported NKD1 and NKD2 binding sequences identified through SAAB assays (Gontarek et al.,
2016). The analysis identified 2,872 and 2,500 putative targets of NKD1 and NKD2,
respectively, including 649 genes as putative common targets of both TFs (Figure 4.5A). GO
term enrichment analysis showed that the putative direct targets of NKD1 alone were most
significantly enriched for “organonitrogen compound biosynthetic process” (Supplemental Data
Set 4.9). Moreover, among the putative direct targets of NKD1 and/or NKD2 (i.e., NKD1 alone,
NKD2 alone, or both) were many genes associated with the nkd mutant phenotype. These
included O2 as a predicted direct target of NKD1; the SCARECROW-LIKE1 (SCL1) gene,
suggested as a regulator of AL versus SE cell fate specification in early endosperm development
(Gontarek et al., 2016; Gontarek and Becraft, 2017), as a direct target of NKD2; and Floury-3
and R1, TF genes linked with the opaque/floury endosperm and anthocyanin deficiency of the
nkd mutants, respectively (Selinger and Chandler, 1999; Li et al., 2017), as direct targets of both
NKD1 and NKD2 (Supplemental Data Set 4.9). Analysis of overlaps between the putative NKD
target gene sets with the O2-modulated and/or bound gene sets (Chapter 3) showed that the three
gene sets, which are likely regulated directly by individual or both NKD TFs, were significantly
enriched for indirect O2 targets (both the O2-activated and repressed genes), whereas only the
putative direct targets of NKD1 significantly overlapped with O2-activated direct targets
(hypergeometric P < 10-5) (Figure 4.5B). These data support a role for NKD1 in regulating the
O2-regulated genes both as a cofactor of O2 and as a downstream-acting TF of the O2 network,
whereas NKD2 seems to predominantly function in the latter context only.

Gene coexpression network analysis of the NKD-regulated genes

In order to further understand the nature of the gene networks under NKD regulation, we first
identified co-regulated gene sets and the corresponding TFs among the NKD-regulated genes by
performing a weighted gene coexpression network analysis (WGCNA) using the RPKM values
of all the DEGs, and looked for any enrichment of TF families and cis-motifs in the coexpression
modules. The analysis assigned 11,544 of the 11,554 NKD-regulated genes into 12 statistically
robust coexpression modules, with NKD1 and NKD2 assigned to M7 and M5, respectively
(Figure 4.6A; Supplemental Table 4.4 and Supplemental Data Set 4.10). Each of these modules
contained 134 to 2,332 genes (Figure 4.7B). Interestingly, a comparison of the WGCNA-
204

identified modules with all the DEG sets collapsed the 12 coexpression modules into two distinct
classes based on up- or downregulation patterns in the mutants. The first class included modules
1, 5, 7, 9, 10, and 11 that were predominantly enriched for genes downregulated in the mutants
(i.e., NKD-activated modules), whereas the second class included modules 2, 3, 4, 6, 8, and 12
that were predominantly enriched for genes upregulated in the mutants (i.e., NKD-repressed
modules) (hypergeometric P < 10-5) (Supplemental Figure 4.12). GO term enrichment analysis
showed that each of the 12 coexpression modules except M3 was associated in main with a
unique set of related BPs: the NKD-activated modules were enriched for BPs related to RNA
processing (M1), transposition (M5), primary metabolism (M7), starch biosynthesis (M9),
transport (M10), and development (M11), whereas the NKD-repressed modules were enriched
for BPs associated with abiotic stress response (M2), superoxide metabolism (M4), hormone
response (M6), cell cycle process (M8), and stimulus response (M12) (Figure 4.6B;
Supplemental Data Set 4.11).

Each module contained from two (M3) to 136 (M7) TF genes (except for M4 with no TF genes;
Supplemental Data Set 4.10). O2 and VP1 were assigned to M2 and M1, respectively
(Supplemental Data Set 4.10). Both these TF genes were shown to be regulated directly by
NKD1 and NKD2 (Gontarek et al., 2016), suggesting that M2 and M1 modules contain genes
that are in turn regulated by O2 and VP1, respectively. In support of this hypothesis, M2
contained 10 genes determined previously as direct targets and 167 determined as indirect targets
of O2 (Supplemental Figure 4.13) (Chapter 3). A TF family enrichment analysis showed that
modules 6 and 11 were enriched for ARF family and MIKC-type MADS-box family TF genes,
respectively (P < 10-4) (Supplemental Figure 4.14). Many TFs in the ARF and the MIKC-type
MADS-box families have previously been shown to be involved in regulation of auxin
signaling/response and floral organ identity, respectively (Causier et al., 2010; Roosjen et al.,
2018). These TF family enrichments were consistent with GO term enrichments of the two
modules, showing that M6 and M11 were significantly enriched for BPs associated with auxin
signaling/response and floral organ development, respectively (Supplemental Data Set 4.11). In
addition, a close inspection of genes associated with the enriched GO terms showed that M6
contained many non-ARF genes associated with auxin biosynthesis/signaling/response-related
GO terms, including an indole-3-acetic acid amido synthetase gene (GRMZM2G378106), a
205

putative auxin efflux carrier gene (PIN1), and several Aux/IAA family genes that putatively
function in auxin signaling (Supplemental Data Set 4.11 and Supplemental Table 4.5).

We carried out an enrichment analysis of the reported binding sites of NKD1 and NKD2 in the
promoter regions of genes in each coexpression module and found that modules 1, 2, 6, 8, 9, and
11 were significantly enriched for NKD1 binding sites, and modules 1, 2, 4, 5, 6, 7, 8, and 12
were enriched for NKD2 binding sites (adjusted P < 0.05) (Supplemental Table 4.6), indicating
that a large fraction of the genes in each of these modules are likely direct targets of the NKDs.
Moreover, a de novo motif enrichment analysis and comparison to reported plant cis-motifs
showed that each coexpression module was significantly enriched for one to several cis-motifs
(E < 0.05) that were similar to known TF binding sites (q < 0.05) (Supplemental Table 4.7). For
example, modules 2, 7, 8, and 10 were significantly enriched for motifs similar to reported
binding sites of DOF family TFs [e.g., the Arabidopsis (Arabidopsis thaliana) OBF-binding
protein 3 (OBP3) (O'Malley et al., 2016)], and, correspondingly, these modules contained at least
one TF of the maize DOF family. These observations suggest that the DOF TF(s) may directly
regulate genes in the same coexpression module through binding to the enriched DOF-like cis-
motifs. Therefore, our coexpression network analysis showed that the NKD proteins coordinately
regulate downstream BPs, including storage compound synthesis and hormone response, through
direct or indirect regulation of the coexpressed gene sets, and that many NKD-regulated
downstream TF genes are likely involved in regulation of these coordinately expressed gene sets.

4.4 DISCUSSION

We identified gene sets likely regulated by individual NKD1 and NKD2 proteins through a
transcriptome profiling of 8-, 12-, and 16-DAP endosperm of the single Ds mutants of NKD1 and
NKD2 as well as a double Ds mutant combination. The RNA-Seq analysis and the subsequent
RT-qPCR assays detected extremely low mRNA levels of NKD1 and NKD2 in their
corresponding single mutant and in the double mutant (Supplemental Figure 4.3). The results are
consistent with the previous report, which showed that the Ds transposon insertion mutants of
NKD1 and NKD2 use in this study represent strong loss-of-function (perhaps null) alleles (Yi et
al., 2015). Consistent with the previous observation that the nkd1-Ds mutant resembled the
kernel phenotype of a double Ds mutant, whereas the nkd2-Ds mutant did not show any
206

detectable phenotype (Yi et al., 2015), comparison of the endosperm transcriptome profiles
revealed high similarity in dysregulated gene sets between the nkd1-Ds mutant and the double
mutant. Nonetheless, the nkd2-Ds mutant exhibited substantial differences from the WT,
indicating that NKD2 plays an important role in regulation of endosperm gene expression.

Our differential gene expression analysis detected 11,554 genes that were dysregulated in at least
one mutant at one or more stages. At each stage, the NKD1-regulated gene set strongly
overlapped with the NKD2-regulated gene set (Figures 4.2G to 4.2J), suggesting that NKD1 and
NKD2 regulate overlapping gene networks. Correspondingly, GO term enrichment analysis of
all the NKD-regulated gene sets revealed that NKD1 and NKD2 are commonly involved in the
regulation of a wide range of BPs, including regulation of gene expression (Supplemental Data
Set 4.3). On the other hand, the analysis also identified gene sets that were likely regulated by
NKD1 or NKD2 individually (Figures 4.2G to 4.2J), with the former enriched for BPs associated
with rRNA processing and translation (Supplemental Data Set 4.5). These observations support
distinct roles for NKD1 and NKD2 in regulation of endosperm gene expression. Moreover, many
downstream genes were dysregulated only in the double mutant, and were significantly enriched
for BPs related to DNA replication (Figure 4.2J; Supplemental Data Set 4.5), suggesting that
NKD1 and NKD2 regulate these genes and the associated BPs in a redundant fashion.

Further analysis of genes that are co-regulated by the NKDs revealed that in most cases NKD1
and NKD2 function cooperatively (Figure 4.3). Consistent with this, a previous study has shown
that NKD1 and NKD2 can form heterodimers to co-regulate downstream genes (Gontarek et al.,
2016). Therefore, it is likely that a subset of the genes, possibly including the anthocyanin-
biosynthesis regulator R1 gene, that are cooperatively regulated by NKD1 and NKD2 are
regulated by the NKD1-NKD2 heterodimers, either as activators or as repressors. Moreover, as
exemplified by several genes involved in the anthocyanin biosynthesis pathway
(GRMZM2G036409, ANTHOCYANINLESS2, and GRMZM2G155911; Supplemental Figure
4.9), the regulatory function of the combination of NKD1 and NKD2 can also change temporally
from activators to repressors as the endosperm develops. Therefore, the dual (and temporally
changing) roles of the NKD proteins as both activators and repressors likely constitute a
molecular basis for fine co-regulation of the downstream gene networks by NKD1 and NKD2.
207

Our previous study of the O2-regulated gene network has suggested that O2, NKDs, and VP1 co-
regulate a downstream gene network (Chapter 3). The analysis here showed that NKD2 was
significantly downregulated in the nkd1-Ds mutant at all three stages, O2 was downregulated in
nkd1-Ds at 8 DAP, and VP1 was downregulated in nkd1-Ds at 12 DAP. Consistent with these
observations, previous studies have shown that NKD1 and NKD2 bind the promoter regions of
O2 and VP1, and that VP1 is indirectly activated by O2 (Gontarek et al., 2016) (Chapter 3).
These data together suggest a central role for NKD1 in the gene network regulated by the NKDs,
O2, and VP1, together regulating cell proliferation, differentiation, synthesis of storage
products/anthocyanin/carotenoids, maturation, and pathogen defense of the maize endosperm. In
support of this model, enrichment analysis of putative direct NKD target gene sets among the
O2-modulated and/or bound gene sets suggest that NKD1 participates in regulation of the O2
gene network through direct interaction with a portion of the O2 target genes. This is in contrast
to NKD2 which likely acts on the O2 network genes that are further removed from the direct
regulation of O2 and regulated by downstream TFs (Figure 4.4B).

Our gene coexpression network analysis of the NKD-regulated genes led to the identification of
gene modules that are coordinately regulated through the activity of NKD TFs alone or in
conjunction with other TFs (Figure 4.6). Based on enrichment analysis of reported NKD-binding
sites, each module likely contains many direct target genes of NKD1 and/or NKD2, including
many coexpressed TF genes, which are associated with key BPs relevant to the nkd phenotype
specifically and endosperm development in general. These would include cell cycle processes
(M8), starch biosynthesis (M9), hormone responses (M6), and abiotic stress responses (M2)
(Supplemental Data Sets 4.9 to 4.11). However, whether the network-associated TF genes are
directly regulated by the NKDs remains to be determined. Future understanding the NKD-
regulated gene network will likely require further identification of direct and indirect target
genes of the NKD TFs, and analysis of the downstream TFs and their respective target genes,
using genome-wide approaches such as RNA-Seq and chromatin immunoprecipitation followed
by high-throughput sequencing (ChIP-Seq).

Analyses of NKD-regulated genes associated with the nkd mutant phenotype revealed that they
generally display a dysregulated pattern in nkd1-Ds and the double mutant (Supplemental
Figures 4.5 to 4.11). This suggests that the mutant phenotype can be attributed primarily to the
208

dysregulation of these genes and is likely due to the loss of NKD1 function. Nonetheless, a
subset of these genes were also significantly dysregulated in the nkd2-Ds mutant, suggesting that
NKD2 also plays a role in their proper regulation. However, a detailed cytological analysis of
endosperm cell types for the nkd2-Ds mutant is still lacking. The stage-biased patterns of gene
dysregulation observed in our study (Figure 4.2A; Supplemental Figures 4.5 to 4.11) suggest that
the NKD genes are more crucially required for the normal regulation of endosperm gene
expression programs at the later stages (12 and 16 DAP) than at the earlier stage (8 DAP) of
endosperm development. Moreover, the previously reported analysis of the nkd-R mutant had
shown that NKDs regulated distinct sets of genes in AL and SE (Gontarek et al., 2016).
Therefore, a time-course ultrastructural analysis coupled with a spatiotemporal transcriptome
analysis of the endosperm cell types (particularly AL and SE) of nkd1-Ds, nkd2-Ds, and the
double mutant would be necessary to describe fully the manifestation of the nkd mutant
phenotype and the associated gene expression programs regulated by the NKD proteins.

4.5 METHODS

Plant Materials and Growth

All plants were grown in the field at Iowa State University and self-pollinated or sibling-crossed
to generate staged endosperm tissues, which were collected in August 2016. Two to four batches
of at least two endosperms were collected from a single ear and each batch was handled as one
biological replicate (Supplemental Table 4.1).

RNA-Seq and data analysis

At least 6 µg total RNA was isolated from each of the 45 biological samples as described
previously (Li et al., 2014) and were quality checked using Agilent RNA ScreenTape (Agilent
Technologies). About 100 to 500 ng total RNA of each sample was used by the High Throughput
Genomics Core Facility, University of Utah to construct multiplexed sequencing libraries using
the Illumina TruSeq Stranded mRNA Library Preparation Kit with poly(A) selection. The
resulting libraries were sequenced using the HiSeq 125 Cycle Paired-End Sequencing v4
protocol on 5 flow cell lanes (Supplemental Table 4.1) of the Illumina HiSequation 2500
platform at the same facility.
209

Raw reads were quality-checked using FastQC (Andrews, 2010) and mapped to the maize
reference genome (B73 RefGen_v3) using Tophat v2.1.1 (Trapnell et al., 2009) essentially as
described before (Zhan et al., 2015), except that the maximum number of mismatches per read
was set to 4. Reads mapped to each gene were counted using featureCounts (Liao et al., 2014),
and normalization and differential gene expression analysis were carried out using edgeR v3.18.1
(Robinson et al., 2010); the genes with raw counts per million > 0.65 in at least 2 of the 45
sequenced samples were tested for differential expression using a generalized linear model-based
method. Fold change > 1.5 and FDR < 0.05 were used as the cutoff criteria to call DEGs. Gene
coexpression network analysis was performed using WGCNA v1.61 (Langfelder and Horvath,
2008) on the log2-transformed RPKM values [i.e., log2(RPKM+1)] of all the DEGs using a
previously described method (Zhan et al., 2015) with the following modifications: first, the
matrix of pairwise SCCs between genes was transformed into the connection strengths matrix by
raising the correlation matrix to the power (β) of 22; second, after the hierarchical clustering tree
was cut, modules with fewer than 100 genes were merged into their closest neighboring module
to generate the final modules.

Functional annotations of maize genes were obtained from Ensembl Plants (plants.ensembl.org),
Gramene (gramene.org), MaizeGDB (maizegdb.org), PlantTFDB v3.0
(http://planttfdb.cbi.pku.edu.cn/) and GRASSIUS (http://grassius.org/) (Yilmaz et al., 2009; Jin
et al., 2014). GO term enrichment analyses were performed using Blast2GO v3.0.11 (Gotz et al.,
2008), with only the most specific GO terms with FDR < 0.05 reported. DEGs encoding
enzymes involved in the pathways of carbohydrate biosynthesis and degradation, carotenoid
biosynthesis, anthocyanin biosynthesis, and ABA biosynthesis were identified using CornCyc
v8.0 on Plant Metabolic Network (plantcyc.org) (Walsh et al., 2016). Motif enrichment analyses
and comparison with known plant cis-motifs available from the JASPAR 2018 Plantae non-
redundant motif database were performed using the AME, MEME, and Tomtom of the MEME
Suite (Bailey et al., 2009). The promoter sequences subject to all cis-motif analyses were 0.5 kb
upstream the annotated transcription start sites.

RT-qPCR

Quantitative PCR assays were performed as described previously (Li et al., 2014)(Chapter 3)
using the primers listed in Supplemental Table 4.8. The raw threshold cycle (Ct) values were
210

normalized against the Ct values of the Thioredoxin gene, which showed relatively stable
expression levels across all the samples based on RNA-Seq (Supplemental Data Set 4.1).
Normalized Ct values were cut off at 36. Log2FC was determined as described in Chapter 3.
211

Figure 4.1. Relationship between endosperm mRNA populations of all analyzed mutants
and the WT.

(A) PCA of mRNA populations from the 45 individual biological replicates of samples. PCs
were calculated based on log2-transformed RPKM data of the 34,348 genes tested for differential
212

expression. Principal components one and two (PC1 and PC2) collectively explains 69.66% of
the variance in the mRNA populations. Error bars indicate standard deviation. (B) SCC analysis
of biological samples based on log2-transformed mean RPKM values of biological replicates.
The clustering dendrogram was inferred by applying (1 - SCC) as the distance function.
213

Figure 4.2. Differential gene expression analysis.

(A) Number of DEGs detected in each mutant versus the WT of the same stage. (B) A chord
diagram showing numbers of overlapping genes among all 9 DEG sets. (C-E) Venn diagrams of
DEG sets identified in each genotype versus WT. (F) Venn diagram of united DEG sets of each
developmental stage. (G-I) Venn diagrams of DEG sets identified at each developmental stage.
(J) Venn diagram of united DEG sets in each mutant versus WT. In (C) through (J), number of
genes in each gene set are shown in parentheses.
214

Figure 4.3. Analysis of gene sets cooperatively co-regulated by NKD1 and NKD2.

(A and B) Venn diagrams of genes that are co-activated by NKD1 and NKD2 (A) and co-
repressed by NKD1 and NKD2 (B). (C and D) Log2FC heat maps of genes that are co-regulated
by NKD1 and NKD2 in a given manner at all three stages.
215

Figure 4.4. Comparison of the NKD-regulated gene sets with the previously identified O2-
modulated and/or bound gene sets.

(A) Venn diagrams of number of DEGs in the endosperm of the mutants versus WT (union of
three stages) and number of DEGs in the o2 mutant endosperm versus WT (15 DAP) (Chapter
3). (B) Relationships of the NKD-regulated gene sets with the O2-modulated and/or bound gene
sets. The heat map indicates P values (-log10) of hypergeometric tests of overrepresentation of
genes in a given pair of gene sets. Boxes contain the numbers of overlapping genes. The O2-
modulated and/or bound gene sets and number of genes in each set are noted on the x axis, and
the NKD-regulated gene sets and number of genes in each set are noted on the y axis.
216

Figure 4.5. Putative direct target genes of NKD1 and NKD2.

(A) Venn diagram of putative target genes of NKD1 and NKD2. (B) Relationships of the
putative direct NKD target gene sets with the O2-modulated and/or bound gene sets (Chapter 3).
The heat map indicates P values (-log10) of hypergeometric tests of overrepresentation of genes
in a given pair of gene sets. Boxes contain the numbers of overlapping genes. The putative direct
NKD target gene sets and number of genes in each set are noted on the x axis, and the O2-
modulated and/or bound gene sets and number of genes in each set are noted on the y axis.
217

Figure 4.6. Coexpression networks identified using WGCNA.


218

(A) Clustering of genes on topological overlap-based dissimilarity with module colors and a
log2FC heat map clustered accordingly. (B) Module eigengenes (MEs, defined as the first
principal component of expression levels of module members) and enriched BPs for the 12
coexpression modules. The number of genes assigned to each module is shown in parentheses.
219

Table 4.1. Abbreviations for genotypes and developmental stages.

8 DAP 12 DAP 16 DAP


Wild type (W22) w.8 w.12 w.16
nkd1-Ds/nkd1-Ds; +/+ n1.8 n1.12 n1.16
+/+; nkd2-Ds0766/nkd2-Ds0766 n2.8 n2.12 n2.16
nkd1-Ds/nkd1-Ds; nkd2-Ds0766/nkd2-Ds0766 d.8 d.12 d.16
220

4.6 SUPPLEMENTAL DATA

Supplemental Figure 4.1. SCC heat map of biological samples based on log2-transformed
RPKM values of all the 34,348 genes tested for differential expression.

The clustering dendrogram was inferred by applying (1 - SCC) as distance. The SCC between
biological replicates of were all ≥ 0.96. Biological replicates are indicated by -1, -2, -3, and -4.
221

Supplemental Figure 4.2. Scatter plots of log2FC versus average log2CPM of all the genes
tested for differential expression.

DEGs are represented by red dots, and the other genes black. Horizontal blue lines indicate 1.5-
fold changes (i.e., |log2FC| = 0.59).
222

Supplemental Figure 4.3. RT-qPCR validation of DEGs.

The mRNA levels of NKD1, NKD2, O2, and VP1 based on RNA-Seq (Left) and RT-qPCR
(Right) are shown in comparison to a non-differentially expressed 22-kD α-zein gene (azs22-4).
Data are log2RPKM (Left) and normalized Ct values (mean ± SEM; n = 2 to 4) (Right),
respectively.
223

Supplemental Figure 4.4. Analysis of gene sets antagonistically co-regulated by NKD1 and
NKD2.

Venn diagrams of genes that are activated by NKD1 and repressed by NKD2 (A), and repressed
by NKD1 and activated by NKD2 (B).
224

Supplemental Figure 4.5. Log2FC heat map of NKD-regulated cell cycle control-related
genes. Asterisks indicate significant differential expression.
225

Supplemental Figure 4.6. Log2FC heat map of NKD-regulated storage-protein genes and
Opaque/Floury genes. Asterisks indicate significant differential expression.
226

Supplemental Figure 4.7. Log2FC heat map of NKD-regulated carbohydrate biosynthesis


or degradation-related genes. Asterisks indicate significant differential expression.
227

Supplemental Figure 4.8. Log2FC heat map of NKD-regulated carotenoid biosynthesis-


related genes. Asterisks indicate significant differential expression.
228

Supplemental Figure 4.9. Log2FC heat map of NKD-regulated anthocyanin biosynthesis-


related genes. Asterisks indicate significant differential expression.
229

Supplemental Figure 4.10. Log2FC heat map of NKD-regulated ABA


biosynthesis/degradation/signaling-related genes. Asterisks indicate significant differential
expression.
230

Supplemental Figure 4.11. Log2FC heat map of NKD-regulated pathogen defense-related


genes. Asterisks indicate significant differential expression.
231

Supplemental Figure 4.12. Relationships of the WGCNA-identified coexpression modules


with the DEG sets.

The heat map indicates P values (-log10) of hypergeometric tests of overrepresentation of genes
in a given pair of gene sets. Boxes contain the numbers of overlapping genes. The coexpression
modules and number of genes in each module are noted on the x axis, and the DEG sets and
number of genes in each set are noted on the y axis.
232

Supplemental Figure 4.13. Relationships of the coexpression modules with the O2-
modulated and/or bound gene sets.

The heat map indicates P values (-log10) of hypergeometric tests of overrepresentation of genes
in a given pair of gene sets. Boxes contain the numbers of overlapping genes. The coexpression
modules and number of genes in each module are noted on the x axis, and the O2-modulated
and/or bound gene sets and number of genes in each set are noted on the y axis.
233

Supplemental Figure 4.14. Enrichments of TF families among the coexpression modules.

The heat map indicates P values (-log10) of hypergeometric tests of overrepresentation of TF


families in the coexpression modules. Boxes contain the numbers of overlapping genes. The
coexpression modules and number of genes in each module are noted on the x axis, and the TF
families and number of annotated family numbers are noted on the y axis.
234

Supplemental Table 4.1. The origins of RNAs and experimental design for RNA-Seq.
HiSeq 2500 HiSeq 2500
Sample Origin Sample Origin
lane # lane #
w.8-1 Batch 1 from Ear 1 2 n2.8-2 Batch 2 from Ear 11 2
w.8-2 Batch 2 from Ear 1 3 n2.8-3 Batch 1 from Ear 12 3
w.8-3 Batch 1 from Ear 2 4 n2.8-4 Batch 2 from Ear 12 5
w.8-4 Batch 2 from Ear 2 5 n2.12-1 Batch 1 from Ear 13 1
w.12-1 Batch 1 from Ear 3 2 n2.12-2 Batch 2 from Ear 13 2
w.12-2 Batch 2 from Ear 3 3 n2.12-3 Batch 3 from Ear 13 4
w.12-3 Batch 1 from Ear 4 4 n2.16-1 Batch 1 from Ear 14 1
w.12-4 Batch 2 from Ear 4 5 n2.16-2 Batch 2 from Ear 14 2
w.16-1 Batch 1 from Ear 5 2 n2.16-3 Batch 1 from Ear 15 4
w.16-2 Batch 2 from Ear 5 3 n2.16-4 Batch 2 from Ear 15 5
w.16-3 Batch 1 from Ear 6 4 d.8-1 Batch 1 from Ear 16 1
w.16-4 Batch 2 from Ear 6 5 d.8-2 Batch 2 from Ear 16 3
n1.8-1 Batch 1 from Ear 7 1 d.8-3 Batch 1 from Ear 17 4
n1.8-2 Batch 2 from Ear 7 2 d.8-4 Batch 2 from Ear 17 5
n1.12-1 Batch 1 from Ear 8 1 d.12-1 Batch 1 from Ear 18 1
n1.12-2 Batch 2 from Ear 8 2 d.12-2 Batch 2 from Ear 18 3
n1.12-3 Batch 1 from Ear 9 3 d.12-3 Batch 1 from Ear 19 4
n1.12-4 Batch 2 from Ear 9 4 d.12-4 Batch 2 from Ear 19 5
n1.16-1 Batch 1 from Ear 10 1 d.16-1 Batch 1 from Ear 20 1
n1.16-2 Batch 2 from Ear 10 2 d.16-2 Batch 2 from Ear 20 3
n1.16-3 Batch 3 from Ear 10 3 d.16-3 Batch 1 from Ear 21 4
n1.16-4 Batch 4 from Ear 10 5 d.16-4 Batch 2 from Ear 21 5
n2.8-1 Batch 1 from Ear 11 1
235

Supplemental Table 4.2. Summary statistics of RNA-Seq reads and mapping.


