new regulators of xylem lignification in arabidopsis

84
New regulators of xylem lignification in Arabidopsis Bernadette Sztojka Umeå Plant Science Centre Department of Plant Physiology Umeå 2020

Upload: others

Post on 15-Feb-2022

1 views

Category:

Documents


0 download

TRANSCRIPT

New regulators of xylem

lignification in Arabidopsis

Bernadette Sztojka

Umeå Plant Science Centre

Department of Plant Physiology

Umeå 2020

This work is protected by the Swedish Copyright Legislation (Act 1960:729)

Dissertation for PhD

ISBN (printed version): 978-91-7855-428-7

ISBN (digital version): 978-91-7855-429-4

Cover design: Schematic representation of an Arabidopsis thaliana plant; Illustration of

transverse sections of the xylem in the base of the stem and in the hypocotyl of

Arabidopsis thaliana; Schematic view of xylem cells. (by Bernadette Sztojka)

Electronic version available at: http://umu.diva-portal.org/

Printed by: KBC Service Centre, Umeå University

Umeå, Sweden 2020

Szüleimnek.

To my parents.

i

Table of Contents

Table of Contents i

Abstract iii

Sammanfattning iv

Abbreviations v

List of publications ix

Introduction 1

1. The importance of wood 1

2. Xylem 2

2.1. Vascular cambium 2

Vascular cambium and secondary growth 2

Vascular cambium specification and maintenance 2

2.2. Arabidopsis as a model to study the basic molecular mechanisms of

secondary growth 6

2.3. The function of the xylem 7

2.4. Different cell types building up the xylem 8

Tracheids and vessel elements 8

Fibers 9

Parenchyma cells and ray cells 10

3. Secondary cell wall 11

3.1. Cellulose 11

3.2. Hemicellulose 13

3.3. Cell wall proteins 14

3.4. Lignin 15

The evolution of lignin 15

Major monolignols 15

Distribution of lignin in different species and cell types 16

The function of lignin 16

3.4.1. Biosynthesis of the lignin monomers 16

ii

3.4.1.a. The lignin biosynthetic pathway 16

3.4.1.b. Regulation of lignin biosynthesis 18

Transcriptional regulation 19

Chromatin-level regulation 22

Circadian regulation 22

Additional levels of regulation 23

3.4.2. Transport of the lignin monomers 23

3.4.3. Lignin polymerization 25

Laccases and peroxidases 25

Secondary substrates 27

Assembly of the polymer 28

3.4.4. Cell-autonomous vs. non-cell-autonomous lignification 29

3.4.5. Lignin as a resource for biotechnological applications 31

Research aims and objectives 34

Results and discussion 35

PIRIN2 is a non-cell-autonomous regulator of S-type lignin accumulation 35

HISTONE MONOUBIQUITINATION2 (HUB2) affects the lignin

composition of xylem vessels 37

PIB is a potential modulator of the diurnal timing of lignin biosynthesis 39

Conclusions and perspectives 41

Acknowledgements 43

References 45

iii

Abstract

The ability of land plants to grow upright, bear their own weight and withstand

adverse environmental conditions is largely dependent on the secondary

xylem tissues of the stem. The xylem cells acquire thick secondary cell walls

which are composed of cellulose, hemicellulose and lignin. The chemical

structure of lignin renders the secondary cell wall rigid and waterproof,

facilitating the transport of water and solutes through the vascular system.

Lignin is a polyphenolic polymer composed of three different types of lignin

units, guaiacyl (G), syringyl (S) and p-hydroxyphenyl (H), derived from the

coniferyl, sinapyl and p-coumaryl alcohol, respectively. Lignin biosynthesis,

monolignol transport and lignin polymerization (collectively called as

”lignification”) are controlled by numerous transcription factors and other

regulators.

This thesis work uncovers three novel regulators of lignification in the

secondary xylem tissues of Arabidopsis (Arabidopsis thaliana) stem and

hypocotyl. The cupin domain containing protein PIRIN2 (PRN2) suppresses

S-type lignin accumulation. PRN2 functions in a non-cell-autonomous

fashion: it is expressed in the cells next to the xylem vessel elements, but

affects the lignin composition of the vessel and fiber cell walls of the

neighbouring cells. Two protein interactors of PRN2 are characterized here in

connection to lignification. Opposite to the function of PRN2, the chromatin-

modifying protein HISTONE MONOUBIQUITINATION2 (HUB2)

promotes S-type lignin deposition. In line with this, PRN2 and HUB2

antagonistically regulate the expression of FERULATE-5-HYDROXYLASE1

which encodes the key S-type lignin-biosynthetic enzyme. Possibly, PRN2

antagonizes the S-lignin promoting function of HUB2 to ensure that the cell

walls of the vessel elements get enriched in G-type lignin. Finally,

identification of a potential diurnal modulator of lignin biosynthesis is

described in this work. The PRN2-interacting basic helix-loop-helix

transcription factor (PIB) does not influence the lignin content or composition

of the secondary cell walls. However, PIB affects the diurnal expression

pattern and promoter activity of some lignin-biosynthetic genes. Altogether,

PRN2, HUB2 and PIB highlight the importance of intercellular co-operation

in lignification, and uncover novel regulatory aspects of this process.

iv

Sammanfattning

Växternas förmåga att växa upprätt, bära sin egen vikt och tåla ogynnsamma

miljöförhållanden styrs till stor del av stammens vaskulär vävnad som består

av floem- och xylemceller. Xylemceller ackumulerar tjocka sekundära

cellväggar som är sammansatta av cellulosa, hemicellulosa och lignin. Den

kemiska strukturen hos lignin gör cellväggen styv och vattentät, vilket

möjliggör transport av vatten och näringsämnen genom det vaskulära

systemet. Lignin är en polyfenol som består av tre olika typer av lignin

enheter: guaiacyl (G), syringyl (S) och p-hydroxyfenol (H). Lignin biosyntes,

monolignol transport och lignin polymerisering (kollektivt kallad som

"lignifiering") styrs av ett flertal transkriptionsfaktorer och andra regulatorer.

Denna avhandling avslöjar tre nya regulatorer av lignifiering i xylemvävnader

av Arabidopsis (Arabidopsis thaliana) stam och hypokotyl. Genetiska och

kemiska analyser avslöjade ett nytt protein, PIRIN2 (PRN2), som dämpar S-

typ lignin ackumulering. PRN2 fungerar på ett icke-cell-autonomt sätt: det

uttrycks i cellerna bredvid xylemkärlselement, men påverkar

ligninsammansättningen hos de angränsande kärlens cellväggar. I

avhandlingsarbetet identifierades och karakteriserades också två andra

proteiner som interagerar med PRN2 i lignifiering. I motsats till funktionen

av PRN2, det kromatin-modifierande HISTONE

MONOUBIQUITINATION2 (HUB2) främjar S-typ lignin ackumulering. I

linje med detta reglerar PRN2 och HUB2 antagonistiskt uttrycket av

FERULATE-5-HYDROXYLASE1 som kontrollerar biosyntes av S-typ lignin.

Det verkar troligt att PRN2 motverkar S-lignin främjande funktion av HUB2

för att säkerställa anrikning av G-typ lignin i kärlelementens cellväggar.

Slutligen beskrivs här en helix-loop-helix transkriptionsfaktor (PIB) som

potentiellt modulerar lignin biosyntes enligt dygnsrytmen. PIB påverkar inte

lignin innehållet eller sammansättningen av de sekundära cellväggarna men

påverkar tajmningen av genuttrycket och promotoraktiviteten hos vissa

lignin-biosyntetiska gener. Sammantaget belyser PRN2, HUB2 och PIB

vikten av intercellulärt samarbete i lignifiering och avslöjar nya

regleringsaspekter av denna process.

v

Abbreviations

ABC ATP-binding cassette

AGP Arabinogalactan-rich glycoproteins

AHP6 ARABIDOPSIS HISTIDINE PHOSPHOTRANSFER

PROTEIN6

ANT AINTEGUMENTA

Arabidopsis Arabidopsis thaliana

ARF Auxin response factor

ARR Arabidopsis response regulator

BES1 BRI1-EMS SUPPRESSOR1

BIL1 BRASSINOSTEROID-INSENSITIVE2-LIKE1

CAD Cinnamyl alcohol dehydrogenase

CCoAOMT Caffeoyl-CoA O-methyltransferase

CCR Cinnamoyl-CoA reductase

CLE CLAVATA3/EMBRYO SURROUNDING REGION-

related

COMT Caffeic acid O-methyltransferase

CSC Cellulose synthase protein complex

C3H P-coumarate 3-hydroxylase

C4H Cinnamate 4-hydroxylase

4CL 4-coumarate:CoA ligase

C-type lignin Caffeyl alcohol

ER Endoplasmic reticulum

ESB1/DIR10 ENHANCED SUBERIN1/DIRIGENT10

FAO Food and Agriculture Organization of the United Nations

FLA Fasciclin-like arabinogalactan-protein

F5H Ferulate 5-hydroxylase

GlcA Glucuronic acid

vi

Abbreviations (2/4)

GRP Glycine-rich glycoproteins

GSK3 Glycogen synthase kinase

GT43 Glycosyltransferase family 43

GUX Glucuronic acid substitution of xylan

GXMT Glucuronoxylan methyltransferase

G-type lignin Guaiacyl-type lignin

HCT Hydroxycinnamoyltransferase

HD-ZIP III Class III homeodomain-leucine zipper

HRGP Hydroxyproline-rich glycoproteins

H3K4me3 Histone H3 lysine K4 methylation

H3K27me3 Histone H3 lysine K27 methylation

H2O2 Hydrogen peroxide

H-type lignin P-hydroxyphenyl-type lignin

IRX Irregular xylem

IRX-L Irregular xylem-like

KFB Kelch-repeat F-box

KNOX Knotted-like homeobox

LAC Laccase

LHW LONESOME HIGHWAY

LOG LONELY GUY

MED Mediator

[Me]GlcA Methyl-glucuronic acid

MOL1 MORE LATERAL GROWTH1

MP MONOPTEROS

MYB Myeloblastosis

vii

Abbreviations (3/4)

NAC NAM, ATAF1/2, CUC2

NADPH oxidase Nicotineamide adenine dinucleotide phosphate hydrogen

oxidase

NST NAC SECONDARY WALL THICKENING

PROMOTING FACTOR

O2 Molecular oxygen

PAL Phenylalanine ammonia-lyase

PCW Primary cell wall

PEAR PEAR PHLOEM EARLY DOF1

PIN PIN-FORMED

PRP Proline-rich glycoproteins

PRX Peroxidase

PXY PHLOEM INTERCALATED WITH XYLEM

TDR TRACHEARY ELEMENT DIFFERENTIATION

INHIBITORY FACTOR RECEPTOR

RUL1 REDUCED IN LATERAL GROWTH1

RWA REDUCED WALL ACETYLATION

SCW Secondary cell wall

SERKs SOMATIC EMBRYOGENESIS RECEPTOR KINASEs

SMRE Secondary wall MYB-responsive element

SMXL5 SUPPRESSOR OF MAX2 1-LIKE5

SNBE Secondary wall NAC-binding element

SND SECONDARY WALL-ASSOCIATED NAC DOMAIN

SOD Superoxide dismutase

S-type lignin Syringyl-type lignin

TDIF TRACHEARY ELEMENT DIFFERENTIATION

INHIBITORY FACTOR

viii

Abbreviations (4/4)

TE Tracheary element

TED Tracheary element differentiation-related

TF Transcription factor

TMO TARGET OF MONOPTEROS

T5L1 TMO5-LIKE1

VND VASCULAR RELATED NAC DOMAIN

VNI2 VND-INTERACTING2

VNS VND, NST/SND, and SMB-related protein

WOX WUSCHEL-related homeobox

XND1 XYLEM NAC DOMAIN1

YUC4 YUCCA4

Other abbreviations are explained in the text, when they first appear.

ix

List of publications

Paper I

Zhang B*, Sztojka B*, Escamez S, Vanholme R, Hedenström M, Wang Y,

Turumtay H, Gorzsás A, Boerjan W, Tuominen H. 2020. PIRIN2 suppresses S-

type lignin accumulation in a noncell-autonomous manner in Arabidopsis xylem

elements. New Phytologist 225: 1923-1935.

* These authors contributed equally.

Paper II

Zhang B*, Sztojka B*, Seyfferth C, Escamez S, Miskolczi P, Chantreau M, Bakó

L, Delhomme N, Gorzsás A, Bhalerao RP, Tuominen H. 2020. The chromatin-

modifying protein HUB2 is involved in the regulation of lignin composition in

xylem vessels. Journal of Experimental Botany 71: 5484-5494.

* These authors contributed equally.

Paper III

Sztojka B, Escamez S, Seyfferth C, Wang Y, Zhang B, Gorzsás A, Demedts B,

Boerjan W, Tuominen H. (manuscript). PIB – a potential modulator of

lignification in Arabidopsis.

Publication not included in the PhD thesis

Escamez S, André D, Sztojka B, Bollhöner B, Hall H, Berthet B, Voß U, Lers A,

Maizel A, Andersson M, Bennett M, Tuominen H. 2020. Cell death in cells

overlying lateral root primordia facilitates organ growth in Arabidopsis. Current

Biology 30: 455-464.

x

Author’s contribution

Paper I: Obtained plant material and prepared samples for FT-IR and Raman

microspectroscopy, took part in the data analysis and interpretation. Performed

GUS staining of the Populus tremula x tremuloides proPttPRN2::GUS lines.

Obtained plant material for the PRN2 overexpression lines and prepared samples

for pyrolysis-GC/MS, performed the data analysis. Contributed to data

interpretation, the writing and formatting of the manuscript.

Paper II: Obtained plant material for the PRN2OE6 and hub2-1 sample set,

prepared samples for pyrolysis-GC/MS, performed the data analysis. Obtained

plant material and prepared samples for FT-IR and Raman microspectroscopy,

took part in the data analysis and interpretation. Performed the greenhouse

experiment, collected plant material, and prepared RNA for RNA-sequencing,

took part in the data analysis. Contributed to data interpretation, the writing and

formatting of the manuscript.

Paper III: Cloned the proPIB:GUS constructs, generated and selected transgenic

lines, and performed GUS staining. Cloned the proPIB::PIB:mCherry construct,

generated and selected transgenic lines. Performed DAPI staining and confocal

microscopic imaging of the proPIB::PIB:mCherry seedlings. Obtained plant

material, prepared samples for pyrolysis-GC/MS, performed the data analysis.

Performed the greenhouse experiment, collected plant material, and prepared

RNA for RNA-sequencing, took part in the data analysis. Performed the diurnal

experiment, analyzed the data. Assessed the gametophytic lethality of the pib-1

pib-like-1 double mutants. Obtained plant material and determined the acetyl

bromide soluble lignin content. Obtained plant material and prepared samples for

FT-IR microspectroscopy, took part in the data analysis. Prepared samples for the

monosaccharide analysis by acidic methanolysis and TMS derivatization,

performed the data analysis. Carried out the growth assay of inflorescence stems.

Performed two of the experiments to assay root elongation in response to different

sucrose concentrations. Wrote the manuscript with contribution from the co-

authors.