Sample Total # reads Unique-map Multi-map Unmapped
w.8-1 67867886 44562568 (65.66%) 13007069 (19.17%) 10298249 (15.17%)
w.8-2 55339944 36967972 (66.80%) 10852401 (19.61%) 7519571 (13.59%)
w.8-3 56633648 36790486 (64.96%) 10728916 (18.94%) 9114246 (16.09%)
w.8-4 54917182 35722935 (65.05%) 10613702 (19.33%) 8580545 (15.62%)
n1.8-1 58607914 36332101 (61.99%) 13830060 (23.60%) 8445753 (14.41%)
n1.8-2 57120078 37257292 (65.23%) 10596363 (18.55%) 9266423 (16.22%)
n2.8-1 54570050 36263356 (66.45%) 10393178 (19.05%) 7913516 (14.50%)
n2.8-2 64481678 42705378 (66.23%) 12741084 (19.76%) 9035216 (14.01%)
n2.8-3 65501686 43218789 (65.98%) 12775144 (19.50%) 9507753 (14.52%)
n2.8-4 57727564 37508569 (64.98%) 11152005 (19.32%) 9066990 (15.71%)
d.8-1 59720402 38902459 (65.14%) 11454982 (19.18%) 9362961 (15.68%)
d.8-2 53794606 35737133 (66.43%) 10331433 (19.21%) 7726040 (14.36%)
d.8-3 56103760 36281530 (64.67%) 10302137 (18.36%) 9520093 (16.97%)
d.8-4 67014954 44782939 (66.83%) 12738260 (19.01%) 9493755 (14.17%)
w.12-1 60718944 35913141 (59.15%) 15089849 (24.85%) 9715954 (16.00%)
w.12-2 52077302 31588054 (60.66%) 12940566 (24.85%) 7548682 (14.50%)
w.12-3 51614338 32079744 (62.15%) 11560711 (22.40%) 7973883 (15.45%)
w.12-4 65032982 40130611 (61.71%) 14887378 (22.89%) 10014993 (15.40%)
n1.12-1 61772074 38892368 (62.96%) 13683655 (22.15%) 9196051 (14.89%)
n1.12-2 61232106 38413509 (62.73%) 13381596 (21.85%) 9437001 (15.41%)
n1.12-3 72728112 44885581 (61.72%) 16460963 (22.63%) 11381568 (15.65%)
n1.12-4 54427108 33271701 (61.13%) 11304604 (20.77%) 9850803 (18.10%)
d.12-1 58849710 37880347 (64.37%) 11972578 (20.34%) 8996785 (15.29%)
d.12-2 68927682 44201315 (64.13%) 14825568 (21.51%) 9900799 (14.36%)
d.12-3 49937390 30890031 (61.86%) 10656937 (21.34%) 8390422 (16.80%)
d.12-4 66595980 42576175 (63.93%) 14435623 (21.68%) 9584182 (14.39%)
w.16-1 54937070 30518594 (55.55%) 14283809 (26.00%) 10134667 (18.45%)
w.16-2 62791026 34632218 (55.15%) 16545782 (26.35%) 11613026 (18.49%)
w.16-3 57955334 30621132 (52.84%) 17001219 (29.34%) 10332983 (17.83%)
w.16-4 65320404 33845799 (51.82%) 19250604 (29.47%) 12224001 (18.71%)
n1.16-1 57328216 34955180 (60.97%) 12179016 (21.24%) 10194020 (17.78%)
n1.16-2 62612500 38093136 (60.84%) 13895761 (22.19%) 10623603 (16.97%)
n1.16-3 62562394 38925377 (62.22%) 13707830 (21.91%) 9929187 (15.87%)
n1.16-4 59742256 36506331 (61.11%) 13023668 (21.80%) 10212257 (17.09%)
n2.12-1 53602418 33040732 (61.64%) 12346402 (23.03%) 8215284 (15.33%)
n2.12-2 52637368 32263107 (61.29%) 12693311 (24.11%) 7680950 (14.59%)
n2.12-3 58402708 34413360 (58.92%) 13680902 (23.43%) 10308446 (17.65%)
n2.16-1 62358820 34465909 (55.27%) 17051558 (27.34%) 10841353 (17.39%)
n2.16-2 53329532 30377234 (56.96%) 13955664 (26.17%) 8996634 (16.87%)
236

Supplemental Table 4.2 Continued


n2.16-3 53123074 30498980 (57.41%) 12985491 (24.44%) 9638603 (18.14%)
n2.16-4 54875560 32313066 (58.88%) 13006476 (23.70%) 9556018 (17.41%)
d.16-1 56383106 33736811 (59.83%) 12848787 (22.79%) 9797508 (17.38%)
d.16-2 61421858 37082891 (60.37%) 14389956 (23.43%) 9949011 (16.20%)
d.16-3 55451710 31232094 (56.32%) 14243046 (25.69%) 9976570 (17.99%)
d.16-4 52938182 31511712 (59.53%) 12739317 (24.06%) 8687153 (16.41%)
237

Supplemental Table 4.3. Normalized Ct values for genes analyzed using RT-qPCR.
Gene NKD1 NKD2 O2 VP1 azs22-4
mean Ct w.8 28.01 26.69 23.18 27.43 27.26
w.12 28.06 26.87 20.77 27.43 15.81
w.16 27.95 27.77 20.56 29.41 13.99
n1.8 36.00 36.00 24.08 28.40 30.59
n1.12 36.00 36.00 19.46 28.44 14.71
n1.16 36.00 36.00 19.44 28.76 14.46
n2.8 28.24 36.00 22.50 26.86 24.93
n2.12 28.19 36.00 20.75 27.56 15.14
n2.16 28.32 36.00 19.99 28.70 14.27
d.8 36.00 36.00 24.36 30.22 26.48
d.12 36.00 36.00 19.90 27.52 14.86
d.16 36.00 36.00 19.60 29.09 13.80
SEM of Ct w.8 0.48 0.58 0.45 0.62 1.89
w.12 0.18 0.18 0.17 0.13 0.20
w.16 0.31 0.84 0.07 0.37 0.22
n1.8 0 0 0.52 2.04 0.22
n1.12 0 0 0.06 0.28 0.11
n1.16 0 0 0.11 0.21 0.03
n2.8 1.23 0 0.36 0.88 1.65
n2.12 0.58 0 0.17 0.47 0.03
n2.16 0.30 0 0.07 0.35 0.20
d.8 0 0 0.49 1.35 1.19
d.12 0 0 0.15 0.28 0.16
d.16 0 0 0.12 0.59 0.09
238

Supplemental Table 4.4. Robustness of coexpression modules based on a permutation test of


average topological overlap.
Module Mean TO observeda Mean TO randomb Maximal TO randomc P valued
1 0.149748634 0.028131294 0.031158976 < 10-5
2 0.078259886 0.028172181 0.031459441 < 10-5
3 0.101435532 0.032000924 0.043832489 < 10-5
4 0.102286711 0.032284812 0.046148874 < 10-5
5 0.092245573 0.03109811 0.041610692 < 10-5
6 0.069554915 0.028533966 0.033097827 < 10-5
7 0.12472178 0.028137199 0.031202723 < 10-5
8 0.177877953 0.028186402 0.031646581 < 10-5
9 0.069714411 0.031419646 0.043617328 < 10-5
10 0.059263414 0.031405512 0.043372765 < 10-5
11 0.062374245 0.032168614 0.047049369 < 10-5
12 0.079978046 0.034972848 0.052824411 < 10-5
a the average TO for the genes in an observed module.
b the average TO of 100,000 iterations.
c the maximum of the average TO for the 100,000 iterations.
d the empirical probability of finding an average TO greater than or equal to the observed TO in 100,000

collections of modules comprised of randomly selected genes.


239

Supplemental Table 4.5. Functional annotation of genes associated with GO terms "response to
auxin" and/or "auxin-activated signaling pathway" enriched for M6.
Gene ID Gene Symbol TF Family Functional Annotation
GRMZM2G001799 IAA41 -- Uncharacterized protein
GRMZM2G030465 IAA24 -- IAA9-auxin-responsive Aux/IAA family
member
GRMZM2G033359 -- -- Uncharacterized protein
GRMZM2G048131 IAA33 -- IAA26-auxin-responsive Aux/IAA family
member
GRMZM2G050550 MYB153 MYB Sucrose responsive element binding
protein
GRMZM2G053338 GH3 -- Uncharacterized protein
GRMZM2G054361 ACS6 -- Uncharacterized protein
GRMZM2G077356 IAA21 -- Uncharacterized protein
GRMZM2G086949 ARFTF29 ARF Uncharacterized protein
GRMZM2G087955 MYB139 MYB Uncharacterized protein
GRMZM2G098643 PIN1 -- Putative auxin efflux carrier
GRMZM2G128421 IAA22 -- Uncharacterized protein
GRMZM2G147243 IAA28 -- IAA17-auxin-responsive Aux/IAA family
member; Uncharacterized protein
GRMZM2G148074 HB78 Homeobox Uncharacterized protein
GRMZM2G149449 IAA31 -- --
GRMZM2G152796 IAA19 -- Uncharacterized protein
GRMZM2G166944 -- -- Uncharacterized protein
GRMZM2G169820 ARFTF1 ARF Auxin response factor 1
GRMZM2G366373 IAA3 -- Uncharacterized protein
GRMZM2G378106 GH3 -- Indole-3-acetic acid amido synthetase
GRMZM2G410710 PHB3 -- Uncharacterized protein
GRMZM2G460861 -- -- SAUR33-auxin-responsive SAUR family
member
GRMZM5G894619 ACS7 -- Uncharacterized protein
240

Supplemental Table 4.6. Bonferroni-corrected P values for enrichment analysis of reported


NKD-binding motifs in 0.5 kb promoter regions determined using AME.
NKD1 motifa NKD2 motifa

Module

1 6.80e-4 5.36e-15
2 1.18e-10 7.32e-9
3 >0.05 >0.05
4 >0.05 3.68e-4
5 >0.05 1.95e-4
6 4.95e-3 2.78e-6
7 >0.05 8.71e-8
8 8.62e-3 8.38e-9
9 4.98e-2 >0.05
10 >0.05 >0.05
11 1.05e-3 >0.05
12 >0.05 3.30e-2

aThe motifs were generated from the SAAB-identified NKD1/2-binding sites using the sites2meme
and ceqlogo programs of the MEME suite.
241

Supplemental Table 4.7. Results of de novo motif enrichment analysis of the coexpression
modules.
Module Motif E value sequence Matching TFa
1 1 1.20E-225 GGCGGCGG ERF4
2 1.50E-222 AAAAAAAA OBP3
3 4.00E-136 SCSCGCSC -
4 2.50E-106 GCYGCTGC -
5 3.20E-90 GGAGGRRG -
6 6.00E-49 SCCGGCCS -
7 5.70E-44 TAWATATA -
8 6.40E-136 AAATAAAA -
9 1.30E-89 AAMCAAAC -
10 2.20E-88 TATTTTTT At5g52660
2 1 5.10E-251 CSSCCGCC ERF4
2 2.20E-229 TTTTTTTT OBP3
3 5.20E-123 TATWTTTT At5g52660
4 4.20E-97 WTTTATTT -
5 5.90E-82 GCGCGGSS -
6 7.40E-79 CCYCCTCC -
7 1.50E-84 GAGAGAGA RAMOSA1
8 4.10E-80 AAASAAAA -
9 1.00E-57 TATATATA -
10 2.80E-50 GCMGCAGC -
3 1 2.30E-28 GGMGGNGG RAP26
4 1 5.80E-34 GGGCKGSG ERF008
2 7.30E-35 TCCGAGGG -
3 1.40E-06 CYCTSCYC -
4 1.90E-02 TGCAGGGC -
6 1.70E-06 CTYCMTCC -
5 1 3.30E-45 SCCGCSSC -
2 3.60E-35 AAAAWAAA NUC
3 6.20E-19 CCYCCTCC -
4 1.60E-09 CAGCMGCM -
5 2.60E-03 CGYCSKCS RAP211
6 1 4.80E-295 TWTTTTTT NUC
2 3.70E-83 YCYCTCYC RAMOSA1
3 3.30E-106 TTTTVTTT -
4 4.40E-74 CCGCCSSC ERF4
5 3.00E-49 GCGCRCGS -
6 1.30E-47 TATAWATA -
7 4.90E-40 GCRGCMGC -
242

Supplemental Table 4.7 Continued


8 1.00E-34 CCTCCTCC -
9 4.70E-16 GTTTGKTT -
7 1 3.50E-245 TTTTTTTT OBP3
2 1.10E-204 GGCGGCGG ERF4
3 6.20E-103 CCTCYCYC -
4 1.30E-172 TTTATTTT -
5 2.60E-121 TTTSTTTT -
6 9.90E-73 SCGCSCGC -
7 1.10E-67 GCTGCWGC -
8 1.50E-77 GTGGGSTG -
9 5.10E-56 ATATATAT -
10 4.40E-51 GTTTGGTT -
8 1 3.9E-351 AAAAAAAA OBP3
2 5.40E-171 AAAWAAAT -
3 8.00E-135 CGGCSGCG ERF094
4 1.40E-107 GGGSBGGG -
5 2.80E-106 TCTCTCTC RAMOSA1
6 3.80E-54 ATATATAW -
7 4.40E-129 AAATAAAA -
8 2.20E-56 AAAAWATA -
9 5.00E-54 CKCGGCGC -
10 6.60E-62 GRGGAGGR -
9 1 5.80E-56 TTTWTTTT NUC
2 1.30E-16 GGCGGSGG ERF4
3 2.40E-14 CACCTGCC -
5 3.90E-02 ATATTATT -
10 1 1.90E-24 AAAAAAAA AT2G28810
2 7.20E-17 CCYCCKCC -
4 5.20E-04 RGARRRAG -
11 1 1.20E-28 AWAWAWAW -
2 9.60E-21 CSSCSGCC ERF4
3 2.00E-08 GAGAGAGA ROMOSA1
12 1 9.10E-71 AAAAWAAA NUC
2 3.50E-19 CCBCCKCC ERF4
3 1.10E-04 GCAGCAGC -
a Transcription factors of which the known binding motif most significantly matches the identified motif (q <
0.05).
243

Supplemental Table 4.8. Sequences of primers used for RT-qPCR.


Gene Forward Primer Reverse Primer
NKD1 TCTGGTGATAGAACCAATAAGGAT AGCTTGCACGATTAATGACA
NKD2 CATGTGATCACCCATTTGC AGCTAGCTACAAATAACATCCT
VP1 TCGTGTGTGCTAAGCATGTA ACTAGTATAGCTAATTAACCACGAC
O2 same as used in Chapter 3
azs22-4 same as used in Chapter 3
Thioredoxin same as used in Chapter 3
244

Supplemental Data Set 4.1. Normalized expression levels in RPKMs (mean of biological
replicates) and functional annotation for the 34,348 genes tested for differential expression.

Supplemental Data Set 4.2. Results of differential expression analysis and functional annotation
for all the DEGs.

Supplemental Data Set 4.3. GO term enrichments of the NKD-regulated gene sets.

Supplemental Data Set 4.4. Functional annotation and GO term enrichments of the stage-
specific NKD-regulated genes.

Supplemental Data Set 4.5. Functional annotation and GO term enrichments of the genotype-
specific NKD-regulated genes.

Supplemental Data Set 4.6. GO term enrichments of genes co-regulated by NKD1 and NKD2.

Supplemental Data Set 4.7. Metabolic pathways associated with NKD-regulated genes.

Supplemental Data Set 4.8. NKD-regulated genes that are direct targets of O2.

Supplemental Data Set 4.9. GO term enrichments of putative direct targets of NKDs.

Supplemental Data Set 4.10. WGCNA-identified coexpression modules of NKD-regulated


genes, and module memberships and functional annotations of genes.

Supplemental Data Set 4.11. GO term enrichments of the coexpression modules.


245

CHAPTER 5

Future questions and experiments

The cereal endosperm contains large quantities of carbohydrates and proteins needed for seed
germination, and is an important source of human food, animal feed, and feedstock for numerous
industrial products including biofuels (Lopes and Larkins, 1993; Sabelli and Larkins, 2009;
FAO, 2015). The goal of my dissertation research has been to understand the gene regulatory
mechanisms controlling the differentiation of the basal endosperm transfer layer (BETL), starchy
endosperm (SE), and aleurone (AL) of the maize endosperm, which are cell types that are
involved in nutrient uptake from the surrounding maternal tissues and storage within the
endosperm during kernel development (Sabelli and Larkins, 2009; Becraft and Gutierrez-
Marcos, 2012). The spatial transcriptome analysis of an 8-day-after-pollination (DAP) kernel
identified a gene regulatory module associated with BETL differentiation, which is a sub-circuit
of the MYB-Related Protein-1 (MRP-1) gene regulatory network (GRN) (Chapter 2). The
identification of direct and indirect target genes of Opaque-2 (O2) at 15 DAP followed by a
temporal analysis of O2 target-gene expression in the developing endosperm of o2 versus WT
uncovered the complexity and temporal activities of the O2 GRN, and an essential role of O2 in
regulation of AL differentiation (Chapter 3). As a continuation of the analysis of the O2 GRN,
the genes that are regulated by each individual NAKED ENDOSPERM (NKD) protein at three
stages of endosperm development (8, 12, and 16 DAP) were analyzed (Chapter 4). The analysis
showed that NKD1 plays a more critical role than O2 in regulation of genes associated with the
nkd mutant phenotype, and uncovered cooperative and antagonistic activities between the two
TFs. To further understand the GRNs that control the cell differentiation of the maize
endosperm, in addition to the future questions that have been discussed in Chapters 1 through 4,
the following questions should be answered:

1. What are the genome-wide target genes of MRP-1 and the NKDs and how are they
regulated developmentally?

Through the spatial analysis of the 8-DAP kernel compartments and the subsequent cis-motif
enrichment analysis and yeast one-hybrid assays, a 93-gene sub-circuit of the MRP-1 GRN was
identified (Chapter 2). However, the full spectrum of MRP-1 target genes remains unknown. As
246

exemplified by the analysis of the O2 GRN (Chapter 3), chromatin immunoprecipitation


followed by high-throughput sequencing (ChIP-Seq) is a powerful approach for genome-wide
identification of TF target genes. A ChIP-Seq analysis of MRP-1 targets are currently underway
in our lab. To identify both direct and indirect target genes of a TF protein, ChIP-seq is often
coupled with differential expression analysis of a loss-of-function mutant versus wild type
(Bolduc et al., 2012; Morohashi et al., 2012; Eveland et al., 2014; Li et al., 2015) (Chapter 3).
Since no loss-of-function mutant of MRP-1 has been reported, approaches such as the CRISPR-
Cas9 technology (Hannon, 2002; Bortesi and Fischer, 2015) are being considered in our lab.

Similarly, for the NKDs, with the available transcriptome data described in Chapter 3, a ChIP-
Seq, ideally performed on endosperm material of the same developmental stages, could identify
the GRNs that are directly or indirectly regulated by NKD1 and NKD2 individually. For this
purpose, given the available Ds insertion mutants of both NKD genes, transgenic maize plants
with tagged NKD proteins in a mutant background is currently being generated by our
collaborators so that a commercially available antibody against the tag (e.g., an HA tag) could be
used to perform the ChIP assays (Pautler et al., 2015). Alternatively, antibodies could be raised
against the endogenous NKD proteins as discussed above for MRP-1. The NKD GRN could be
integrated with the O2 GRN (Chapter 3) to further understand the larger GRN that is co-
regulated by the NKDs and O2.

The analyses of the O2 and NKD networks (Chapters 3 and 4) suggest that each TF likely
regulates a temporally active set of genes during endosperm development; therefore, the
differential gene expression and ChIP-Seq assays should be performed using a developmental
series of endosperm to gain a comprehensive understanding of the temporal dynamics of the
MRP-1/O2/NKD GRNs and the temporally changing roles of the TFs.

2. Do the close paralogs of MRP-1 play a role in regulation of BETL differentiation?

The analysis of MRP-1 GRN led to identification of several close paralogs of MRP-1 that are
spatially coexpressed with MRP-1 (data not shown). These MYBRs potentially function
redundantly with MRP-1 to regulate BETL differentiation. In order to fully understand the role
of these MYBR proteins including MRP-1 itself, the CRISPR-Cas9 tool could be used to knock
them out individually or in combination. If defects in BETL differentiation are observed for the
single knockout mutants, it would suggest that a MYBR gene alone is required for BETL
247

differentiation. Otherwise, if defects in BETL differentiation were observed only in a higher-


order mutant combination rather than the single mutants, it would suggest that two or more
MYBRs are redundantly required for BETL differentiation. Nonetheless, even if no obvious
phenotype is detected in some of those mutants, a transcriptome analysis in comparison to wild
type may detect perturbations in gene expression programs, which in turn would provide insights
into the function of the TFs, as exemplified by the transcriptome analysis of the nkd2-Ds mutant
(Chapter 4).

3. Does O2 and the NKD proteins each play distinct roles in AL versus SE?

The analysis of the O2 GRN (Chapter 3) and a previous report (Gontarek et al., 2016) together
suggest that O2 and the NKDs regulate AL versus SE differentiation. A laser-capture-
microdissection (LCM)-based transcriptome analysis of the nkd-R mutant showed that O2 was
significantly down-regulated in AL but not in SE (Gontarek et al., 2016), suggesting that O2 and
the NKDs play distinct roles in AL versus SE. To further understand their roles in these cell
types, cell-type-specific analysis of gene expression programs and DNA-binding patterns would
be required. For profiling of the mRNA populations in AL and SE, I have collected (in
collaboration with another colleague in our lab) from the single/double Ds insertion mutants of
NKD1 and NKD2 (same genotypes as the three mutants analyzed in Chapter 4) and a W22o2
mutant at three stages of endosperm development (8, 12, and 16 DAP), and the LCM of AL and
SE cell types are in progress. The experiments would uncover the downstream genes that are
regulated in SE and AL by NKD1, NKD2, and O2 either individually or in combination. In the
future, if the NKD antibodies or transgenic lines expressing tagged NKD proteins are available,
to understand the cell-type-specific binding patterns of O2, NKD1, and NKD2, approaches such
as nuclear sorting followed by ChIP-Seq (Wang and Deal, 2015), as described in Chapter 1,
could be used. For this purpose, transgenic maize plants expressing fluorescent proteins driven
by promoters of SE and AL-specific genes [e.g., the maize 27-kD γ-zein promoter and the barley
Ltp2 promoter (Gruis et al., 2006)] could potentially be used to obtain a number of SE/AL nuclei
that are sufficient for ChIP assays. Together, the combination of the cell type-specific
transcriptome and genome-wide TF-DNA binding profiles would provide valuable insights into
the cell type-specific gene regulatory mechanisms associated with endosperm cell differentiation
and its overall development.
248

APPENDIX A: Dissertation author’s main contributions to each chapter

Chapter 1 – The early period of endosperm development is a critical phase for establishing the
storage capacity of maize kernel. In this chapter, I reviewed the literature on early maize
endosperm development, with an emphasis on our current knowledge of molecular mechanisms
that control cell differentiation, which is the main question addressed in this dissertation. Where
necessary, the relevant developmental processes in other cereal species and Arabidopsis were
also discussed. The main conclusions that I drew from the literature were that: (a) the cell fate
specification and differentiation of the maize endosperm cells are regulated by a combination of
endogenous cues and positional information, (b) the gene regulatory programs associated with
cell differentiation exhibit extensive interplay with the programs underlying other key
endosperm developmental processes, and (c) future understanding of the gene regulatory
mechanisms controlling endosperm cell differentiation will require a comprehensive
understanding of the spatiotemporal gene expression programs in the developing endosperm. In
writing the manuscript, I generated the original draft and created the schematic diagrams of the
maize kernel for Figure 1.2. My co-author, Joanne Dannenhoffer (Central Michigan Univ.)
provided Figure 1.1. The manuscript was then edited by Joanne Dannenhoffer and Ramin
Yadegari for publication in the book Maize Kernel Development [2017, B.A. Larkins, ed
(CABI), pp. 28-43].

Chapter 2 – As a first step toward understanding the molecular mechanisms underlying


endosperm cell differentiation, I identified gene sets (modules) coexpressed in the maize kernel
compartments (both of maternal and zygotic origins) using the 8-day-after-pollination (DAP)
transcriptome data. These data were generated through the efforts of my co-first author, Dhiraj
Thakare (Yadegari Lab), and co-author Chuang Ma (Wang Lab). The identified gene
coexpression modules included modules associated with each of the endosperm cell types. By
comparing the modules with the previously published temporal patterns of gene expression in
maize endosperm, I observed distinct temporal dynamics of expression for each endosperm-cell-
type-associated module, which reflects the active gene regulatory processes underlying cell
differentiation at 8 DAP. The significant enrichments of previously reported endosperm-
imprinted genes in several of the coexpression modules and the detection of many transcription-
factor (TF) genes suggested extensive interplay between the TF-based regulatory programs and
249

epigenetic programs underlying endosperm cell differentiation. Each coexpression module that I
identified contained TF genes and non-TF genes, including genes with reported functions in
regulation of cell differentiation, and therefore can be used as starting points to decipher the gene
regulatory networks (GRNs) associated with endosperm cell differentiation. As a proof-of-
concept, I identified a sub-circuit of the MYB-Related Protein-1 (MRP-1) GRN based on
detection a number of MRP-1-binding-site-like cis-motifs in the promoter regions of genes in the
BETL-associated module and validations of MRP-1 binding to a subset of those cis-motifs by
yeast one-hybrid (Y1H) assays. The Y1H assays (and the mRNA in situ hybridizations) were
carried out by my co-authors Alan Lloyd, Neesha Nixon, Angela Arakaki, William Burnett, Kyle
Logan and Gary Drews (Univ. Utah). The analysis suggested that the sub-circuit of the MRP-1
GRN and the larger BETL-associated coexpression module is likely sufficient for regulation of
BETL differentiation. I created the first draft of the manuscript, except for the first section of
Results and the methods for laser-capture microdissection and RNA sequencing, which were
drafted by Dhiraj Thakare, and the methods of Y1H assay were written by Gary Drews. The
manuscript was then edited by Ramin Yadegari for publication in The Plant Cell (2015, vol. 27,
513-531).

Chapter 3 – To understand the molecular mechanisms that link the cell differentiation programs
with the storage program in maize endosperm, I identified the Opaque-2 (O2)-modulated and/or
bound gene sets and their associated gene ontology (GO) terms using the transcriptome data and
genome-wide O2-binding data that were generated by my co-first author Guosheng Li (Yadegari
Lab). To understand the timing of target gene regulation by O2, I performed a temporal analysis
of O2 target gene expression in an o2 mutant as compared to wild-type endosperm and observed
two distinct modes of O2-mediated target gene activation, which suggested that the O2 GRN is
finely modulated depending on the stage of endosperm development. Through a phylogenetic
analysis of the maize bZIP family proteins, I identified a novel putative O2-heterodimerizing
protein (bZIP17) that is encoded by an O2-acticated direct target gene, and detected target-gene
co-activation activities between bZIP17 and O2. The protein-protein interaction between O2 and
bZIP17 was verified by my co-author Choong-Hwan Ryu (Yadegari Lab) using in vitro pull-
down assays and biomolecular fluorescence complementation assays. To understand the role of
O2 in regulation of endosperm cell differentiation, I detected target-gene co-activation activities
between O2 and the INDETERMINATE DOMAIN family protein NAKED ENDOSPERM2
250

(NKD2), which I identified as encoded by an O2-activated direct target gene. Based on these
data, my co-authors Carla Spears, Devon Leroux, and Joanne Dannenhoffer (Central Michigan
Univ.) examined the aleurone cells of an o2 mutant in comparison to a wild type using
transmission electron microscopy and showed that the o2 aleurone cells differ from the wild type
in anticlinal cell wall thickness and in the abundance as well as size distribution of lipid bodies
and protein storage vacuoles, indicating that O2 is required for proper aleurone differentiation.
Together, the results support a model that O2, bZIP17, and the NKD proteins co-regulate a
complex network associated with multiple key aspects of endosperm/kernel development,
including cell differentiation, storage, maturation, germination, as well as endosperm’s responses
to abiotic stresses. I wrote the first draft of the manuscript, except for the methods of chromatin
immunoprecipitation and immunoblot assays which were written by Guosheng Li, the results and
methods of ultrastructural analysis which were written by Joanne Dannenhoffer, and the methods
of electrophoretic mobility assays which were written by Gary Drews. The manuscript was then
edited by Ramin Yadegari in preparation for submission to a journal.

Chapter 4 – To understand the molecular mechanisms underlying the coordinated regulation of


cell cycle control, cell differentiation, storage function, and maturation of endosperm by the
NKD TFs, and the regulatory role of each individual NKD protein, I performed a transcriptome
analysis of endosperm dissected from the single mutants and a double mutant of NKD1 and
NKD2. The endosperm materials were collected by my co-authors Hao Wu, Bryan Gontarek, and
Philip Becraft (Iowa State Univ.). Consistent with the reported nkd mutant phenotype, I detected
a more critical role for NKD1 than NKD2 in regulation of endosperm gene expression, and
showed that the nkd mutant phenotype is primarily associated with the loss of NKD1 function. In
addition, my analyses suggested strongly that the NKD proteins function as both activators and
repressors, and uncovered evidence for cooperative and antagonistic activities between NKD1
and NKD2. My comparative analysis of the NKD-regulated genes versus O2-modulated and/or
bound genes provided further insights into the downstream gene networks that are associated
with multiple key aspects of endosperm development (see above) and are co-regulated by O2 and
NKDs, with NKD1 likely playing a central role. To discover additional key players in the gene
networks regulated by NKDs, I identified coexpressed gene sets from all the NKD-regulated
genes and analyzed the GO terms and TFs enriched in each gene set. The data could be used to
further decipher the NKD-regulated gene networks through analysis of TF-target gene
251

interactions in the future. I wrote the first draft of the manuscript, which was then edited by
Ramin Yadegari in preparation for submission to a journal.
APPENDIX B: REPRINT OF “ENDOSPERM DEVELOPMENT AND CELL SPECIALIZATION”

252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
APPENDIX C: REPRINT OF “RNA SEQUENCING OF LASER-CAPTURE MICRODISSECTED
COMPARTMENTS OF THE MAIZE KERNEL IDENTIFIES REGULATORY MODULES
ASSOCIATED WITH ENDOSPERM CELL DIFFERENTIATION” 268
The Plant Cell, Vol. 27: 513–531, March 2015, www.plantcell.org ã 2015 American Society of Plant Biologists. All rights reserved.

LARGE-SCALE BIOLOGY ARTICLE

RNA Sequencing of Laser-Capture Microdissected


Compartments of the Maize Kernel Identifies Regulatory
Modules Associated with Endosperm Cell Differentiation OPEN

Junpeng Zhan,a,1 Dhiraj Thakare,a,1 Chuang Ma,a,2 Alan Lloyd,b Neesha M. Nixon,b Angela M. Arakaki,b
William J. Burnett,b Kyle O. Logan,b Dongfang Wang,a,3 Xiangfeng Wang,a,4 Gary N. Drews,b and Ramin Yadegaria,5
a School of Plant Sciences, University of Arizona, Tucson, Arizona 85721
b Department of Biology, University of Utah, Salt Lake City, Utah 84112
ORCID IDs: 0000-0001-7353-7608 (J.Z.); 0000-0002-4870-5064 (D.T.); 0000-0002-0975-2984 (R.Y.)

Endosperm is an absorptive structure that supports embryo development or seedling germination in angiosperms. The
endosperm of cereals is a main source of food, feed, and industrial raw materials worldwide. However, the genetic networks
that regulate endosperm cell differentiation remain largely unclear. As a first step toward characterizing these networks, we
profiled the mRNAs in five major cell types of the differentiating endosperm and in the embryo and four maternal compartments
of the maize (Zea mays) kernel. Comparisons of these mRNA populations revealed the diverged gene expression programs
between filial and maternal compartments and an unexpected close correlation between embryo and the aleurone layer of
endosperm. Gene coexpression network analysis identified coexpression modules associated with single or multiple kernel
compartments including modules for the endosperm cell types, some of which showed enrichment of previously identified
temporally activated and/or imprinted genes. Detailed analyses of a coexpression module highly correlated with the basal
endosperm transfer layer (BETL) identified a regulatory module activated by MRP-1, a regulator of BETL differentiation and
function. These results provide a high-resolution atlas of gene activity in the compartments of the maize kernel and help to
uncover the regulatory modules associated with the differentiation of the major endosperm cell types.

INTRODUCTION materials (Lopes and Larkins, 1993; Olsen, 2001, 2004; Sabelli
and Larkins, 2009; FAO, 2012).
Seed development is initiated by double fertilization of the In most flowering plants, endosperm development begins with
haploid egg cell and the dikaryotic central cell to produce two the formation of a coenocyte, as the fertilized central cell undergoes
filial structures, a diploid embryo and a triploid endosperm, re- multiple rounds of nuclear divisions without cytokinesis. The mul-
spectively (Faure, 2001; Hamamura et al., 2012). Endosperm func- tinucleated coenocyte then undergoes cellularization and cell dif-
tions as an absorptive structure that supports embryo development ferentiation (Olsen, 2004; Sabelli and Larkins, 2009). In dicots, the
or seedling germination in angiosperms (Lopes and Larkins, 1993). endosperm is mostly absorbed by the developing embryo shortly
Recent evidence also indicates that endosperm plays a critical role after cellularization. By contrast, in monocots, and particularly in
in regulation of seed development through interaction with the cereals, the endosperm enlarges significantly after cellularization
embryo and the seed coat (Berger et al., 2006; Lafon-Placette and through many rounds of cell division accompanied by cell enlarge-
Köhler, 2014). The endosperm of cereal grains occupies a large ment and organelle proliferation. Consequently, the cereal endo-
portion of the mature seed, holds large amounts of proteins and sperm acquires a high storage capacity of carbohydrates and
carbohydrates required for seedling development, and is an proteins prepared for mobilization upon seedling germination (Lopes
important source of food, feed, and renewable industrial raw and Larkins, 1993; Sreenivasulu and Wobus, 2013). The acquisition
of endosperm storage capacity is enabled in part through the activity
1 These authors contributed equally to this work.
of specialized cell types or compartments that mediate uptake of
2 Current address: College of Life Sciences, Northwest A&F University, nutrients from the maternal structures and their storage in the inner
Yangling, Shaanxi 712100, China. compartments of the endosperm. Therefore, elucidating how cell
3 Current address: Department of Biology, Spelman College, Atlanta, GA
differentiation is regulated during endosperm development is central
30314.
4 Current address: Department of Plant Genetics and Breeding, China
to understanding endosperm structure and function.
Agricultural University, Beijing 100193, China. Because of its relatively large size and economic importance,
5 Address correspondence to yadegari@email.arizona.edu. the maize (Zea mays) endosperm represents an excellent model
The author responsible for distribution of materials integral to the findings system to study early regulatory processes that regulate regional
presented in this article in accordance with the policy described in the and cellular differentiation events. The initial coenocytic phase of
Instructions for Authors (www.plantcell.org) is: Ramin Yadegari (yadegari@
email.arizona.edu).
endosperm growth in maize occurs during the first 2 d after polli-
OPEN
Articles can be viewed without a subscription. nation (DAP), and this is followed by a period of cellularization
www.plantcell.org/cgi/doi/10.1105/tpc.114.135657 during 3 to 4 DAP. Following cellularization, the endosperm cells
269
514 The Plant Cell

undergo two major phases of mitotic proliferation, an early phase RESULTS


that lasts until 8 to 12 DAP in the central region, and a late phase
that continues until 20 to 25 DAP in the outer endosperm layers. Capture and Analysis of mRNA Populations of Filial and
Starting at ;8 to 10 DAP, the central portion of endosperm cells Maternal Compartments of 8-DAP Kernel
gradually switches from mitosis to endoreduplication and be-
comes filled with starch and storage proteins (Brink and Cooper, To identify the genes active in each of the endosperm cell types,
1947; Olsen, 2001, 2004; Sabelli and Larkins, 2009; Becraft and we used LCM to isolate and profile mRNA populations of five
Gutierrez-Marcos, 2012; Olsen and Becraft, 2013; Leroux et al., endosperm compartments (cell types) of maize inbred line B73
2014). at 8 DAP. The compartments analyzed included AL, BETL, ESR,
Differentiation of maize endosperm cells occurs primarily at 4 and two subregions of SE, the CSE and the CZ. To compare
to 6 DAP (following endosperm cellularization and before the endosperm gene expression programs with embryonic and mater-
initiation of mitotic proliferation), resulting in four main cell types, nal programs, we also captured the embryo (EMB), nucellus (NU),
including the starchy endosperm (SE), the aleurone (AL), the placento-chalazal region (PC), pericarp (PE), and the vascular region
embryo-surrounding region (ESR), and the basal endosperm of the pedicel (PED) at 8 DAP (Figure 1A; Supplemental Figures 1 to
transfer layer (BETL) (Olsen, 2001; Becraft and Gutierrez-Marcos, 3 and Supplemental Table 1). We selected this time point because it
2012; Leroux et al., 2014). The SE is the cell type that accumulates follows differentiation of the main cell types of the endosperm, which
starch and storage proteins. The SE itself contains at least three occurs at 4 to 6 DAP, and precedes developmental programs as-
subregions, including the central starchy endosperm (CSE), the sociated with endosperm function, including the activation of stor-
conducting zone (CZ), and the subaleurone (Becraft, 2001; Olsen, age product synthesis and deposition program, which initiates at
2001, 2004; Sabelli and Larkins, 2009). The AL is a single peripheral ;8 to 10 DAP (Becraft, 2001; Olsen, 2001, 2004; Sabelli and
layer of cells that produces hydrolytic enzymes to mobilize the Larkins, 2009; Becraft and Gutierrez-Marcos, 2012; Olsen and
storage products in the SE when activated during seed germina- Becraft, 2013; Leroux et al., 2014). Total RNA extracted from bi-
tion. The ESR is believed to act as a physical barrier and mes- ological triplicates for the endosperm compartments and embryo,
senger between endosperm and embryo (Olsen, 2004). The BETL and single replicates of maternal compartments (22 samples) were
is a transfer cell layer that transports nutrients from the maternal reverse-transcribed to cDNA using oligo(dT) and random primers,
tissue into the inner endosperm cells, including the developing SE, amplified, and paired-end sequenced using an Illumina HiSequation
in order to enable starch and protein synthesis (Sabelli and Larkins, 2000 platform.
2009; Becraft and Gutierrez-Marcos, 2012). Recent studies have The resulting reads were quality checked and mapped to the
identified many genes expressed specifically in the BETL, including maize reference genome (B73 RefGen_v3). Of the resulting
multiple genes encoding cysteine-rich proteins that are thought to mapped reads (8.9 to 34.7 million, 52.3 to 89.7% of total reads),
act as antimicrobial or intercellular signal molecules (Tailor et al., 3.2 to 11.3 million (22.1 to 39.5%) were mapped to exonic
1997; Marshall et al., 2011), and MRP-1 (Myb-Related Protein-1), sequences (Supplemental Table 2). The exonic reads were nor-
a MYB-related (MYBR) transcription factor previously shown to malized using Cufflinks (Trapnell et al., 2012) and reported as frag-
activate a number of these genes in the BETL (Gómez et al., 2002, ments per kilobase of transcript per million mapped reads (FPKM). A
2009; Gutiérrez-Marcos et al., 2004). Moreover, ectopic expression gene was considered expressed in a given sample if the lower
of MRP-1 in the AL has been shown to produce a transient BETL- boundary of its FPKM 95% confidence interval (FPKM_conf_lo) was
like structure (Gómez et al., 2009). Additional recent efforts have greater than zero (Hansey et al., 2012). Based on this criterion,
enabled genome-wide identification of gene expression during 29,369 genes were identified as expressed in at least one of the 22
nearly all stages of endosperm development (Sekhon et al., 2013; samples (Supplemental Data Set 1). Using pairwise Spearman
Chen et al., 2014; Li et al., 2014). However, little is known about the correlation coefficient (SCC) analysis, the triplicate FPKM values
gene regulatory networks (GRNs) that regulate the differentiation from each of the endosperm compartments and the embryo were
and determine the function of the individual cell types or com- shown to be highly correlated (r = 0.87 to 0.91; Supplemental
partments of the endosperm in maize. Figure 4). Accordingly, we pooled each triplicate set of exonic reads
Here, we used a coupled laser-capture microdissection (LCM) and renormalized the data using Cufflinks. Using the same cutoff as
and RNA sequencing (RNA-Seq) strategy to comprehensively indicated above, we detected 30,665 genes expressed in at least
profile the mRNA populations present in each of the main cell one of the 10 compartments (Supplemental Data Set 2 and
types of the maize endosperm, as well as the embryo and four Supplemental Figure 5), 10,725 genes expressed in all 10 com-
maternal compartments of the kernel at 8 DAP. We identified partments (Supplemental Figure 6A), and between 15,910 and
mRNAs that specifically accumulate in each of the captured 23,853 (NU and ESR, respectively) expressed in individual com-
compartments. Also, using an unbiased network analysis tool, we partments (Figure 1B; Supplemental Table 3). In all cases, the
detected modules of coexpressed genes that are either pre- proportion of transcription factor (TF) genes detected as expressed
dominantly expressed in a single compartment or expressed in tracked closely with the total number of expressed genes (be-
multiple compartments, including several endosperm-correlated tween ;5.3 and 6.0% for CZ and PED, respectively; Figure 1B;
modules that are enriched for temporally upregulated genes and/or Supplemental Table 3). The proportions of high-expressing
imprinted genes that we previously identified. By focusing on the (FPKM $ 10), medium-expressing (2 # FPKM < 10), and low-
analysis of genes in a BETL-correlated coexpression module, we expressing (FPKM < 2) genes were relatively similar in all com-
identified and experimentally validated a regulatory module of the partments (Figure 1C; Supplemental Table 4). Collectively, 22,703
BETL GRN that is activated by MRP-1. genes were detected as expressed in EMB, and 28,078 and
270
Cell Differentiation Modules in Maize Endosperm 515

Figure 1. Profiles of Sequenced RNAs from the Captured Filial and Maternal Compartments of 8-DAP Maize Kernel.

(A) Graphic representation of an 8-DAP maize kernel showing the relative position of the 10 captured filial and maternal compartments used for RNA
sequencing.
(B) Numbers of TF genes, non-TF protein-coding genes, microRNA genes, transposable elements, and pseudogenes expressed in the 10 captured
compartments.
(C) Proportions of genes expressed at different levels (based on FPKM) in the 10 kernel compartments.

22,989 genes expressed in at least one captured endosperm and Identification of Gene Sets Specifically Expressed in Each of
maternal compartment, respectively. The three sources of captured the 8-DAP Kernel Compartments
tissues shared 19,009 expressed genes in total (Supplemental
Figure 6B). Taken together, our analysis of LCM-derived RNA-Seq To discover the gene expression programs that characterize
data indicates that we have obtained sufficient coverage of the each kernel compartment, we identified mRNAs that specifically
transcriptome of the filial and maternal compartments of the kernel accumulate in each compartment at 8 DAP by applying a com-
for subsequent analysis of gene networks. partment specificity (CS) scoring algorithm (see Methods) to the
genes with FPKM $ 2 in at least one compartment. In this
Filial and Maternal Compartments of the Kernel Exhibit analysis, we defined the corresponding genes with CS score
Distinct mRNA Populations > 0.3 as being expressed in a compartment-specific pattern. Using
this cutoff, 13,009 compartment-specific genes were identified in
To understand the relationships between the mRNA populations total for all captured compartments (Supplemental Data Set 3;
isolated from the individual compartments, we performed a princi- Figure 3). In contrast to the similarity of the overall mRNA profiles
pal component analysis (PCA) (Figure 2A; Supplemental Table 5) detected among the compartments as described above (Figure
and a hierarchical clustering of data from the SCC analysis (Figure 1C), the numbers of detected compartment-specific genes showed
2B) of normalized expression levels for the 30,665 genes expressed dramatic differences among the 10 compartments (Figure 3B). The
in at least one compartment. Those compartments with highest endosperm cell types showed the lowest number of compartment-
overlap in mRNA populations are expected to be more closely specific genes, ranging from a low of 331 in the CSE (1.6% of all
associated in such analyses and are likely to share functions. Both CSE-expressed genes) to a high of 912 in the BETL (4.4% of all
analyses showed high correlation among endosperm cell types and BETL-expressed genes) compared with the maternal compart-
a distinct clustering of these cell types in comparison to the ma- ments that ranged from 1390 in the PC (8.6% of all PC-expressed
ternal compartments (Figures 2A and 2B). As expected, the filial genes) to 2432 in the PE (13.2% of all PE-expressed genes) (Figure
EMB showed closer correlation with endosperm cell types as 3B). For the EMB, 2235 genes were identified as compartment
compared with the maternal compartments. However, the endo- specific, which corresponds to 9.8% of all EMB-expressed genes
sperm AL showed a closer correlation with EMB (r = 0.80) than with (Figure 3B). The proportion of TF genes among the compartment-
any other endosperm cell type (Figure 2B). An analysis of our data specific genes did not track uniformly across all compartments,
for expression of two AL marker genes, namely, VPP1 (VACUOLAR varying from a low of 3.9% (CZ) to a high of 9.9% (EMB) (Figure
H + -TRANSLOCATING INORGANIC PYROPHOSPHATASE1) 3B). The variable number and proportion of compartment-specific
(Wisniewski and Rogowsky, 2004) and AL-9 (Gómez et al., 2009), genes and the associated variation in the proportion of TF genes
indicated that the captured EMB RNA sample was not contaminated suggest that most of the endosperm cell types captured express
by the AL RNAs (Supplemental Data Set 2). Therefore, our obser- less complex gene sets compared with the maternal compartments
vation suggests that the AL is distinct in some zygotic functions or the embryo. Alternatively, the complexity of expression may
typically not found in the other endosperm cell types. Together, our simply reflect the complexity of the captured compartments, as the
data indicate that maternal and filial gene expression programs are captured EMB and maternal compartments likely contain more
divergent as they arise from distinct genetic origins and that the than one cell type.
captured compartments show sufficient diversity at the mRNA level We used three sets of expression localization data to validate
to allow identification of unique gene sets for each compartment. the cell type-specific patterns of mRNA accumulation in the
271
516 The Plant Cell

Figure 2. Relationship between the RNA Populations Obtained from the Filial and Maternal Compartments of 8-DAP Kernel.
(A) PCA of genes expressed in the captured kernel compartments. Principal components one through three (PC1 to PC3) collectively explained 61.9%
of the variance in the mRNAs obtained from the 10 compartments.
(B) SCC analysis of the mRNA data for the 10 kernel compartments using log2-transformed FPKM values of the 30,665 expressed genes. The
hierarchical clustering dendrogram was inferred by applying (1 2 SCC) as distance function.

endosperm. First, we performed a series of mRNA in situ hy- 2005; Langfelder and Horvath, 2008) to the expressed genes
bridizations for genes that were shown to be highly specific to after excluding the ones with low FPKM (average FPKM < 1)
a single compartment based on CS scores ranging from 0.74 to and/or low coefficient of variation (<1) across all 10 compartments.
0.99, including 20 genes expressed in the CSE (2), ESR (1), AL (3), The 9361 genes that fulfilled these stringent criteria fell into 18
and BETL (14) (Supplemental Figure 7). Second, we previously coexpression modules (M1 to M18), containing from 83 (M1) to
performed in situ hybridization for 10 genes specifically expressed 1209 (M4) genes, including 3 (M1) to 111 (M8) coexpressed TF
in CSE (3), CZ (1), ESR (3), AL (1), and BETL (2) (Li et al., 2014). genes (Figure 4B; Supplemental Data Set 4). Trend-plot analysis of
These genes showed CS scores in the given compartments Z-scores of genes in each module showed that these gene sets
ranging from 0.47 to 0.99. Third, we summarized previously re- were expressed in a highly coordinated manner (Supplemental
ported in situ hybridization or promoter activity data for cell-specific Figure 8). Significantly, a permutation test showed that the average
genes from the literature (Hueros et al., 1995, 1999; Magnard et al., topological overlap of the 18 observed modules was greater than
2000, 2003; Serna et al., 2001; Woo et al., 2001; Gómez et al., randomly sampled modules of the same size (P value < 1025;
2002, 2009; Gutiérrez-Marcos et al., 2004; Wisniewski and Supplemental Table 7), suggesting that the assignment of gene
Rogowsky, 2004; Balandín et al., 2005; Massonneau et al., 2005; sets to each of the modules was highly robust. Association of each
Muñiz et al., 2006, 2010; Royo et al., 2014) and also found these coexpression module with each compartment or cell type was
genes to have a relatively high range of CS scores (0.51 to 0.99). quantified by Pearson correlation coefficient (PCC) analysis and
Altogether, these comprise 44 genes showing cell-specific ex- visualized using a hierarchically clustered heat map (Figure 4A).
pression in the endosperm (Supplemental Table 6). All of these Interestingly, the 18 coexpression modules fell into two distinct
genes showed highly specific mRNA localization patterns within the categories showing a relatively high correlation (r $ 0.35) with either
cell types with high CS scores, indicating that our RNA-Seq data the filial (i.e., M1, M2, M8, M9, M10, M12, M15, M17, and M18) or
accurately reflect the accumulation of endogenous mRNAs in the the maternal (i.e., M3-M6, M11, M13, M14, and M16) compart-
endosperm. ments, except for only one module (M7) that was highly correlated
with both EMB (r = 0.68) and PE (r = 0.64). Furthermore, 10 of the
Identification of Gene Coexpression Modules of 8-DAP 18 modules, including M3, M4, M6, M8, M10, M11, M12, M15,
Maize Kernel M17, and M18, specifically correlated with individual compartments
(r $ 0.85 for one compartment and r < 0.35 for other compart-
To begin to understand the nature of the GRNs in each of the ments), indicating that the expression of genes in these modules
captured compartments or cell types of the 8-DAP kernel, we are highly compartment or cell type specific. The other eight
identified coexpressed gene sets by applying weighted gene modules showed relatively high correlation to at least two com-
coexpression network analysis (WGCNA) (Zhang and Horvath, partments (r $ 0.35). Consistent with the high correlation of mRNAs
272
Cell Differentiation Modules in Maize Endosperm 517

Figure 3. Compartment-Specific Gene Sets Identified Using the CS Scoring Method.


(A) Heat map of scaled FPKM values of the 13,009 compartment-specific genes identified in all 10 kernel compartments.
(B) Numbers of TF and non-TF genes in each compartment-specific gene set.

detected between EMB and AL described above (Figure 2B), three expression at 0 to 4 DAP and high expression at 6 to 12 DAP).
modules including M1, M2, and M9 showed high correlation with Comparison of the spatial coexpression modules described here
both EMB and AL (r $ 0.35). with the temporally upregulated gene sets showed that all of the
To confirm that the assignment of genes to these coexpression five endosperm compartment-correlated modules significantly
modules using WGCNA reflected valid compartment-based patterns overlapped with the up@6DAP gene set (P value < 1025), while four
of mRNA accumulation, we examined the number of overlapping of them significantly overlapped with the up@8DAP gene set
genes between these modules and the compartment-specific gene (Figure 4D). Consistent with this, an analysis of the overall ex-
sets identified above (Supplemental Data Set 3). This analysis pression levels for each coexpression module using the available
showed that from ;52 to 94% of genes in the compartment- developmental RNA-Seq data generated by us and others (Chen
correlated coexpression modules were in fact detected within et al., 2014; Li et al., 2014) from whole kernel and whole endo-
the corresponding compartment-specific gene sets and that the sperm material showed that nearly all endosperm-correlated
overlap between each compartment-correlated module and the modules (M10, M12, M15, M17, and M18) showed upregulation
corresponding compartment-specific gene set was greater than at 6 to 8 DAP (Supplemental Figures 9 and 10). However, the
expected by chance (hypergeometric test, P value < 1025) (Figure extent of the upregulated patterns varied among the endosperm-
4C). These data indicate that a large proportion of coexpressed correlated modules with the expression of genes in the AL-, ESR-,
genes in each module detected through WGCNA is related to and BETL-correlated modules (M12, M15, and M18, respectively),
compartment- or cell type-specific functions. Taken together, showing a more rapid decline by 10 DAP, whereas the expression
these results indicate that each of the filial and maternal com- of CSE- and CZ-correlated modules (M10 and M17, respectively)
partments of the maize kernel is associated with one or more exhibited a more gradual decline beyond 22 DAP (Supplemental
coexpression modules that reflect the gene regulatory processes Figures 9 and 10). Interestingly, genes in modules M1, M2, and
specific to each compartment and are indicators of the differen- M9 with high correlations with both EMB and AL also showed
tiation programs functioning within each compartment. a similar temporal expression pattern as those of the endosperm-
correlated modules (Supplemental Figure 11). Together, these
The Endosperm-Associated Coexpression Modules Are data indicate that the coexpression modules associated with the
Associated with Distinct Temporal Programs of Expression major endosperm compartments at 8 DAP are regulated in a highly
coordinated manner in both space and time.
We previously described a set of temporal programs of gene
activity during early kernel and endosperm development in maize The Endosperm-Associated Modules Are Enriched for
and suggested that some of these programs correlated with cell Endosperm-Imprinted Genes
differentiation (Li et al., 2014). The identified temporal programs
included gene sets exhibiting temporal upregulation at (“up@”) Gene imprinting has been suggested to be involved in the regula-
specific stages (e.g., up@6DAP refers to a gene set exhibiting low tion of nutrient allocation from the maternal tissues to endosperm
273
518 The Plant Cell

Figure 4. Gene Coexpression Modules Detected Using WGCNA.


(A) Heat map of the correlations between detected modules (M1 to M18) and kernel compartments hierarchically clustered based on Euclidean
distance. The PCC values are quantitative indicators of relative expression levels of all genes in each module.
(B) Numbers of TF and non-TF genes in each coexpression module.
(C) Relationships of the compartment-specific gene sets with the corresponding compartment-correlated coexpression modules obtained using
WGCNA. The heat map indicates P values of hypergeometric tests of overrepresentation of genes in a given tested pair of gene sets. Compartment-
specific gene sets are noted on the x axis and the corresponding WGCNA coexpression modules (those with a specific pattern for a single com-
partment) on the y axis. Boxes contain the numbers of overlapping genes and proportions (in parentheses) of these genes in the WGCNA-identified
modules.
(D) Relationships of the temporal gene sets (Li et al., 2014) with the coexpression modules obtained using WGCNA. The heat map indicates P values
(-log10) of hypergeometric tests of overrepresentation of genes in a given tested pair of gene sets. Temporal gene sets are noted on the x axis and all
WGCNA coexpression modules on the y axis. Boxes contain the numbers of overlapping genes. Numbers of genes in each temporal gene set:
up@2DAP, 54; up@3DAP, 68; up@4DAP, 92; up@6DAP, 523; up@8DAP, 1,402; up@10DAP, 552; and up@12DAP, 241.
274
Cell Differentiation Modules in Maize Endosperm 519

(Haig and Westoby, 1989; Moore and Haig, 1991; Costa et al., In the case of zein-related genes, the mRNAs for only four
2012). To investigate the relationship between the spatial programs genes encoding the 15-kD b-zein, the 16-kD g-zein, the 27-kD
of gene expression and gene imprinting, we examined the overlap g-zein, and the 18-kD d-zein were detected in our analysis
between the WGCNA-generated coexpression modules with the (FPKM_conf_lo > 0) in at least one cell type (Supplemental Data
imprinted genes that we previously identified in developing endo- Set 2). Interestingly, all four zein genes, as well as the Floury-1
sperm (Xin et al., 2013). This analysis showed that a subset of gene, which encodes an endoplasmic reticulum membrane
genes in the CZ- and BETL-correlated coexpression modules (M17 protein involved in the targeted localization of an 22-kD a-zein in
and M18, respectively) significantly overlapped with a subset of protein body formation (Holding et al., 2007), were contained
the previously described paternally expressed gene sets (PEGs); the within the M10 module (Supplemental Data Set 4). This obser-
CSE-correlated module M10 and the modules associated with the vation correlates well with previous reports that the formation of
maternal compartments (M3, M5, M6, M11, M13, M14, and M16) zein-containing protein bodies start as small accretions con-
showed significant overlap with the maternally expressed gene sets sisting primarily of b- and g-zeins (Woo et al., 2001), suggesting
(MEGs); and a subset of the ESR-correlated coexpression module that the M10 module contains the key early genes necessary for
exhibited significant overlap with both the previously described storage protein body biogenesis. The M10 module also included
PEGs and MEGs (hypergeometric P value < 1025; Figure 4E). In the TF genes Opaque-2 (O2) and PBF (PROLAMIN-BOX BIND-
support of these data, a similar pattern of overlaps was observed ING FACTOR), with the relatively high M10 module membership
when we applied less stringent criteria to identify genes with al- (MM) scores of 0.93 and 0.97, respectively (Supplemental Data
lele-biased expression patterns using the same set of normalized Set 4). Furthermore, visualization of M10 using VisANT (Hu et al.,
RNA-Seq data (Supplemental Figure 12). Interestingly, many of 2004) showed that these two TFs are among the most highly
the imprinted/allele-biased genes assigned by WGCNA to the connected intramodular hubs of this module (Supplemental
endosperm-associated coexpression modules were TF genes from Figure 14A). O2 and PBF have previously been shown to regu-
multiple families (Supplemental Table 8). These results indicate an late storage program gene expression in maize endosperm
extensive interplay between epigenetic programs that regulate al- (Schmidt et al., 1990, 1992; Marzábal et al., 2008). For example,
lelic expression and the transcriptional regulatory programs in- the 15-kD b-zein and the 27-kD g-zein have been shown to be
volved in the cellular differentiation and function of endosperm. regulated by O2 and PBF, respectively (Cord Neto et al., 1995;
Marzábal et al., 2008). Therefore, the M10 coexpression module
Biological Processes Enriched in Coexpression Modules of likely includes a number of direct gene targets of both TFs.
the Filial Compartments of the Kernel For the CZ-correlated M17 module, key functional over-
representations included “glycolysis,” “response to hydrogen
The identified coexpression modules (Figure 4A) are likely associ- peroxide,” “response to cadmium ion,” and “response to heat”
ated with specific biological processes or pathways involved in the (Supplemental Figure 13). This module showed a modest correla-
development or function of each compartment. To identify the tion with the CSE (r = 0.27; Figure 4A), suggesting that the CZ may
major biological processes associated with the filial coexpression express an overlapping set of genes that function similarly to those
modules, we used Blast2GO (Conesa et al., 2005; Conesa and detected in the CSE. Conversely, the captured CSE cells are ex-
Götz, 2008; Götz et al., 2008) to identify the processes that were pected to have contained a portion of the CZ cells (Supplemental
significantly enriched (false discovery rate < 0.05) in the modules Figure 1) as the latter likely extend basally and centrally into the
that showed high correlation with endosperm compartments and/ captured CSE region (Cooper, 1951; Charlton et al., 1995; Becraft,
or the EMB. These included modules M1, M2, M8, M9, M10, M17, 2001). This may explain the modest level of correlation of M10 with
M12, M15, and M18. As expected, the CSE-correlated M10 was CZ (r = 0.29; Figure 4A).
shown to be enriched for “starch biosynthetic process” and “gly- The AL-correlated module M12 was enriched for “single-organism
cogen biosynthetic process” (Supplemental Figure 13), with the process” (Supplemental Figure 13), and the two previously de-
former Gene Ontology (GO) category including the Shrunken-2 scribed markers of the aleurone, VPP1 (Wisniewski and Rogowsky,
(Sh2), Brittle-2 (Bt2), STARCH-BRANCHING ENZYME1 (SBE1), 2004) and AL-9 (Gómez et al., 2009), were assigned within this
and Waxy1 genes, which all have well characterized functions in module with high MM values (0.91 and 0.85, respectively).
starch biosynthesis (Shure et al., 1983; Giroux et al., 1994; Blauth Likely due to the greater structural complexity of the EMB
et al., 2002). Additionally, close inspection of genes in M10 revealed compared with the captured endosperm compartments, the EMB-
that this module also contained other starch synthesis-related specific module M8 was enriched for more diverse GO categories
genes without any current GO annotation, including the Sh1, in comparison to the endosperm-specific modules. The four bi-
Sugary1 (Su1), and STARCH SYNTHASE1 (SS1) genes (Chourey ological processes that were most significantly enriched for this
and Nelson, 1979; James et al., 1995; Commuri and Keeling, 2001). module included “regulation of transcription, DNA-templated,”

Figure 4. (continued).
(E) Relationships of the imprinted gene sets (Xin et al., 2013) with the coexpression modules obtained using WGCNA. The heat map indicates P values
(-log10) of hypergeometric tests of overrepresentation of genes in a given tested pair of gene sets. Imprinted gene sets are noted on the x axis and all
WGCNA coexpression modules on the y axis. Boxes contain the numbers of overlapping genes. Numbers of genes in each imprinted gene set: 7-DAP
MEG, 37; 10-DAP MEG, 185; 15-DAP MEG, 15; 7-DAP PEG, 80; 10-DAP PEG, 50; and 15-DAP PEG, 48.
275
520 The Plant Cell