1

Introduction

1. The importance of wood

Wood is an essential, renewable bioresource with massive ecological and

economic importance. As of 2015, 30,6 % of the global land area is covered by

forest (UN, 2017). Wood is utilized in numerous ways, for example as raw

material for construction, furnishings, paper and packaging, toolmaking, and art

(Meents et al., 2018). According to data from the Food and Agriculture

Organization of the United Nations (FAO), one third of the world’s population,

more than 2.4 billion inhabitants, use wood as energy source in their daily lives

(FAO, 2014; 2017A). Indigent rural populations worldwide base their livelihood

on wood and non-wood forest products. Forest products constitute 40 % of the

global renewable energy, the same proportion as wind, solar and hydroelectric

power combined (FAO, 2017B). Furthermore, in the era of accelerating climate

change, the crucial function of forest as a carbon sink cannot be emphasized

enough. Trees act as a carbon reservoir, by absorbing about 2 billion tonnes of

carbon dioxide yearly (FAO, 2018).

While forests have beneficial role in binding greenhouse gases, deforestation

scores as the second prominent cause of climate change. To mitigate the harmful

effect of deforestation, sustainable forest management as well as biotechnological

developments for improved tree performance offer solutions. The combination of

natural variation, molecular and genetic tools can generate trees with beneficial

properties for the forest industry. Such properties may be enhanced growth,

climate resilience, wood characteristics which allow more efficient processing

and thus lessen the environmental repercussions. Hence, the speedy development

of basic and applied biology tools which address the abovementioned features is

crucial to facilitate forest improvement. The Paris Climate Agreement (2015)

established the shared responsibility in utilizing forests for the good of the

climate, and biotechnological innovations can help in addressing this.

2

2. Xylem

2.1. Vascular cambium

Vascular cambium and secondary growth

Secondary (lateral) growth is a developmental process characterized by the

activity of the post-embryonic meristems, vascular cambium and phellogen,

which generate cylindrical tissue layers in the plant body. These two meristems

produce tissues in a bidirectional manner, a feature which is strictly orchestrated

by molecular and hormonal factors (Ragni and Greb, 2018).

The vascular cambium is responsible for the secondary growth of stems and roots,

and the formation of the woody tissue in dicots and gymnosperms. Two types of

vascular stem cells exist in the vascular cambium: the long and narrow fusiform

initials, and the short and small ray initials. Through periclinal cell divisions

(parallel to the surface) fusiform initials give rise to the specialized cell pool for

xylem and phloem, the vital tissues for mechanical support and long-distance

transport. The ray initials are the precursors of ray parenchyma cells, which form

the storage and transport network between phloem and xylem (Fosket, 1994;

Fischer et al., 2019).

The phellogen, also called cork cambium, generates phelloderm inwards and cork

outwards. These tissues have protective role against environmental stresses, like

water loss, biotic and abiotic attacks (Campilho et al., 2020).

Vascular cambium specification and maintenance

Whether xylem and phloem precursors within the vascular cambium are derived

from one or more radial stem cell layers, has been long debated. The 150-year-

old uniseriate cambium organization theory (Sanio, 1873) was supported recently

by lineage-tracing experiments, revealing that the vascular cambium consists of

a single layer of bifacial stem cells (Bossinger and Spokevicius, 2018; Smetana

et al., 2019; Shi et al., 2019). Furthermore, the work of Smetana et al. (2019)

demonstrated that the cambial cell division and differentiation programs are

orchestrated by the meristem organizer cells that are located adjacent to the

cambial stem cells. Local auxin response maxima promote the xylem identity and

quiescence of the organizer cells by the upregulation of CLASS III

HOMEODOMAIN-LEUCINE ZIPPER (HD-ZIP III) transcription factors.

Positive feedback regulation between auxin, AUXIN RESPONSE FACTORS

(ARF) and HD-ZIP III transcription factors assures vascular cambium

maintenance. The stem cell organizer of the vascular cambium is dynamic; the

3

organizer differentiates into a xylem vessel and an adjacent cambial cell becomes

the new organizer. Members of the WUSCHEL RELATED HOMEOBOX

(WOX) transcription factor family mark all three stem cell organizers;

WUSCHEL in the shoot apical meristem, WOX5 in the root meristem and WOX4

in the vascular cambium (Mayer et al., 1998; Sarkar et al., 2007; Smetana et al,

2019), and they are all downstream of CLE peptide signaling (Schoof et al., 2000;

Stahl et al., 2009; Hirakawa et al., 2010).

Smetana et al. (2019) and Shi et al. (2019) dissected and defined the cambial

regions based on the activity of molecular markers. Bifacial stem cells are

characterized by the overlapping activity of WOX4, PHLOEM

INTERCALATED WITH XYLEM/TRACHEARY ELEMENT

DIFFERENTIATION INHIBITORY FACTOR RECEPTOR (PXY/ TDR),

AINTEGUMENTA (ANT) and SUPPRESSOR OF MAX2 1-LIKE5 (SMXL5).

The xylem cell precursors are marked by WOX4 and PXY, the stem cells by high

levels of ANT, while those destined for phloem development express SMXL5.

Furthermore, it was recently uncovered that a concentration gradient formed by

six mobile PEAR PHLOEM EARLY DOF1 (PEAR) transcription factors

demarks the procambial zone and promotes cell division. A negative feedback

loop exists between the cytokinin-inducible PEAR proteins and the xylem cell

fate-promoting HD-ZIP III transcription factors which are induced by auxin. This

regulatory network establishes the foundation of radial growth, restricting the

stem cell niche by transcriptional regulation and the control of protein movement

(Fig. 1) (Miyashima et al., 2019).

A fundamental mechanism regulating cambium activity and maintenance is the

receptor-ligand complex formed between PXY/TDR and the

CLAVATA3/EMBRYO SURROUNDING REGION-related 41

(CLE41)/CLE44 peptides (Hirakawa et al., 2008; Etchells and Turner, 2010).

CLE41 and CLE44 encode the 12-amino-acid-long peptide ligand called

TRACHEARY ELEMENT DIFFERENTIATION INHIBITORY FACTOR

(TDIF). The CLE peptides are expressed mainly in the phloem and they travel to

the dividing cambial cells and bind to PXY/TDR, a plasma membrane-bound

leucine-rich repeat receptor-like kinase (Fig. 1). SOMATIC EMBRYOGENESIS

RECEPTOR KINASEs (SERKs) are coreceptors in CLE41/TDIF-PXY/TDR

signaling (Zhang et al., 2016). Signaling pathways downstream of CLE41 and

PXY/TDR promote cambial cell division and cell fate maintenance, inhibit

differentiation of xylem cells and control vascular patterning (Hirakawa et al.,

4

2008; 2010; Whitford et al., 2008; Etchells and Turner, 2010; reviewed by

Fischer et al., 2019).

The CLE41-PXY/TDR interaction is crucial to promote procambial cell

proliferation, by enhancing WOX4/WOX14 expression (Hirakawa et al., 2010;

Etchells et al., 2013). Simultaneously, CLE41-PXY/TDR maintains homeostasis

between xylem and phloem production by the activation of the GLYCOGEN

SYNTHASE KINASE3 (GSK3) pathway and the brassinosteroid-dependent

signaling cascade (Kondo et al., 2014; Han et al., 2018). GSK3s negatively

regulate xylem differentiation by suppression of BRI1-EMS SUPPRESSOR1

(BES1) which itself promotes xylem cell fate (Kondo et al., 2014). One member

of the GSK3s, BRASSINOSTEROID-INSENSITIVE2-LIKE1 (BIL1), inhibits

cambial activity downstream of CLE41-PXY via ARF5/MONOPTEROS (MP)

(Han et al., 2018). ARF5/MP promotes the transition of stem cells to xylem cells

by suppressing cytokinin response through the negative regulators of cytokinin

signaling ARABIDOPSIS RESPONSE REGULATOR7 (ARR7)/ARR15, and

BIL1 enhances this by the phosphorylation of ARF5/MP. PXY blocks the BIL1-

ARF5-cytokinin pathway by inhibiting BIL1 and its negative regulation of

cambial activity (Han et al., 2018). Notably, BIL1 is key in cambium

maintenance since it connects peptide signaling with auxin and cytokinin

signaling. Despite the complex regulations encompassing vascular cambium

maintenance, Lebovka et al. (2020) demonstrated by computational modelling

that the CLE41-PXY complex is sufficient to define tissue organization in the

cambium.

Additionally, REDUCED IN LATERAL GROWTH1 (RUL1) and MORE

LATERAL GROWTH1 (MOL1) are two receptor-like kinases regulating

cambial cell proliferation independent of PXY, RUL1 promotes and MOL1

suppresses the process (Agustí et al., 2011; Gursanscky et al., 2016).

As mentioned above, cambium organization is regulated by hormonal cross talks.

Mutual inhibition takes place between auxin and cytokinin signaling, which is

essential to define the developing xylem and phloem regions (Bishopp et al.,

2011A; Mellor et al., 2017). High cytokinin levels promote auxin flow towards

the xylem via the PIN-FORMED (PIN) auxin efflux carriers (Bishopp et al.,

2011B). In contrast, in xylem auxin promotes the expression of the cytokinin

signaling inhibitor, ARABIDOPSIS HISTIDINE PHOSPHOTRANSFER

PROTEIN6 (AHP6) (Bishopp et al., 2011A). In general, high levels of auxin

promote xylem development, while elevated cytokinin signaling stimulates cell

proliferation (Mellor et al., 2017; Ruonala et al., 2017).

5

The auxin-dependent ARF3, ARF4 and ARF5/MP transcription factors have

distinct roles in cambium regulation. ARF3 and ARF4 stimulate cambium

activity from outside the stem cell region, while ARF5/MP promotes the

differentiation of stem cells to xylem cells by suppressing WOX4 expression and

activating xylem-related genes (Brackmann et al., 2018).

Figure 1. Regulatory pathways of the vascular cambium activity.

The CLE41/44 peptides are expressed in the phloem and they move to the cambium,

where they bind to the PXY receptor. CLE41-PXY promotes cambial proliferation

through WOX4/WOX14. In addition, CLE41-PXY maintains xylem-phloem

homeostasis by activating GSK3, which suppresses xylem differentiation through

BES1. Downstream of CLE41-PXY, BIL inhibits cambial activity via ARF5/MP.

ARF5/MP suppresses cytokinin signaling through ARR7/ARR15. The auxin-

dependent ARF3 and ARF4 promote cambium activity from outside the stem cell

region. RUL1 promotes cambial cell proliferation and MOL1 suppresses the process.

The negative feedback loop between the cytokinin-inducible PEAR proteins and the

auxin-inducible HD-ZIP III transcription factors restricts the stem cell niche by

transcriptional regulation and protein movement control. High cytokinin levels

stimulate auxin flow towards the xylem via the PINs. ARF5/MP connects cytokinin

and auxin signaling by targeting TMO5. The TMO5-LHW heterodimer induces

AHP6 expression, which inhibits cytokinin signaling thus promoting xylem

formation. The TMO5-LHW complex can promote cytokinin biosynthesis through

the upregulation of LOG3 and LOG4, and auxin biosynthesis via YUC4, resulting in

a positive feedback loop.

6

MP is an important connection between auxin and cytokinin, MP targets

TARGET OF MONOPTEROS3 (TMO3)/TMO5/TMO7 which are upstream of

the cytokinin signaling cascade (Schlereth et al., 2010). TMO5/TMO5-LIKE1

(T5L1) form heterodimers with LONESOME HIGHWAY (LHW) (De Rybel et

al., 2013) and induce the expression of AHP6 (Ohashi-Ito et al., 2014), which in

turn promotes xylem formation by inhibiting cytokinin signaling (Mähönen et al.,

2006). Strikingly, the TMO5-LHW transcription factor complex can also

promote cytokinin production by the upregulation of the biosynthetic genes,

LONELY GUY (LOG3) and LOG4, which leads to enhanced cell proliferation

(Ohashi-Ito et al., 2014; Vera-Sirera et al., 2015). A key transcriptional hub for

cell proliferation downstream of TMO5-LHW is the transcription factor DOF2.1

which together with its close homologs control procambial cell divisions by

regulating a subset of cytokinin-dependent genes (Smet et al., 2019).

Furthermore, the TMO5-LHW dimer is able to upregulate auxin biosynthesis

through YUCCA4 (YUC4), leading to a positive feedback loop since elevated

auxin levels maintain the expression of TMO5, LHW and their downstream

targets (Ohashi-Ito et al., 2019). On the other hand, a negative feedback loop,

involving the polyamine signaling molecule thermospermine, confines the

TMO5-LHW levels (Katayama et al., 2015; Vera-Sirera et al., 2015).

Such intricate control mechanisms are essential to initiate vascular development

and modulate that based on the needs of the indefinitely growing plant body and

the changing environment, without compromising the integrity of the organism.

2.2. Arabidopsis as a model to study the basic molecular

mechanisms of secondary growth

Arabidopsis thaliana (from here on “Arabidopsis”) is a small, herbaceous annual

plant, a native species of Western Eurasia, which has colonized habitats across

the globe with vast climatic and environmental fluctuations. Arabidopsis not only

colonized our geographical world, but also the world of plant science, becoming

the central model plant and reference organism. The first study of this species

was published more than a century ago in 1907 by the German botanist, Friedrich

Laibach (Laibach, 1907; Krämer, 2015). Since then, research fields from cell

biology to ecology have adopted Arabidopsis as a model plant thanks to the small

size of the plant, its small genome and all the available toolkits. Aware of the

limitations of Arabidopsis, we shall acknowledge that this model plant has

enabled the molecular understanding of key biological concepts. One such

concept is secondary growth. Despite its herbaceous nature, the root, hypocotyl

7

and stem of Arabidopsis undergo secondary growth, making it a simple and handy

model. The root and the stem display a secondary growth gradient, with only

primary growth present close to the apical meristems and extensive secondary

growth in the proximity of the rosette. In contrast, the secondary growth of the

hypocotyl is longitudinally quite uniform, forming a cylinder of wood similar to

angiosperm trees (reviewed by Ragni and Greb, 2018), which makes it an

attractive model for secondary xylem development (Chaffey et al., 2002). During

secondary growth, a continuous cambial ring develops in the root and hypocotyl,

producing xylem on the inner and phloem on the outer side. Initially the

secondary xylem consists of vessel elements and parenchyma cells (phase I). A

developmental shift towards xylem expansion takes place when floral transition

occurs, and the cell type repertoire broadens with the emergence of fiber cells

(phase II) (Chaffey et al., 2002). In the inflorescence stem, uniform vascular

cambium is formed at the very base of the stem where the fascicular cambia of

the vascular bundles are linked together by the interfascicular cambia (Altamura

et al., 2001). A wide array of molecular genetics findings from Arabidopsis have

been corroborated in tree species (reviewed by Barra-Jiménez and Ragni, 2017).

Some of the features which make Arabidopsis a popular model system, can

occasionally become disadvantageous. For example, the small size of the plant

can make dissection of specific xylem cell types difficult. Due to its annual

nature, long-term research of the same individual is limited. Arabidopsis has a

poor biomass yield compared to trees, limiting biomass studies and the

characterization of certain traits important for the forest industry, such as the

mechanical properties of wood and sugar release upon enzymatic pretreatment.

2.3. The function of the xylem

Early in the evolution of vascular plants the development of xylem and phloem

enabled long-distance transport, leading to the gradual colonization of terrestrial

habitats. However, it should be noted that non-vascular plants also exist,

demonstrating that the presence of the vascular system is more of an evolutionary

advantage than a prerequisite for survival (Agustí and Blázquez, 2020).