“floral organ development,” “phyllome development,” and “re- development (Marshall et al., 2011), while the TCRR genes encode
sponse to hormone.” Modules M1, M2, and M9 showed relatively type-A response regulators (Muñiz et al., 2006, 2010). Significantly,
high correlation with both the AL and the EMB (r $ 0.35). Among all of these genes showed high MM (>0.90) to module M18, and
these, M9 also showed a slightly lower correlation with CSE (r = many of them were among the top-scoring hubs (Supplemental
0.33). Among these modules, M2 contained the largest number Figure 14E and Supplemental Data Set 4). Furthermore, the BETL-
(477) of genes. Similar to M8, this module was also enriched for specific TF gene MRP-1, previously shown to be involved in reg-
many GO categories, most of which were biological processes ulation of BETL differentiation (Gómez et al., 2002, 2009), was also
related to DNA replication and mitotic cell division (e.g., “DNA detected in M18 with a high MM score (1.00; Supplemental Data
replication,” “DNA repair,” “mitotic spindle assembly checkpoint,” Set 4). This is consistent with its role as an activator of many BETL-
and “microtubule-based movement,” etc.). Similarly, M9 (contain- specific genes, including BETL-1, BETL-9, BETL-10, BAP-2, MEG-1,
ing 194 genes) was enriched for “cell cycle process,” “chromosome TCRR-1, and TCRR-2 (Gómez et al., 2002, 2009; Gutiérrez-Marcos
segregation,” “regulation of DNA replication,” and “single-organism et al., 2004; Muñiz et al., 2006, 2010). Therefore, these data suggest
organelle organization,” etc. Furthermore, M1, M2, and M9 shared that MRP-1 acts as a major regulator of a subset of genes in the
enrichment for “nucleosome assembly” and “DNA duplex un- M18 coexpression module. Together, our results suggest that the
winding,” which are also biological processes involved in DNA WGCNA-identified coexpression modules can be used as starting
replication and cell division (Supplemental Figure 13). These GO points for identification of GRNs functioning in each endosperm
enrichments suggest that the EMB and AL, and to some degree the compartment.
CSE, are programmed to undergo extensive mitotic cell pro-
liferation at 8 DAP via the coordinated expression of a relatively De Novo Identification of cis-Motifs Associated with the
extensive gene network. Endosperm Coexpression Modules
Consistent with the presumptive function of the ESR in mediating
endosperm-embryo interaction and expressing antimicrobial prod- As a first step toward identification of the endosperm GRNs, we
ucts (Sabelli and Larkins, 2009), the ESR-specific module M15 was used MEME software (Bailey and Elkan, 1994) to detect putative
enriched for “cell-cell signaling involved in cell fate commitment” cis-regulatory elements in upstream gene sequences from each
(Supplemental Figure 13). This module also included many genes of the five endosperm coexpression modules identified using
isolated previously by virtue of their highly specific ESR expression WGCNA, including M10, M12, M15, M17, and M18 (Figure 4A).
pattern, including ESR-1, ESR-2, ESR-3, ESR-6, ESR-6B, and AE-3 We searched for 10- to 12-bp sequence motifs overrepresented
(Schel et al., 1984; Opsahl-Ferstad et al., 1997; Bonello et al., 2000; within 21 to +0.5 kb (relative to transcription start site) of genes
Balandín et al., 2005; Sosso et al., 2010). In our analysis, many of the in each module and further identified motifs that were signifi-
same genes, including ESR-1, ESR-2, ESR-3, and ESR-6, were cantly similar (q-value < 0.05) to the known plant cis-motifs
positioned within the intramodular hubs with high MM (e.g., 0.99) to available in the JASPAR CORE database using TOMTOM program
M15 (Supplemental Figure 14C and Supplemental Data Set 4). (Gupta et al., 2007). We found that most of the detected motifs
As expected from the BETL’s reported role as mediator of were shared among the majority of the modules as exemplified by
sugar and metabolite uptake into the endosperm through its the motifs that contained exclusively CG- or AT-rich sequences
interaction with the underlying maternal placenta-chalazal region (Supplemental Table 9) with significant similarity to the reported
(Sabelli and Larkins, 2009), the BETL-correlated module M18 was binding sites of ABI4 (a maize AP2-EREBP protein) and SOC1 (an
found to be enriched for “transmembrane transport,” “ion trans- Arabidopsis thaliana MIKC-type MADS box protein), respectively
port,” and “sucrose transport” functions (Supplemental Figure 13). (Riechmann et al., 1996; Niu et al., 2002). Coincidently, four of the
The M18 module contained nearly all of the previously identified five endosperm modules (M10, M12, M15, and M17) included at
BETL-expressed genes (Hueros et al., 1995, 1999; Cheng et al., least one MIKC-type MADS box gene, while all five modules in-
1996; Doan et al., 1996; Serna et al., 2001; Magnard et al., 2003; cluded at least one AP2-EREBP gene based on the current maize
Gutiérrez-Marcos et al., 2004; Massonneau et al., 2005; Gruis et al., genome annotation (Supplemental Data Set 4), indicating that the
2006; Muñiz et al., 2006, 2010; Brugière et al., 2008; Gómez et al., MIKC-type MADS box and AP2-EREBP families may play broad
2009), including BETL-1, 3, 4, 9, and 10; BAP-1A, 1B, 2, 3A, and regulatory roles in maize endosperm development.
3B; TCRR-1 and 2; and INCW2 (CELL WALL INVERTASE2), In contrast, a small number of motifs were detected specifically
EBE-2 (EMBRYO SAC/BASAL ENDOSPERM-LAYER/EMBRYO- in single modules. In one instance, the Motif 10 that was enriched in
SURROUNDING REGION-2), IPT-2 (ISOPENTENYL TRANS- the CSE-correlated module M10 showed significant similarity to
FERASE-2), and CC-8 (CORN CYSTATIN-8) (Supplemental Data a cis-motif that has been shown to bind a bZIP TF in snapdragon
Set 4). In addition, this module contained 11 of the 13 MEG genes (Antirrhinum majus) (Supplemental Table 9) (Martínez-García et al.,
(with MEG-4 and MEG-14 being the two exceptions) that have 1998). Accordingly, two bZIP TF genes, bZIP46 and the storage
been identified in the B73 genome (Supplemental Data Set 4), in- program regulator O2, were detected within M10 (Supplemental
cluding MEG-1, which is a BETL-specific gene that has been Data Set 4). These results suggest that a subset of genes in module
shown to be important for the development and differentiation of M10 may be regulated by the bZIP genes in the same module. In
the BETL (Costa et al., 2012). The BETL-, BAP-, and MEG-type a second case, Motifs 3, 6, and 8 that were detected in upstream
genes encode small, secreted, cysteine-rich proteins that have sequences of M18 genes each contained a repeated GATA se-
been suggested to protect the embryo from maternally transmitted quence (Figure 5A) similar to a sequence previously shown to be
pathogens (Tailor et al., 1997) and to serve as signaling molecules involved in binding and activation of target genes by MRP-1
that coordinate the supply of nutrients to the embryo during kernel (Baranowskij et al., 1994; Barrero et al., 2006), whereas no
276
Cell Differentiation Modules in Maize Endosperm 521

GATA-containing motifs were detected in any of the other endo-


sperm coexpression modules. In support of a transcriptional regu-
latory role for the motifs, our analysis indicated a biased distribution
of Motifs 3, 6, and 8 upstream to the transcription start site (Figure
5B). A close inspection of the submotifs, namely, the variant se-
quences of each motif, of Motifs 3, 6, and 8 revealed that at least
148 genes in the M18 module contained one or more submotifs
within their upstream sequences. Collectively, our data suggest that
a large subset of genes in the M18 coexpression module is likely
regulated by MRP-1 through binding at GATA-rich sequences.

Identification of a Gene Regulatory Module Associated with


BETL Cell Differentiation

We focused on the BETL-associated coexpression module M18


to decipher a portion of the BETL GRN. The BETL transports
nutrients from the maternal tissue into the endosperm and is
important for the proper development of the endosperm and the
endosperm’s capacity as a storage organ (Thompson et al., 2001;
Costa et al., 2012). Based on the available data (Baranowskij et al., Figure 5. De Novo GATA-Rich Sequence Motifs Overrepresented in the
1994; Gómez et al., 2002, 2009; Gutiérrez-Marcos et al., 2004; Upstream Sequences of the Coexpression Module M18 Genes Identified
Barrero et al., 2006; Muñiz et al., 2006, 2010) including our identi- by the MEME Program.
fication of the coexpression module M18 and its association with
(A) Motifs 3, 6, and 8 enriched among the M18 genes.
GATA-rich sequence motifs (discussed above), we hypothesized
(B) Position frequencies of the same motifs determined as number of
that a GRN for BETL differentiation is minimally composed of MRP-1 motifs per 50-bp bins in the upstream gene regions.
and a large set of target genes that are activated upon binding of
MRP-1 to Motifs 3, 6, and 8. A close examination of all the sub-
motifs of the 10 motifs enriched for M18 showed that each of the BAP-2, BAP-3A, and BAP-3B, were present in this regulatory
motifs is represented by two to 21 submotifs that appeared one to module (Supplemental Table 11). Based on the available functional
90 times within the promoters of the M18 genes (Supplemental annotation of maize genes, this regulatory module contained at least
Table 10). 17 cysteine-rich protein-coding genes (including 3 BETL-, 4 BAP-,
We tested for binding of MRP-1 to Motifs 3, 6, and 8 using and 10 MEG-type genes; Supplemental Table 11), among which
directed yeast one-hybrid (Y1H) assays. We introduced two BETL-1, BETL-2, BETL-10, and MEG-1 have previously been
constructs into yeast cells, one that resulted in expression of shown to be regulated by MRP-1. Notably, six TF genes were de-
MRP-1 and a second that comprised the test sequence fused tected within this regulatory module, including three MYBR-, two
upstream of the yeast AUR1-C gene. AUR1-C confers resistance C2C2-GATA, one AP2-EREBP, and one DBB family genes, sug-
to aureobasidin A (AbA) (Heidler and Radding, 1995). In this assay, gesting strongly that MRP-1 indirectly regulates a subset of the
MRP-1 binding to the test sequence results in AUR1-C activation BETL-expressed genes by regulating these TFs. In addition, this
and, thus, growth of cells on plates containing AbA, whereas ab- regulatory module also included a gene (GRMZM2G406552) that
sence of MRP-1 binding results in absence of growth on plates encodes a putative nonspecific lipid transfer protein (Supplemental
containing AbA. Testing for MRP-1 binding to all of the 24 sub- Table 11). Furthermore, comparison of the genes in the regulatory
motifs comprising Motifs 3, 6, and 8 indicated strong binding to module to the GenBank nonredundant (nr) and Swissprot
four of the submotifs. These submotifs included Motifs 3a/8a, 6a/ protein databases (BLASTX, E-value < 1026) revealed that
8b, 8c, and 8f, with Motifs 3a and 6a identical to 8a and 8b, re- a number of genes encoding putative transporters of peptide
spectively (Table 1; Supplemental Figure 15). Comparison of these (GRMZM2G156794), calcium (GRMZM5G836886), phosphate
submotifs with the previously reported MRP-1 binding sites (GRMZM2G466545), and magnesium (GRMZM2G054632) were also
showed that Motifs 6a/8b and 8c were identical to the reported included in this regulatory module (Supplemental Data Sets 5 and 6).
MRP-1 binding site in the promoters of BETL-1 (Motif IV) and To validate this putative network further, we used Y1H assays
BETL-2, respectively (Barrero et al., 2006), while Motifs 3a/8a and to test for binding of MRP-1 to 22 gene promoters from the M18
8f represent newly identified MRP-1 binding sites. module containing motifs positive for MRP-1 binding. We also
We identified 93 genes that contain at least one of the four tested 19 gene promoters lacking these motifs. Of the 41 promoters
submotifs that were shown to be bound by MRP-1 in the Y1H tested, 31 were shown to bind MRP-1 and 10 did not. Significantly,
assays (Supplemental Table 11). All the 93 genes showed relatively all promoters containing the MRP-1 binding submotifs were
high MM values to M18, ranging from 0.62 to 1.00, with 78 of them shown to bind MRP-1 (Supplemental Figure 16 and Supplemental
higher than 0.95 (Supplemental Table 11). These results suggest Table 12). Interestingly, 7 of the 19 promoters lacking the MRP-1
that these 93 genes constitute a regulatory module of the MRP-1- binding submotifs were found to bind MRP-1 in our assays, indi-
regulated GRN. Significantly, 7 of the 14 BETL-expressed genes cating that MRP-1 binds to additional sequences not identified in
described previously, including MEG-1, BETL-1, BETL-10, BAP-1A, the MEME analysis.
277
522 The Plant Cell

These results suggest strongly that the 93 genes containing a large subset of M18 genes constitutes a portion of the
the MRP-1 binding submotifs, as well as many additional genes MRP-1-regulated gene network and that these genes are acti-
(at least seven), are directly regulated by MRP-1 and that this vated by MRP-1 through binding to specific upstream cis-regulatory
gene set constitutes a regulatory module within the BETL GRN sequences.
(Figure 6A). If so, these genes should exhibit a similar temporal
pattern of expression to that of MRP-1. An analysis of the ex-
DISCUSSION
pression of the 93 putative MRP-1 target genes throughout
development using the RNA-Seq data generated and/or processed We used an LCM RNA-Seq profiling approach to comprehensively
by Chen et al. (2014) showed that a large subset of these genes detect mRNA populations for 10 filial and maternal compartments
(including most of the cysteine-rich protein-coding genes) were of an 8-DAP maize kernel, and subsequently identified highly cor-
primarily expressed in the endosperm peaking at 6 to 8 DAP, related gene expression programs associated with each compart-
mirroring the pattern of MRP-1 expression (Figure 6B). Interestingly, ment using WGCNA. The endosperm coexpression modules are
this set included genes that displayed restricted patterns of mRNA expected to reflect the state of cellular differentiation within in-
accumulation throughout development, with most showing low dividual endosperm compartments or cell types. Our data indicate
mRNA prevalence in the vegetative organs, but showing a wide that the timing and extent of these differentiation processes are
range of mRNA levels in the 6- to 8-DAP endosperm/kernel unique to each compartment as suggested previously (Olsen,
(Supplemental Figure 17). In comparison to the rest of the M18 2001, 2004; Sabelli and Larkins, 2009; Becraft and Gutierrez-
genes, the MRP-1 regulatory module genes showed a significantly Marcos, 2012; Leroux et al., 2014; Li et al., 2014). As a test case,
higher level of expression in the 8-DAP BETL (Figure 6C; P value = we deciphered an MRP-1 regulatory module containing 93 genes
6.4e-15 based on unpaired t test). As the M18 genes showed that are likely involved in BETL cellular differentiation.
significant overlap with the genes in the two temporal programs The high-quality RNAs isolated from the laser-captured cells
up@6DAP and up@8DAP (Figure 4D), we examined the overlap (Supplemental Figures 2 and 3 and Supplemental Table 1) and
between the 93 genes containing the MRP-1 binding submotifs the resulting highly reproducible RNA-Seq data (Supplemental
and the two temporal clusters. The result showed that 49 of the 93 Figure 4) enabled us to detect mRNAs of 30,666 genes accumu-
genes lie within the up@6DAP gene set (Figure 6D; hypergeometric lated in at least one captured compartment (Supplemental Data Set
P value = 8.1e-89), while only three genes were found to overlap 2) and 28,078 genes expressed in at least one endosperm com-
with the up@8DAP gene set, indicating that a significant portion of partment (Supplemental Figure 6B). The latter is similar to the
the MRP-1 regulatory module is coordinately upregulated by 33,084 genes that we previously detected as expressed in the
MRP-1 at 6 DAP in BETL. Taken together, these results suggest that 8-DAP whole endosperm (Li et al., 2014). The difference is likely
due to the use of different cutoff criteria for defining genes as ex-
pressed in the two studies and to the fact that we likely did not
Table 1. Results of Y1H Assays for Binding of MRP-1 to the MEME-
collect every portion of the 8-DAP endosperm in our LCM analysis.
Identified Sequence Motifs of M18
Application of a CS scoring method to the 30,666 expressed
Submotif Sequencea Y1H Resultsb genes identified 13,009 genes that were predominantly expressed
in single compartments (Supplemental Data Set 3). These cell-
3a/8a TAGATAGATAGA +
3b TAGATATATAGA – specific patterns were validated with 20 genes using in situ hy-
3c TAGATAAATAGA – bridization and by comparisons with the expression patterns of
6a/8b TAGATATAGATA + previously reported endosperm-expressed genes (Supplemental
6b TAGATATATATA – Figure 7 and Supplemental Table 6). In all 44 cases tested, the CS
6c TAAATATAAATA – patterns closely matched the experimentally observed patterns,
6d TAGATATAGAAA – indicating that our RNA-Seq data accurately reflect the accumu-
6e TAGATATAAATA – lation of endogenous mRNAs in the endosperm.
6f TAAATATAAAAA – As expected, the PCA and SCC analysis showed that each
6g TAAATATATAAA –
captured compartment exhibited a distinct mRNA population,
6h TATATATATATA –
with the maternal and filial compartments forming two separate
6i TAAATATATATA –
6j TAGATAAAAAAA – groups. Consistent with this, WGCNA identified both coexpression
6k TAAATATAGAAA – modules that were correlated with multiple compartments (r $ 0.35)
6l TAGATATATAAA – and modules that were specifically correlated with each of the
6m TAGATATAAAAA – captured compartments (r $ 0.85 for one compartment and r <
8c TAGATAGAGATA + 0.35 for other compartments), with the filial and maternal com-
8d TAGAGAGAGAGA – partment-correlated coexpression modules falling into two nearly
8e TAGATATAGAGA – distinct groups (Figure 4A). Notably, the SCC analysis also showed
8f TAGATAGATATA + that the AL was more closely related to the EMB than to any of the
8g AAGATAGAGAGA –
other endosperm compartments (r = 0.80; Figure 2B). Accordingly,
8h TAGATAGAGAGA –
WGCNA identified three coexpression modules (M1, M2, and M9)
a
Flanking sequence information is provided in Supplemental Table 14. that showed relatively high correlation (r $ 0.35) to both AL and
b
Scoring: + indicates strong positive; – indicates negative. Images of the EMB, with M9 also exhibiting a modest correlation to CSE (r = 0.27)
Y1H growth assays are shown in Supplemental Figure 15.
(Figure 4A). GO analysis indicated that these modules were
278
Cell Differentiation Modules in Maize Endosperm 523

Figure 6. A Regulatory Module of the MRP-1-Regulated GRN Based on the Analysis of MRP-1-Bound Submotifs.

(A) A network of the 93 M18 genes associated with at least one MRP-1-bound submotifs visualized using Cytoscape (Shannon et al., 2003). Thickness
of the arrows indicates the number of motifs associated with each gene.
279
524 The Plant Cell

enriched for genes involved in mitotic cell proliferation (Supplemental Analysis of endosperm-imprinted genes (Xin et al., 2013) in the
Figure 13). This suggests that partially overlapping sets of cell pro- WGCNA-identified coexpression modules (Figure 4E) revealed that
liferative programs distinguish the AL and EMB from the rest of the the BETL- and CZ-associated modules (M17 and M18, re-
filial compartments of an 8-DAP kernel. spectively) were enriched for PEGs, the CSE-correlated module
GO enrichment of the coexpression modules correlated with M10 was enriched for MEGs, while the ESR-associated module
single endosperm compartments (M10, M12, M15, M17, and M15 was enriched for both PEGs and MEGs. These observations
M18) were generally consistent with the presumptive functions indicate that the five endosperm compartment-correlated modules
of each compartment as described previously (Olsen, 2001, identified in this study can be differentiated in part by memberships
2004; Sabelli and Larkins, 2009; Becraft and Gutierrez-Marcos, of imprinted genes that are likely a reflection of compartment’s
2012). This indicates that these coexpression modules can be function. For example, BETL and CZ presumably function in the
used to decipher the key regulatory programs associated with nutrient transport from the maternal tissue to the inner endosperm
cellular differentiation and related functions of each compart- cells including the developing SE (Becraft, 2001; Sabelli and
ment. In support of this notion, these modules were shown to Larkins, 2009). The association of these compartments with the
contain many cell type-specific genes that had been described expression of PEGs is consistent with the parental conflict model,
previously and also identified by us using the CS scoring method which predicts opposite roles for MEGs and PEGs in regulating
(Figure 4C; Supplemental Data Set 4). The latter, in effect, con- nutrient allocation from the mother to offspring (Haig and Westoby,
stitutes a high-resolution atlas of spatially specific gene expression 1989; Moore and Haig, 1991). On the other hand, the association of
programs in the differentiating maize endosperm. CSE with some MEGs and the dual association of ESR with PEGs
We previously performed RNA-Seq with whole kernels or iso- and MEGs suggest a more complex relationship between gene
lated endosperm at eight temporal stages and identified gene sets imprinting and endosperm cell function.
exhibiting temporal patterns of gene expression (Li et al., 2014). The CSE, as a major subregion within the SE, is responsible
Among these were temporally upregulated genes exhibiting rela- for the storage of most starch and storage proteins in the endosperm
tively low expression at early stages and relatively high expression (Olsen, 2001, 2004; Sabelli and Larkins, 2009). Correspondingly, as
at later stages. Analysis of the temporally upregulated genes revealed by the GO enrichment analysis, the CSE-correlated
among the WGCNA coexpression modules detected significant module M10 included many starch biosynthetic genes (Supplemental
enrichment of the up@6DAP and/or up@8DAP temporal programs Figure 13). In addition, a number of early zein genes, as well as the
in the endosperm compartment-specific modules including the storage program regulators O2 and PBF, were also detected in this
BETL-correlated M18 (Figure 4D). Similarly, a broad survey of the module (Supplemental Data Set 4). Because the expression of
endosperm-correlated modules for expression during development a large subset of the storage protein genes, considered to be
indicated that these modules follow distinct dynamics of expres- regulated by O2 and/or PBF, is not fully activated by 8 DAP, only
sion with an onset of activation at ;6 DAP, yet the ESR, BETL, and a few known targets of O2 and PBF were detected in our data set.
AL (M15, M18, and M12, respectively) show a more rapid down- These included the 15-kD b-zein gene and the cyPPDK1 gene
regulation pattern compared with the CZ and CSE (M17 and M10, (encoding a cytoplasmic pyruvate orthophosphate dikinase) known
respectively) (Supplemental Figures 9 to 11). These data suggest to be regulated by O2, and the 27-kD g-zein gene regulated by PBF
that although the endosperm compartments captured for this study (Cord Neto et al., 1995; Gallusci et al., 1996; Maddaloni et al., 1996;
already exhibit characteristics of the differentiated state, the asso- Marzábal et al., 2008). Therefore, although our data may not allow
ciated coexpression modules nonetheless reflect active regulatory us to fully decipher the GRNs regulating the storage function of the
processes that may underlie continuous differentiation and speciali- SE, they provide an insight into the early phase of the storage
zation of the relevant cell types. However, the fact that the previously program activation.
identified temporal programs lack spatial resolution within the en- The BETL-correlated module M18 contained numerous pre-
dosperm limits our ability to accurately correlate the spatial programs viously described BETL-specific genes (Supplemental Data Set 4),
with the developmental dynamics of individual cell types at this time. including MRP-1 and seven genes regulated by MRP-1 (Gómez
Therefore, further capture and analysis of endosperm compartments et al., 2002, 2009; Gutiérrez-Marcos et al., 2004; Muñiz et al., 2006,
from multiple early stages of endosperm will enable a more com- 2010). The identification of GATA-containing motifs among a sub-
prehensive understanding of these regulatory processes. set of M18 genes (Figure 5) and confirmation of binding of MRP-1

Figure 6. (continued).

(B) Expression pattern of the MRP-1 target genes in seed tissues compared with vegetative and reproductive tissues based on the log2-transformed
RPKM data from Chen et al. (2014). The fraction of MRP-1 target genes that have available expression data is indicated in parentheses. Blue line
indicates the expression of MRP-1 itself. The selected data include vegetative (Veg) tissues shoots (Sh), roots (R), leaf (L), shoot apical meristem (SAM;
replicate 1); reproductive (Rep) tissues ear (E), tassel (T; replicate 1), preemergence cob (C), silk (Si), anther (A), ovule (O), and pollen (P); and whole
kernels, endosperm, and embryos of different developmental stages (in DAP).
(C) Expression level of genes within the MRP-1-regulated regulatory module in BETL compared with all other genes in M18. Asterisk indicates P value <
1025 (unpaired t test).
(D) Venn diagram showing overlaps between the genes that are upregulated at 6 DAP (Li et al., 2014) with those of the M18 coexpression module and
those with detected MRP-1 binding sites. Asterisk indicates hypergeometric P value < 1025.
The boxes in (B) and (C) represent the interquantile range, green lines the median, red dots the mean, and whiskers 1.5 times the interquantile range.
280
Cell Differentiation Modules in Maize Endosperm 525

to these sequences using Y1H assays (Supplemental Table 11) this data set can be used as a starting point to dissect the modules
allowed us to propose a 93-gene, direct-target regulatory module regulating endosperm cell differentiation. The analyses provided
for MRP-1 (Figure 6). Functional annotation of the 93-gene regu- here constitutes a significant step toward the identification of GRNs
latory module indicates a diverse array of gene functions activated that regulate maize endosperm cell differentiation and determine its
by MRP-1, including putative signaling and nutrient transport function.
(Supplemental Table 11). The larger M18 gene set also exhibits
a wide range of putative functions including those expected for METHODS
a transfer cell layer (Supplemental Figure 13). These gene functions
are likely sufficient to support BETL differentiation, as a recent Plant Materials and Growth
study showed that the ectopic expression of MRP-1 via an aleu-
rone-specific gene promoter was capable of inducing differentia- Plants of the reference maize (Zea mays) genotype, B73, were grown
tion of an ectopic BETL in the aleurone albeit in a transient manner under greenhouse conditions (16-h day) at the University of Arizona during
April to July 2012 and self-pollinated to obtain 8-DAP kernels (Li et al., 2014)
(Gómez et al., 2009). On the other hand, ectopic activation of these
for LCM. The kernels for morphological analysis shown in Supplemental
genes may not produce viable cells outside the endosperm con-
Figure 1 were collected during October and November 2013. Kernel
text, as attempts with overexpressing MRP-1 in maize using compartments were delineated for capture using tissue sections ob-
ubiquitously expressing promoters produced no transformants tained from Farmer’s fixed (see below) or from paraformaldehyde-fixed
(Gómez et al., 2009). Further understanding of the MRP-1-regulated material that was further stained with Toluidine Blue as described
network and the associated gene functions will likely require previously (Drews, 1998). The tissue sections for each biological replicate
characterization of complete or partial loss-of-function mutants. of a given compartment were captured from multiple kernels of a single ear
The gene set activated by MRP-1 is likely significantly larger (Supplemental Table 1).
than the 93-gene set discussed above for two reasons. First, our
Y1H assays showed activation of seven genes that lack motifs LCM, RNA Isolation, and cDNA Synthesis and Amplification
identified in our MEME analysis (Supplemental Figure 16 and
Kernels were harvested from plants mid-day and cut at the pedicel,
Supplemental Table 12), indicating that MRP-1 binds to additional punctured through the pericarp, vacuum infiltrated with cold Farmer’s
sequences not identified in our MEME analysis and that M18 fixative (ethanol:glacial acetic acid, 3:1) (Kerk et al., 2003), and stored in
contains additional MRP-1-regulated genes. Second, MRP-1 likely cold fixative overnight. Fixed kernels were then dehydrated in a graded
directly regulates six TFs, including MYBR24, MYBR33, GATA7, ethanol series, cleared in n-butanol, embedded in Paraplast X-tra (Mc-
GATA33, and EREB137, and a DBB family TF (Supplemental Table Cormick Scientific Leica) using microwave (Takahashi et al., 2010), cut to
11). Each of these TFs may activate a gene set within M18. Addi- 10-mm sections, and mounted on PEN-coated slides. Shortly before
tional protein-DNA interaction studies, such as electrophoretic capture, sections were deparaffinized in Xylenes and air-dried (Takahashi
mobility shift assays, and transient directed-expression assays et al., 2010). Individual cell types or kernel compartments were captured
directly into an aliquot of Arcturus PicoPure RNA extraction buffer (Ap-
would be necessary to further characterize the interaction between
plied Biosystems) using a Leica LMD6500 laser microdissection system
MRP-1 and the full spectrum of its direct targets. Thus, the entire
(Leica Microsystems). RNA was extracted following the manufacturer’s
regulatory module activated by MRP-1 (i.e., both directly and in- suggested protocol (Arcturus PicoPure kit), checked for quality, DNase
directly) probably encompasses a much larger proportion of the treated using TURBO DNase (Life Technologies), and further purified
M18 module than what we report here. Studies devoted to identi- using the Arcturus PicoPure columns (Applied Biosystems). For each of
fying the full spectrum of MRP-1 binding sites and to the identifi- the 22 samples (Supplemental Table 1), 10 ng purified RNA was used for
cation of the target genes activated by the TFs activated by MRP-1 cDNA synthesis and amplification. cDNA synthesis with oligo(dT) and
are in progress to fully characterize this module. random primers and cDNA amplifications were performed using an
Furthermore, it is notable that in addition to MRP-1 itself, M18 Ovation RNA-Seq System V2 kit (Nugen Technologies) following the
contains at least 48 coexpressed TF genes including six MYBR manufacturer’s protocols with minor modifications. The quality and profile
of the RNA and amplified cDNA samples were checked on an Agilent 2100
genes (Supplemental Data Set 4). The latter may regulate over-
Bioanalyzer (Agilent Technologies) using an Agilent RNA 6000 Pico Kit
lapping gene sets with MRP-1, possibly through binding to the
(Agilent Technologies) and an Agilent High Sensitivity DNA Kit (Agilent
identified GATA-containing motifs. Limitations of the de novo cis- Technologies), respectively.
motif detection approaches utilized here and an absence of an
extensive cis-motif-TF database for plants have precluded identi-
RNA-Seq Library Construction and Sequencing
fication of additional TF targets in M18. Therefore, further ap-
proaches including directed Y1H assays for specific TF-target Construction and sequencing of RNA-Seq libraries were performed at the
interactions in combination with transient expression studies of the University of Arizona Genetics Core. Using an Illumina TruSeq DNA
TFs and analysis of any available TF gene mutants will be required sample preparation kit (v2; Illumina), nearly 1 µg amplified cDNA for each
to determine the full extent of the BETL GRN. of the 22 samples was used to generate multiplexed RNA-Seq libraries
(mean size 350 to 380 bp, including 120-bp adapters) by following the
In summary, our data set provides a high-resolution atlas of
manufacturer’s suggested procedures. The samples were sequenced in
gene expression in differentiating endosperm compartments
batches on four flow cell lanes of an Illumina HiSequation 2000 platform
and maternal compartments of an early maize kernel. This data using a TruSeq SBS kit (v3) to produce 2 3 100-nucleotide paired-end
set provides insights into the functions of the endosperm cell reads. Two of the four lanes each contained libraries for a single replicate
types and into the coexpressed gene sets that establish the dif- of the captured endosperm compartments (AL, BETL, ESR, CSE, and CZ);
ferentiated states and functions of these cell types. Furthermore, as the third lane contained libraries for a single replicate of the endosperm
exemplified by our initial analysis of the MRP-1 regulatory module, compartments plus the library for a single replicate of the maternal
281
526 The Plant Cell

compartment PC; and the fourth lane contained libraries for EMB (three compartments) or low coefficient of variation of FPKM (CV < 1 among 10
replicates) and the other three maternal compartments (NU, PE, and PED, compartments) were filtered out. Using the FPKM values of the remaining
one replicate each). After the raw reads were generated, adapter se- 9361 genes, a matrix of pairwise SCCs between all pairs of genes was
quences were trimmed using the Trimmomatic program (Bolger et al., created and transformed into a matrix of connection strengths (an ad-
2014). Quality of the trimmed reads was checked using the FastQC jacency matrix) by raising the correlation matrix to the power
  b 
program (Andrews, 2010).
b ¼ 12 connection  strength ¼ 1þcorrelation
2 . The power b was in-

Reads Mapping and Analysis terpreted as a soft threshold of the correlation matrix. The resulting
adjacency matrix was then converted to a topological overlap (TO) matrix
RNA-Seq reads were aligned to the maize reference genome version 3 by the TOMsimilarity algorithm. Genes were hierarchically clustered
(B73 RefGen_v3) (Hubbard et al., 2002) using TopHat v2.0.9 (Trapnell based on TO similarity. The Dynamic Tree Cut algorithm was used to cut
et al., 2009). Intron length was set to 30 to 8000 nucleotides, while the hierarchal clustering tree, and modules were defined as branches from
maximum number of mismatches per read was set to 3. To eliminate the the tree cutting. Modules with fewer than 30 genes were merged into their
effect of reads mapping in intergenic and/or repeated genomic regions on closest larger neighbor module. Each module was summarized by the first
the estimation of effective library size that may be caused by the cDNA principal component of the scaled module expression profiles (referred to
amplification method, reads mapped to exonic regions were extracted as module eigengene [ME]). MM (also known as module eigengene-based
using the intersect function of BEDTools v2.17.0 (Quinlan and Hall, 2010) connectivity kME) of a gene to a given module was calculated as PCC
and were provided as input to Cufflinks v2.1.1 (Trapnell et al., 2010) for between the expression levels (FPKMs, in 10 compartments) of the gene
normalization and estimation of gene expression level. The multi-mapped and the ME of the module using the signedKME algorithm. Finally, genes
reads correction and fragment bias correction options of Cufflinks were were reassigned using the moduleMergeUsingKME algorithm to ensure
used. Gene expression levels were reported as FPKM (Mortazavi et al., each gene possesses the highest MM in its own assigned module.
2008). The upper and lower bound FPKM values (FPKM_conf_hi and Module-compartment associations were quantified by PCC analysis
FPKM_conf_lo, respectively) for the 95% confidence interval of each gene where each module was represented by its ME, and each compartment
were also provided by Cufflinks. A gene was defined as expressed in was represented with a numeric vector with “1” for the compartment of
a sample if the FPKM_conf_lo was greater than zero. Information re- interest, and “0” for all other compartments.
garding maize genome annotation used in these analyses was obtained
from Ensembl Plants (plants.ensembl.org, release 19). Computational Validation of Module Robustness
SCC analysis (Zar, 1972; Hollander et al., 2013) was used to quantify
the reproducibility of data between the triplicates of endosperm com- Average TO for each identified coexpression module was calculated and
partments and EMB. SCCs was calculated from log2-transformed FPKM compared with the average TO of modules of the same size generated by
values [i.e., log2 (FPKM + 1)] of the expressed genes. Based on the high randomly assigning the 9361 tested genes to 18 modules; 100,000
correlation of gene expression profiles among the replicated samples, permutations of randomly sampled modules were tested. The observed
exonic reads were merged to create a union of each triplicate, and FPKMs modules were considered robust if the average TOs were significantly
were recalculated for the six merged samples. With the resulting FPKM higher than the randomly generated modules (P value < 1025).
data (log2-transformed) of the expressed genes (FPKM_conf_lo > 0 after
renormalization), PCA and SCC analyses were used to compare gene Visualization of Hub Genes
expression profiles among all 10 compartments. The prcomp and cor.test
functions in R were used for PCA and SCC analysis, respectively. Genes with highest degree of connectivity within a module are referred to
as intramodular hub genes (Langfelder and Horvath, 2008). The top 200
connections (based on topological overlap) among the top 100 genes in
Identification of Compartment-Specific Gene Sets
each module ranked by kME was visualized by VisANT (Hu et al., 2004).
The genes specifically expressed in each kernel compartment were
identified using a CS scoring algorithm that compares the expression level
Gene Annotation and Functional Enrichment Analysis
of a gene in a given compartment with its maximal expression level in the
other nine compartments (Ma and Wang, 2012; Ma et al., 2014). For Locus names and functional annotation of maize genes were obtained
a given gene i, its expression values in 10 compartments are denoted as from Ensembl Plants. The recently annotated MEG family members (Xiong
EVi ¼ ðE1i ; E2i ; E3i ; .; E10i Þ, and the CS score of this gene in compartment j et al., 2014) were also incorporated. Annotation of TF family members
max  E i
is defined as: CSði; jÞ ¼ 1 2 E i k , where 1 # k #10, k  j. Thus, CS scores were based on information from Plant Transcription Factor Database v3.0
j
range from 0 to 1, and the higher the CS score of a gene for a com- (Jin et al., 2014) and GrassTFDB of GRASSIUS (Gray et al., 2009; Yilmaz
partment, the more likely the gene is specifically expressed in that et al., 2009). Annotation of zein genes in the B73 genome was based on
compartment. the information as summarized by Chen et al. (2014). Putative functions of
genes of interest were identified and inspected manually. cDNA se-
quences of the longest isoform of each gene were obtained from Ensembl
In Situ Hybridization Localization of mRNAs
Plants and used for homology searches against the NCBI nr and Swis-
Staged kernels were obtained from B73 plants grown at the University of sprot protein databases, respectively, using the BLASTX program. Only
Utah. Kernels were harvested, fixed, processed, and hybridized to probes the top five hits for each gene (E-value < 1026) were considered in our
as previously described (Li et al., 2014). Primers used to generate the analysis of putative gene functions.
probe clones are listed in Supplemental Table 13. GO term enrichment analyses of the WGCNA-identified coexpression
modules were performed using a modified Fisher’s exact test in Blast2GO
software (false discovery rate < 0.05). GO annotations for maize genes
Identification of Coexpression Modules
were obtained from Gramene (gramene.org, release 40). For each module,
The R package WGCNA (Zhang and Horvath, 2005; Langfelder and only protein-coding genes (i.e., transposable elements, microRNA genes,
Horvath, 2008) was used to identify modules of highly correlated genes and pseudogenes excluded) were subject to GO enrichment analysis,
based on the FPKM data. Genes with low FPKM (mean FPKM < 1 for 10 with the most specific biological processes reported.
282
Cell Differentiation Modules in Maize Endosperm 527

Identification of cis-Motifs images of the yeast cells were captured. This procedure was performed
with two colonies per strain for each experiment, and each experiment
MEME program was used to identify de novo motifs in the promoter
was performed twice on separate days with independently transformed
regions of genes in each coexpression module. We defined the promoter
strains and independently prepared plates (i.e., four independent colonies
regions as 1 kb upstream and 500 bp downstream of transcription start
tested per strain). In all cases, all four colonies exhibited very similar
sites and obtained the genomic sequences using a customized Perl
growth patterns.
script. For each module, 10 motifs (Motifs 1 through 10) were reported by
Yeast growth was scored as follows. Cell growth fell into three general
MEME, and the ones with E-value higher than 1026 were excluded manually.
categories. In the first category, the +MRP-1 and 2MRP-1 strains both
Using the TOMTOM motif comparison tool, the resulting motifs were aligned
grew well on plates lacking AbA but failed to grow on plates containing
with motifs in the JASPAR CORE Plantae database (Mathelier et al., 2014) to
AbA. This growth pattern indicated no binding of MRP-1 to the test
identify significantly similar known cis-motifs (q-value < 0.05). A customized
sequence and these strains were scored as negative. In the second
Perl script was used to identify the genes that contain at least one of the four
category, the 2MRP-1 strains grew well on plates lacking AbA but failed
submotifs shown to be bound by MRP-1 in the Y1H assays.
to grow on plates containing AbA, and the +MRP-1 strains grew equally
well (or nearly so) on plates lacking and containing AbA. This growth
Y1H Assays pattern indicated strong binding of MRP-1 to the test sequence and these
strains were scored as positive. In the third category, growth of the +MRP-1
The Matchmaker Gold kit from Clontech for Y1H assays was used to strains was reduced on plates containing AbA relative to plates lacking AbA
validate MRP-1-target sequence interactions following the manu- but this growth was significantly greater than that of the –MRP-1 strains on
facturer’s procedures with modifications, including the use of the yeast plates containing AbA. This growth pattern suggested weak binding of MRP-1
CUP1 promoter (Etcheverry, 1990) in place of the yeast ADH promoter to to the test sequence and these strains were scored as weak positives.
drive MRP-1 expression in yeast cells. The plasmid containing the MRP-1
coding region driven by the CUP1 promoter was generated as follows. Accession Numbers
The CUP1 promoter was provided in plasmid pMB465 by Marcus Babst.
A NotI/SacI fragment containing the CUP1 promoter fragment was Sequence data for the genes cited in this article can be found in the
subcloned into yeast vector pRS425 to make pCUP425. The MRP-1 open Gramene database or GenBank/EMBL libraries under the following accession
reading frame was generated by PCR using primers MRP1-FEco numbers: 15-kD b-zein, GRMZM2G086294; 16-kD g-zein, GRMZM2G060429;
(59-GGCCGAATTCAATCCCAACTTCAACAGTGTG-39) and MRP1-RBam 18-kD d-zein, GRMZM2G100018; 27-kD g-zein, GRMZM2G138727; AL-9,
(59-GGCCGGATCCTCGGTTATATATCTGGCTCTCC-39). The resulting PCR GRMZM2G091054; BAP-1A, GRMZM2G008271; BAP-1B, GRMZM2G008403;
product was cloned into pGADT7 (Clontech) using the EcoRI and BamHI BAP-2, GRMZM2G152655; BAP-3A, GRMZM2G133382; BAP-3B,
restriction sites introduced during PCR. The resulting plasmid was called GRMZM2G133370; BETL-1, GRMZM2G082785; BETL-3, GRMZM2G175976;
MRP1/pGADT7. Using primers NoAD-FNot (59-GATCGCGGCCGCATG- BETL-4, GRMZM2G073290; BETL-9, GRMZM2G087413; BETL-10,
GAGTACCCATACGACG-39) and PGAD-R2Sal (59-GATCGTCGACGGAA- GRMZM2G091445; Bt2, GRMZM2G068506; cyPPDK1, GRMZM2G306345;
TATGTTCATAGGGTAG-39), fragments containing an HA tag, the MRP-1 EBE-2, GRMZM2G167733; ESR-1, GRMZM2G046086; ESR-2,
open reading frame, and the ADH1 terminator were amplified. The resulting GRMZM2G315601; ESR-3, GRMZM2G140302; ESR-6, GRMZM2G048353;
PCR product was cloned into pCUP425 through the use of the NotI and SalI ESR-6B, GRMZM2G437040; Floury-1, GRMZM2G094532; INCW2,
restriction sites introduced during PCR. The final plasmid was called MRP1/ GRMZM2G119689; IPT-2, GRMZM2G084462; MEG-1, GRMZM2G354335;
pCUP425. The identical plasmid lacking the MRP-1 coding region was called MEG-2, GRMZM2G502035; MEG-3, GRMZM2G344323; MEG-4,
pCUP425. GRMZM2G137959; MEG-6, GRMZM2G094054; MEG-7, GRMZM2G116212;
The motif:AUR1-C constructs included 147 bp of sequence upstream MEG-8, GRMZM2G123153; MEG-9, GRMZM2G088896; MEG-10,
of the translational start codon of TCRR-1. The motif-promoter fragments GRMZM2G086827; MEG-11, GRMZM2G181051; MEG-12, GRMZM2G175896;
were generated by PCR amplification using the primers listed in Supplemental MEG-13, GRMZM2G175912; MEG-14, GRMZM2G145466; MRP-1,
Table 14. The promoter:AUR1-C constructs included 257 to 1187 bp of GRMZM2G111306; O2, GRMZM2G015534; PBF, GRMZM2G146283; Sh1,
sequence upstream of the translational start codon. The promoter fragments GRMZM2G089713; Sh2, GRMZM2G429899; SS1, GRMZM2G129451; Su1,
were generated by PCR amplification using the primers listed in Supplemental GRMZM2G138060; TCRR-1, GRMZM2G016145; TCRR-2, GRMZM2G090264;
Table 15. The resulting PCR products were cloned into pAbAi (Clontech) VPP1, GRMZM2G069095; Waxy1, GRMZM2G024993; bZIP46,
through the use of unique 59 and 39 restriction sites introduced during PCR GRMZM2G037910; EREB137, GRMZM2G028386; GATA33, GRMZM2G048850;
(Supplemental Tables 14 and 15). The resulting plasmids then were then GATA7, GRMZM2G118214; MYBR24, GRMZM2G049695; and MYBR33,
integrated into the yeast genome of strain Y1HGold following the manu- GRMZM2G422083. The data reported in this article have been deposited in
facturer’s recommended protocol (Clontech). the Gene Expression Omnibus database (www.ncbi.nlm.nih.gov/geo; acces-
We introduced the MRP1/pCUP425 plasmid into each of the promoter: sion number GSE62778).
AUR1-C and motif:AUR1-C yeast strains; these were called +MRP-1
strains in the Y1H figures. To generate control strains, we also introduced Supplemental Data
pCUP425 into each of the promoter:AUR1-C yeast strains and motif:AUR1-C;
Supplemental Figure 1. Structure of an 8-DAP B73 maize kernel.
these were called -MRP-1 (or empty vector) strains in the Y1H figures. Yeast
transformations were performed using the lithium acetate procedure. Supplemental Figure 2. Representative images of kernel compart-
Yeast growth assays were performed as follows. Equal numbers of ments that were marked and collected for the LCM RNA-Seq analyses
cells from single colonies were spotted in a 1:5 dilution series (;625 cells/ reported in this study.
spot, ;125 cells/spot, ;25 cells/spot, and ;5 cells/spot) onto plates. The Supplemental Figure 3. Representative quality assessments of LCM-
plate media was deficient in leucine and contained 750 to 2000 ng/mL derived RNAs used in sequencing.
AbA (variation in the concentration of AbA used was due to manufacturer
Supplemental Figure 4. Reproducibility of RNA-Seq reads for each
batch variability) and 0 or 0.1 mM CuSO4. As a growth control, each cell
triplicate of the filial compartments.
suspension was also spotted onto plates lacking AbA. This procedure
was followed for both the +MRP1 and 2MRP strains. The plates con- Supplemental Figure 5. Distribution profile of RNA-Seq reads along
taining the spotted yeast were incubated at 30°C for ;48 h and then the length of the gene models.
283
528 The Plant Cell

Supplemental Figure 6. Profiles of sequenced RNAs from the Supplemental Table 13. Sequences of the primers used to generate
captured kernel compartments. clones for the in situ hybridization probes.
Supplemental Figure 7. Validation of the compartment-specific Supplemental Table 14. Sequences of the primers used to generate
expression patterns using in situ hybridization localization of the mRNAs. the constructs to test the submotifs in Y1H assays.
Supplemental Figure 8. Z-score plots showing expression profiles of Supplemental Table 15. Sequences of the primers used to generate
all genes in the WGCNA-generated coexpression modules M1-M18. the constructs to test the promoters in Y1H assays.
Supplemental Figure 9. Expression pattern of genes in the endo- Supplemental Data Set 1. Normalized expression levels in FPKM for
sperm-associated modules identified by WGCNA based on previously the 29,369 expressed genes expressed in at least one of the 22
reported RNA-Seq data. sequenced samples.
Supplemental Figure 10. Expression pattern of genes in the endo- Supplemental Data Set 2. Normalized expression levels in FPKM for
sperm-associated modules identified by WGCNA based on RNA-Seq the 30,665 genes expressed in at least one of the ten compartments
data from Chen et al. (2014). after merge of replicates.
Supplemental Figure 11. Expression pattern of genes in the WGCNA Supplemental Data Set 3. Compartment specificity scores for the
coexpression modules (M1, M2, and M9) associated with both AL and 13,009 compartment-specific genes.
EMB (in comparison to the other two EMB-associated modules M7
Supplemental Data Set 4. Module assignment and module member-
and M8) based on RNA-Seq data from Chen et al. (2014).
ships for the 9361 genes selected for WGCNA.
Supplemental Figure 12. Relationships of new sets of allele-biased
Supplemental Data Set 5. BLASTX search of the 93 M18 genes that
genes with the coexpression modules obtained using WGCNA.
contain at least one MRP-1 binding submotifs against the NCBI nr
Supplemental Figure 13. Enrichment of biological processes in the database (E-value < 1026).
filial compartment-correlated modules.
Supplemental Data Set 6. BLASTX search of the 93 M18 genes that
Supplemental Figure 14. Visualization of the five endosperm compart- contain at least one MRP-1 binding submotifs against the Swissprot
ment-correlated coexpression modules using the VisANT program. database (E-value < 1026).
Supplemental Figure 15. Yeast one-hybrid assays for binding of
MRP-1 to the sequence motifs listed in Table 1.
ACKNOWLEDGMENTS
Supplemental Figure 16. Yeast one-hybrid assays for binding of
MRP-1 to the promoters of the genes listed in Supplemental Table 12. We thank John Harada, Julie Pelletier, and Bob Goldberg for suggestions
and advice on improving the protocols for LCM, RNA isolation, and
Supplemental Figure 17. Expression pattern of the 93 genes con-
cDNA amplification. J.Z. was supported by a graduate fellowship from
taining MRP-1 binding submotifs based on RNA-Seq data from Chen
the China Scholarship Council. This work was supported by National
et al. (2014).
Science Foundation Grant IOS-0923880 (to G.N.D. and R.Y.).
Supplemental Table 1. The origins, quality, and quantities of RNAs
isolated using laser-capture microdissection.
AUTHOR CONTRIBUTIONS
Supplemental Table 2. Summary statistics of RNA-Seq reads and
mapping. D.T., D.W., G.N.D., and R.Y. designed the project. D.T., A.L., N.M.N.,
Supplemental Table 3. Number of genes expressed in each of the ten A.M.A., W.J.B., and K.O.L. performed research. J.Z., C.M., and X.W.
compartments. conducted bioinformatics analysis. J.Z., D.T., C.M., X.W., G.N.D., and
R.Y. analyzed the data. J.Z., D.T., G.N.D., and R.Y. wrote the article.
Supplemental Table 4. Number of genes expressed at different FPKM
levels in the ten compartments.
Received December 18, 2014; revised January 28, 2015; accepted
Supplemental Table 5. Summary of principal component analysis of
February 26, 2015; published March 17, 2015.
the ten compartments.
Supplemental Table 6. Summary of the available in situ hybridization
and promoter analysis data for endosperm cell-specific genes.
REFERENCES
Supplemental Table 7. Robustness of WGCNA-generated coexpres-
sion modules based on a permutation test of average topological overlap. Andrews, S. (2010). FastQC: A quality control tool for high-throughput
sequence data. (http://www.bioinformatics.babraham.ac.uk/projects/fastqc).
Supplemental Table 8. Allele-biased genes enriched in the endo-
Bailey, T.L. and Elkan, C. (1994). Fitting a Mixture Model by Expec-
sperm-associated coexpression modules M10, M15, M17, and M18.
tation Maximization to Discover Motifs in Bipolymers. (San Diego,
Supplemental Table 9. Motif analysis of the endosperm compart- CA: University of California).
ment-correlated modules M10, M12, M15, M17, and M18. Balandín, M., Royo, J., Gómez, E., Muniz, L.M., Molina, A., and
Supplemental Table 10. Occurrences of submotifs of each motif Hueros, G. (2005). A protective role for the embryo surrounding
enriched in upstream sequences of M18 genes as identified by MEME. region of the maize endosperm, as evidenced by the character-
isation of ZmESR-6, a defensin gene specifically expressed in this
Supplemental Table 11. Putative functions of the 93 M18 genes that
region. Plant Mol. Biol. 58: 269–282.
contain at least one MRP-1 binding submotifs (Motifs 3a/8a, 6a/8b, 8c,
Baranowskij, N., Frohberg, C., Prat, S., and Willmitzer, L. (1994). A
and 8f).
novel DNA binding protein with homology to Myb oncoproteins
Supplemental Table 12. Results of Y1H assays for binding of MRP-1 containing only one repeat can function as a transcriptional acti-
to the promoters of genes within M18. vator. EMBO J. 13: 5383–5392.
284
Cell Differentiation Modules in Maize Endosperm 529

Barrero, C., Muñiz, L.M., Gómez, E., Hueros, G., and Royo, J. Drews, G.N. (1998). In situ hybridization. Methods Mol. Biol. 82: 353–371.
(2006). Molecular dissection of the interaction between the tran- Etcheverry, T. (1990). Induced expression using yeast copper met-
scriptional activator ZmMRP-1 and the promoter of BETL-1. Plant allothionein promoter. Methods Enzymol. 185: 319–329.
Mol. Biol. 62: 655–668. FAO (2012). FAO Statistical Yearbook 2012. (http://www.fao.org/
Becraft, P.W. (2001). Cell fate specification in the cereal endosperm. docrep/015/i2490e/i2490e00.htm).
Semin. Cell Dev. Biol. 12: 387–394. Faure, J.-E. (2001). Double fertilization in flowering plants: discovery,
Becraft, P.W., and Gutierrez-Marcos, J. (2012). Endosperm de- study methods and mechanisms. C. R. Acad. Sci. III 324: 551–558.
velopment: dynamic processes and cellular innovations underlying Gallusci, P., Varotto, S., Matsuoko, M., Maddaloni, M., and Thompson,
sibling altruism. Dev. Biol. 1: 579–593. R.D. (1996). Regulation of cytosolic pyruvate, orthophosphate
Berger, F., Grini, P.E., and Schnittger, A. (2006). Endosperm: an dikinase expression in developing maize endosperm. Plant Mol. Biol.
integrator of seed growth and development. Curr. Opin. Plant Biol. 31: 45–55.
9: 664–670. Giroux, M.J., Boyer, C., Feix, G., and Hannah, L.C. (1994). Co-
Blauth, S.L., Kim, K.-N., Klucinec, J., Shannon, J.C., Thompson, D., and ordinated transcriptional regulation of storage product genes in the
Guiltinan, M. (2002). Identification of Mutator insertional mutants of starch- maize endosperm. Plant Physiol. 106: 713–722.
branching enzyme 1 (sbe1) in Zea mays L. Plant Mol. Biol. 48: 287–297. Gómez, E., Royo, J., Guo, Y., Thompson, R., and Hueros, G. (2002).
Bolger, A.M., Lohse, M., and Usadel, B. (2014). Trimmomatic: Establishment of cereal endosperm expression domains: identifi-
a flexible trimmer for Illumina sequence data. Bioinformatics 30: cation and properties of a maize transfer cell-specific transcription
2114–2120. factor, ZmMRP-1. Plant Cell 14: 599–610.
Bonello, J.F., Opsahl-Ferstad, H.G., Perez, P., Dumas, C., and Gómez, E., Royo, J., Muñiz, L.M., Sellam, O., Paul, W., Gerentes,
Rogowsky, P.M. (2000). Esr genes show different levels of expression in D., Barrero, C., López, M., Perez, P., and Hueros, G. (2009). The
the same region of maize endosperm. Gene 246: 219–227. maize transcription factor myb-related protein-1 is a key regulator
Brink, R.A., and Cooper, D.C. (1947). The endosperm in seed de- of the differentiation of transfer cells. Plant Cell 21: 2022–2035.
velopment. Bot. Rev. 13: 479–541. Götz, S., García-Gómez, J.M., Terol, J., Williams, T.D., Nagaraj,
Brugière, N., Humbert, S., Rizzo, N., Bohn, J., and Habben, J.E. S.H., Nueda, M.J., Robles, M., Talón, M., Dopazo, J., and Conesa,
(2008). A member of the maize isopentenyl transferase gene family, A. (2008). High-throughput functional annotation and data mining with
Zea mays isopentenyl transferase 2 (ZmIPT2), encodes a cytokinin the Blast2GO suite. Nucleic Acids Res. 36: 3420–3435.
biosynthetic enzyme expressed during kernel development. Cyto- Gray, J., et al. (2009). A recommendation for naming transcription
kinin biosynthesis in maize. Plant Mol. Biol. 67: 215–229. factor proteins in the grasses. Plant Physiol. 149: 4–6.
Charlton, W.L., Keen, C.L., Merriman, C., Lynch, P., Greenland, Gruis, D.F., Guo, H., Selinger, D., Tian, Q., and Olsen, O.-A. (2006).
A.J., and Dickinson, H.G. (1995). Endosperm development in Zea Surface position, not signaling from surrounding maternal tissues,
mays; implication of gametic imprinting and paternal excess in specifies aleurone epidermal cell fate in maize. Plant Physiol. 141:
regulation of transfer layer development. Development 121: 3089–3097. 898–909.
Chen, J., Zeng, B., Zhang, M., Xie, S., Wang, G., Hauck, A., and Lai, Gupta, S., Stamatoyannopoulos, J.A., Bailey, T.L., and Noble, W.S.
J. (2014). Dynamic transcriptome landscape of maize embryo and (2007). Quantifying similarity between motifs. Genome Biol. 8: R24.
endosperm development. Plant Physiol. 166: 252–264. Gutiérrez-Marcos, J.F., Costa, L.M., Biderre-Petit, C., Khbaya, B.,
Cheng, W.-H., Taliercio, E.W., and Chourey, P.S. (1996). The Miniature1 O’Sullivan, D.M., Wormald, M., Perez, P., and Dickinson, H.G.
seed locus of maize encodes a cell wall invertase required for normal (2004). maternally expressed gene1 Is a novel maize endosperm
development of endosperm and maternal cells in the pedicel. Plant Cell 8: transfer cell-specific gene with a maternal parent-of-origin pattern
971–983. of expression. Plant Cell 16: 1288–1301.
Chourey, P.S., and Nelson, O.E. (1979). Interallelic complementation Haig, D., and Westoby, M. (1989). Parent-specific gene expression
at the sh locus in maize at the enzyme level. Genetics 91: 317–325. and the triploid endosperm. Am. Nat. 134: 147–155.
Commuri, P.D., and Keeling, P.L. (2001). Chain-length specificities Hamamura, Y., Nagahara, S., and Higashiyama, T. (2012). Double
of maize starch synthase I enzyme: studies of glucan affinity and fertilization on the move. Curr. Opin. Plant Biol. 15: 70–77.
catalytic properties. Plant J. 25: 475–486. Hansey, C.N., Vaillancourt, B., Sekhon, R.S., de Leon, N., Kaeppler,
Conesa, A., and Götz, S. (2008). Blast2GO: A comprehensive suite for S.M., and Buell, C.R. (2012). Maize (Zea mays L.) genome diversity as
functional analysis in plant genomics. Int. J. Plant Genomics 2008: revealed by RNA-sequencing. PLoS ONE 7: e33071.
619832. Heidler, S.A., and Radding, J.A. (1995). The AUR1 gene in Saccharomyces
Conesa, A., Götz, S., García-Gómez, J.M., Terol, J., Talón, M., and cerevisiae encodes dominant resistance to the antifungal agent aur-
Robles, M. (2005). Blast2GO: a universal tool for annotation, visualization eobasidin A (LY295337). Antimicrob. Agents Chemother. 39: 2765–2769.
and analysis in functional genomics research. Bioinformatics 21: 3674– Holding, D.R., Otegui, M.S., Li, B., Meeley, R.B., Dam, T., Hunter,
3676. B.G., Jung, R., and Larkins, B.A. (2007). The maize floury1 gene
Cooper, D.C. (1951). Caryopsis development following matings between encodes a novel endoplasmic reticulum protein involved in zein
diploid and tetraploid strains of Zea mays. Am. J. Bot. 38: 702–708. protein body formation. Plant Cell 19: 2569–2582.
Cord Neto, G., Yunes, J.A., da Silva, M.J., Vettore, A.L., Arruda, P., Hollander, M., and Wolfe, D.A. Chicken, E. (2013). Nonparametric
and Leite, A. (1995). The involvement of Opaque 2 on beta-pro- Statistical Methods. (New York: John Wiley & Sons).
lamin gene regulation in maize and Coix suggests a more general Hu, Z., Mellor, J., DeLisi, C. (2004). Analyzing networks with VisANT.
role for this transcriptional activator. Plant Mol. Biol. 27: 1015–1029. In Current Protocols in Bioinformatics, Andreas D. Baxevanis, ed
Costa, L.M., Yuan, J., Rouster, J., Paul, W., Dickinson, H., and (New York: John Wiley & Sons), pp. 8.8.1–8.8.24.
Gutierrez-Marcos, J.F. (2012). Maternal control of nutrient allocation in Hubbard, T., et al. (2002). The Ensembl genome database project.
plant seeds by genomic imprinting. Curr. Biol. 22: 160–165. Nucleic Acids Res. 30: 38–41.
Doan, D.N., Linnestad, C., and Olsen, O.-A. (1996). Isolation of Hueros, G., Royo, J., Maitz, M., Salamini, F., and Thompson, R.D.
molecular markers from the barley endosperm coenocyte and the (1999). Evidence for factors regulating transfer cell-specific ex-
surrounding nucellus cell layers. Plant Mol. Biol. 31: 877–886. pression in maize endosperm. Plant Mol. Biol. 41: 403–414.
285
530 The Plant Cell