Xylem has two key functions; transport and mechanical support. The transport of

xylem sap (water, mineral nutrients and hormones) from the root towards the

aboveground organs occurs through dead xylem cells which form hollow pipes.

Vascular plants invest resources into xylem development, which in turn allows

water transport at low energetic cost under negative pressure. Water transport is

driven by the transpiration rate of the leaves. In the negative pressure environment

8

of the xylem conduits water is metastable, thus susceptible for cavitation. Air-

blockage or embolism of the water-conducting cells may lead to the impediment

of water flow, thus compromising plant health. Environmental factors, such as

drought and frost, affect resistance to embolism (Lens et al., 2013; Brodersen et

al., 2019). To perform water transport efficiently and maintain integrity, plants

have developed strategies to prevent and withstand the formation of air bubbles

under negative pressure.

The biomechanical demands of up-right growth and endurance of the water flow-

derived tension are assured by specific morphological features of the xylem and

mechanical support of the secondary cell walls. The cellulose, hemicellulose and

lignin content of the secondary cell walls and the interaction of these polymers

determine the chemical and physical properties of the xylem. These properties

also affect the degree at which a species is able to sequester carbon from the

atmosphere (Myburg et al., 2013).

2.4. Different cell types building up the xylem

Environmental factors, hormone levels and cross talk among the different plant

hormones promote the wood anatomical diversity seen between and within

species (Carlquist, 2013). The morphology, distribution and characteristics of the

specialized cell types present in the xylem determine the properties of wood

(Myburg et al., 2013). The three main cell types are tracheids or vessel elements

(depending on the plant species), fibers and parenchyma cells.

Tracheids and vessel elements

Tracheids and vessel elements, collectively called as tracheary elements, are the

water-conducting cells in the xylem. Tracheary elements are able to withstand the

negative pressure associated with sap flow thanks to their patterned and heavily

lignified secondary cell walls (Turner et al., 2007). Tracheids are narrow (8-80

µm), elongated (0.5-10 mm) cells, essential for xylem sap transport and structural

support, predominantly present in the xylem of gymnosperms (Panshin and de

Zeeuw, 1970; Pittermann, 2010; Wiedenhoeft, 2013). Vessel elements differ in

morphology, being shorter (0.1-1.2 mm) and wider in diameter (typically 50-200

µm). Vessel elements are specialized on sap transport and they are the

characteristic of angiosperms (Tyree and Zimmermann, 2002; Wiedenhoeft,

2013). The final steps of differentiation for these cell types are programmed cell

death, clearance of cell content through autolysis and the formation of bordered

pits and perforation plates (Fukuda, 1996). Bordered pits are lined by

9

semipermeable membrane-like structure made of the primary wall and middle

lamella at the pits of neighbouring cells, permitting fluid transport but blocking

the passage of large air bubbles and pathogens. The tracheids in gymnosperms

are highly connected through bordered pits; every tracheid is connected to several

others via pits along the radial walls (Choat et al., 2008). As tracheids overlap

with their lower and upper neighbouring cells 20-30% of their length, water needs

to take a zigzag route to pass through the pits. Tracheids are relatively inefficient

conduits due to their small diameter and the presence of the pit membrane which

imposes resistance to the water flow (Wiedenhoeft, 2013). Furthermore, the level

of connectivity within the xylem affects sap flow efficiency and susceptibility to

embolism (Brodersen et al., 2019). During vessel element development,

secondary cell wall is not deposited at the apical and basal ends of the cells.

Instead, membrane-free perforation plates are formed by partial digestion of the

primary walls (Turner et al., 2007). When differentiation is completed, multiple

vertically aligned vessel elements form hollow pipes called vessels, which

facilitate the free movement of water. Vessels are much longer than tracheids,

and may reach several meters in length (Brodersen et al., 2019). The individual

vessel elements are connected with each other through intervessel pits located on

their lateral walls, while half-bordered pits form between vessel elements and ray

cells (Wiedenhoeft, 2013). These properties make vessels far more effective in

water transport than tracheids (Lewis and Boose, 1995). Tracheids also occur

outside the gymnosperms, for example in the Fagaceae family, but due to their

smaller diameter the water transport performed by individual tracheids is

negligible compared to vessels. Primitive angiosperms, such as those in the genus

Amborella, lack vessels (Hacke et al., 2007; Sperry et al., 2007). Overall, the

general evolutionary trend within vascular plants was to increase xylem conduit

diameter and length, thus improving water conductance and capability to adapt

to a wide variety of environments (Lewis and Boose, 1995).

Fibers

Fibers provide mechanical support in the xylem. Generally, fibers are shorter than

tracheids (0.2-1.2 mm) but longer than the vessel elements of the same species

(Wiedenhoeft, 2013). Their thick lignified secondary cell walls are the main

determinants of the mechanical strength and density of wood. The model plant,

Arabidopsis, has long-living fibers (Bollhöner et al., 2012), in contrast to most

angiosperms where fibers, similarly to tracheary elements, undergo cell death.

However, this developmental program differs greatly between the two cell types;

DNA degradation in the fibers’ nuclei begins much prior to cell death, while in

10

tracheary elements DNA breakage seems to occur only shortly prior to the

vacuolar collapse. Upon vacuolar burst the cell content of xylem vessels is rapidly

degraded, in contrast to the gradual hydrolysis observed in fibers (Courtois-

Moreau et al., 2009; reviewed by Bollhöner et al., 2012).

Parenchyma cells and ray cells

Xylem parenchyma cells are the facilitators of a functional xylem by their

involvement in a multitude of processes. Tridimensional parenchyma cell

networks spam throughout the secondary xylem and phloem, ensuring high level

of interconnectivity. Xylem parenchyma cells often deposit secondary cell walls,

which have simpler structure than the cell walls of tracheary elements, resembling

thickened primary walls (Panshin and de Zeeuw, 1980). The bulk of the living

cells in the wood are the ray and axial parenchyma cells. Axial parenchyma cells

are derived from fusiform cambial initials, and as their name suggests, they are

axially elongated (Carlquist, 1988), while ray cells develop from ray initials. Ray

parenchyma cells are radially connected and they extend across the phloem,

cambium and xylem, enabling xylem-phloem exchange. A subset of parenchyma

cells is in direct contact with water-conducting cells through cell wall pits or

indirectly via the apoplast (reviewed by Spicer, 2014). The cells within the rays

are connected through plasmodesmata, enabling the continuous flow of resources

(Sauter and Kloth, 1986; Chaffey and Barlow, 2001).

Parenchyma cells have a wide range of functions from storage and transport of

non-structural carbohydrates, lipids and storage proteins (Höll, 2000; Plavcová

and Jansen, 2015) to water storage (Holbrook, 1995; Borchert and Pockman,

2005) and defense against biotic factors (Schmitt and Liese, 1993). Thanks to the

interconnection between tracheary elements and parenchyma cells, the secondary

cell walls of dead, functional TEs are structurally reinforced by the lignin

monomers coming from the parenchyma cells (Smith et al., 2013; Pesquet et al.,

2013; Zhang et al., 2020A). Furthermore, parenchyma cells are implicated in long

distance water transport, whereby they alter xylem hydraulic conductance and

sap flow rate through ion exchange into the transpiration stream (Jansen et al.,

2011; Nardini et al., 2011; reviewed by Ménard and Pesquet, 2015). Parenchyma

is also involved in restoring water transport in embolized tracheary elements by

the active export of sucrose, ions and H+ into the embolized cell, thus reducing

the water potential and triggering sap influx (Salleo et al., 2004; 2009; Brodersen

et al., 2010; Secchi and Zwieniecki, 2011; 2012). When a water-conducting cell

gets fatally damaged, the neighbouring xylem parenchyma cells will fill the

lumen of the dysfunctional cell with tyloses, pectin-rich gels and gums, thus

11

sealing the cell off from the conducting network (Kitin et al., 2010; Sun et al.,

2008).

The fraction, anatomy and spatial distribution of parenchyma cells are highly

variable in different species, and this variability has been exploited as a diagnostic

feature for wood identification (IAWA Committee, 1989; 2004). The quantitative

and qualitative features related to parenchyma cells are influenced by genetics

and environmental factors, such as temperature, precipitation and seasonality. In

general, coniferous gymnosperms contain very few parenchyma cells, while stem

succulents and lianas exhibit the highest proportions of xylem parenchyma within

the plant kingdom (Morris et al., 2016). In the xylem of the model plant

Arabidopsis, parenchyma cells are interspersed between fibers and vessel

elements (Smith et al., 2013). Under normal growth, Arabidopsis does not form

rays, but artificial weight treatment may induce this developmental process,

highlighting the plasticity of xylogenesis (Mazur and Kurczynska, 2012).

3. Secondary cell wall

Generally, secondary cell wall deposition takes place after plant cells have

reached their final size by expansion. By definition, a “secondary cell wall” has

a specific chemical composition, characterized by the presence of cellulose,

hemicellulose and lignin, with less contribution from cell wall proteins and

enzymes (Sjöström, 1993). The primary cell wall is mainly composed of pectin,

hemicellulose, cellulose and glycoproteins (Fry, 2004). The secondary cell wall

is deposited inside the primary cell wall, between the primary cell wall and the

plasma membrane. In order to begin secondary cell wall formation, the primary

cell wall biosynthetic machinery is replaced by new enzymes to meet the

biosynthetic and regulatory demands imposed by the secondary cell walls. First,

the cell wall polysaccharides, hemicellulose and cellulose, are deposited,

followed by lignification (Meents et al., 2018).

3.1. Cellulose

With a global annual yield of approx. 100 billion tonnes, cellulose is the most

abundant natural polymer on earth (de Souza Lima and Borsali, 2004). Cellulose

contributes to the load-bearing scaffold of the cell, accounting for 40-50% of the

SCW in land plants (Timell, 1967). Cellulose is composed of a linear chain of β-

1,4 linked glucose units, organized into microfibrils through intra- and

intermolecular bonds and Van der Waals forces (Kim et al., 2013). In contrast to

12

cellulose present in the primary cell walls, the cellulose of the secondary cell

walls displays higher degree of polymerization and crystallinity.

Cellulose microfibrils are synthesized at the plasma membrane by the large,

mobile, rosette-shaped cellulose synthase protein complex (CSC). The CSC is a

complex structure, built of multiple individual CesA subunits. The enzymes of

CSC specialized on secondary cell walls are CesA4/IRX5, CesA7/IRX3 and

CesA8/IRX1 (Turner and Sommerville, 1997; Taylor et al., 2003). Several recent

studies provided evidence that the CSC is likely to be a hexamer composed of

trimers or tetramers of 18-24 CesAs instead of the previously proposed structure

of the CSC composed of a minimum of 36 CesAs (reviewed by Meents et al.,

2018). The precise stoichiometry of the complex differs between species. In

Arabidopsis, both primary and secondary cell wall CesAs are present in

equimolar (1:1:1) amounts, likely composing a CSC with 18 subunits (Gonneau

et al., 2014; Hill et al., 2014). Using cytosolic UDP-glucose as a substrate, the 18

subunits simultaneously synthesize 18 glucan chains bound into a cellulose

microfibril with a diameter of approx. ~2.5 nm (reviewed by Polko and Kieber,

2019). Zhang et al. (2018) demonstrated that the gymnosperm Norway spruce

(Picea abies) has the same equimolar CesA stoichiometry as in Arabidopsis,

while the angiosperm aspen (Populus tremula) had a CesA ratio of 3:2:1 for

PtCesA8a/b:PtCesA4:PtCesA7a/b.

While ejecting glucan chains which polymerize and crystallize to cellulose

microfibrils, the CSCs move along cortical microtubule rails in the plasma

membrane (Paredez et al., 2006). The linker protein CSI1/POM2 physically

connects CSCs to the microtubules (Li et al., 2012; Bringmann et al., 2012;

Derbyshire et al., 2015; Schneider et al., 2017). Recently, the microtubule-

independent movement of CSCs was demonstrated, where the CSCs progress on

trails left by previous complexes, likely by interacting with recently formed

cellulose microfibrils. The microtubule-guided movement of CSCs is more

common, and can override the microtubule-independent CSC movement (Chan

and Coen, 2020). The CSCs in the SCW are present in higher density than in

PCW (Watanabe et al., 2015). They form clusters which move on coordinated

trajectories, depositing aggregates of cellulose microfibrils characteristic of the

SCW (reviewed by Meents et al., 2018).

Beyond the core cellulose synthase machinery, several non-CesA proteins have

been linked to SCW cellulose synthesis although their mode of action remains to

be scrutinized. KORRIGAN, a membrane-bound endoglucanase, contributes to

proper cellulose biosynthesis, likely by regulating cellulose crystallinity

13

(Szyjanowicz et al., 2004; Maloney and Mansfield, 2010). The transmembrane

proteins Tracheary Element Differentiation-Related6 (TED6) and TED7 bind to

the SCW CSCs, suggesting that they may be accessory components of the

complex (Endo et al., 2009; Rejab et al., 2015). COBRA-LIKE4, a

phosphatidylinositol-anchored glycoprotein, binds to cellulose microfibrils and

affects cellulose crystallinity (Persson et al., 2005; Liu et al., 2013A).

3.2. Hemicellulose

10-40 % of the SCW is composed of hemicelluloses. Hemicelluloses contribute

to the load-bearing capacity of the SCW by their interaction with cellulose and

lignin. A variety of hemicelluloses occur in terrestrial plants, including

xyloglucans, xylans, mannans, glucomannans and β-(1-3, 1-4)-glucans. Xylans

are the predominant noncellulosic polysaccharide of dicot secondary cell walls.

Mannans, in the form of galactoglucomannans, are the major hemicellulose in the

SCW of gymnosperms (reviewed by Scheller and Ulvskov, 2010).

In contrast to cellulose, hemicelluloses are synthesized in the medial section of

the Golgi apparatus and subsequently transported to the plasma membrane. For

efficient hemicellulose biosynthesis the enzymes involved in the process are

likely to form complexes. In planta assays done in wheat (Triticum aestivum) and

asparagus (Asparagus officialis) revealed that the xylan biosynthetic enzymes

form homo- and heterodimers in the Golgi (Zeng et al., 2010; 2016). There are

three key enzymes which generate the xylan backbone; the Glycosyltransferase

Family 43 (GT43) members IRX9 and IRX14, and a GT47 member IRX10. They

all have functionally redundant paralogs; IRX9-L, IRX14-L and IRX10-L (Lee

et al., 2007; Brown et al., 2007; Wu et al., 2010). Since IRX10 does not have a

transmembrane domain, the interaction and complex formation with IRX9 and

IRX14 may be necessary for its proper Golgi localization and function (reviewed

by Meents et al., 2018). Functional characterization studies revealed that

IRX9/IRX9-L may have structural rather than enzymatic function,

IRX14/IRX14-L plays a role in substrate binding, while IRX10/IRX10L

presumably has xylosyl transferase activity (Ren et al., 2014; Urbanowicz et al.,

2014; Zeng et al., 2016)

The glucuronic acid (GlcA) substitutions on the xylan backbone are introduced

by two members of the GT8 family, Glucuronic Acid Substitution of Xylan

(GUX) 1 and GUX2 (Rennie et al., 2012). GUX1 adds the GlcA side chains

preferentially to evenly spaced xylose residues, always on the same side of the

backbone. GUX2 generates more tightly clustered decorations, substituting

14

xylose residues on either side of the backbone. The activity of GUX1 and GUX2

yields two different xylan domains which are likely to be part of the same

heterogenous xylan molecule. These distinct xylan domains may contribute to the

crosslinking capacity of the molecule with other cell wall polymers (Bromley et

al., 2013). Next, methyl groups may be added to the GlcA residues by members

of the glucuronoxylan methyltransferase (GXMT) family, which are DUF579

domain-containing proteins. GXMT1 methylates 75% of the GlcA residues in

Arabidopsis (Urbanowicz et al., 2012). IRX15 and IRX15L are putative

methyltransferases with no detected catalytic activity, likely playing a structural

role in the xylan synthase complex, similarly to IRX9/IRX9-L (Brown et al.,

2011; Jensen et al., 2011).

Initially, the DUF231 domain-containing acetyltransferase, ESK1/TBL29, was

shown to perform the monoacetylation of xylan at O-2 and O-3 position (Yuan et

al., 2013; Xiong et al., 2013). ESK1/TBL29 is required for evenly patterned xylan

acetylation, which then guides the activity of GUX1 for the evenly distributed

GlcA depositions (Grantham et al., 2017). Recently, several studies dissected the

role of other TBL family members in xylan acetylation. TBL3 and TBL31

specifically affect O-3 acetylation (Yuan et al., 2016A), while TBL32, TBL33,

TBL34 and TBL35 mediate the acetyl substitutions of xylosyl residues at O-2

and O-3 (Yuan et al., 2016B; 2016C). These TBL enzymes may also be essential

for the establishment of uniform xylan decorations. The four REDUCED WALL

ACETYLATION (RWA) genes in Arabidopsis also contribute to xylan

acetylation and proper secondary cell wall formation, but they have lower

substrate specificity than the TBL proteins since they can acetylate mannans and

xyloglucans as well. RWA and TBL proteins may form complexes where the

specificity of TBL drives the substrate preference. Alternatively, RWA might

perform xylan acetylation first, thus providing the substrate to TBL (Manabe et

al., 2013).

The even pattern of xylan substitution is a characteristic of vascular plants and

essential for the binding of xylan to the hydrophilic faces of cellulose microfibrils

to ensure normal secondary cell wall development (Grantham et al., 2017).

3.3. Cell wall proteins

Cell wall proteins account for a minor fraction of the SCW and their exact

function is poorly understood (Liu et al., 2013A). Cell wall proteins typically

belong to the following groups: glycine-rich (GRPs), proline-rich (PRPs),

arabinogalactan-rich (AGPs) and hydroxyproline-rich glycoproteins (HRGPs or

15

extensins) (Showalter, 1993). Fasciclin-like AGP (FLA) proteins are enriched in

the SCWs, and supposedly involved in cellulose deposition on the basis of the

presence of a GPI-anchor on several FLA and reverse genetic studies (reviewed

by Kumar et al., 2016).

3.4. Lignin

Lignin is the product of the phenylpropanoid pathway, an aromatic polymer,

which constitutes 15-36% of the woody cell wall biomass (Zobel and van

Buijtenen, 1989). Lignification, the final step of SCW biosynthesis, occurs prior

as well as after cells undergo programmed cell death (Smith et al., 2013; Pesquet

et al., 2013; Zhang et al., 2020A).

The evolution of lignin

Early in the evolution of the monolignol biosynthetic pathway its role might have

been to produce UV-protectant molecules (reviewed by Weng & Chapple, 2010).

Eight key enzymes required for the biosynthesis of the monolignols p-coumaryl

alcohol and coniferyl alcohol were most likely recruited from the primary

metabolism. The pathways within the primary metabolism are vital for plant

survival, and their products serve as precursors for the diverse reactions of the

secondary metabolism (reviewed by Pott et al., 2019). Homologs of the lignin-

biosynthetic enzymes can be identified within the primary metabolism. Besides

lignin, other hydrophobic polymers such as cutin, suberin, sporopollenin and

soluble metabolites such as flavonoids, tannins, stilbenes and lignans are the end

products of the phenylpropanoid pathway (reviewed by Weng & Chapple, 2010).

Major monolignols

To date 35 different naturally occurring lignin monomers have been identified,

most of which are very species specific. The three most widely occurring

monomers, the p-hydroxyphenyl (H), guaiacyl (G) and syringyl (S) units, are

derived through the polymerization of p-coumaryl, coniferyl and sinapyl alcohol,

respectively. Branching and interconnections curtail the flexibility of the

polymer. The formation of many chain-starting monomers results in lignin with

low molecular weight (reviewed by Vanholme et al., 2019). The more rare

compounds have potentially been under selective pressure, therefore the three

major lignin units contribute to the bulk of the lignin polymer.

16

Distribution of lignin in different species and cell types

Lignin content, monomer composition and interunit linkage distribution vary

depending on the species, cell type, cell wall layer as well as upon environmental

stimuli (Campbell and Sederoff, 1996; Studer et al., 2011). Lignin has a very

wide pattern for deposition, being present in tracheary elements, sclerenchyma

cells, endodermal cells, seed coat and silique cells (Barros et al., 2015).

Gymnosperm species contain mostly G units with minor amounts of H units,

while angiosperm lignin is primarily formed of G and S units. The H unit content

is higher in monocots than in dicots (Boerjan et al., 2003). S-type lignin is specific

to flowering plants and the lycophyte Selaginella, while H- and G-type lignin are

elemental for all tracheophytes (reviewed by Weng and Chapple, 2010).

The water-conducting xylem vessels contain mostly G units and lignify first,

while the structural fibers containing both G and S units lignify later. H units are

incorporated into the middle lamella of vessel elements and fibers early on during

SCW formation (Fukushima and Terashima, 1990). H units are typically

deposited at the beginning of the lignification process in the cell corners and

middle lamella, G units are first deposited in the middle lamella and later in the

SCW layers, and finally S units are incorporated into the SCW (Terashima and

Fukushima, 1988; Terashima et al., 1988; Fukushima and Terashima, 1991;

reviewed by Donaldson, 2001). The relative abundance of each lignin unit affects

the overall structure and thus the physical properties of the polymer (reviewed by

Bonawitz and Chapple, 2010).

The function of lignin

The chemical structure of lignin confers rigidity and hydrophobicity to the cell

wall, which facilitates the transport of water and solutes through the vascular

system (Boerjan et al., 2003; Vanholme et al., 2010; Barros et al., 2015).

Moreover, lignin protects the plants against various environmental stresses, such

as pathogen attacks (reviewed by Miedes et al., 2014). Different developmental

and environmental factors, such as wounding, metabolic stress or perturbation in

the cell wall integrity, can trigger lignin biosynthesis (Cano-Delgado et al., 2003;

Tronchet et al., 2010).

3.4.1. Biosynthesis of the lignin monomers

3.4.1.a. The lignin biosynthetic pathway

Monolignols are synthesized most often from phenylalanine, in grasses also from

tyrosine (Barros et al., 2016), by sequential steps involving eleven enzymes of

17

the phenylpropanoid pathway. The H-, G- and S-lignin units are distinguished by

their degree of methoxylation.

Phenylalanine, the precursor of the phenylpropanoid pathway, is derived from the

shikimate pathway and converted in the first enzymatic step to cinnamic acid by

phenylalanine ammonia-lyase (PAL). Subsequently, the aromatic ring is p-

hydroxylated by the enzyme cinnamate 4-hydroxylase (C4H) and further

converted to p-coumaryl-CoA by the 4-coumarate:CoA ligase (4CL) enzyme.

After this, two possible directions can be taken; if H-lignin unit is being produced

then the p-coumaryl-CoA is converted to p-coumaryl alcohol by cinnamoyl-CoA

reductase (CCR) and cinnamyl alcohol dehydrogenase (CAD). If G or S-lignin

unit synthesis is taking place, p-coumaroyl-CoA is meta-hydroxylated by p-

hydroxycinnamoyl-CoA shikimate/quinate hydroxycinnamoyltransferase (HCT)

and p-coumarate 3-hydroxylase (C3H). The meta-hydroxyl residue is then

converted into methoxy residue by caffeoyl-CoA O-methyltransferase

(CCoAOMT), which is followed by the conversion of the CoA-conjugate of the

propene residue to an aldehyde by CCR. At this step CCR redirects the

phenylpropanoid pathway towards lignin monomer biosynthesis. The aldehyde

moiety of the coniferaldehyde is converted to an alcohol by CAD, which leads to

G unit production. The synthesis of S unit monomers requires the hydroxylation

of the coniferaldehyde by the enzyme ferulate 5-hydroxylase (F5H), which is

followed by a methoxylation by caffeic acid O-methyltransferase (COMT) (Fig.

2).

The hydroxylation enzymes, C4H, C3H and F5H are bound to the endoplasmic

reticulum (ER), while all other lignin biosynthetic enzymes are located in the

cytosol (Takabe et al., 2001; Ruelland et al., 2003). The interactions found

between the ER bound and cytosolic enzymes suggest that the site of lignin

monomer biosynthesis is in the cytosol, close to the ER (Achnine et al., 2004).

The enzymes 4CL, CCR, CAD, HCT, COMT and F5H have multiple substrates,

hence the pathway is composed of multiple parallel routes towards coniferyl and

sinapyl alcohol production (reviewed by Vanholme et al., 2019). Several

biosynthetic enzymes form protein complexes to increase activity and efficiency;

CAD1 and CCR2 form heterodimers (Yan et al., 2019), the enzymes performing

the 4- and 3-hydroxylation of cinnamic acid, C4H1, C4H2 and C3H, form

heterodimers and trimers (Chen et al., 2011), and 4CL3 and 4CL5 form

heterotetrameric complexes (Chen et al., 2014). Furthermore, the cytochrome

P450 enzymes, C4H, C3H and F5H, form clusters through interactions with

scaffold proteins (Gou et al., 2018).

18

Figure 2. Monolignol biosynthesis through the phenylpropanoid pathway.

Each arrow represents a reaction, the abbreviated names of the enzymes catalyzing

the reaction are shown in orange. The name of the intermediates of the biosynthetic

pathway are shown in bold black. The light orange background highlights the core

part of the pathway as it is known to this date. PAL, Phenylalanine ammonia-lyase;

C4H, Cinnamate 4-hydroxylase; 4CL, 4-coumarate:CoA ligase; HCT,

Hydroxycinnamoyltransferase; C3H, P-coumarate 3-hydroxylase; CSE, Caffeoyl

shikimate esterase; CCoAOMT, Caffeoyl-CoA O-methyltransferase; CCR,

Cinnamoyl-CoA reductase; CAD, Cinnamyl alcohol dehydrogenase; HCALDH,

hydroxycinnamaldehyde dehydrogenase; F5H, Ferulate 5-hydroxylase; COMT,

Caffeic acid O-methyltransferase.

3.4.1.b. Regulation of lignin biosynthesis

Secondary cell wall formation including lignin deposition is strictly controlled in

order to direct and restrict lignin deposition to the proper cell types and

developmental stages, and to avoid abnormal growth and development (Behr et

al., 2019). Multiple regulatory steps underlie the production of lignin polymer

biosynthetic enzymes, production of lignin polymer precursors, transport of

lignin polymer enzymes and precursors, and finally building of the highly

organized lignin structure (reviewed by Kumar et al., 2016).

19

Transcriptional regulation

More than 2000 genes in Arabidopsis are possibly related to cell wall biogenesis

(Carpita et al., 2001). The activity of the genes involved in cell wall synthesis and

modification is under transcriptional and post-transcriptional control. Taylor-

Teeples et al. (2015) characterized the complex gene regulatory network of

Arabidopsis transcription factors and secondary cell wall metabolic genes,

illustrating that xylem cell differentiation, including phenylpropanoid

biosynthesis, is controlled by a highly branched rather than a linear regulatory

pathway. Therefore, perturbation of a side pathway can lead to massive changes

in the whole network. This network analysis detected feed forward loops as a

very frequent form of regulation. E2Fc was identified as a key upstream regulator

of the master regulators as well as the structural genes. E2Fc, a negative regulator

of endoreduplication (Del Pozo et al., 2007), belongs to the E2F family of

transcription factors, and can act as a transcriptional activator as well as repressor

in a dose-dependent manner.

The majority of the transcription factors controlling secondary cell wall formation

belong to the NAC (NAM, ATAF1/2, CUC2) domain or to the MYB

(Myeloblastosis) transcription factor family, and they act in hierarchical order

(Fig. 3). The first level of transcription factors are the master regulators of

secondary cell wall formation. There are five master regulator NAC domain

transcription factors that belong to the VNS subfamily (VND, NST/SND, and

SMB-related protein): NST1 (NAC SECONDARY WALL THICKENING

PROMOTING FACTOR1), SND1 (SECONDARY WALL-ASSOCIATED

NAC DOMAIN1), NST2, VND6 (VASCULAR RELATED NAC DOMAIN6)

and VND7. The VNS proteins act as molecular master switches of xylem cell

differentiation by regulating genes encoding for cell wall biosynthetic enzymes

and genes related to programmed cell death (Ohtani and Demura, 2019). NST1 is

expressed in interfascicular fibers and differentiating vessels (Mitsuda et al.,

2005; 2007), and SND1 is expressed in interfascicular and xylary fibers (Zhong

et al., 2006). NST1 and SND1 regulate secondary cell wall formation in fibers

(Mitsuda et al., 2005; 2007; Zhong et al., 2006). In contrast, the expression and

regulatory activity of VND6 and VND7 are xylem vessel specific (Kubo et al.,

2005; Yamaguchi et al., 2008). Several molecular factors have been reported to

act as negative regulators of the master switches. VND7 interacts with VND-

INTERACTING2 (VNI2) which leads to the repression of vessel-specific genes

regulated by VND7, making VNI2 a transcriptional repressor of xylem cell fate

(Yamaguchi et al., 2010). XYLEM NAC DOMAIN1 (XND1) inhibits tracheary

element differentiation and premature SCW formation (Zhao et al., 2008; 2017).

20

XND1 interacts with NST1 thereby modulating SCW deposition (Zhang et al.,

2020B). WRKY12 directly binds to the NST2 promoter, thus suppressing the

SCW activator role of NST2 (Wang et al., 2010). Several of these so-called first

level transcription factors bind to a conserved cis-acting element, the secondary

wall NAC-binding element (SNBE), present in the promoter of all their targets

(Zhong et al., 2010).

Figure 3. Transcriptional regulation of secondary cell wall formation. Transcription

factors coordinate the transcriptional regulation, and E2Fc is the key upstream

regulator. The first and second layer master switches can regulate the transcription

factors below them or the secondary cell wall biosynthetic genes directly. The

downstream regulators directly bind to the secondary cell wall biosynthetic genes.

γMYB2 represses some of the second layer master switches and downstream

regulators. Subunits of the transcriptional co-regulator Mediator complex regulate

lignification through homeostatic repression.

MYB (MYB46, MYB83 and MYB55) and NAC (SND3 and XND1)

transcription factors are among the second layer of transcriptional regulators.