Hueros, G., Varotto, S., Salamini, F., and Thompson, R.D. (1995). Mathelier, A., et al. (2014) JASPAR 2014: an extensively expanded
Molecular characterization of BET1, a gene expressed in the en- and updated open-access database of transcription factor binding
dosperm transfer cells of maize. Plant Cell 7: 747–757. profiles. Nucleic Acids Res. 42: D142–D147.
James, M.G., Robertson, D.S., and Myers, A.M. (1995). Character- Moore, T., and Haig, D. (1991). Genomic imprinting in mammalian
ization of the maize gene sugary1, a determinant of starch com- development: a parental tug-of-war. Trends Genet. 7: 45–49.
position in kernels. Plant Cell 7: 417–429. Mortazavi, A., Williams, B.A., McCue, K., Schaeffer, L., and Wold,
Jin, J., Zhang, H., Kong, L., Gao, G., and Luo, J. (2014). PlantTFDB B. (2008). Mapping and quantifying mammalian transcriptomes by
3.0: a portal for the functional and evolutionary study of plant RNA-Seq. Nat. Methods 5: 621–628.
transcription factors. Nucleic Acids Res. 42: D1182–D1187. Muñiz, L.M., Royo, J., Gómez, E., Baudot, G., Paul, W., and Hueros,
Kerk, N.M., Ceserani, T., Tausta, S.L., Sussex, I.M., and Nelson, G. (2010). Atypical response regulators expressed in the maize
T.M. (2003). Laser capture microdissection of cells from plant tissues. endosperm transfer cells link canonical two component systems
Plant Physiol. 132: 27–35. and seed biology. BMC Plant Biol. 10: 84.
Lafon-Placette, C., and Köhler, C. (2014). Embryo and endosperm, Muñiz, L.M., Royo, J., Gómez, E., Barrero, C., Bergareche, D., and
partners in seed development. Curr. Opin. Plant Biol. 17: 64–69. Hueros, G. (2006). The maize transfer cell-specific type-A response
Langfelder, P., and Horvath, S. (2008). WGCNA: an R package regulator ZmTCRR-1 appears to be involved in intercellular signal-
for weighted correlation network analysis. BMC Bioinformatics ling. Plant J. 48: 17–27.
9: 559. Niu, X., Helentjaris, T., and Bate, N.J. (2002). Maize ABI4 binds
Leroux, B.M., Goodyke, A.J., Schumacher, K.I., Abbott, C.P., Clore, coupling element1 in abscisic acid and sugar response genes. Plant
A.M., Yadegari, R., Larkins, B.A., and Dannenhoffer, J.M. (2014). Maize Cell 14: 2565–2575.
early endosperm growth and development: from fertilization through cell Olsen, O.A. (2001). Endosperm development: cellularization and cell fate
type differentiation. Am. J. Bot. 101: 1259–1274. specification. Annu. Rev. Plant Physiol. Plant Mol. Biol. 52: 233–267.
Li, G., et al. (2014). Temporal patterns of gene expression in de- Olsen, O.-A. (2004). Nuclear endosperm development in cereals and
veloping maize endosperm identified through transcriptome se- Arabidopsis thaliana. Plant Cell 16 (Suppl): S214–S227.
quencing. Proc. Natl. Acad. Sci. USA 111: 7582–7587. Olsen, O.-A., and Becraft, P.W. (2013). Endosperm development. In
Lopes, M.A., and Larkins, B.A. (1993). Endosperm origin, de- Seed Genomics, P.W. Becraft, ed (New York: John Wiley & Sons),
velopment, and function. Plant Cell 5: 1383–1399. pp. 43–62.
Ma, C., and Wang, X. (2012). Application of the Gini correlation co- Opsahl-Ferstad, H.G., Le Deunff, E., Dumas, C., and Rogowsky,
efficient to infer regulatory relationships in transcriptome analysis. P.M. (1997). ZmEsr, a novel endosperm-specific gene expressed in
Plant Physiol. 160: 192–203. a restricted region around the maize embryo. Plant J. 12: 235–246.
Ma, C., Xin, M., Feldmann, K.A., and Wang, X. (2014). Machine Quinlan, A.R., and Hall, I.M. (2010). BEDTools: a flexible suite of
learning-based differential network analysis: a study of stress-responsive utilities for comparing genomic features. Bioinformatics 26: 841–842.
transcriptomes in Arabidopsis. Plant Cell 26: 520–537. Riechmann, J.L., Wang, M., and Meyerowitz, E.M. (1996). DNA-
Maddaloni, M., Donini, G., Balconi, C., Rizzi, E., Gallusci, P., binding properties of Arabidopsis MADS domain homeotic proteins
Forlani, F., Lohmer, S., Thompson, R., Salamini, F., and Motto, APETALA1, APETALA3, PISTILLATA and AGAMOUS. Nucleic Acids
M. (1996). The transcriptional activator Opaque-2 controls the ex- Res. 24: 3134–3141.
pression of a cytosolic form of pyruvate orthophosphate dikinase-1 Royo, J., Gómez, E., Sellam, O., Gerentes, D., Paul, W., and Hueros, G.
in maize endosperms. Mol. Gen. Genet. 250: 647–654. (2014). Two maize END-1 orthologs, BETL9 and BETL9like, are tran-
Magnard, J.-L., Le Deunff, E., Domenech, J., Rogowsky, P.M., scribed in a non-overlapping spatial pattern on the outer surface of the
Testillano, P.S., Rougier, M., Risueño, M.C., Vergne, P., and developing endosperm. Front. Plant Sci. 5: 180.
Dumas, C. (2000). Genes normally expressed in the endosperm are Sabelli, P.A., and Larkins, B.A. (2009). The development of endo-
expressed at early stages of microspore embryogenesis in maize. sperm in grasses. Plant Physiol. 149: 14–26.
Plant Mol. Biol. 44: 559–574. Schel, J., Kieft, H., and Lammeren, A.v. (1984). Interactions between
Magnard, J.L., Lehouque, G., Massonneau, A., Frangne, N., embryo and endosperm during early developmental stages of maize
Heckel, T., Gutierrez-Marcos, J.F., Perez, P., Dumas, C., and caryopses (Zea mays). Can. J. Bot. 62: 2842–2853.
Rogowsky, P.M. (2003). ZmEBE genes show a novel, continuous Schmidt, R.J., Burr, F.A., Aukerman, M.J., and Burr, B. (1990). Maize
expression pattern in the central cell before fertilization and in regulatory gene opaque-2 encodes a protein with a “leucine-zipper” motif
specific domains of the resulting endosperm after fertilization. Plant that binds to zein DNA. Proc. Natl. Acad. Sci. USA 87: 46–50.
Mol. Biol. 53: 821–836. Schmidt, R.J., Ketudat, M., Aukerman, M.J., and Hoschek, G.
Marshall, E., Costa, L.M., and Gutierrez-Marcos, J. (2011). Cyste- (1992). Opaque-2 is a transcriptional activator that recognizes
ine-rich peptides (CRPs) mediate diverse aspects of cell-cell com- a specific target site in 22-kD zein genes. Plant Cell 4: 689–700.
munication in plant reproduction and development. J. Exp. Bot. 62: Sekhon, R.S., Briskine, R., Hirsch, C.N., Myers, C.L., Springer, N.M.,
1677–1686. Buell, C.R., de Leon, N., and Kaeppler, S.M. (2013). Maize gene
Martínez-García, J.F., Moyano, E., Alcocer, M.J., and Martin, C. atlas developed by RNA sequencing and comparative evaluation of
(1998). Two bZIP proteins from Antirrhinum flowers preferentially transcriptomes based on RNA sequencing and microarrays. PLoS
bind a hybrid C-box/G-box motif and help to define a new sub- ONE 8: e61005.
family of bZIP transcription factors. Plant J. 13: 489–505. Serna, A., Maitz, M., O’Connell, T., Santandrea, G., Thevissen, K.,
Marzábal, P., Gas, E., Fontanet, P., Vicente-Carbajosa, J., Torrent, Tienens, K., Hueros, G., Faleri, C., Cai, G., Lottspeich, F., and
M., and Ludevid, M.D. (2008). The maize Dof protein PBF activates Thompson, R.D. (2001). Maize endosperm secretes a novel anti-
transcription of gamma-zein during maize seed development. Plant fungal protein into adjacent maternal tissue. Plant J. 25: 687–698.
Mol. Biol. 67: 441–454. Shannon, P., Markiel, A., Ozier, O., Baliga, N.S., Wang, J.T.,
Massonneau, A., Condamine, P., Wisniewski, J.-P., Zivy, M., and Ramage, D., Amin, N., Schwikowski, B., and Ideker, T. (2003).
Rogowsky, P.M. (2005). Maize cystatins respond to developmental Cytoscape: a software environment for integrated models of bio-
cues, cold stress and drought. Biochim. Biophys. Acta 1729: 186–199. molecular interaction networks. Genome Res. 13: 2498–2504.
286
Cell Differentiation Modules in Maize Endosperm 531

Shure, M., Wessler, S., and Fedoroff, N. (1983). Molecular identifi- Trapnell, C., Roberts, A., Goff, L., Pertea, G., Kim, D., Kelley, D.R.,
cation and isolation of the Waxy locus in maize. Cell 35: 225–233. Pimentel, H., Salzberg, S.L., Rinn, J.L., and Pachter, L. (2012).
Sosso, D., Wisniewski, J.-P., Khaled, A.-S., Hueros, G., Gerentes, Differential gene and transcript expression analysis of RNA-seq
D., Paul, W., and Rogowsky, P.M. (2010). The Vpp1, Esr6a, Esr6b experiments with TopHat and Cufflinks. Nat. Protoc. 7: 562–578.
and OCL4 promoters are active in distinct domains of maize en- Wisniewski, J.-P., and Rogowsky, P.M. (2004). Vacuolar H+-trans-
dosperm. Plant Sci. 179: 86–96. locating inorganic pyrophosphatase (Vpp1) marks partial aleurone
Sreenivasulu, N., and Wobus, U. (2013). Seed-development pro- cell fate in cereal endosperm development. Plant Mol. Biol. 56:
grams: a systems biology-based comparison between dicots and 325–337.
monocots. Annu. Rev. Plant Biol. 64: 189–217. Woo, Y.M., Hu, D.W.N., Larkins, B.A., and Jung, R. (2001). Ge-
Tailor, R.H., Acland, D.P., Attenborough, S., Cammue, B.P., Evans, nomics analysis of genes expressed in maize endosperm identifies
I.J., Osborn, R.W., Ray, J.A., Rees, S.B., and Broekaert, W.F. novel seed proteins and clarifies patterns of zein gene expression.
(1997). A novel family of small cysteine-rich antimicrobial peptides Plant Cell 13: 2297–2317.
from seed of Impatiens balsamina is derived from a single precursor Xin, M., et al. (2013). Dynamic expression of imprinted genes asso-
protein. J. Biol. Chem. 272: 24480–24487. ciates with maternally controlled nutrient allocation during maize
Takahashi, H., Kamakura, H., Sato, Y., Shiono, K., Abiko, T., endosperm development. Plant Cell 25: 3212–3227.
Tsutsumi, N., Nagamura, Y., Nishizawa, N.K., and Nakazono, Xiong, Y., Mei, W., Kim, E.D., Mukherjee, K., Hassanein, H.,
M. (2010). A method for obtaining high quality RNA from paraffin Barbazuk, W.B., Sung, S., Kolaczkowski, B., and Kang, B.H.
sections of plant tissues by laser microdissection. J. Plant Res. 123: (2014). Adaptive expansion of the maize maternally expressed gene
807–813. (MEG) family involves changes in expression patterns and protein
Thompson, R.D., Hueros, G., Becker, H., and Maitz, M. (2001). secondary structures of its members. BMC Plant Biol. 14: 204.
Development and functions of seed transfer cells. Plant Sci. 160: Yilmaz, A., Nishiyama, M.Y., Jr., Fuentes, B.G., Souza, G.M., Janies, D.,
775–783. Gray, J., and Grotewold, E. (2009). GRASSIUS: a platform for com-
Trapnell, C., Pachter, L., and Salzberg, S.L. (2009). TopHat: dis- parative regulatory genomics across the grasses. Plant Physiol. 149:
covering splice junctions with RNA-Seq. Bioinformatics 25: 1105–1111. 171–180.
Trapnell, C., Williams, B.A., Pertea, G., Mortazavi, A., Kwan, G., Zar, J.H. (1972). Significance testing of the Spearman rank correlation
van Baren, M.J., Salzberg, S.L., Wold, B.J., and Pachter, L. coefficient. J. Am. Stat. Assoc. 67: 578–580.
(2010). Transcript assembly and quantification by RNA-Seq reveals Zhang, B., and Horvath, S. (2005). A general framework for weighted
unannotated transcripts and isoform switching during cell differ- gene co-expression network analysis. Stat. Appl. Genet. Mol. Biol.
entiation. Nat. Biotechnol. 28: 511–515. 4: e17.
287

REFERENCES

Abramowitz, L.K., and Bartolomei, M.S. (2012). Genomic imprinting: recognition and
marking of imprinted loci. Current opinion in genetics & development 22, 72-78.
Adams, S., Vinkenoog, R., Spielman, M., Dickinson, H.G., and Scott, R.J. (2000). Parent-of-
origin effects on seed development in Arabidopsis thaliana require DNA methylation.
Development 127, 2493-2502.
Andorf, C.M., Cannon, E.K., Portwood, J.L., 2nd, Gardiner, J.M., Harper, L.C., Schaeffer,
M.L., Braun, B.L., Campbell, D.A., Vinnakota, A.G., Sribalusu, V.V., Huerta, M.,
Cho, K.T., Wimalanathan, K., Richter, J.D., Mauch, E.D., Rao, B.S., Birkett, S.M.,
Sen, T.Z., and Lawrence-Dill, C.J. (2016). MaizeGDB update: new tools, data and
interface for the maize model organism database. Nucleic acids research 44, D1195-1201.
Andrews, S. (2010). FastQC: A quality control tool for high throughput sequence data.
Reference Source.
Bailey, T.L., and Elkan, C. (1994). Fitting a mixture model by expectation maximization to
discover motifs in bipolymers (Department of Computer Science and Engineering,
University of California, San Diego).
Bailey, T.L., Boden, M., Buske, F.A., Frith, M., Grant, C.E., Clementi, L., Ren, J., Li,
W.W., and Noble, W.S. (2009). MEME SUITE: tools for motif discovery and searching.
Nucleic acids research 37, W202-208.
Balandin, M., Royo, J., Gomez, E., Muniz, L.M., Molina, A., and Hueros, G. (2005). A
protective role for the embryo surrounding region of the maize endosperm, as evidenced
by the characterisation of ZmESR-6, a defensin gene specifically expressed in this region.
Plant molecular biology 58, 269-282.
Baranowskij, N., Frohberg, C., Prat, S., and Willmitzer, L. (1994). A novel DNA binding
protein with homology to Myb oncoproteins containing only one repeat can function as a
transcriptional activator. The EMBO journal 13, 5383-5392.
Barrero, C., Muniz, L.M., Gomez, E., Hueros, G., and Royo, J. (2006). Molecular dissection
of the interaction between the transcriptional activator ZmMRP-1 and the promoter of
BETL-1. Plant molecular biology 62, 655-668.
Bartel, P. (1993). Using the two-hybrid system to detect protein-protein interactions. Cellular
Interactions in Development: Practical Approach, 153-179.
Bate, N.J., Niu, X., Wang, Y., Reimann, K.S., and Helentjaris, T.G. (2004). An invertase
inhibitor from maize localizes to the embryo surrounding region during early kernel
development. Plant physiology 134, 246-254.
Baud, S., Wuilleme, S., Lemoine, R., Kronenberger, J., Caboche, M., Lepiniec, L., and
Rochat, C. (2005). The AtSUC5 sucrose transporter specifically expressed in the
endosperm is involved in early seed development in Arabidopsis. The Plant journal : for
cell and molecular biology 43, 824-836.
Becraft, P.W. (2001). Cell fate specification in the cereal endosperm. Seminars in cell &
developmental biology 12, 387-394.
Becraft, P.W. (2007). Aleurone cell development. In Endosperm (Springer), pp. 45-56.
Becraft, P.W., and Asuncion-Crabb, Y. (2000). Positional cues specify and maintain aleurone
cell fate in maize endosperm development. Development 127, 4039-4048.
Becraft, P.W., and Yi, G. (2011). Regulation of aleurone development in cereal grains. Journal
of experimental botany 62, 1669-1675.
288

Becraft, P.W., and Gutierrez-Marcos, J. (2012). Endosperm development: dynamic processes


and cellular innovations underlying sibling altruism. Wiley interdisciplinary reviews.
Developmental biology 1, 579-593.
Becraft, P.W., Li, K., Dey, N., and Asuncion-Crabb, Y. (2002). The maize dek1 gene
functions in embryonic pattern formation and cell fate specification. Development 129,
5217-5225.
Berger, F., Grini, P.E., and Schnittger, A. (2006). Endosperm: an integrator of seed growth
and development. Current opinion in plant biology 9, 664-670.
Bernard, L., Ciceri, P., and Viotti, A. (1994). Molecular analysis of wild-type and mutant
alleles at the Opaque-2 regulatory locus of maize reveals different mutations and types of
O2 products. Plant molecular biology 24, 949-959.
Bhat, R.A., Borst, J.W., Riehl, M., and Thompson, R.D. (2004). Interaction of maize Opaque-
2 and the transcriptional co-activators GCN5 and ADA2, in the modulation of
transcriptional activity. Plant molecular biology 55, 239-252.
Bhat, R.A., Riehl, M., Santandrea, G., Velasco, R., Slocombe, S., Donn, G., Steinbiss, H.H.,
Thompson, R.D., and Becker, H.A. (2003). Alteration of GCN5 levels in maize reveals
dynamic responses to manipulating histone acetylation. The Plant journal : for cell and
molecular biology 33, 455-469.
Blauth, S.L., Kim, K.N., Klucinec, J., Shannon, J.C., Thompson, D., and Guiltinan, M.
(2002). Identification of Mutator insertional mutants of starch-branching enzyme 1 (sbe1)
in Zea mays L. Plant molecular biology 48, 287-297.
Boisnard-Lorig, C., Colon-Carmona, A., Bauch, M., Hodge, S., Doerner, P., Bancharel, E.,
Dumas, C., Haseloff, J., and Berger, F. (2001). Dynamic analyses of the expression of
the HISTONE::YFP fusion protein in arabidopsis show that syncytial endosperm is
divided in mitotic domains. The Plant cell 13, 495-509.
Bolduc, N., Yilmaz, A., Mejia-Guerra, M.K., Morohashi, K., O'Connor, D., Grotewold, E.,
and Hake, S. (2012). Unraveling the KNOTTED1 regulatory network in maize
meristems. Genes & development 26, 1685-1690.
Bolger, A.M., Lohse, M., and Usadel, B. (2014). Trimmomatic: a flexible trimmer for Illumina
sequence data. Bioinformatics 30, 2114-2120.
Bonello, J.F., Opsahl-Ferstad, H.G., Perez, P., Dumas, C., and Rogowsky, P.M. (2000). Esr
genes show different levels of expression in the same region of maize endosperm. Gene
246, 219-227.
Bonello, J.F., Sevilla-Lecoq, S., Berne, A., Risueno, M.C., Dumas, C., and Rogowsky, P.M.
(2002). Esr proteins are secreted by the cells of the embryo surrounding region. Journal
of experimental botany 53, 1559-1568.
Bortesi, L., and Fischer, R. (2015). The CRISPR/Cas9 system for plant genome editing and
beyond. Biotechnology advances 33, 41-52.
Brady, S.M., Orlando, D.A., Lee, J.Y., Wang, J.Y., Koch, J., Dinneny, J.R., Mace, D.,
Ohler, U., and Benfey, P.N. (2007). A high-resolution root spatiotemporal map reveals
dominant expression patterns. Science 318, 801-806.
Braybrook, S.A., Stone, S.L., Park, S., Bui, A.Q., Le, B.H., Fischer, R.L., Goldberg, R.B.,
and Harada, J.J. (2006). Genes directly regulated by LEAFY COTYLEDON2 provide
insight into the control of embryo maturation and somatic embryogenesis. Proceedings of
the National Academy of Sciences of the United States of America 103, 3468-3473.
289

Brink, R.A., and Cooper, D.C. (1947). The Endosperm in Seed Development. Bot Rev 13, 479-
541.
Brown, R., Lemmon, B., and Olsen, O.-A. (1996a). Polarization predicts the pattern of
cellularization in cereal endosperm. Protoplasma 192, 168-177.
Brown, R.C., Lemmon, B.E., and Olsen, O.A. (1994). Endosperm Development in Barley:
Microtubule Involvement in the Morphogenetic Pathway. The Plant cell 6, 1241-1252.
Brown, R.C., Lemmon, B.E., and Olsen, O.A. (1996b). Development of the endosperm in rice
(Oryza sativa L): Cellularization. J Plant Res 109, 301-313.
Brown, R.C., Lemmon, B.E., and Nguyen, H. (2003). Events during the first four rounds of
mitosis establish three developmental domains in the syncytial endosperm of Arabidopsis
thaliana. Protoplasma 222, 167-174.
Brown, R.C., Lemmon, B.E., Nguyen, H., and Olsen, O.A. (1999). Development of
endosperm in Arabidopsis thaliana. Sex Plant Reprod 12, 32-42.
Brugiere, N., Humbert, S., Rizzo, N., Bohn, J., and Habben, J.E. (2008). A member of the
maize isopentenyl transferase gene family, Zea mays isopentenyl transferase 2 (ZmIPT2),
encodes a cytokinin biosynthetic enzyme expressed during kernel development.
Cytokinin biosynthesis in maize. Plant molecular biology 67, 215-229.
Burdo, B., Gray, J., Goetting-Minesky, M.P., Wittler, B., Hunt, M., Li, T., Velliquette, D.,
Thomas, J., Gentzel, I., dos Santos Brito, M., Mejia-Guerra, M.K., Connolly, L.N.,
Qaisi, D., Li, W., Casas, M.I., Doseff, A.I., and Grotewold, E. (2014). The Maize
TFome--development of a transcription factor open reading frame collection for
functional genomics. The Plant journal : for cell and molecular biology 80, 356-366.
Bushell, C., Spielman, M., and Scott, R.J. (2003). The basis of natural and artificial
postzygotic hybridization barriers in Arabidopsis species. The Plant cell 15, 1430-1442.
Causier, B., Schwarz-Sommer, Z., and Davies, B. (2010). Floral organ identity: 20 years of
ABCs. Seminars in cell & developmental biology 21, 73-79.
Charlton, W.L., Keen, C.L., Merriman, C., Lynch, P., Greenland, A.J., and Dickinson,
H.G. (1995). Endosperm Development in Zea-Mays - Implication of Gametic Imprinting
and Paternal Excess in Regulation of Transfer Layer Development. Development 121,
3089-3097.
Chen, J., Zeng, B., Zhang, M., Xie, S., Wang, G., Hauck, A., and Lai, J. (2014a). Dynamic
transcriptome landscape of maize embryo and endosperm development. Plant physiology
166, 252-264.
Chen, Y., Zhou, Z., Zhao, G., Li, X., Song, L., Yan, N., Weng, J., Hao, Z., Zhang, D., Li, M.,
and Zhang, S. (2014b). Transposable element rbg induces the differential expression of
opaque-2 mutant gene in two maize o2 NILs derived from the same inbred line. PloS one
9, e85159.
Cheng, W.H., Taliercio, E.W., and Chourey, P.S. (1996). The Miniature1 Seed Locus of
Maize Encodes a Cell Wall Invertase Required for Normal Development of Endosperm
and Maternal Cells in the Pedicel. The Plant cell 8, 971-983.
Chourey, P.S., and Nelson, O.E. (1979). Interallelic Complementation at the Sh-Locus in
Maize at the Enzyme Level. Genetics 91, 317-325.
Chourey, P.S., and Hueros, G. (2017). The Basal Endosperm Transfer Layer (BETL): Gateway
to the Maize Kernel. In Maize Kernel Development, B.A. Larkins, ed (CAB
International), pp. 68-80.
290

Cifuentes-Rojas, C., Kannan, K., Tseng, L., and Shippen, D.E. (2011). Two RNA subunits
and POT1a are components of Arabidopsis telomerase. Proceedings of the National
Academy of Sciences of the United States of America 108, 73-78.
Clark, S.J., Lee, H.J., Smallwood, S.A., Kelsey, G., and Reik, W. (2016). Single-cell
epigenomics: powerful new methods for understanding gene regulation and cell identity.
Genome biology 17, 72.
Cock, J.M., and McCormick, S. (2001). A large family of genes that share homology with
CLAVATA3. Plant physiology 126, 939-942.
Coleman, C.E., and Larkins, B.A. (1999). The prolamins of maize. In Seed Proteins (Springer),
pp. 109-139.
Commuri, P.D., and Keeling, P.L. (2001). Chain-length specificities of maize starch synthase I
enzyme: studies of glucan affinity and catalytic properties. The Plant journal : for cell and
molecular biology 25, 475-486.
Conesa, A., and Gotz, S. (2008). Blast2GO: A comprehensive suite for functional analysis in
plant genomics. International journal of plant genomics 2008, 619832.
Conesa, A., Gotz, S., Garcia-Gomez, J.M., Terol, J., Talon, M., and Robles, M. (2005).
Blast2GO: a universal tool for annotation, visualization and analysis in functional
genomics research. Bioinformatics 21, 3674-3676.
Cooper, D.C. (1951). Caryopsis Development Following Matings between Diploid and
Tetraploid Strains of Zea-Mays. American journal of botany 38, 702-708.
Cord Neto, G., Yunes, J.A., da Silva, M.J., Vettore, A.L., Arruda, P., and Leite, A. (1995).
The involvement of Opaque 2 on beta-prolamin gene regulation in maize and Coix
suggests a more general role for this transcriptional activator. Plant molecular biology 27,
1015-1029.
Costa, L.M., Gutierrez-Marcos, J.F., and Dickinson, H.G. (2004). More than a yolk: the short
life and complex times of the plant endosperm. Trends in plant science 9, 507-514.
Costa, L.M., Gutierrez-Marcos, J.F., Brutnell, T.P., Greenland, A.J., and Dickinson, H.G.
(2003). The globby1-1 (glo1-1) mutation disrupts nuclear and cell division in the
developing maize seed causing alterations in endosperm cell fate and tissue
differentiation. Development 130, 5009-5017.
Costa, L.M., Yuan, J., Rouster, J., Paul, W., Dickinson, H., and Gutierrez-Marcos, J.F.
(2012). Maternal control of nutrient allocation in plant seeds by genomic imprinting.
Current biology : CB 22, 160-165.
Costa, L.M., Marshall, E., Tesfaye, M., Silverstein, K.A., Mori, M., Umetsu, Y., Otterbach,
S.L., Papareddy, R., Dickinson, H.G., Boutiller, K., VandenBosch, K.A., Ohki, S.,
and Gutierrez-Marcos, J.F. (2014). Central cell-derived peptides regulate early embryo
patterning in flowering plants. Science 344, 168-172.
Danilevskaya, O.N., Hermon, P., Hantke, S., Muszynski, M.G., Kollipara, K., and Ananiev,
E.V. (2003). Duplicated fie genes in maize: expression pattern and imprinting suggest
distinct functions. The Plant cell 15, 425-438.
de Vetten, N.C., and Ferl, R.J. (1995). Characterization of a maize G-box binding factor that is
induced by hypoxia. The Plant journal : for cell and molecular biology 7, 589-601.
Deal, R.B., and Henikoff, S. (2010). A simple method for gene expression and chromatin
profiling of individual cell types within a tissue. Developmental cell 18, 1030-1040.
291