MYB46 and MYB83 were shown to bind to the secondary wall MYB-responsive

element (SMRE), and thus directly control an array of not only transcription

factors (MYB58, MYB63), but secondary cell wall biosynthetic genes as well

(Zhong and Ye, 2012). The third level comprises transcription factors which are

downstream targets of the master regulators and the members of the second layer

transcription factors. They control specifically the expression of cell wall

biosynthetic genes by binding to their promoters. MYB20, MYB69, MYB79,

21

MYB85, MYB58, MYB63, MYB103 and KNAT1 control the expression of the

lignin-biosynthetic genes (reviewed by Kumar et al., 2016; Öhman et al., 2013).

MYB58 and MYB63 transcriptionally activate all lignin-biosynthetic genes

(except F5H) as well as LAC4, through binding to the AC cis-regulatory elements

of the promoters. C4H and COMT were also directly activated, even though their

promoter may only contain deteriorated AC elements. Similar AC element

dependent activation of lignin-biosynthetic genes was shown for MYB85 (Zhou

et al., 2009). MYB103 affects S-type lignin content and S/G lignin ratio by

altering the expression of F5H through an unknown mechanism (Öhman et al.,

2013). The knotted-like homeobox (KNOX) protein, KNAT1, directly regulates

several lignin-biosynthetic genes by binding to their promoters (Mele et al.,

2003).

Furthermore, MYB4 and its two close homologs MYB7 and MYB32 are negative

regulators of the lignin-biosynthetic pathway (reviewed by Zhang et al., 2018;

Behr et al., 2019). Recently, MYB20, MYB42, MYB43 and MYB85 were shown

to act redundantly in the activation of phenylalanine and lignin biosynthesis. This

transcription factor quartet also suppressed flavonoid biosynthesis through

MYB4, thus ensuring sufficient phenylalanine supply for the lignin-biosynthetic

pathway (Geng et al., 2020). MYB52 is another negative regulator of

lignification. It is co-expressed with several SCW related TFs and biosynthetic

genes, proposed to have a role in maintaining the balance between the production

of different SCW components (Cassan-Wang et al., 2013). MYB75 promotes

anthocyanin biosynthesis and, together with KNAT7, suppresses SCW formation

(Bhargava et al., 2010; 2013). A member of the MYB coiled-coil family, γMYB2,

is also involved in SCW establishment by repressing SND3, MYB46, MYB63 and

MYB103 (Nguyen et al., 2019).

Besides the evolutionarily highly conserved NAC-MYB gene regulatory network

(Xu et al., 2014; Bowman et al., 2017), there are likely to be additional molecular

regulators and modulators involved in the control of lignification. These

molecular factors might be less conserved, but more species specific (Ohtani and

Demura, 2019).

In addition to the transcription factors, the transcription of lignin-biosynthetic

genes is also controlled by the Mediator complex (Fig. 3). The Mediator complex

is a transcriptional co-regulator in eukaryotes containing around 30 subunits,

accounting for the core head, middle and tail modules, and the dissociable kinase

module (Asturias et al., 1999; Tsai et al., 2014). MED5a/5b, two paralogs located

in the tail module of Mediator, are homeostatic repressors of the phenylpropanoid

22

metabolism (Bonawitz et al., 2012; 2014). MED5a/5b genetically interact with

the other subunits of the tail module MED2, MED16 and MED23 (Dolan et al.,

2017), as well as the subunits of the kinase module CDK8 and MED12 (Mao et

al., 2019A), thereby regulating lignin biosynthesis and plant growth (Mao et al.,

2019B). The role of Mediator in the regulation of lignification may be to integrate

various signals, ensuring appropriate developmental transitions.

Chromatin-level regulation

To fine-tune plant developmental processes, the transcriptional activity of genes

is commonly regulated at the chromatin level. The state of the chromatin affects

its accessibility for the transcription factor machinery. While a plethora of

transcription factors have been linked to lignification, knowledge on its

chromatin level regulation is sporadic. Studies done in the xylem of the

Eucalyptus tree model found correlations between the abundance of the

transcriptionally activating (H3K4me3) and repressive (H3K27me3) chromatin

marks and the expression level of some lignin-biosynthetic genes (Hussey et al.,

2015; 2017). The H3K27me3 enrichment of lignin and phenylpropanoid

biosynthetic genes during early xylem development corresponds to the lack of

lignification at that point (Hussey et al., 2017). Plants likely utilize the toolbox

of chromatin modifications to coordinate the developmental transitions within the

xylem. Genes involved in secondary cell wall development might be repressed in

the cell expansion zone and activated later on, however this aspect of lignification

needs to be elucidated in future studies.

Circadian regulation

The circadian clock is present in all living organisms. In plants the biological

timing coordinates the metabolism, growth, molecular and cellular processes with

environmental changes. Secondary growth is also likely to be under circadian

regulation.

It was demonstrated in Populus tremula x tremuloides trees that the circadian

clock differentially influences the auxin and cytokinin flow, which in turn affects

cell elongation and xylem formation in a diurnal fashion (Edwards et al., 2018).

Genes encoding for 23 enzymes from the phenylpropanoid pathway exhibit a

synchronized expression peak before dawn (Harmer et al., 2000). Furthermore,

starch turnover and carbohydrate availability influence the magnitude of

circadian oscillations of lignin biosynthetic gene expression. Independent of the

circadian clock these same biosynthetic genes may be impacted by light

perception as well (Rogers et al., 2005). 13C labelling experiments of the

23

developing wood of Populus tremula x tremuloides trees showed that the

incorporation of labelled carbon to lignin was highest at the end of the night,

supporting the early findings (Harmer et al., 2000) that lignin biosynthesis takes

place before dawn (Mahboubi et al., 2015).

Lignin is a major carbon sink and carbon availability is cyclic during the day.

Involvement of diurnal aspect in the control of lignin biosynthesis is plausible

due to the fact that optimal biomass accumulation in the woody tissues should be

coordinated with resource availability and starch turnover (Rogers et al., 2005).

Additional levels of regulation

The enzyme catalysing the first step of lignin biosynthesis, PAL, controls the flux

into the pathway. PAL is known to be under post-translational control. Kelch-

repeat F-box (KFB) proteins physically interact with PAL isozymes and mediate

their proteolysis via the ubiquitin-26S proteasome pathway (Zhang et al., 2013).

Furthermore, the Populus trichocarpa COMT is regulated by reversible

phosphorylation, enabling the cell to adjust the metabolic flux to developmental

and environmental needs (Wang et al., 2015).

Other factors affecting the flux through the pathway are the compartmentalization

of pathway intermediates as well as their inhibitory activity towards specific

biosynthetic enzymes, and the access to cofactors and co-substrates like

shikimate, ATP and S-adenosylmethionine (reviewed by Vanholme et al., 2019).

3.4.2. Transport of the lignin monomers

The detection of only scarce pools of free monolignols by targeted metabolomics

(Jaini et al., 2017) suggests that, upon synthesis, the lignin monomers are rapidly

transported across the plasma membrane for polymerization in the secondary cell

wall. Three different transport mechanisms have been proposed: exocytosis via

ER-Golgi derived vesicles, passive diffusion, and export via membrane-bound

transporters and proton-coupled antiporters (Meents et al., 2018). Observations

of early autoradiographic studies using radiolabelled phenylalanine suggested

that monolignols are present in the rough ER, Golgi and plasma membrane

associated vesicles (Pickett-Heaps, 1968; Fujita and Harada, 1979; Takabe et al.,

1985). The more recent findings of Kaneda et al. (2008) showed that the

phenylalanine incorporation was primarily into proteins rather than monolignols.

Hence the contribution of the ER-Golgi-vesicle mediated exocytosis as a lignin

monomer transport mechanism remains unclear. The small size of monolignols

may facilitate their passive diffusion across membranes. This theory was

24

supported by in vitro observation of the diffusion of lignin precursor analogs into

liposomes and lipid bilayer discs (Boija et al., 2007; 2008), but there is no in vivo

evidence for passive diffusion yet. According to this theory, the concentration

gradient of monolignols between the cytosol and the cell wall would determine

the diffusion rate. The secondary cell wall localized lignin-polymerizing laccases

and peroxidases may create a monolignol sink and an energetically favourable

gradient by the quick polymerization of free monolignols (Perkins et al., 2019).

In addition, Vermaas et al. (2019) demonstrated that most lignin-related

compounds can passively cross membranes at adequate translocation rate to meet

the need of the polymerization machinery.

The third mechanism for lignin monomer transport across the plasma membrane

to the cell wall is mediated by ATP-binding cassette (ABC) transporters and

proton-coupled antiporters (PCA) (Miao and Liu, 2010; Kaneda et al., 2011;

Alejandro et al., 2012; Tsuyama et al., 2013). Based on co-expression with

phenylpropanoid biosynthetic genes, several ABC transporters were identified as

candidates in monolignol export (Ehlting et al., 2005). To date only one,

ABCG29, has been demonstrated as a monolignol transporter (Alejandro et al.,

2012). ABCG29 is a candidate exporter for p-coumaryl alcohol units which are

however only found in small amounts in the angiosperm secondary cell wall.

Yeast cells expressing the Arabidopsis ABCG29 were capable of p-coumaryl

alcohol transport, while the lack of ABCG29 in Arabidopsis increased the

sensitivity to this alcohol (Alejandro et al., 2012). Furthermore, ABCG11,

ABCG22, ABCG29 and ABCG36 are co-expressed with MYB58 in Arabidopsis

tracheary element cell cultures, indirectly suggesting their involvement in

lignification (Takeuchi et al., 2018). In addition, proton-coupled antiporters were

proposed as a transport mechanism of lignin precursors in both angiosperms and

gymnosperms. Membrane vesicles prepared from the differentiating xylem

displayed coniferin (the glucoside of coniferyl alcohol) transport activity in an

ATP-dependent fashion (Tsuyama et al., 2013). Such conserved, proton-gradient

dependent transport mechanism was demonstrated for p-glucocoumaryl alcohol

as well (the glucoside of p-coumaryl alcohol) (Tsuyama et al., 2019). Currently,

the passive diffusion of monolignols without a need for transporters seems to be

the strongest theory for monolignol movement across the plasma membrane

(Vermaas et al., 2019), however this needs to be confirmed with more in vivo

studies.

25

3.4.3. Lignin polymerization

Laccases and peroxidases

Polymerization of lignin typically follows the deposition of polysaccharides

during secondary cell wall formation. It has been postulated that the

polysaccharide matrix provides a scaffold for lignin polymerization (Tobimatsu

and Schuetz, 2019). In the cell wall, each monolignol produces the monomer

radical form by losing one H atom. The free radicals then assemble by oxidative

coupling, forming the lignin polymer. Laccases and peroxidases are the

phenoloxidases implicated in the activation of monolignols to radicals (Fig. 4).

Upon polymerization, monolignols become syringyl (S), guaiacyl (G) and p-

hydroxyphenyl (H) lignin units (Boerjan et al., 2003).

Laccases are multi-copper containing extracellular glycoproteins that oxidize

phenolic, inorganic and aromatic amine substrates, using molecular oxygen (O2)

as a secondary substrate. A large number of functionally redundant laccase

isoforms are present in angiosperms. Out of the 17 laccases found in Arabidopsis,

reverse genetic studies showed that LACCASE4 (LAC4), LAC11, LAC15 and

LAC17 have roles in lignin polymerization. While LAC15 is required for seed

coat lignification (Liang et al., 2006), LAC4, LAC11 and LAC17 are involved in

the lignification of tracheary elements and xylem fibers (Berthet et al., 2011;

Zhao et al., 2013). Several members of the LAC gene family are under the post-

transcriptional control of miRNAs. miRNA397 targets members of the LAC gene

family in Populus trichocarpa, Chinese pear, flax (Linum usitatissimum) and

hemp (Behr et al., 2019). Moreover, the Arabidopsis miRNA857 and

miRNA397b regulate LAC7 and LAC4 transcripts, respectively (Zhao et al.,

2015; Wang et al., 2014).

Another group of phenoloxidases expressed in lignifying tissues are the heme

containing class III peroxidases. These enzymes polymerize lignin monomers in

vitro and their secondary substrate is hydrogen peroxide (H2O2). Similar to

laccases, peroxidases are encoded by a multigene family. Arabidopsis has 73

class III peroxidases. Only a subset of these enzymes has been implicated in

lignification of different cell types: PEROXIDASE2 (PRX2), PRX4, PRX25,

PRX47, PRX52, PRX64, PRX66, PRX71 and PRX72 (Tokunaga et al., 2009;

Herrero et al., 2013; Lee et al., 2013; Fernández-Pérez et al., 2015; Shigeto et al.,

2013; 2015). Peroxidases display different affinity towards the lignin precursors.

For example, some only use sinapyl alcohol as a substrate (Østergaard et al.,

2000), while others can catalyze the polymerization of sinapyl and coniferyl

alcohol as well (Demont-Caulet et al., 2010).

26

Several studies suggest that laccases and peroxidases do not act redundantly, but

rather in a complementary manner (Lee et al., 2013; Zhao et al., 2013). Laccases

are expressed in the Casparian strip forming endodermis, but they do not

contribute to the lignification of the Casparian strip (Rojas-Murcia et al., 2020).

Lignin polymerization in the Casparian strip is dependent on PRX64 (Lee et al.,

2013). The establishment of functional TEs involves both phenoloxidase

families, and it has been proposed that laccases and peroxidases may act

sequentially in TE lignification (Sterjiades et al., 1993). Barros et al. (2015)

discussed three hypothetical models how the phenoloxidase activities could be

determined by different substrate specificities, distinct spatio-temporal

expression or differential complex formation. Complementary activity of

laccases and peroxidases is supported by the subcellular localization study which

found that fluorescently-tagged LAC4 is localized in all secondary cell wall

layers of fibers and vessels, while PRX64 is found only in the middle lamella and

cell corners of fibers (Chou et al., 2018). Recently, Hoffmann et al. (2020)

demonstrated that, depending on the cell type and the developmental stage,

specific LACs and PRXs colocalized within specific cell wall domains. High

resolution localization analyses in Arabidopsis revealed the presence of LAC4,

LAC17, and PRX72 in the thick secondary cell wall of xylem vessel elements

and fibers, whereas LAC4, PRX64, and PRX71 were localized to fiber cell

corners. Early in fiber development LAC4 was briefly expressed in the fiber cell

corners, possibly having a role in the initiation of lignification.

The establishment of the characteristic patterned secondary cell walls requires the

spatial restriction of lignification. Once in the apoplast, monolignols are highly

mobile (Lion et al., 2017; reviewed by Tobimatsu and Schuetz, 2019), while

fluorescence recovery after photobleaching (FRAP) analysis revealed that LAC4

is immobile, likely being anchored to the secondary cell wall (Schuetz et al.,

2014; Chou et al., 2018). PRXs are localized throughout the whole cell wall, but

their oxidative activity is specific to the lignifying regions (Hoffmann et al.,

2020). Presumably, the controlled production of reactive oxygen species and their

relative amount in the apoplastic region restrict PRX activity (Hoffmann et al.,

2020).

Altogether, it seems clear that the patterning of lignin deposition is determined

by the localization of the phenoloxidases and the availability of the monolignols

and the reactive oxygen species (Tobimatsu and Schuetz, 2019; Hoffmann et al.,

2020).

27

Figure 4. Monolignol transport and the formation of the lignin polymer. The

monolignols are moved outside the cell through passive diffusion across the plasma

membrane, by vesicle-mediated transport or by ABC transporters and proton-coupled

antiporters (PCA). O2 and H2O2 are the secondary substrates of phenoloxidases. O2

can freely diffuse from the air, while H2O2 is generated by the sequential activity of

NADPH oxidase and superoxide dismutase (SOD). Laccases and peroxidases

activate the monolignols to radicals. The free monolignol radicals react with a radical

in a growing lignin polymer via oxidative coupling.

Secondary substrates

The secondary substrates of the phenoloxidases, O2 and H2O2, are necessary for

the production of the lignin monomer radicals. O2 availability is not limited, since

oxygen can diffuse from the air and it is therefore present in the xylem sap

(Gansert, 2003). H2O2 has a very short half-life (10-9 – 10-3 s) (D’Autréaux and

Toledano, 2007), implying its continuous production in lignifying tissues. H2O2

is produced by the sequential activity of two enzymes: nicotineamide adenine

dinucleotide phosphate hydrogen oxidase (NADPH oxidase) and superoxide

dismutase (SOD) (Fig. 4) (Ogawa et al., 1997). NADPH oxidase catalyzes the

production of apoplastic superoxide from cytoplasmic NADPH (Sagi and Fluhr,

2006), and SOD derives O2 and H2O2 from the toxic superoxide radicals (Liochev

and Fridovich, 1994). In the xylem, H2O2 production was localized to tracheary

elements and parenchyma cells (Ros Barceló et al., 2005; Gómez Ros et al.,

2006). Furthermore, SOD was shown to be present in xylem vessels, fibers and

28

parenchyma cells (Karpinska et al., 2001), indicating that all xylem cell types are

a source of phenoloxidase secondary substrates.

Assembly of the polymer

The dehydrogenation of the monolignols by phenoloxidases is followed by cross

coupling to the growing lignin polymer. Presumably, the monomer-polymer

coupling reactions are the result of a redox shuttle system. Önnerud et al. (2002)

identified manganese as a radical carrier. Upon oxidation by phenoloxidases, the

manganese ions diffuse to the lignification site in the cell wall, where the

monolignols and the growing end of the lignin polymer are oxidized. Manganese

and calcium are both homogenously distributed in wood; the latter forms calcium

pectate at the lignification initiation sites, in the middle lamella and the cell

corners. Calcium was proposed to form calcium superoxide complexes that

oxidize the growing lignin polymer (Boerjan et al., 2003).

The structure of the highly branched lignin polymer forms by the cross-coupling

of monomer and oligomer radicals through condensed (C-C) linkages (5-5

biphenyl, β-β resinol, β-5 phenylcoumaran, β-1 diarylpropane) and non-

condensed (C-O-C) ether linkages (β-O-4 β-aryl-ether, 5-O-4 biphenyl ether)

(Terashima and Fukushima, 1988). The coupling of two monolignols,

dimerization, preferentially occurs at the β positions, resulting in β-O-4, β-β and

β-5 bonds. The β-O-4 ether bond is generated most frequently (Vanholme et al.,

2010). The formed dimer is then dehydrogenated in order to couple with the next

monomer radical. The gradual extension of the polymer with one radical is

through endwise polymerization. Most of the 5-5 and 5-O-4 bonds are formed by

coupling of preformed oligomers (Boerjan et al., 2003). 5-5 linkages can be

observed in the coupling of G-lignin oligomers (Vanholme et al., 2010). Lignin

with high G unit content is more highly crosslinked, more rigid and more

hydrophobic than lignin rich in S units. Interestingly, if only β-O-4, β-β and β-5

bonds were present in the polymer, it would result in a completely linear structure

(Bonawitz and Chapple, 2010). The abundance of the various linkage types is

highly dependent on species, on the availability of each lignin subunit and on the

cell type. Also, the polymer length is determined by the availability and reactivity

of different lignin monomer radicals (Ralph et al., 2008). In Populus tremula x

alba the average length of a linear lignin chain may be between 13 and 20 units

(reviewed by Vanholme et al., 2010).

The initiation site for lignin polymerization is generally at the cell corner of the

middle lamella and in the S1 region of the secondary cell wall. Several hypotheses

29

have been put forward about what defines the exact location where

polymerization begins. Dirigent proteins were proposed to be the landmarks for

lignin initiation sites (Davin and Lewis, 2000). Currently ENHANCED

SUBERIN1/DIRIGENT10 (ESB1/DIR10) is the only dirigent protein which has

been shown to be required for the formation and lignification of the Casparian

strip (Hosmani et al., 2013). In grasses ferulates, wall-bound phenolics, can

dimerize or polymerize by forming ester-to-ether linkages to cross-link lignins

and polysaccharides (Ralph et al., 1995). These phenolics can provide initiation

sites for lignin polymerization (reviewed by Ralph et al., 2004). Furthermore

glycine-rich proteins and pectin-bound peroxidases were also postulated to form

nucleation sites (Boerjan et al., 2003).

3.4.4. Cell-autonomous vs. non-cell-autonomous lignification

Lignin biosynthesis and deposition are tightly regulated but at the same time

flexible and dynamic processes, occurring in response to specific developmental,

metabolic or environmental changes. A better understanding of the mechanisms

regulating lignification is necessary to enable lignin engineering (Ragauskas et

al., 2014). Lignification of certain cell types can occur in cell-autonomous as well

as non-cell-autonomous fashion (Barros et al., 2015). It was suggested early on

that lignification of the vessel walls continues after their cell death since they

differentiate and die rapidly to become functional in water transport (Stewart,

1966; Donaldson, 2001). The post-mortem, non-cell-autonomous lignification

would require the cooperative action of vessel-neighbouring, living cells. In line

with this theory, several genes encoding for enzymes from the lignin biosynthetic

pathway are not exclusively expressed in TEs, but also in the cambium and the

xylem parenchyma cells (Hu et al., 1998; Chen et al., 2000; Lacombe et al., 2000;

Lauvergeat et al., 2002; Baghdady et al., 2006; Blokhina et al., 2019).

Additionally, metabolomics analysis of Norway spruce tracheids and ray

parenchyma cells revealed the presence of monolignols and glycoconjugates in

both cell types (Blokhina et al., 2019). The xylogenic Zinnia elegans cell culture

system, containing a mixture of TEs and non-lignifying parenchymatic cells, has

been practical to dissect the cooperative aspect of lignification. The dead TEs in

this in vitro system continued to lignify thanks to the action of the living

parenchymatic cells (Hosokawa et al., 2001; Pesquet et al., 2013). Manipulation

and analyses of the cell culture media revealed that monolignols are supplied to

the TEs from the parenchymatic cells through the medium (Hosokawa et al.,

2001). Lignin-biosynthetic genes are expressed in the parenchymatic cells which

therefore can both synthesize and transport monolignols and reactive oxygen

30

species to the TEs (Pesquet et al., 2013). The limitation of the cell culture system

is the artificial setup and the lack of developmental and tissue context.

Smith et al. (2013) demonstrated in planta that vessel lignification is a

cooperative process where the lignin monomers are supplied to the lignifying

cells by the neighbouring cells. Lignin biosynthesis was suppressed in lignifying

xylem cells by an artificial microRNA targeting CCR1 driven by the CESA7

promoter. Still, the vessel elements lignified normally, thanks to the contribution

of the neighbouring xylem parenchyma cells. However, when lignin biosynthesis

was specifically blocked in the parenchyma cells the vessels exhibited collapsed,

irregular xylem phenotype (Smith et al., 2017). The cell wall of these partially

collapsed vessels was not completely devoid of lignin, suggesting that the vessels

initiate cell-autonomous lignification while they are still alive. Moreover, non-

native monolignol conjugates were expressed in the parenchyma cells and these

compounds were incorporated into the cell wall (Smith et al., 2017). Later in stem

development xylary fibers also contribute lignin precursors to vessel lignification

but less than the parenchyma cells, and this lignin pool is not essential for proper

vessel formation (Smith et al., 2017). Xylary fibers also have the ability to

undergo non-cell-autonomous lignification, while interfascicular fibers fully

lignify in a cell-autonomous manner (Smith et al., 2013). The vessel-

neighbouring fibers of Populus tremula exhibited lignin composition

intermediate between that of vessels and fibers, having relatively high levels of

G lignin units (Gorzsás et al., 2011). It is conceivable that the function of the

xylem cell and its tissue context determines the type of lignification the cell

undergoes.

Early in the evolution many plant taxa did not contain parenchyma cells within

the xylem, possibly relying exclusively on cell-autonomous lignification. The

ability of the parenchymatic cells to cooperatively supply lignin precursors to the

water-transporting vessels may be a more recently evolved trait (Smith et al.,

2017). However, the evolution of the rays might not be the only determinant of

non-cell-autonomous lignification, since some of the fibers also have the ability

to take part in the process. Altogether, the reason for the appearance of non-cell-

autonomous lignification is not clear. This developmental strategy of the land

plants could have evolved to cope with the increasing tension pressure created by

water transport and with the increase in the height and weight of the growing

plant body. Continuous fortification of the cell walls throughout the plant’s

lifespan certainly allows more efficient and undisturbed flow of the xylem sap

through the vascular system. Furthermore, non-cell-autonomous lignification

31

could enrich the ability of xylem elements to adapt to changing environment and

pathogens.

3.4.5. Lignin as a resource for biotechnological applications

The secondary cell walls are a rich source of lignocellulosic plant biomass, with

the potential to partially replace fossil fuel usage. The effectiveness at which such

biomass pool can be utilized depends greatly on the chemical properties and

ultrastructure of the cell wall (Meents et al., 2018). Currently, the large-scale

industrial use of plant biomass is limited due to high processing costs. Lignin is

the major barrier to efficient extraction of cellulose fibers and to saccharification

for production of pulp and liquid biofuel, respectively (Baucher et al., 2003; Zeng

et al., 2014), and its removal is pricey, laborious and environmentally hazardous.

Lignin also curtails the nutritional value of crops meant for human consumption

or livestock feed (Wang and Frei, 2011). Unfortunately, extreme environmental

conditions, which will become more common due to climate change, enhance

lignin biosynthesis (Moura et al., 2010; Ployet et al., 2018; Behr et al., 2019).

Biotechnological modifications aim to improve lignocellulose utilization by

increasing the calorific values, the extractability of cell wall polymers, as well as

saccharification and fermentation yields (Boerjan et al, 2003; Weng et al., 2008;

Mansfield, 2009; Novaes et al., 2010; Vanholme et al., 2010). In field crops,

lignin content could be decreased by traditional breeding methods, natural and

induced mutations, and transgenesis. Ideally, one would like to reduce the content

of the lignin polymer to a level where efficient saccharification could be achieved

without pretreatment. Notably, the pretreatment of the lignocellulosic biomass

imposes major environmental hazards. However, the straightforward reduction of

the total lignin content by genetic engineering has frequently led to reduced

overall biomass production even under controlled growth conditions (Piquemal

et al., 1998; Franke et al., 2002; Leplé et al., 2007; Voelker et al., 2010).

Typically, reduction of lignin affects negatively plant fitness and productivity

(Wagner et al., 2009). Nonetheless, it was demonstrated in Arabidopsis, tobacco

(Nicotiana tabacum) and Populus tremula x alba that changes in the ratio of

certain lignin units do not compromise plant performance and the water-

conducting capacity of the vessel elements. Specifically, increased S/G ratio

could improve the bioprocessing properties of wood without having a negative

influence on plant growth. (Franke et al., 2000; Huntley et al., 2003; Sibout et

al., 2002; Stewart et al., 2009; Studer et al., 2011). The methoxy groups of S-

lignin present at the 3 and 5 positions of the aromatic ring enhance reactivity to

32

thermal and catalytic fragmentation. In the native lignin polymer, the β-ether

bonds are the most labile, and their frequency is higher in the lignin polymer

when there is increased S unit content. Consequently, the S unit rich lignin has

less 5-5 and β-5 bonds, which makes it more easily extractable (Mahon and

Mansfield, 2019). Increasing the S/G ratio of the lignin polymer is however not

feasible in softwood species which do not have S-type lignin. Nevertheless,

Wagner et al. (2015) demonstrated that S-type monolignol units could be

generated in Pinus radiata by metabolic engineering of the phenylpropanoid

pathway. Furthermore, esther linkages were introduced in Populus alba ×

grandidentata by expressing a non-native monolignol transferase; this

modification improved the digestibility of the lignin polymer (Wilkerson et al.,

2014). Altogether, the plasticity of the lignin biosynthetic pathway is a promising

asset to generate easily exploitable biomass.

In the xylem of Arabidopsis, along with most angiosperms, fibers contain the

bulk of the secondary cell walls. In Arabidopsis the lignin content of the

interfascicular fiber cell walls can be significantly reduced without affecting plant

fitness. Growth is not at all or only hardly compromised as long as vessels lignify

normally (Smith et al., 2017). Using synthetic biology tools Yang et al. (2013)

managed to direct lignin biosynthesis to vessels and increase the polysaccharide

deposition in the fiber cell walls. This combined strategy resulted in higher sugar

yields. Tissue-specific alteration of the lignin profile, with the aim to maintain

vessel integrity, is a feasible biotechnological approach for biomass amelioration.

The biosynthesis of an alternative lignin monomer, curcumin, was successfully

introduced into Arabidopsis, resulting in a polymer more prone for chemical

changes. Plants containing curcumin and phenylpentanoids in their cell walls

exhibited normal growth and greatly increased saccharification efficiency,

indicating the potential of non-native monomers as a way to improve biomass

properties (Oyarce et al., 2019).

Likely, crosslinking and interaction between lignin and the cell wall carbohydrate

matrix limits the sugar release efficiency after enzymatic saccharification (Mahon

and Mansfield, 2019). For instance, xylan might form linkages with lignin via the

methyl-glucuronic acid ([Me]GlcA) branching groups, since the genetic removal

of [Me]GlcA lead to increased xylose and glucose release upon enzymatic

hydrolysis (Lyczakowski et al., 2017). Such findings demonstrate that

downstream applications may be eased by genetically distancing the cell wall

polymers from each other.

33

Lignin is often referred to as the recalcitrant in the cell wall we all want to get rid

of, but this rudimentary view should be revised. Lignin is also a highly attractive

polymer due to its high energy content and structure that could be utilized. A

sustainable and profitable biorefinery concept requires maximal and diversified

use of the feedstock (Liao et al., 2020). The approaches to turn lignin into a

resource include the generation of homogenous and linear polymers consisting

exclusively of caffeyl alcohol (C-lignin) for carbon-fiber production, introduction

of valuable novel monomers like tricin or increasing the aldehyde group content

(Ralph et al., 2019; Mahon and Mansfield, 2019). Tricin is a flavonoid which can

only function as polymerization initiator, and it can be used to generate shorter

polymers (Lan et al., 2015). Aldehyde-rich lignin has less non-covalent

interactions with hemicelluloses and in turn it is more suitable for downstream

applications, such as saccharification (Zhao et al., 2013; Carmona et al., 2015;

Van Acker et al., 2017).

Recent advances in lignin valorization include its use in thermoplastics,

blends/composites, as carbon fiber precursors, heavy metal absorbents and

nanoparticles, as wells as the incorporation of lignin-based copolymers into

biomedical materials (reviewed by Wang et al., 2020). These emerging fields

provide diverse novel opportunities for the utilization of the lignocellulosic

biomass.