Doan, D.N., Linnestad, C., and Olsen, O.A. (1996). Isolation of molecular markers from the
barley endosperm coenocyte and the surrounding nucellus cell layers. Plant molecular
biology 31, 877-886.
Doll, N.M., Depege-Fargeix, N., Rogowsky, P.M., and Widiez, T. (2017). Signaling in Early
Maize Kernel Development. Molecular plant 10, 375-388.
Dong, X., Zhang, M., Chen, J., Peng, L., Zhang, N., Wang, X., and Lai, J. (2016). Dynamic
and Antagonistic Allele-Specific Epigenetic Modifications Controlling the Expression of
Imprinted Genes in Maize Endosperm. Molecular plant.
Downs, G.S., Bi, Y.M., Colasanti, J., Wu, W., Chen, X., Zhu, T., Rothstein, S.J., and
Lukens, L.N. (2013). A developmental transcriptional network for maize defines
coexpression modules. Plant physiology 161, 1830-1843.
Drews, G.N. (1998). In situ hybridization. Methods in molecular biology 82, 353-371.
Du, Z., Zhou, X., Ling, Y., Zhang, Z., and Su, Z. (2010). agriGO: a GO analysis toolkit for the
agricultural community. Nucleic acids research 38, W64-70.
Emmert-Buck, M.R., Bonner, R.F., Smith, P.D., Chuaqui, R.F., Zhuang, Z., Goldstein,
S.R., Weiss, R.A., and Liotta, L.A. (1996). Laser capture microdissection. Science 274,
998-1001.
Etcheverry, T. (1990). Induced Expression Using Yeast Copper-Metallothionein Promoter.
Methods in enzymology 185, 319-329.
Eveland, A.L., Goldshmidt, A., Pautler, M., Morohashi, K., Liseron-Monfils, C., Lewis,
M.W., Kumari, S., Hiraga, S., Yang, F., Unger-Wallace, E., Olson, A., Hake, S.,
Vollbrecht, E., Grotewold, E., Ware, D., and Jackson, D. (2014). Regulatory modules
controlling maize inflorescence architecture. Genome research 24, 431-443.
Evrard, A., Bargmann, B.O., Birnbaum, K.D., Tester, M., Baumann, U., and Johnson, A.A.
(2012). Fluorescence-activated cell sorting for analysis of cell type-specific responses to
salinity stress in Arabidopsis and rice. Methods in molecular biology 913, 265-276.
FAO. (2012). FAO statistical Yearbook 2012.
(http://www.fao.org/docrep/015/i2490e/i2490e00.htm).
FAO. (2015). FAO Statistical Pocketbook 2015.
(http://www.fao.org/documents/card/en/c/383d384a-28e6-47b3-a1a2-2496a9e017b2/).
Faure, J.E. (2001). Double fertilization in flowering plants: discovery, study methods and
mechanisms. Comptes rendus de l'Academie des sciences. Serie III, Sciences de la vie
324, 551-558.
Feng, J., Liu, T., Qin, B., Zhang, Y., and Liu, X.S. (2012). Identifying ChIP-seq enrichment
using MACS. Nature protocols 7, 1728-1740.
Finnegan, E.J., and Dennis, E.S. (1993). Isolation and identification by sequence homology of
a putative cytosine methyltransferase from Arabidopsis thaliana. Nucleic acids research
21, 2383-2388.
Fletcher, J.C., Brand, U., Running, M.P., Simon, R., and Meyerowitz, E.M. (1999).
Signaling of cell fate decisions by CLAVATA3 in Arabidopsis shoot meristems. Science
283, 1911-1914.
Formstecher, E., Aresta, S., Collura, V., Hamburger, A., Meil, A., Trehin, A., Reverdy, C.,
Betin, V., Maire, S., Brun, C., Jacq, B., Arpin, M., Bellaiche, Y., Bellusci, S.,
Benaroch, P., Bornens, M., Chanet, R., Chavrier, P., Delattre, O., Doye, V., Fehon,
R., Faye, G., Galli, T., Girault, J.A., Goud, B., de Gunzburg, J., Johannes, L.,
Junier, M.P., Mirouse, V., Mukherjee, A., Papadopoulo, D., Perez, F., Plessis, A.,
292

Rosse, C., Saule, S., Stoppa-Lyonnet, D., Vincent, A., White, M., Legrain, P.,
Wojcik, J., Camonis, J., and Daviet, L. (2005). Protein interaction mapping: a
Drosophila case study. Genome research 15, 376-384.
Friedman, W.E., Madrid, E.N., and Williams, J.H. (2008). Origin of the fittest and survival of
the fittest: Relating female gametophyte development to endosperm genetics. Int J Plant
Sci 169, 79-92.
Frizzi, A., Caldo, R.A., Morrell, J.A., Wang, M., Lutfiyya, L.L., Brown, W.E., Malvar,
T.M., and Huang, S. (2010). Compositional and transcriptional analyses of reduced zein
kernels derived from the opaque2 mutation and RNAi suppression. Plant molecular
biology 73, 569-585.
FromontRacine, M., Rain, J.C., and Legrain, P. (1997). Toward a functional analysis of the
yeast genome through exhaustive two-hybrid screens. Nat Genet 16, 277-282.
Gallusci, P., Varotto, S., Matsuoko, M., Maddaloni, M., and Thompson, R.D. (1996).
Regulation of cytosolic pyruvate, orthophosphate dikinase expression in developing
maize endosperm. Plant molecular biology 31, 45-55.
Garner, A.G., Kenney, A.M., Fishman, L., and Sweigart, A.L. (2016). Genetic loci with
parent-of-origin effects cause hybrid seed lethality in crosses between Mimulus species.
The New phytologist 211, 319-331.
Gehl, C., Waadt, R., Kudla, J., Mendel, R.R., and Hansch, R. (2009). New GATEWAY
vectors for high throughput analyses of protein-protein interactions by bimolecular
fluorescence complementation. Molecular plant 2, 1051-1058.
Gehring, M. (2013). Genomic imprinting: insights from plants. Annual review of genetics 47,
187-208.
Gehring, M., and Satyaki, P.R. (2017). Endosperm and Imprinting, Inextricably Linked. Plant
physiology 173, 143-154.
Giroux, M.J., Boyer, C., Feix, G., and Hannah, L.C. (1994). Coordinated Transcriptional
Regulation of Storage Product Genes in the Maize Endosperm. Plant physiology 106,
713-722.
Gomez, E., Royo, J., Guo, Y., Thompson, R., and Hueros, G. (2002). Establishment of cereal
endosperm expression domains: identification and properties of a maize transfer cell-
specific transcription factor, ZmMRP-1. The Plant cell 14, 599-610.
Gomez, E., Royo, J., Muniz, L.M., Sellam, O., Paul, W., Gerentes, D., Barrero, C., Lopez,
M., Perez, P., and Hueros, G. (2009). The maize transcription factor myb-related
protein-1 is a key regulator of the differentiation of transfer cells. The Plant cell 21, 2022-
2035.
Gontarek, B.C., and Becraft, P.W. (2017). Aleurone. In Maize Kernel Development, B.A.
Larkins, ed (CAB International), pp. 68-80.
Gontarek, B.C., Neelakandan, A.K., Wu, H., and Becraft, P.W. (2016). NKD Transcription
Factors Are Central Regulators of Maize Endosperm Development. The Plant cell 28,
2916-2936.
Gotz, S., Garcia-Gomez, J.M., Terol, J., Williams, T.D., Nagaraj, S.H., Nueda, M.J.,
Robles, M., Talon, M., Dopazo, J., and Conesa, A. (2008). High-throughput functional
annotation and data mining with the Blast2GO suite. Nucleic acids research 36, 3420-
3435.
Gray, J., Bevan, M., Brutnell, T., Buell, C.R., Cone, K., Hake, S., Jackson, D., Kellogg, E.,
Lawrence, C., McCouch, S., Mockler, T., Moose, S., Paterson, A., Peterson, T.,
293

Rokshar, D., Souza, G.M., Springer, N., Stein, N., Timmermans, M., Wang, G.L.,
and Grotewold, E. (2009). A recommendation for naming transcription factor proteins in
the grasses. Plant physiology 149, 4-6.
Grimault, A., Gendrot, G., Chamot, S., Widiez, T., Rabille, H., Gerentes, M.F., Creff, A.,
Thevenin, J., Dubreucq, B., Ingram, G.C., Rogowsky, P.M., and Depege-Fargeix, N.
(2015). ZmZHOUPI, an endosperm-specific basic helix-loop-helix transcription factor
involved in maize seed development. The Plant journal : for cell and molecular biology
84, 574-586.
Gruis, D.F., Guo, H., Selinger, D., Tian, Q., and Olsen, O.A. (2006). Surface position, not
signaling from surrounding maternal tissues, specifies aleurone epidermal cell fate in
maize. Plant physiology 141, 898-909.
Guitton, A.E., Page, D.R., Chambrier, P., Lionnet, C., Faure, J.E., Grossniklaus, U., and
Berger, F. (2004). Identification of new members of Fertilisation Independent Seed
Polycomb Group pathway involved in the control of seed development in Arabidopsis
thaliana. Development 131, 2971-2981.
Gunning, B.E.S., and Pate, J.S. (1969). Transfer Cells Plant Cells with Wall Ingrowths,
Specialized in Relation to Short Distance Transport of Solutes - Their Occurrence,
Structure, and Development. Protoplasma 68, 107-&.
Gupta, S., Stamatoyannopoulos, J.A., Bailey, T.L., and Noble, W.S. (2007). Quantifying
similarity between motifs. Genome biology 8.
Gutierrez-Marcos, J.F., Costa, L.M., and Evans, M.M. (2006). Maternal gametophytic
baseless1 is required for development of the central cell and early endosperm patterning
in maize (Zea mays). Genetics 174, 317-329.
Gutierrez-Marcos, J.F., Costa, L.M., Biderre-Petit, C., Khbaya, B., O'Sullivan, D.M.,
Wormald, M., Perez, P., and Dickinson, H.G. (2004). maternally expressed gene1 Is a
novel maize endosperm transfer cell-specific gene with a maternal parent-of-origin
pattern of expression. The Plant cell 16, 1288-1301.
Haig, D. (2014). Coadaptation and conflict, misconception and muddle, in the evolution of
genomic imprinting. Heredity 113, 96-103.
Haig, D., and Westoby, M. (1989). Parent-Specific Gene-Expression and the Triploid
Endosperm. Am Nat 134, 147-155.
Hamamura, Y., Nagahara, S., and Higashiyama, T. (2012). Double fertilization on the move.
Current opinion in plant biology 15, 70-77.
Handley, A., Schauer, T., Ladurner, A.G., and Margulies, C.E. (2015). Designing Cell-Type-
Specific Genome-wide Experiments. Mol Cell 58, 621-631.
Hannon, G.J. (2002). RNA interference. Nature 418, 244-251.
Hansey, C.N., Vaillancourt, B., Sekhon, R.S., de Leon, N., Kaeppler, S.M., and Buell, C.R.
(2012). Maize (Zea mays L.) Genome Diversity as Revealed by RNA-Sequencing. PloS
one 7.
Hartings, H., Lauria, M., Lazzaroni, N., Pirona, R., and Motto, M. (2011). The Zea mays
mutants opaque-2 and opaque-7 disclose extensive changes in endosperm metabolism as
revealed by protein, amino acid, and transcriptome-wide analyses. BMC genomics 12,
41.
Haun, W.J., Laoueille-Duprat, S., O'Connell M, J., Spillane, C., Grossniklaus, U., Phillips,
A.R., Kaeppler, S.M., and Springer, N.M. (2007). Genomic imprinting, methylation
294

and molecular evolution of maize Enhancer of zeste (Mez) homologs. The Plant journal :
for cell and molecular biology 49, 325-337.
Hehenberger, E., Kradolfer, D., and Kohler, C. (2012). Endosperm cellularization defines an
important developmental transition for embryo development. Development 139, 2031-
2039.
Heidler, S.A., and Radding, J.A. (1995). The AUR1 gene in Saccharomyces cerevisiae encodes
dominant resistance to the antifungal agent aureobasidin A (LY295337). Antimicrobial
agents and chemotherapy 39, 2765-2769.
Hoecker, U., Vasil, I.K., and McCarty, D.R. (1995). Integrated control of seed maturation and
germination programs by activator and repressor functions of Viviparous-1 of maize.
Genes & development 9, 2459-2469.
Holding, D.R., Otegui, M.S., Li, B., Meeley, R.B., Dam, T., Hunter, B.G., Jung, R., and
Larkins, B.A. (2007). The maize floury1 gene encodes a novel endoplasmic reticulum
protein involved in zein protein body formation. The Plant cell 19, 2569-2582.
Holding, D.R., Hunter, B.G., Chung, T., Gibbon, B.C., Ford, C.F., Bharti, A.K., Messing,
J., Hamaker, B.R., and Larkins, B.A. (2008). Genetic analysis of opaque2 modifier loci
in quality protein maize. TAG. Theoretical and applied genetics. Theoretische und
angewandte Genetik 117, 157-170.
Hollander, M., Wolfe, D.A., and Chicken, E. (2013). Nonparametric statistical methods. (John
Wiley & Sons).
Hu, Z., Mellor, J., and DeLisi, C. (2004). Analyzing networks with VisANT. Current protocols
in bioinformatics / editoral board, Andreas D. Baxevanis ... [et al.] Chapter 8, Unit 8 8.
Hubbard, T., Barker, D., Birney, E., Cameron, G., Chen, Y., Clark, L., Cox, T., Cuff, J.,
Curwen, V., Down, T., Durbin, R., Eyras, E., Gilbert, J., Hammond, M.,
Huminiecki, L., Kasprzyk, A., Lehvaslaiho, H., Lijnzaad, P., Melsopp, C., Mongin,
E., Pettett, R., Pocock, M., Potter, S., Rust, A., Schmidt, E., Searle, S., Slater, G.,
Smith, J., Spooner, W., Stabenau, A., Stalker, J., Stupka, E., Ureta-Vidal, A.,
Vastrik, I., and Clamp, M. (2002). The Ensembl genome database project. Nucleic
acids research 30, 38-41.
Hueros, G., Varotto, S., Salamini, F., and Thompson, R.D. (1995). Molecular characterization
of BET1, a gene expressed in the endosperm transfer cells of maize. The Plant cell 7,
747-757.
Hueros, G., Royo, J., Maitz, M., Salamini, F., and Thompson, R.D. (1999). Evidence for
factors regulating transfer cell-specific expression in maize endosperm. Plant molecular
biology 41, 403-414.
Hunter, B.G., Beatty, M.K., Singletary, G.W., Hamaker, B.R., Dilkes, B.P., Larkins, B.A.,
and Jung, R. (2002). Maize opaque endosperm mutations create extensive changes in
patterns of gene expression. The Plant cell 14, 2591-2612.
Hwang, Y.S., Ciceri, P., Parsons, R.L., Moose, S.P., Schmidt, R.J., and Huang, N. (2004).
The maize O2 and PBF proteins act additively to promote transcription from storage
protein gene promoters in rice endosperm cells. Plant & cell physiology 45, 1509-1518.
Ingouff, M., Haseloff, J., and Berger, F. (2005). Polycomb group genes control developmental
timing of endosperm. The Plant journal : for cell and molecular biology 42, 663-674.
Ishikawa, R., Ohnishi, T., Kinoshita, Y., Eiguchi, M., Kurata, N., and Kinoshita, T. (2011).
Rice interspecies hybrids show precocious or delayed developmental transitions in the
295

endosperm without change to the rate of syncytial nuclear division. The Plant journal :
for cell and molecular biology 65, 798-806.
James, M.G., Robertson, D.S., and Myers, A.M. (1995). Characterization of the maize gene
sugary1, a determinant of starch composition in kernels. The Plant cell 7, 417-429.
Jia, H., Nettleton, D., Peterson, J.M., Vazquez-Carrillo, G., Jannink, J.L., and Scott, M.P.
(2007). Comparison of transcript profiles in wild-type and o2 maize endosperm in
different genetic backgrounds. Crop Science 47, S45-S59.
Jia, M., Wu, H., Clay, K.L., Jung, R., Larkins, B.A., and Gibbon, B.C. (2013). Identification
and characterization of lysine-rich proteins and starch biosynthesis genes in the opaque2
mutant by transcriptional and proteomic analysis. BMC plant biology 13, 60.
Jin, J., Zhang, H., Kong, L., Gao, G., and Luo, J. (2014). PlantTFDB 3.0: a portal for the
functional and evolutionary study of plant transcription factors. Nucleic acids research
42, D1182-1187.
Johnson, D.R., and Tanner, J.W. (1972). Calculation of the rate and duration of grain filling in
corn (Zea mays L.). Crop Science 12, 485-486.
Jones, R.J., Roessler, J., and Ouattar, S. (1985). Thermal Environment during Endosperm
Cell-Division in Maize - Effects on Number of Endosperm Cells and Starch Granules.
Crop Science 25, 830-834.
Jones, R.J., Schreiber, B.M.N., and Roessler, J.A. (1996). Kernel sink capacity in maize:
Genotypic and maternal regulation. Crop Science 36, 301-306.
Jurgens, G. (2005). Cytokinesis in higher plants. Annual review of plant biology 56, 281-299.
Kang, H.J., Kawasawa, Y.I., Cheng, F., Zhu, Y., Xu, X., Li, M., Sousa, A.M., Pletikos, M.,
Meyer, K.A., Sedmak, G., Guennel, T., Shin, Y., Johnson, M.B., Krsnik, Z., Mayer,
S., Fertuzinhos, S., Umlauf, S., Lisgo, S.N., Vortmeyer, A., Weinberger, D.R., Mane,
S., Hyde, T.M., Huttner, A., Reimers, M., Kleinman, J.E., and Sestan, N. (2011).
Spatio-temporal transcriptome of the human brain. Nature 478, 483-489.
Kang, I.H., Steffen, J.G., Portereiko, M.F., Lloyd, A., and Drews, G.N. (2008). The AGL62
MADS domain protein regulates cellularization during endosperm development in
Arabidopsis. The Plant cell 20, 635-647.
Kawakatsu, T., and Takaiwa, F. (2010). Cereal seed storage protein synthesis: fundamental
processes for recombinant protein production in cereal grains. Plant biotechnology
journal 8, 939-953.
Kemper, E.L., Neto, G.C., Papes, F., Moraes, K.C., Leite, A., and Arruda, P. (1999). The
role of opaque2 in the control of lysine-degrading activities in developing maize
endosperm. The Plant cell 11, 1981-1994.
Kent, W.J. (2002). BLAT--the BLAST-like alignment tool. Genome research 12, 656-664.
Kerk, N.M., Ceserani, T., Tausta, S.L., Sussex, I.M., and Nelson, T.M. (2003). Laser capture
microdissection of cells from plant tissues. Plant physiology 132, 27-35.
Khoo, U., and Wolf, M.J. (1970). Origin and Development of Protein Granules in Maize
Endosperm. American journal of botany 57, 1042-&.
Kiesselbach, T.A. (1999). The structure and reproduction of corn. (Cold spring harbor
laboratory press).
Kodrzycki, R., Boston, R.S., and Larkins, B.A. (1989). The opaque-2 mutation of maize
differentially reduces zein gene transcription. The Plant cell 1, 105-114.
296

Kradolfer, D., Wolff, P., Jiang, H., Siretskiy, A., and Kohler, C. (2013). An imprinted gene
underlies postzygotic reproductive isolation in Arabidopsis thaliana. Developmental cell
26, 525-535.
Kusano, T., Berberich, T., Harada, M., Suzuki, N., and Sugawara, K. (1995). A maize DNA-
binding factor with a bZIP motif is induced by low temperature. Molecular & general
genetics : MGG 248, 507-517.
Kyle, D.J., and Styles, E.D. (1977). Development of aleurone and sub-aleurone layers in maize.
Planta 137, 185-193.
Lafon-Placette, C., and Kohler, C. (2014). Embryo and endosperm, partners in seed
development. Current opinion in plant biology 17, 64-69.
Langfelder, P., and Horvath, S. (2008). WGCNA: an R package for weighted correlation
network analysis. BMC bioinformatics 9, 559.
Langmead, B., and Salzberg, S.L. (2012). Fast gapped-read alignment with Bowtie 2. Nature
methods 9, 357-359.
Lara, P., Onate-Sanchez, L., Abraham, Z., Ferrandiz, C., Diaz, I., Carbonero, P., and
Vicente-Carbajosa, J. (2003). Synergistic activation of seed storage protein gene
expression in Arabidopsis by ABI3 and two bZIPs related to OPAQUE2. The Journal of
biological chemistry 278, 21003-21011.
Lending, C.R., and Larkins, B.A. (1989). Changes in the zein composition of protein bodies
during maize endosperm development. The Plant cell 1, 1011-1023.
Leroux, B.M., Goodyke, A.J., Schumacher, K.I., Abbott, C.P., Clore, A.M., Yadegari, R.,
Larkins, B.A., and Dannenhoffer, J.M. (2014). Maize early endosperm growth and
development: from fertilization through cell type differentiation. American journal of
botany 101, 1259-1274.
Li, C., Qiao, Z., Qi, W., Wang, Q., Yuan, Y., Yang, X., Tang, Y., Mei, B., Lv, Y., Zhao, H.,
Xiao, H., and Song, R. (2015). Genome-wide characterization of cis-acting DNA targets
reveals the transcriptional regulatory framework of opaque2 in maize. The Plant cell 27,
532-545.
Li, G., Wang, D., Yang, R., Logan, K., Chen, H., Zhang, S., Skaggs, M.I., Lloyd, A.,
Burnett, W.J., Laurie, J.D., Hunter, B.G., Dannenhoffer, J.M., Larkins, B.A.,
Drews, G.N., Wang, X., and Yadegari, R. (2014). Temporal patterns of gene expression
in developing maize endosperm identified through transcriptome sequencing.
Proceedings of the National Academy of Sciences of the United States of America 111,
7582-7587.
Li, J., and Berger, F. (2012). Endosperm: food for humankind and fodder for scientific
discoveries. The New phytologist 195, 290-305.
Li, J., Nie, X., Tan, J.L., and Berger, F. (2013). Integration of epigenetic and genetic controls
of seed size by cytokinin in Arabidopsis. Proceedings of the National Academy of
Sciences of the United States of America 110, 15479-15484.
Li, Q., Wang, J., Ye, J., Zheng, X., Xiang, X., Li, C., Fu, M., Wang, Q., Zhang, Z.Y., and
Wu, Y. (2017). The Maize Imprinted Gene Floury3 Encodes a PLATZ Protein Required
for tRNA and 5S rRNA Transcription Through Interaction with RNA Polymerase III.
The Plant cell.
Liao, Y., Smyth, G.K., and Shi, W. (2014). featureCounts: an efficient general purpose
program for assigning sequence reads to genomic features. Bioinformatics 30, 923-930.
297

Lid, S.E., Gruis, D., Jung, R., Lorentzen, J.A., Ananiev, E., Chamberlin, M., Niu, X.,
Meeley, R., Nichols, S., and Olsen, O.A. (2002). The defective kernel 1 (dek1) gene
required for aleurone cell development in the endosperm of maize grains encodes a
membrane protein of the calpain gene superfamily. Proceedings of the National Academy
of Sciences of the United States of America 99, 5460-5465.
Locatelli, S., Piatti, P., Motto, M., and Rossi, V. (2009). Chromatin and DNA modifications in
the Opaque2-mediated regulation of gene transcription during maize endosperm
development. The Plant cell 21, 1410-1427.
Lohmer, S., Maddaloni, M., Motto, M., Di Fonzo, N., Hartings, H., Salamini, F., and
Thompson, R.D. (1991). The maize regulatory locus Opaque-2 encodes a DNA-binding
protein which activates the transcription of the b-32 gene. The EMBO journal 10, 617-
624.
Lopes, M.A., and Larkins, B.A. (1993). Endosperm origin, development, and function. The
Plant cell 5, 1383-1399.
Lu, X., Chen, D., Shu, D., Zhang, Z., Wang, W., Klukas, C., Chen, L.L., Fan, Y., Chen, M.,
and Zhang, C. (2013). The differential transcription network between embryo and
endosperm in the early developing maize seed. Plant physiology 162, 440-455.
Ma, C., and Wang, X. (2012). Application of the Gini correlation coefficient to infer regulatory
relationships in transcriptome analysis. Plant physiology 160, 192-203.
Ma, C., Xin, M., Feldmann, K.A., and Wang, X. (2014). Machine learning-based differential
network analysis: a study of stress-responsive transcriptomes in Arabidopsis. The Plant
cell 26, 520-537.
Macaulay, I.C., and Voet, T. (2014). Single cell genomics: advances and future perspectives.
PLoS genetics 10, e1004126.
Machanick, P., and Bailey, T.L. (2011). MEME-ChIP: motif analysis of large DNA datasets.
Bioinformatics 27, 1696-1697.
Maddaloni, M., Di Fonzo, N., Hartings, H., Lazzaroni, N., Salamini, F., Thompson, R., and
Motto, M. (1989). The sequence of the zein regulatory gene opaque-2 (O2) of Zea Mays.
Nucleic acids research 17, 7532.
Maddaloni, M., Donini, G., Balconi, C., Rizzi, E., Gallusci, P., Forlani, F., Lohmer, S.,
Thompson, R., Salamini, F., and Motto, M. (1996). The transcriptional activator
Opaque-2 controls the expression of a cytosolic form of pyruvate orthophosphate
dikinase-1 in maize endosperms. Molecular & general genetics : MGG 250, 647-654.
Magnard, J.L., Le Deunff, E., Domenech, J., Rogowsky, P.M., Testillano, P.S., Rougier, M.,
Risueno, M.C., Vergne, P., and Dumas, C. (2000). Genes normally expressed in the
endosperm are expressed at early stages of microspore embryogenesis in maize. Plant
molecular biology 44, 559-574.
Magnard, J.L., Lehouque, G., Massonneau, A., Frangne, N., Heckel, T., Gutierrez-Marcos,
J.F., Perez, P., Dumas, C., and Rogowsky, P.M. (2003). ZmEBE genes show a novel,
continuous expression pattern in the central cell before fertilization and in specific
domains of the resulting endosperm after fertilization. Plant molecular biology 53, 821-
836.
Makarevitch, I., Eichten, S.R., Briskine, R., Waters, A.J., Danilevskaya, O.N., Meeley,
R.B., Myers, C.L., Vaughn, M.W., and Springer, N.M. (2013). Genomic distribution
of maize facultative heterochromatin marked by trimethylation of H3K27. The Plant cell
25, 780-793.
298

Mangan, S., and Alon, U. (2003). Structure and function of the feed-forward loop network
motif. Proceedings of the National Academy of Sciences of the United States of America
100, 11980-11985.
Marshall, E., Costa, L.M., and Gutierrez-Marcos, J. (2011). Cysteine-rich peptides (CRPs)
mediate diverse aspects of cell-cell communication in plant reproduction and
development. Journal of experimental botany 62, 1677-1686.
Martinez-Garcia, J.F., Moyano, E., Alcocer, M.J., and Martin, C. (1998). Two bZIP proteins
from Antirrhinum flowers preferentially bind a hybrid C-box/G-box motif and help to
define a new sub-family of bZIP transcription factors. The Plant journal : for cell and
molecular biology 13, 489-505.
Marzabal, P., Gas, E., Fontanet, P., Vicente-Carbajosa, J., Torrent, M., and Ludevid, M.D.
(2008). The maize Dof protein PBF activates transcription of gamma-zein during maize
seed development. Plant molecular biology 67, 441-454.
Massonneau, A., Condamine, P., Wisniewski, J.P., Zivy, M., and Rogowsky, P.M. (2005).
Maize cystatins respond to developmental cues, cold stress and drought. Biochimica et
biophysica acta 1729, 186-199.
Mathelier, A., Zhao, X., Zhang, A.W., Parcy, F., Worsley-Hunt, R., Arenillas, D.J.,
Buchman, S., Chen, C.Y., Chou, A., Ienasescu, H., Lim, J., Shyr, C., Tan, G., Zhou,
M., Lenhard, B., Sandelin, A., and Wasserman, W.W. (2014). JASPAR 2014: an
extensively expanded and updated open-access database of transcription factor binding
profiles. Nucleic acids research 42, D142-147.
McCarty, D.R., Carson, C.B., Stinard, P.S., and Robertson, D.S. (1989). Molecular Analysis
of viviparous-1: An Abscisic Acid-Insensitive Mutant of Maize. The Plant cell 1, 523-
532.
McCarty, D.R., Hattori, T., Carson, C.B., Vasil, V., Lazar, M., and Vasil, I.K. (1991). The
Viviparous-1 developmental gene of maize encodes a novel transcriptional activator. Cell
66, 895-905.
Mertz, E.T., Nelson, O.E., and Bates, L.S. (1964). Mutant Gene That Changes Protein
Composition + Increases Lysine Content of Maize Endosperm. Science 145, 279-&.
Monjardino, P., Machado, J., Gil, F.S., Fernandes, R., and Salema, R. (2007). Structural and
ultrastructural characterization of maize coenocyte and endosperm cellularization. Can J
Bot 85, 216-223.
Monjardino, P., Rocha, S., Tavares, A.C., Fernandes, R., Sampaio, P., Salema, R., and da
Camara Machado, A. (2013). Development of flange and reticulate wall ingrowths in
maize (Zea mays L.) endosperm transfer cells. Protoplasma 250, 495-503.
Moore, T., and Haig, D. (1991). Genomic imprinting in mammalian development: a parental
tug-of-war. Trends in genetics : TIG 7, 45-49.
Moreno-Risueno, M.A., Gonzalez, N., Diaz, I., Parcy, F., Carbonero, P., and Vicente-
Carbajosa, J. (2008). FUSCA3 from barley unveils a common transcriptional regulation
of seed-specific genes between cereals and Arabidopsis. The Plant journal : for cell and
molecular biology 53, 882-894.
Morohashi, K., Casas, M.I., Falcone Ferreyra, M.L., Mejia-Guerra, M.K., Pourcel, L.,
Yilmaz, A., Feller, A., Carvalho, B., Emiliani, J., Rodriguez, E., Pellegrinet, S.,
McMullen, M., Casati, P., and Grotewold, E. (2012). A genome-wide regulatory
framework identifies maize pericarp color1 controlled genes. The Plant cell 24, 2745-
2764.
299