Plant biomass represents a valuable and multiuse resource. Enhancing the value

and usability of lignin by exploiting natural variations and the potentials of

genetic engineering may enable the establishment of sustainable biorefineries.

34

Research aims and objectives

The aim of my thesis is to identify and characterize novel regulators of

lignification, using Arabidopsis thaliana as a model system. Lignification is a

multistep process, and the focus of these investigations is on the biosynthesis of

lignin monomers. The overall goal was to address the regulation of lignin

biosynthesis from previously less-explored angles, such as non-cell-autonomous

lignification and the timing of lignin biosynthesis, and provide new molecular

tools for the field.

To address the research goal, a variety of methods was employed, such as cell

wall chemistry analyses, genetic, transcriptomic, microscopic and

microspectroscopic tools. Consequently, three potential regulators of

lignification were identified. In Paper I the first known molecular regulator in

non-cell-autonomous lignification is revealed. To investigate the molecular

network of this novel regulator a large-scale protein-interactor screen was

performed (Paper II). The screen yielded several candidate genes, and based on

cell wall chemistry analysis and literature studies two of them were selected for

further investigation. The candidates were characterized in detail in connection

to lignin biosynthesis (Paper II, Paper III). These findings will help to better

understand secondary cell wall development, provide knowledge pool which can

be transferred into economically important species and potentially be used for

biotechnological purposes in the future.

35

Results and discussion

PIRIN2 is a non-cell-autonomous regulator of S-type lignin

accumulation

A PIRIN (PRN) protein was earlier suggested to play a role in non-cell-

autonomous lignification in Zinnia elegans xylogenic cell cultures (Pesquet et al.,

2013). The concept of non-cell-autonomous lignification implies that xylem

elements are assisted by their neighbouring cells to get their cell walls fully

lignified (Hosokawa et al., 2001; Pesquet et al., 2013; Smith et al., 2013; De

Meester et al., 2018). While this theory has been around for many years, the

molecular players have remained unknown. The initial findings by Pesquet et al.

(2013) prompted us to investigate the PRN gene family and their involvement in

lignification in Arabidopsis.

Promoter activity analyses revealed that out of the four Arabidopsis PRN genes,

PRN2, PRN3 and PRN4 are active in the xylem of seedlings and mature plants

(Paper I, Fig. 1). The PRN2 promoter exhibited a very specific pattern restricted

to the xylem cells neighbouring the vessel elements. Similarly, the promoter of

the PRN2 homologue in Populus was active in the ray cells next to the vessel

elements (Paper I, Fig. S2). The localization of PRN2 next to the vessel elements

was confirmed by confocal laser scanning microscopy imaging of the PRN2:GFP

fusion protein (Paper I, Fig. 2). The promoter activity and protein localization

results revealed that PRN2 is present in the cells surrounding the vessel elements,

but not in the vessel elements themselves. These provided the first indication that

PRN2 is likely to be involved in vessel differentiation or function. Since vessel

elements undergo programmed cell death early on, they might need the support

of neighbouring cells for complete differentiation or to function properly in

different processes.

Next, we performed pyrolysis-gas chromatography/mass spectrometry (py-

GC/MS) on a set of prn single and double mutants to assess their possible role in

lignification (Paper I, Fig. 3). Besides determining total lignin content, py-

GC/MS is also able to distinguish between G-, S- and H-type lignin, which are

important indicators of the processability of the cell wall material. All prn2

mutants exhibited increased S-type lignin content and consequently significantly

higher S/G-type lignin ratio (Paper I, Fig. 3, 4). The effect of PRN2 on the S-

type lignin content of the secondary cell walls was confirmed by two-dimensional

nuclear magnetic resonance spectroscopy (Paper I, Fig. 4).

36

Upon observing the vessel-neighbouring expression pattern of PRN2 and

chemical changes at the whole tissue level, we set on to investigate the lignin

profile on a cellular level. Raman and FT-IR microspectroscopy allowed in situ,

high-resolution chemotyping of individual cell walls (Gorzsás et al., 2011; 2017)

of vessels and fibers next to the predicted PRN2-expressing cells in prn2. The

microspectroscopic analyses revealed chemical changes that correlated with S-

type lignin, G-type lignin, lignin content and with structural properties of the

lignin polymer in the hypocotyl and in the vascular bundle of the inflorescence

stem of the prn2 mutants (Paper I, Fig. 5). Smith et al. (2017) demonstrated that

interfascicular fibers lignify predominantly in a cell-autonomous manner.

Interestingly, the cell wall chemistry of the interfascicular fibers was not altered

in prn2, suggesting the absence of non-cell-autonomous lignification in this

region, or that PRN2 does not play a role there. Altogether, these findings confirm

that PRN2 controls the non-cell-autonomous lignification of xylem elements.

Next, we were curious to address the mode of function of PRN2. The human PRN

protein interacts with several different transcription factors and acts as a

transcriptional co-regulator (Wendler et al., 1997; Dechend et al., 1999; Liu et

al., 2013B). The subcellular localization of PRN2 to the nucleus and the

cytoplasm was demonstrated earlier (Zhang et al., 2014). Thus, we hypothesized

that PRN2 might regulate S-type lignin biosynthesis by the transcriptional control

of lignin-biosynthetic genes. Notably, several lignin-biosynthetic genes and

secondary cell wall-related transcription factors had increased expression in prn2

mutants, and the PRN2-overexpressing lines showed opposite tendencies for

many genes (Paper I, Fig. 6). F5H1 encodes the enzyme required for S-type

lignin biosynthesis. Interestingly, out of all genes analyzed F5H1 displayed the

most consistent changes in expression, with significantly higher expression level

in the prn2 mutants and lower in the PRN2-overexpressing lines. In conclusion,

the gene expression analyses uncovered the function of PRN2 as a transcriptional

suppressor of lignin-biosynthetic genes and secondary cell wall-related

transcription factors.

PRN2 had the most consistent and pronounced effect on S-type lignin

accumulation. Presumably, PRN2 suppresses S-type lignin biosynthesis by

regulating the expression of F5H1, a key S-type lignin-specific biosynthetic gene.

This, together with the vessel-neighbouring expression of PRN2, suggest that

PRN2 is a molecular factor ensuring that vessel elements acquire their typical, G-

type lignin enriched cell walls which are more resistant to the mechanical forces

created by water transport.

37

HISTONE MONOUBIQUITINATION2 (HUB2) affects the

lignin composition of xylem vessels

PRN2 that was described as a suppressor of S-type lignin accumulation in

is not a transcription factor, and has to interact with other proteins in the

regulation of lignin-biosynthetic genes. To identify the molecular partners of

PRN2, we performed a large-scale yeast two-hybrid screen. We retrieved 18

protein interactor candidates (Paper II, Dataset 1 available online at Dryad).

To identify interactors which affect lignin content or composition, py-GC/MS

analysis was performed on T-DNA insertion mutants of the PRN2 interactor

genes. From all genotypes analyzed, the two knockout lines of HUB2 exhibited

significantly lower S-type lignin level and increased G-type lignin content (Paper

II, Fig. 1). These findings implied a previously unreported, lignin-related

function for HUB2.

HUB2 is an E3 ubiquitin ligase catalyzing the histone H2B monoubiquitination

of the chromatin (H2Bub1). HUB2 is involved in various plant developmental

processes, e.g. flowering time (Cao et al., 2008; Xu et al., 2009), circadian clock

(Himanen et al., 2012) and pathogen resistance (Zou et al., 2014; Zhang et al.,

2015; Zhao et al., 2020). Cotton GhHUB2, the functional homolog of AtHUB2,

was implicated in secondary cell wall formation and lignification (Feng et al.,

2018). Based on the PRN2-HUB2 protein interaction finding, we addressed the

possible role of HUB2 in lignification in Arabidopsis.

The protein interaction between PRN2 and HUB2 was confirmed by both in vitro

and in vivo approaches (Paper II, Fig. 2). Next, we analyzed the promoter

activity of HUB2 in the xylem of mature stems and hypocotyls. The HUB2

promoter was active in different xylem cell types, including the PRN2-expressing

cells (Paper I, Fig. 1J-L), but was absent from the vessel elements (Paper II,

Fig. 3C-E).

The initial mutant screen was followed up by genetic interaction study through

py-GC/MS analysis of hypocotyl samples of hub2 and prn2 single, and hub2 prn2

double mutants (Paper II, Fig. 3). Unlike the hub2 single mutants, the hub2 prn2

double mutants exhibited higher S-type lignin content compared to WT,

suggesting that prn2 is epistatic to hub2. The G-type lignin content of hub2 and

the hub2 prn2 double mutants was significantly higher compared to WT, while

the effect of prn2 on G-type lignin content was varying. The opposite effect of

HUB2 and PRN2 on S-type lignin accumulation was confirmed in mature parts

of the inflorescence stem (Paper II, Fig. 4).

38

To elucidate the effect of HUB2 on individual xylem cell walls, Raman and FT-

IR microspectroscopic chemotyping of hub2 mutant hypocotyls was performed

(Paper II, Fig. 5, Fig. S2). The in situ chemical profiling corroborated the py-

GC/MS results obtained at the whole tissue level, revealing elevated levels of G-

type lignin, and consequently lower S/G ratios. Altogether, these findings

indicate that HUB2 promotes S-type lignin accumulation at the expense of G-

type lignin.

HUB2 is part of the enzyme complex mediating H2Bub1 chromatin

modifications in Arabidopsis. Hence, we hypothesized that HUB2, together with

PRN2, affect lignification by mediating the H2Bub1 levels of lignin-biosynthetic

genes. Strikingly, the ChIP-qPCR profiling of prn2 did not reveal differences in

H2Bub1 levels of selected lignin-biosynthetic genes (Paper II, Fig. S3). These

results implied that PRN2 and HUB2 act through an H2Bub1-independent

mechanism (Paper I). To address the effect of HUB2 and PRN2 on gene

expression, transcriptomic analysis was performed on mature hypocotyls of hub2

and prn2 single, and hub2 prn2 double mutants (Paper II, Fig. 6). Since PRN2

and HUB2 had distinctly opposite effect on S-type lignin content, we focused on

the transcriptional changes which could underlie such cell wall chemistry

changes. Strikingly, F5H1 which encodes the enzyme specific for S-type lignin

biosynthesis was up-regulated in prn2 and hub2 prn2, but down-regulated in

hub2. These changes in F5H1 expression correlated with the S-type lignin content

of the genotypes. Taken together, our results suggest that PRN2 and HUB2

modulate S-type lignin levels through the transcriptional regulation of F5H1. The

protein interaction and genetic analyses indicate that PRN2 and HUB2 function

together, whereby PRN2 acts downstream of HUB2 in controlling S-type lignin

accumulation.

HUB2 is likely to stimulate S-type lignin biosynthesis in most xylem cells, except

vessel elements, where it is not expressed (Brady et al., 2007). In contrast, the S-

type lignin-suppressor role of PRN2 is concentrated around the vessel elements.

It seems therefore that PRN2 interacts with HUB2 to prevent its S-type lignin

stimulatory function in the neighbourhood of the vessels to ensure their correct

lignin composition.

39

PIB is a potential modulator of the diurnal timing of lignin

biosynthesis

In Arabidopsis xylem, PRN2 suppresses S-type lignin accumulation in a non-cell-

autonomous manner (Paper I) through a mechanism that is at least in part

mediated by suppression of HUB2 function (Paper II). PRN2, however, interacts

with several other proteins (Paper II) which might also participate in the

regulation of vessel lignification. One of these PRN2 interacting proteins is the

CRYPTOCHROME2-interacting basic helix-loop-helix3 (CIB3)/PRN2-

interacting basic helix-loop-helix (PIB) transcription factor (Paper II). The

human PRN protein interacts with transcription factors and functions as a

transcriptional co-regulator, suggesting that PRN2 might interact with PIB in the

regulation of lignin-biosynthetic genes. In addition, the paralog of CIB3,

CIB2/PIB-LIKE, was previously implicated in secondary cell wall formation

(Oikawa et al., 2010). Thus, we addressed the role of PIB in the regulation of

lignification in Arabidopsis.

First, we confirmed the protein interaction between PRN2 and HUB2 in vitro by

yeast growth and enzymatic assays, and in vivo by co-immunoprecipitation assay

(Paper III, Fig. 1). Next, the promoter activity of PIB was analyzed in 5-day-old

seedlings and in 6-week-old plants (Paper III, Fig. 2). The PIB promoter was

active in different tissues throughout the plant, for example in the parenchymatic

cells around the vessel elements in the phase I secondary xylem and in the

vascular tissues of the leaves. Protein localization analysis by confocal laser

scanning microscopy revealed the presence of PIB in the nucleus (Paper III, Fig.

2F). The results of the promoter activity and protein localization analyses imply

the involvement of PIB in a variety of plant developmental processes, including

vascular development.

PRN2 suppresses S-type lignin accumulation (Paper I). To assess the effect of

PIB on lignification, lignin content and composition were determined. Neither

py-GC/MS analysis at the whole tissue level (Paper III, Fig. 3), nor FT-IR

microspectroscopy at the cellular level revealed any changes in the lignin profile

of pib cell walls (Paper III, Fig. S3). Since PIB-LIKE is co-regulated with xylan

biosynthetic genes (Oikawa et al., 2010), the monosaccharide composition of the

pib secondary cell walls was also determined (Paper III, Fig. S4), but no

significant changes could be detected.

Next, gene expression was analyzed by RNA-sequencing of pib-1, pib-like-1,

prn2-2, pib-4 pib-like-2 and pib-1 prn2-2 mutant lines at two timepoints, night

40

(darkness) and day (light) (Paper III, Fig. 4). Two timepoints were selected to

capture any light-dependent changes in gene expression since PIB and PIB-LIKE

interact with the CRY2 photoreceptor (Liu et al., 2013C). The expression of

lignin-biosynthetic genes was significantly higher in pib during the night.

Furthermore, differentially expressed genes (DEGs) belonging to the circadian

rhythm, plant-type cell wall organization and response to blue light gene ontology

(GO) terms were overrepresented in pib during the day. Taken together, it is

plausible that PIB modulates lignin biosynthesis in a diurnal fashion, potentially

to synchronize lignin biosynthesis with resource availability.

Next, we followed the diurnal expression of lignin-biosynthetic genes over a 24-

hr cycle in the roots of the pib mutant. The diurnal expression of C3H1 was

slightly delayed in pib, but the difference was not significant compared to the

Col-0 WT. The expression level of PAL1 and 4CL1 was significantly lower in

pib during the night, but the other lignin-biosynthetic genes did not exhibit

significant changes in their expression (Paper III, Fig. 5). These results suggest

that PIB has only minor effects on the diurnal expression profile of lignin-

biosynthetic genes. However, it is plausible that alterations on a few genes, such

as C3H1, influence the overall timing of lignin biosynthesis. In addition,

functional redundancy between PIB and PIB-LIKE as well as tissue-specific

function complicate the monitoring of any transcriptomic changes at the whole

organ level. Nevertheless, to test the ability of PIB to regulate lignin-biosynthetic

genes, we performed transient expression analysis in tobacco protoplasts. The

transactivation assay demonstrated that PIB can suppress the promoter activity of

most lignin-biosynthetic genes to varying degrees (Paper III, Fig. 6).