Morrison, I.N., Kuo, J., and O'Brien, T.P. (1975). Histochemistry and fine structure of
developing wheat aleurone cells. Planta 123, 105-116.
Mortazavi, A., Williams, B.A., McCue, K., Schaeffer, L., and Wold, B. (2008). Mapping and
quantifying mammalian transcriptomes by RNA-Seq. Nature methods 5, 621-628.
Muniz, L.M., Royo, J., Gomez, E., Barrero, C., Bergareche, D., and Hueros, G. (2006). The
maize transfer cell-specific type-A response regulator ZmTCRR-1 appears to be involved
in intercellular signalling. The Plant journal : for cell and molecular biology 48, 17-27.
Muniz, L.M., Royo, J., Gomez, E., Baudot, G., Paul, W., and Hueros, G. (2010). Atypical
response regulators expressed in the maize endosperm transfer cells link canonical two
component systems and seed biology. BMC plant biology 10, 84.
Muth, J.R., Muller, M., Lohmer, S., Salamini, F., and Thompson, R.D. (1996). The role of
multiple binding sites in the activation of zein gene expression by Opaque-2. Molecular
& general genetics : MGG 252, 723-732.
Neuffer, M.G., and Sheridan, W.F. (1980). Defective kernel mutants of maize. I. Genetic and
lethality studies. Genetics 95, 929-944.
Niu, X., Helentjaris, T., and Bate, N.J. (2002). Maize ABI4 binds coupling element1 in
abscisic acid and sugar response genes. The Plant cell 14, 2565-2575.
O'Malley, R.C., Huang, S.C., Song, L., Lewsey, M.G., Bartlett, A., Nery, J.R., Galli, M.,
Gallavotti, A., and Ecker, J.R. (2016). Cistrome and Epicistrome Features Shape the
Regulatory DNA Landscape. Cell 166, 1598.
Ogawa, M., Shinohara, H., Sakagami, Y., and Matsubayashi, Y. (2008). Arabidopsis CLV3
peptide directly binds CLV1 ectodomain. Science 319, 294.
Olsen, O.-A., and Becraft, P.W. (2013). Endosperm Development. In Seed Genomics (Wiley-
Blackwell), pp. 43-62.
Olsen, O.A. (2001). ENDOSPERM DEVELOPMENT: Cellularization and Cell Fate
Specification. Annual review of plant physiology and plant molecular biology 52, 233-
267.
Olsen, O.A. (2004). Nuclear endosperm development in cereals and Arabidopsis thaliana. The
Plant cell 16 Suppl, S214-227.
Olsen, O.A., Linnestad, C., and Nichols, S.E. (1999). Developmental biology of the cereal
endosperm. Trends in plant science 4, 253-257.
Oneal, E., Willis, J.H., and Franks, R.G. (2016). Disruption of endosperm development is a
major cause of hybrid seed inviability between Mimulus guttatus and Mimulus nudatus.
The New phytologist 210, 1107-1120.
Opsahl-Ferstad, H.G., Le Deunff, E., Dumas, C., and Rogowsky, P.M. (1997). ZmEsr, a
novel endosperm-specific gene expressed in a restricted region around the maize embryo.
The Plant journal : for cell and molecular biology 12, 235-246.
Orozco-Arroyo, G., Paolo, D., Ezquer, I., and Colombo, L. (2015). Networks controlling seed
size in Arabidopsis. Plant Reprod 28, 17-32.
Pate, J.S., and Gunning, B.E.S. (1972). Transfer Cells. Ann Rev Plant Physio 23, 173-&.
Pautler, M., Eveland, A.L., LaRue, T., Yang, F., Weeks, R., Lunde, C., Je, B.I., Meeley, R.,
Komatsu, M., Vollbrecht, E., Sakai, H., and Jackson, D. (2015). FASCIATED EAR4
encodes a bZIP transcription factor that regulates shoot meristem size in maize. The Plant
cell 27, 104-120.
300

Pennington, P.D., Costa, L.M., Gutierrez-Marcos, J.F., Greenland, A.J., and Dickinson,
H.G. (2008). When genomes collide: aberrant seed development following maize
interploidy crosses. Annals of botany 101, 833-843.
Peters, J. (2014). The role of genomic imprinting in biology and disease: an expanding view.
Nature reviews. Genetics 15, 517-530.
Pysh, L.D., and Schmidt, R.J. (1996). Characterization of the maize OHP1 gene: evidence of
gene copy variability among inbreds. Gene 177, 203-208.
Pysh, L.D., Aukerman, M.J., and Schmidt, R.J. (1993). OHP1: a maize basic domain/leucine
zipper protein that interacts with opaque2. The Plant cell 5, 227-236.
Qi, X., Li, S., Zhu, Y., Zhao, Q., Zhu, D., and Yu, J. (2016). ZmDof3, a maize endosperm-
specific Dof protein gene, regulates starch accumulation and aleurone development in
maize endosperm. Plant molecular biology.
Qiao, Z., Qi, W., Wang, Q., Feng, Y., Yang, Q., Zhang, N., Wang, S., Tang, Y., and Song,
R. (2016). ZmMADS47 Regulates Zein Gene Transcription through Interaction with
Opaque2. PLoS genetics 12, e1005991.
Qu, J., Ma, C., Feng, J., Xu, S., Wang, L., Li, F., Li, Y., Zhang, R., Zhang, X., Xue, J., and
Guo, D. (2016). Transcriptome Dynamics during Maize Endosperm Development. PloS
one 11, e0163814.
Quinlan, A.R., and Hall, I.M. (2010). BEDTools: a flexible suite of utilities for comparing
genomic features. Bioinformatics 26, 841-842.
Randolph, L.F. (1936). Developmental morphology of the caryopsis in maize. (United States
Department of Agriculture).
Rebernig, C.A., Lafon-Placette, C., Hatorangan, M.R., Slotte, T., and Kohler, C. (2015).
Non-reciprocal Interspecies Hybridization Barriers in the Capsella Genus Are Established
in the Endosperm. PLoS genetics 11, e1005295.
Reddy, V.M., and Daynard, T.B. (1983). Endosperm Characteristics Associated with Rate of
Grain Filling and Kernel Size in Corn. Maydica 28, 339-355.
Reidt, W., Wohlfarth, T., Ellerstrom, M., Czihal, A., Tewes, A., Ezcurra, I., Rask, L., and
Baumlein, H. (2000). Gene regulation during late embryogenesis: the RY motif of
maturation-specific gene promoters is a direct target of the FUS3 gene product. The Plant
journal : for cell and molecular biology 21, 401-408.
Reyes, F.C., Chung, T., Holding, D., Jung, R., Vierstra, R., and Otegui, M.S. (2011).
Delivery of prolamins to the protein storage vacuole in maize aleurone cells. The Plant
cell 23, 769-784.
Riechmann, J.L., Wang, M., and Meyerowitz, E.M. (1996). DNA-binding properties of
Arabidopsis MADS domain homeotic proteins APETALA1, APETALA3, PISTILLATA
and AGAMOUS. Nucleic acids research 24, 3134-3141.
Robinson, M.D., McCarthy, D.J., and Smyth, G.K. (2010). edgeR: a Bioconductor package
for differential expression analysis of digital gene expression data. Bioinformatics 26,
139-140.
Rodrigues, J.A., and Zilberman, D. (2015). Evolution and function of genomic imprinting in
plants. Genes & development 29, 2517-2531.
Roosjen, M., Paque, S., and Weijers, D. (2018). Auxin Response Factors: output control in
auxin biology. Journal of experimental botany 69, 179-188.
Royo, J., Gomez, E., Sellam, O., Gerentes, D., Paul, W., and Hueros, G. (2014). Two maize
END-1 orthologs, BETL9 and BETL9like, are transcribed in a non-overlapping spatial
301

pattern on the outer surface of the developing endosperm. Frontiers in plant science 5,
180.
Sabelli, P.A., and Larkins, B.A. (2009). The development of endosperm in grasses. Plant
physiology 149, 14-26.
Sabelli, P.A., Dante, R.A., Leiva-Neto, J.T., Jung, R., Gordon-Kamm, W.J., and Larkins,
B.A. (2005). RBR3, a member of the retinoblastoma-related family from maize, is
regulated by the RBR1/E2F pathway. Proceedings of the National Academy of Sciences
of the United States of America 102, 13005-13012.
Saleh, A., Alvarez-Venegas, R., and Avramova, Z. (2008). An efficient chromatin
immunoprecipitation (ChIP) protocol for studying histone modifications in Arabidopsis
plants. Nature protocols 3, 1018-1025.
Santos-Mendoza, M., Dubreucq, B., Baud, S., Parcy, F., Caboche, M., and Lepiniec, L.
(2008). Deciphering gene regulatory networks that control seed development and
maturation in Arabidopsis. The Plant journal : for cell and molecular biology 54, 608-
620.
Scanlon, M.J., Stinard, P.S., James, M.G., Myers, A.M., and Robertson, D.S. (1994).
Genetic analysis of 63 mutations affecting maize kernel development isolated from
Mutator stocks. Genetics 136, 281-294.
Schel, J.H.N., Kieft, H., and Vanlammeren, A.A.M. (1984). Interactions between Embryo and
Endosperm during Early Developmental Stages of Maize Caryopses (Zea-Mays). Can J
Bot 62, 2842-2853.
Schmidt, R.J., Burr, F.A., and Burr, B. (1987). Transposon tagging and molecular analysis of
the maize regulatory locus opaque-2. Science 238, 960-963.
Schmidt, R.J., Burr, F.A., Aukerman, M.J., and Burr, B. (1990). Maize regulatory gene
opaque-2 encodes a protein with a "leucine-zipper" motif that binds to zein DNA.
Proceedings of the National Academy of Sciences of the United States of America 87,
46-50.
Schmidt, R.J., Ketudat, M., Aukerman, M.J., and Hoschek, G. (1992). Opaque-2 is a
transcriptional activator that recognizes a specific target site in 22-kD zein genes. The
Plant cell 4, 689-700.
Scott, R.J., Spielman, M., Bailey, J., and Dickinson, H.G. (1998). Parent-of-origin effects on
seed development in Arabidopsis thaliana. Development 125, 3329-3341.
Sekhon, R.S., Lin, H., Childs, K.L., Hansey, C.N., Buell, C.R., de Leon, N., and Kaeppler,
S.M. (2011). Genome-wide atlas of transcription during maize development. The Plant
journal : for cell and molecular biology 66, 553-563.
Sekhon, R.S., Briskine, R., Hirsch, C.N., Myers, C.L., Springer, N.M., Buell, C.R., de Leon,
N., and Kaeppler, S.M. (2013). Maize gene atlas developed by RNA sequencing and
comparative evaluation of transcriptomes based on RNA sequencing and microarrays.
PloS one 8, e61005.
Sekhon, R.S., Hirsch, C.N., Childs, K.L., Breitzman, M.W., Kell, P., Duvick, S., Spalding,
E.P., Buell, C.R., de Leon, N., and Kaeppler, S.M. (2014). Phenotypic and
Transcriptional Analysis of Divergently Selected Maize Populations Reveals the Role of
Developmental Timing in Seed Size Determination. Plant physiology 165, 658-669.
Selinger, D.A., and Chandler, V.L. (1999). A mutation in the pale aleurone color1 gene
identifies a novel regulator of the maize anthocyanin pathway. The Plant cell 11, 5-14.
302

Serna, A., Maitz, M., O'Connell, T., Santandrea, G., Thevissen, K., Tienens, K., Hueros, G.,
Faleri, C., Cai, G., Lottspeich, F., and Thompson, R.D. (2001). Maize endosperm
secretes a novel antifungal protein into adjacent maternal tissue. The Plant journal : for
cell and molecular biology 25, 687-698.
Sevilla-Lecoq, S., Deguerry, F., Matthys-Rochon, E., Perez, P., Dumas, C., and Rogowsky,
P.M. (2003). Analysis of ZmAE3 upstream sequences in maize endosperm and
androgenic embryos. Sex Plant Reprod 16, 1-8.
Shannon, J.C., Porter, G.A., and Knievel, D.P. (1986). Phloem unloading and transfer of
sugars into developing corn endosperm. Plant biology (USA).
Shannon, P., Markiel, A., Ozier, O., Baliga, N.S., Wang, J.T., Ramage, D., Amin, N.,
Schwikowski, B., and Ideker, T. (2003). Cytoscape: a software environment for
integrated models of biomolecular interaction networks. Genome research 13, 2498-2504.
Sheridan, W.F., and Neuffer, M.G. (1980). Defective Kernel Mutants of Maize II.
Morphological and Embryo Culture Studies. Genetics 95, 945-960.
Shewry, P.R., and Halford, N.G. (2002). Cereal seed storage proteins: structures, properties
and role in grain utilization. Journal of experimental botany 53, 947-958.
Shewry, P.R., Napier, J.A., and Tatham, A.S. (1995). Seed storage proteins: structures and
biosynthesis. The Plant cell 7, 945-956.
Shure, M., Wessler, S., and Fedoroff, N. (1983). Molecular identification and isolation of the
Waxy locus in maize. Cell 35, 225-233.
Sievers, F., Wilm, A., Dineen, D., Gibson, T.J., Karplus, K., Li, W.Z., Lopez, R.,
McWilliam, H., Remmert, M., Soding, J., Thompson, J.D., and Higgins, D.G. (2011).
Fast, scalable generation of high-quality protein multiple sequence alignments using
Clustal Omega. Mol Syst Biol 7.
Slane, D., Kong, J., Berendzen, K.W., Kilian, J., Henschen, A., Kolb, M., Schmid, M.,
Harter, K., Mayer, U., De Smet, I., Bayer, M., and Jurgens, G. (2014). Cell type-
specific transcriptome analysis in the early Arabidopsis thaliana embryo. Development
141, 4831-4840.
Sorensen, M.B., Chaudhury, A.M., Robert, H., Bancharel, E., and Berger, F. (2001).
Polycomb group genes control pattern formation in plant seed. Current biology : CB 11,
277-281.
Sosso, D., Wisniewski, J.P., Khaled, A.S., Hueros, G., Gerentes, D., Paul, W., and
Rogowsky, P.M. (2010). The Vpp1, Esr6a, Esr6b and OCL4 promoters are active in
distinct domains of maize endosperm. Plant Science 179, 86-96.
Sreenivasulu, N., and Wobus, U. (2013). Seed-Development Programs: A Systems Biology-
Based Comparison Between Dicots and Monocots. Annual Review of Plant Biology, Vol
64 64, 189-+.
Stamatakis, A. (2014). RAxML version 8: a tool for phylogenetic analysis and post-analysis of
large phylogenies. Bioinformatics 30, 1312-1313.
Steward, N., Kusano, T., and Sano, H. (2000). Expression of ZmMET1, a gene encoding a
DNA methyltransferase from maize, is associated not only with DNA replication in
actively proliferating cells, but also with altered DNA methylation status in cold-stressed
quiescent cells. Nucleic acids research 28, 3250-3259.
Tailor, R.H., Acland, D.P., Attenborough, S., Cammue, B.P., Evans, I.J., Osborn, R.W.,
Ray, J.A., Rees, S.B., and Broekaert, W.F. (1997). A novel family of small cysteine-
303

rich antimicrobial peptides from seed of Impatiens balsamina is derived from a single
precursor protein. Journal of Biological Chemistry 272, 24480-24487.
Takahashi, H., Kamakura, H., Sato, Y., Shiono, K., Abiko, T., Tsutsumi, N., Nagamura, Y.,
Nishizawa, N.K., and Nakazono, M. (2010). A method for obtaining high quality RNA
from paraffin sections of plant tissues by laser microdissection. J Plant Res 123, 807-813.
Tanaka, H., Onouchi, H., Kondo, M., Hara-Nishimura, I., Nishimura, M., Machida, C., and
Machida, Y. (2001). A subtilisin-like serine protease is required for epidermal surface
formation in Arabidopsis embryos and juvenile plants. Development 128, 4681-4689.
Taylor-Teeples, M., Lin, L., de Lucas, M., Turco, G., Toal, T.W., Gaudinier, A., Young,
N.F., Trabucco, G.M., Veling, M.T., Lamothe, R., Handakumbura, P.P., Xiong, G.,
Wang, C., Corwin, J., Tsoukalas, A., Zhang, L., Ware, D., Pauly, M., Kliebenstein,
D.J., Dehesh, K., Tagkopoulos, I., Breton, G., Pruneda-Paz, J.L., Ahnert, S.E., Kay,
S.A., Hazen, S.P., and Brady, S.M. (2015). An Arabidopsis gene regulatory network for
secondary cell wall synthesis. Nature 517, 571-575.
Tekleyohans, D.G., Nakel, T., and Gross-Hardt, R. (2017). Patterning the Female
Gametophyte of Flowering Plants. Plant physiology 173, 122-129.
Thakare, D., Yang, R., Steffen, J.G., Zhan, J., Wang, D., Clark, R.M., Wang, X., and
Yadegari, R. (2014). RNA-Seq analysis of laser-capture microdissected cells of the
developing central starchy endosperm of maize. Genom Data 2, 242-245.
Thompson, R.D., and Verdier, J. (2012). Networks of Seed Storage Protein Regulation in
Cereals and Legumes at the Dawn of the Omics Era. In Seed Development: OMICS
Technologies toward Improvement of Seed Quality and Crop Yield (Springer), pp. 187-
210.
Thompson, R.D., Hueros, G., Becker, H., and Maitz, M. (2001). Development and functions
of seed transfer cells. Plant science : an international journal of experimental plant
biology 160, 775-783.
Thorvaldsdottir, H., Robinson, J.T., and Mesirov, J.P. (2013). Integrative Genomics Viewer
(IGV): high-performance genomics data visualization and exploration. Briefings in
bioinformatics 14, 178-192.
Tian, G.W., You, R.L., Guo, F.L., and Wang, X.C. (1998). Microtubular cytoskeleton of free
endosperm nuclei during division in wheat. Cytologia 63, 427-433.
Townsend, J.A., Wright, D.A., Winfrey, R.J., Fu, F., Maeder, M.L., Joung, J.K., and
Voytas, D.F. (2009). High-frequency modification of plant genes using engineered zinc-
finger nucleases. Nature 459, 442-445.
Trapnell, C., Pachter, L., and Salzberg, S.L. (2009). TopHat: discovering splice junctions with
RNA-Seq. Bioinformatics 25, 1105-1111.
Trapnell, C., Williams, B.A., Pertea, G., Mortazavi, A., Kwan, G., van Baren, M.J.,
Salzberg, S.L., Wold, B.J., and Pachter, L. (2010). Transcript assembly and
quantification by RNA-Seq reveals unannotated transcripts and isoform switching during
cell differentiation. Nature biotechnology 28, 511-515.
Trapnell, C., Roberts, A., Goff, L., Pertea, G., Kim, D., Kelley, D.R., Pimentel, H.,
Salzberg, S.L., Rinn, J.L., and Pachter, L. (2012). Differential gene and transcript
expression analysis of RNA-seq experiments with TopHat and Cufflinks. Nature
protocols 7, 562-578.
304

Unger, E., Parsons, R.L., Schmidt, R.J., Bowen, B., and Roth, B.A. (1993). Dominant
Negative Mutants of Opaque2 Suppress Transactivation of a 22-kD Zein Promoter by
Opaque2 in Maize Endosperm Cells. The Plant cell 5, 831-841.
Vicente-Carbajosa, J., Moose, S.P., Parsons, R.L., and Schmidt, R.J. (1997). A maize zinc-
finger protein binds the prolamin box in zein gene promoters and interacts with the basic
leucine zipper transcriptional activator Opaque2. Proceedings of the National Academy
of Sciences of the United States of America 94, 7685-7690.
Vojtek, A.B., and Hollenberg, S.M. (1995). Ras-Raf interaction: two-hybrid analysis. Methods
in enzymology 255, 331-342.
von Wangenheim, K.H., and Peterson, H.P. (2004). Aberrant endosperm development in
interploidy crosses reveals a timer of differentiation. Developmental biology 270, 277-
289.
Walley, J.W., Shen, Z., Sartor, R., Wu, K.J., Osborn, J., Smith, L.G., and Briggs, S.P.
(2013). Reconstruction of protein networks from an atlas of maize seed proteotypes.
Proceedings of the National Academy of Sciences of the United States of America.
Walley, J.W., Sartor, R.C., Shen, Z., Schmitz, R.J., Wu, K.J., Urich, M.A., Nery, J.R.,
Smith, L.G., Schnable, J.C., Ecker, J.R., and Briggs, S.P. (2016). Integration of omic
networks in a developmental atlas of maize. Science 353, 814-818.
Walsh, J.R., Schaeffer, M.L., Zhang, P., Rhee, S.Y., Dickerson, J.A., and Sen, T.Z. (2016).
The quality of metabolic pathway resources depends on initial enzymatic function
assignments: a case for maize. BMC systems biology 10, 129.
Wang, D., and Bodovitz, S. (2010). Single cell analysis: the new frontier in 'omics'. Trends
Biotechnol 28, 281-290.
Wang, D., and Deal, R.B. (2015). Epigenome profiling of specific plant cell types using a
streamlined INTACT protocol and ChIP-seq. Methods in molecular biology 1284, 3-25.
Wang, D., Tyson, M.D., Jackson, S.S., and Yadegari, R. (2006). Partially redundant functions
of two SET-domain polycomb-group proteins in controlling initiation of seed
development in Arabidopsis. Proceedings of the National Academy of Sciences of the
United States of America 103, 13244-13249.
Wang, D., Zhang, C., Hearn, D.J., Kang, I.H., Punwani, J.A., Skaggs, M.I., Drews, G.N.,
Schumaker, K.S., and Yadegari, R. (2010). Identification of transcription-factor genes
expressed in the Arabidopsis female gametophyte. BMC plant biology 10, 110.
Waters, A.J., Makarevitch, I., Eichten, S.R., Swanson-Wagner, R.A., Yeh, C.T., Xu, W.,
Schnable, P.S., Vaughn, M.W., Gehring, M., and Springer, N.M. (2011). Parent-of-
origin effects on gene expression and DNA methylation in the maize endosperm. The
Plant cell 23, 4221-4233.
Whelan, S., and Goldman, N. (2001). A general empirical model of protein evolution derived
from multiple protein families using a maximum-likelihood approach. Molecular biology
and evolution 18, 691-699.
Widiez, T., Ingram, G.C., and Gutierrez-Marcos, J.F. (2017). Embryo-Endosperm-
Sporophyte Interactions in Maize Seeds. In Maize Kernel Development, B.A. Larkins, ed
(CAB International), pp. 95-107.
Wisniewski, J.P., and Rogowsky, P.M. (2004). Vacuolar H+-translocating inorganic
pyrophosphatase (Vpp1) marks partial aleurone cell fate in cereal endosperm
development. Plant molecular biology 56, 325-337.
305

Wolff, P., Jiang, H., Wang, G., Santos-Gonzalez, J., and Kohler, C. (2015). Paternally
expressed imprinted genes establish postzygotic hybridization barriers in Arabidopsis
thaliana. eLife 4.
Woo, Y.M., Hu, D.W., Larkins, B.A., and Jung, R. (2001). Genomics analysis of genes
expressed in maize endosperm identifies novel seed proteins and clarifies patterns of zein
gene expression. The Plant cell 13, 2297-2317.
Wu, Y., and Messing, J. (2012). Rapid divergence of prolamin gene promoters of maize after
gene amplification and dispersal. Genetics 192, 507-519.
Xiao, W., Brown, R.C., Lemmon, B.E., Harada, J.J., Goldberg, R.B., and Fischer, R.L.
(2006). Regulation of seed size by hypomethylation of maternal and paternal genomes.
Plant physiology 142, 1160-1168.
Xin, M., Yang, R., Li, G., Chen, H., Laurie, J., Ma, C., Wang, D., Yao, Y., Larkins, B.A.,
Sun, Q., Yadegari, R., Wang, X., and Ni, Z. (2013). Dynamic expression of imprinted
genes associates with maternally controlled nutrient allocation during maize endosperm
development. The Plant cell 25, 3212-3227.
Xing, Q., Creff, A., Waters, A., Tanaka, H., Goodrich, J., and Ingram, G.C. (2013).
ZHOUPI controls embryonic cuticle formation via a signalling pathway involving the
subtilisin protease ABNORMAL LEAF-SHAPE1 and the receptor kinases GASSHO1
and GASSHO2. Development 140, 770-779.
Xiong, Y., Mei, W., Kim, E.D., Mukherjee, K., Hassanein, H., Barbazuk, W.B., Sung, S.,
Kolaczkowski, B., and Kang, B.H. (2014). Adaptive expansion of the maize maternally
expressed gene (Meg) family involves changes in expression patterns and protein
secondary structures of its members. BMC plant biology 14, 204.
Xu, J.H., and Messing, J. (2008). Diverged copies of the seed regulatory Opaque-2 gene by a
segmental duplication in the progenitor genome of rice, sorghum, and maize. Molecular
plant 1, 760-769.
Xue, Z., Huang, K., Cai, C., Cai, L., Jiang, C.Y., Feng, Y., Liu, Z., Zeng, Q., Cheng, L.,
Sun, Y.E., Liu, J.Y., Horvath, S., and Fan, G. (2013). Genetic programs in human and
mouse early embryos revealed by single-cell RNA sequencing. Nature 500, 593-597.
Yan, D., Duermeyer, L., Leoveanu, C., and Nambara, E. (2014). The Functions of the
Endosperm During Seed Germination. Plant & cell physiology 55, 1521-1533.
Yang, J., Ji, C., and Wu, Y. (2016). Divergent Transactivation of Maize Storage Protein Zein
Genes by the Transcription Factors Opaque2 and OHPs. Genetics 204, 581-591.
Yang, S., Johnston, N., Talideh, E., Mitchell, S., Jeffree, C., Goodrich, J., and Ingram, G.
(2008). The endosperm-specific ZHOUPI gene of Arabidopsis thaliana regulates
endosperm breakdown and embryonic epidermal development. Development 135, 3501-
3509.
Yi, G., Neelakandan, A.K., Gontarek, B.C., Vollbrecht, E., and Becraft, P.W. (2015). The
naked endosperm genes encode duplicate INDETERMINATE domain transcription
factors required for maize endosperm cell patterning and differentiation. Plant physiology
167, 443-456.
Yilmaz, A., Nishiyama, M.Y., Jr., Fuentes, B.G., Souza, G.M., Janies, D., Gray, J., and
Grotewold, E. (2009). GRASSIUS: a platform for comparative regulatory genomics
across the grasses. Plant physiology 149, 171-180.
Young, T.E., and Gallie, D.R. (2000). Regulation of programmed cell death in maize
endosperm by abscisic acid. Plant molecular biology 42, 397-414.
306

Young, T.E., Gallie, D.R., and DeMason, D.A. (1997). Ethylene-Mediated Programmed Cell
Death during Maize Endosperm Development of Wild-Type and shrunken2 Genotypes.
Plant physiology 115, 737-751.
Yuan, J., Bateman, P., and Gutierrez-Marcos, J. (2016). Genetic and epigenetic control of
transfer cell development in plants. Journal of genetics and genomics = Yi chuan xue bao
43, 533-539.
Zar, J.H. (1972). Significance Testing of Spearman Rank Correlation Coefficient. J Am Stat
Assoc 67, 578-580.
Zhan, J., Dannenhoffer, J.M., and Yadegari, R. (2017). Endosperm Development and Cell
Specialization. In Maize Kernel Development, B.A. Larkins, ed (CAB International), pp.
28-43.
Zhan, J., Thakare, D., Ma, C., Lloyd, A., Nixon, N.M., Arakaki, A.M., Burnett, W.J.,
Logan, K.O., Wang, D., Wang, X., Drews, G.N., and Yadegari, R. (2015). RNA
sequencing of laser-capture microdissected compartments of the maize kernel identifies
regulatory modules associated with endosperm cell differentiation. The Plant cell 27,
513-531.
Zhang, B., and Horvath, S. (2005). A general framework for weighted gene coexpression
network analysis. Statistical applications in genetics and molecular biology 4, Article17.
Zhang, M., Zhao, H., Xie, S., Chen, J., Xu, Y., Wang, K., Zhao, H., Guan, H., Hu, X., Jiao,
Y., Song, W., and Lai, J. (2011). Extensive, clustered parental imprinting of protein-
coding and noncoding RNAs in developing maize endosperm. Proceedings of the
National Academy of Sciences of the United States of America 108, 20042-20047.
Zhang, M., Xie, S., Dong, X., Zhao, X., Zeng, B., Chen, J., Li, H., Yang, W., Zhao, H.,
Wang, G., Chen, Z., Sun, S., Hauck, A., Jin, W., and Lai, J. (2014). Genome-wide
high resolution parental-specific DNA and histone methylation maps uncover patterns of
imprinting regulation in maize. Genome research 24, 167-176.
Zhang, N., Qiao, Z., Liang, Z., Mei, B., Xu, Z., and Song, R. (2012). Zea mays Taxilin protein
negatively regulates opaque-2 transcriptional activity by causing a change in its sub-
cellular distribution. PloS one 7, e43822.
Zhang, Y., Zhang, F., Li, X., Baller, J.A., Qi, Y., Starker, C.G., Bogdanove, A.J., and
Voytas, D.F. (2013). Transcription activator-like effector nucleases enable efficient plant
genome engineering. Plant physiology 161, 20-27.
Zhang, Y., Liu, T., Meyer, C.A., Eeckhoute, J., Johnson, D.S., Bernstein, B.E., Nusbaum,
C., Myers, R.M., Brown, M., Li, W., and Liu, X.S. (2008). Model-based analysis of
ChIP-Seq (MACS). Genome biology 9, R137.
Zhang, Z., Yang, J., and Wu, Y. (2015). Transcriptional Regulation of Zein Gene Expression
in Maize through the Additive and Synergistic Action of opaque2, Prolamine-Box
Binding Factor, and O2 Heterodimerizing Proteins. The Plant cell 27, 1162-1172.
Zhang, Z., Zheng, X., Yang, J., Messing, J., and Wu, Y. (2016). Maize endosperm-specific
transcription factors O2 and PBF network the regulation of protein and starch synthesis.
Proceedings of the National Academy of Sciences of the United States of America 113,
10842-10847.

You might also like