In conclusion, PIB, the protein interactor of PRN2, modulates the diurnal

expression of lignin-biosynthetic genes via transcriptional repression. The exact

mechanism and the molecular network of PIB as a diurnal transcriptional

modulator remain to be elucidated. Nonetheless, we propose that PIB affects the

diurnal timing of lignin biosynthesis via transcriptional repression.

41

Conclusions and perspectives

Even though lignification of the plant secondary cell walls has been extensively

studied for decades novel molecular players involved in the process are

continuously identified. This work uncovered three proteins which influence

lignin biosynthesis at the transcriptional level and constitute novel aspects in the

regulation of lignification.

First, PRN2 was identified and characterized as the first known molecular

regulator of non-cell-autonomous lignification in Arabidopsis. PRN2 is

expressed in the cells around the vessel elements and affects the cell wall

composition of the neighbouring vessels and fibers. The promoter activity of the

ortholog in hybrid aspen, PttPRN2, is also next to the vessel elements, suggesting

a similar role for PRN2 in trees. By modulating the expression of PRN2 lignin

composition can be altered without compromising plant health and fitness. This

makes PRN2 and non-cell-autonomous lignification promising tools for

biotechnology. For example, the restriction of lignin biosynthesis to the PRN2-

expressing cells might result in low-lignin lignocellulosic biomass without a yield

penalty.

Furthermore, the molecular network of PRN2 was broadened by the

characterization of two protein interactors. The chromatin modifying enzyme

HUB2 promotes S-type lignin biosynthesis. The exact mechanism of HUB2

function in lignin biosynthesis remained unclear but it did not seem to include

chromatin modification. Single-cell and cell-type-specific techniques will enable

further investigation of the mechanism of PRN2 and HUB2 in lignification.

Finally, we report that the diurnal timing of lignin biosynthesis is modulated by

the PRN2-interacting PIB protein. The diurnal regulation of lignin biosynthesis

was previously unexplored, here we present the first potential modulator of this

process. The effect of PIB on the gene expression changes of lignin-biosynthetic

genes could be light-dependent. The exact molecular mechanism needs to be

elucidated in the future, addressing the possible functional redundancy between

PIB and its paralog PIB-LIKE.

HUB2 and PIB are likely to be part of two independent, PRN2-containing protein

complexes. It is plausible that the activity of the complexes is spatially or

temporally separated. However, the possibility that all three proteins belong to a

single complex cannot be excluded. The finding of additional molecular partners

would help uncover the exact mechanism and makeup of the PRN2-dependent

42

protein complex(es). In Arabidopsis gene duplication events within the PRN gene

family and the sub-family containing PIB and PIB-LIKE, may have contributed

to functional redundancy, thus hampering molecular genetics studies.

Interestingly, in hybrid aspen there are only two PRN genes and one PIB/PIB-

LIKE ortholog. Using hybrid aspen to further study these genes could help to

counteract redundancy, while also gaining understanding on the regulation of

lignification in economically more relevant species.

Figure 5. PRN2, HUB2 and PIB regulate lignin biosynthesis.

In the PRN2-expressing cells next to the vessel elements, PRN2 suppresses the

expression of lignin-biosynthetic genes, and thus monolignol biosynthesis. In

particular, PRN2 negatively regulates S-type lignin biosynthesis by blocking the

promoting effect of HUB2 on F5H1 expression. In the fiber cells where only HUB2

is expressed (PRN2 is not), HUB2 stimulates S-type lignin biosynthesis through the

transcriptional regulation of F5H1. PIB modulates the diurnal expression of the

lignin-biosynthetic genes. Whether PRN2, HUB2 and PIB are part of a single protein

complex or two separate complexes is open for future research.

43

Acknowledgements

My dear PhD supervisor, Hannele Tuominen, I am very grateful to you for these

past five years. It has been a bumpy road, as it is always the case in scientific

research, but the finish line is here. I want to thank you the freedom you gave me

to follow my curiosity, but also taking on the guidance when there was no space

to wonder. I very much appreciate your patience, support, sense of humor and

caring.

I am thankful for the two supervisors I had during my master’s projects: Delphine

Gendre and Kristoffer Jonsson. I learnt a lot from both of you, the time spent

under your guidance prepared me for my PhD research.

I want to thank my colleagues who worked with me in shared projects, or who

helped me in any other way. Thanks to Bo Zhang, Pál Miskolczi, Sacha

Escamez, Carolin Seyfferth, Yin Wang and Maxime Chantreau for all your

expertise, support, and patience.

Furthermore, I want to acknowledge all members of the HT group, Angela,

Anna, Bernard, Caro, Hardy, Isura, Marta, Max, Mikko, Sacha, Shruti and

Yin. I am really grateful for the helpful, supportive, friendly and fun atmosphere

of our group.

I want to thank the three project students I got to work with over the years, Inkeri

Soppa, Hannah Ohm and Sara Ölmelid. I really appreciate all your hard work.

I am grateful to the members of my reference group, László Bakó, Markus

Schmid, Maria Eriksson, Rishi Bhalearo and Urs Fischer for their advises and

that they followed my PhD journey.

A significant amount of time and energy was spent on teaching at bachelor’s and

master’s level courses. I really enjoyed these challenges, but it would not have

been the same without my amazing teaching partners. Thanks to Bernard

Wessels, Daria Chrobok, Jean Claude Nzayisenga, Kerstin Richau, Mikko

Luomaranta and Yan Ji.

Thanks to Junko Takahashi-Schmidt for all the help with the various cell wall

chemistry analyses, and very importantly for her endless patience, hard work and

for the fun moments we shared.

44

I want to express my gratitude to Jan Karlsson, Janne, for all his technical

support and patience. I apologize if my questions were sometimes close to drive

you crazy.

I am very grateful for the work of the personnel in the greenhouse and in the

transgenic facility. Without them our research would not progress, their work is

essential and much appreciated.

I am really glad that we kicked off the Greener UPSC group and I hope that it

will be a long-lasting movement. Thanks to Anne B, Sonja, Nico, Sara R,

Domenique, Alex, Noemi, Lill and Loïc for joining in and I wish you all the

best.

I want to wish everyone at UPSC happy and successful years to come. In my

opinion UPSC is special, since everybody contributes in some way for the

betterment of the work environment, be it greenhouse issues or afterwork events.

Thanks for everything!

To all my friends, in Umeå, in Sweden, in my home and in other corners of the

world: perhaps sometimes unknowingly, but you guys helped me so much along

the way. I am very grateful that I have you in my life.

My special thanks to Karolin, Julia, Hanna néni, Eszter and Szabi for the times

we shared in Umeå and everything else.

I cannot thank enough to my extended family for all their love, support and so

much more.

Az én nagy családom, Annamária, Nénje, Robi, Ancsi és Dóri, tudom azt, hogy

nélkületek ez nem sikerült volna. Nagyon hálás vagyok mindenért.

I want to thank from the bottom of my heart to my parents, who always let me

follow my own path. You are my role models. You have made a difference in so

many people’s life, I am very proud of you.

Drága szüleim! Szívből köszönöm, hogy mindig hagyjátok, hogy a saját utam

járjam. Ti vagytok a példaképeim. Olyan sok emberen segítetek, nagyon büszke

vagyok rátok. Köszönöm, hogy ti mindig bíztok bennem, támogattok és velem

vagytok. Remélem, hogy ez még nagyon sokáig így lesz.

My husband, Anirban, you shared every day of this journey with me, literally

from day one. You made me laugh when things were not great, or when they

were. I learnt so much from your wisdom, professionalism, and straight talk. You

made these years really special, thank you so much for everything!

45

References

Achnine L, Blancaflor EB, Rasmussen S, Dixon RA. 2004. Colocalization of

L-phenylalanine ammonia-lyase and cinnamate 4-hydroxylase for

metabolic channeling in phenylpropanoid biosynthesis. Plant Cell 16:

3098-3109.

Agustí J, Lichtenberger R, Schwarz M, Nehlin L, Greb T. 2011.

Characterization of transcriptome remodeling during cambium formation

identifies MOL1 and RUL1 as opposing regulators of secondary growth.

PLoS Genetics 7: e1001312.

Agustí J, Blázquez MA. 2020. Plant vascular development: mechanisms and

environmental regulation. Cellular and Molecular Life Sciences 77: 3711-

3728.

Alejandro S, Lee Y, Tohge T, Sudre D, Osorio S, Park J, Bovet L, Lee Y,

Geldner N, Fernie AR, Martinoia E. 2012. AtABCG29 is a monolignol

transporter involved in lignin biosynthesis. Current Biology 22: 1207–

1212.

Altamura MM, Possenti M, Matteucci A, Baima S, Ruberti I, Morelli G.

2001. Development of the vascular system in the inflorescence stem of

Arabidopsis. New Phytologist 151: 381–389.

Asturias FJ, Jiang YW, Myers LC, Gustafsson CM, Kornberg RD. 1999.

Conserved structures of mediator and RNA polymerase II holoenzyme.

Science 283: 985–987.

Baghdady A, Blervacq AS, Jouanin L, Grima-Pettenati J, Sivadon P,

Hawkins S. 2006. Eucalyptus gunnii CCR and CAD2 promoters are active

in lignifying cells during primary and secondary xylem formation in

Arabidopsis thaliana. Plant Physiology and Biochemistry 44: 674‐683.

Barra-Jiménez A, Ragni L. 2017. Secondary development in the stem: when

Arabidopsis and trees are closer than it seems. Current Opinion in Plant

Biology 35: 145-151.

Barros J, Serk H, Granlund I, Pesquet E. 2015. The cell biology of

lignification in higher plants. Annals of Botany 115: 1053-1074.

Barros J, Serrani-Yarce JC, Chen F, Baxter D, Venables BJ, Dixon RA.

2016. Role of bifunctional ammonia-lyase in grass cell wall biosynthesis.

Nature Plants 2: 16050.

Baucher M, Halpin C, Petit-Conil M, Boerjan W. 2003. Lignin: genetic

engineering and impact on pulping. Critical Reviews in Biochemistry and

Molecular Biology 38: 305-350.

Behr M, Guerriero G, Grima-Pettenati J, Baucher MA. 2019. Molecular

blueprint of lignin repression. Trends in Plant Science 24: 1052‐1064.

Bhargava A, Mansfield SD, Hall HC, Douglas CJ, Ellis BE. 2010. MYB75

functions in regulation of secondary cell wall formation in the Arabidopsis

inflorescence stem. Plant Physiology 154: 1428–1438.

46

Bhargava A, Ahad A, Wang S, Mansfield SD, Haughn GW, Douglas CJ, Ellis

BE. 2013. The interacting MYB75 and KNAT7 transcription factors

modulate secondary cell wall deposition both in stems and seed coat in

Arabidopsis. Planta 237: 1199–1211.

Bishopp A, Help H, El-Showk S, Weijers D, Scheres B, Friml J, Benkova E,

Mahonen AP, Helariutta Y. 2011A. A mutually inhibitory interaction

between auxin and cytokinin specifies vascular pattern in roots. Current

Biology 21: 917-926.

Bishopp A, Lehesranta S, Vaten A, Help H, El-Showk S, Scheres B,

Helariutta K, Mähönen AP, Sakakibara H, Helariutta Y. 2011B.

Phloem-transported cytokinin regulates polar auxin transport and

maintains vascular pattern in the root meristem. Current Biology 21: 927–

932.

Berthet S, Demont-Caulet N, Pollet B, Bidzinski P, Cézard L, Le Bris P,

Borrega N, Hervé J, Blondet E, Balzergue S, Lapierre C, Jouanin L.

2011. Disruption of LACCASE4 and 17 results in tissue-specific alterations

to lignification of Arabidopsis thaliana stems. Plant Cell 23: 1124-1137.

Blokhina O, Laitinen T, Hatakeyama Y, Delhomme N, Paasela T, Zhao L,

Street NR, Wada H, Kärkönen A, Fagerstedt K. 2019. Ray parenchymal

cells contribute to lignification of tracheids in developing xylem of

Norway spruce. Plant Physiology 181: 1552‐1572.

Boerjan W, Ralph J, Baucher M. 2003. Lignin biosynthesis. Annual Review of

Plant Biology 54: 519-546.

Boija E, Lundquist A, Edwards K, Johansson G. 2007. Evaluation of bilayer

disks as plant cell membrane models in partition studies. Analytical

Biochemistry 364: 145-152.

Boija E, Lundquist A, Nilsson M, Edwards K, Isaksson R, Johansson G.

2008. Bilayer disk capillary electrophoresis: A novel method to study drug

partitioning into membranes. Electrophoresis 29: 3377-3383.

Bollhöner B, Prestele J, Tuominen H. 2012. Xylem cell death: emerging

understanding of regulation and function. Journal of Experimental Botany

63: 1081-1094.

Bonawitz ND, Chapple C. 2010. The genetics of lignin biosynthesis: connecting

genotype to phenotype. Annual Review of Genetics 44: 337–363.

Bonawitz ND, Soltau WL, Blatchley MR, Powers BL, Hurlock AK, Seals LA,

Weng JK, Stout J, Chapple C. 2012. REF4 and RFR1, subunits of the

transcriptional coregulatory complex mediator, are required for

phenylpropanoid homeostasis in Arabidopsis. Journal of Biological

Chemistry 287: 5434–5445.

Bonawitz ND, Kim JI, Tobimatsu Y, Ciesielski PN, Anderson NA, Ximenes

E, Maeda J, Ralph J, Conohoe BS, Ladisch M, Chapple C. 2014.

Disruption of Mediator rescues the stunted growth of a lignin-deficient

Arabidopsis mutant. Nature 509: 376–380.

47

Borchert R, Pockman WT. 2005. Water storage capacitance and xylem tension

in isolated branches of temperate and tropical trees. Tree Physiology 25:

457–466.

Bossinger G, Spokevicius AV. 2018. Sector analysis reveals patterns of

cambium differentiation in poplar tree stems. Journal of Experimental

Botany 69: 4339-4348.

Bowman JL, Kohchi T, Yamato KT, Jenkins J, Shu S, Ishizaki K, Yamaoka

S, Nishihama R, Nakamura Y, Berger F et al. 2017. Insights into land

plant evolution garnered from the Marchantia polymorpha genome. Cell

171: 287-304.

Brackmann K, Qi J, Gebert M, Jouannet V, Schlamp T, Grünwald K,

Wallner ES, Novikova DD, Levitsky VG, Agustí J, Sanchez P,

Lohmann JU, Greb T. 2018. Spatial specificity of auxin responses

coordinates wood formation. Nature Communications 9: 875.

Brady SM, Orlando DA, Lee JY, Wang JY, Koch J, Dinneny JR, Mace D,

Ohler U, Benfey PN. 2007. A high-resolution root spatiotemporal map

reveals dominant expression patterns. Science 318: 801-806.

Bringmann M, Li E, Sampathkumar A, Kocabek T, Hauser MT, Persson S.

2012. POM-POM2/cellulose synthase interacting1 is essential for the

functional association of cellulose synthase and microtubules in

Arabidopsis. Plant Cell 24: 163-177.

Brodersen CR, McElrone A, Choat B, Matthews M, Shackel K. 2010.

Dynamics of embolism repair in grapevine: in vivo visualisations using

HRCT. Plant Physiology 154: 1088–1095.

Brodersen CR, Roddy AB, Wason JW, McElrone AJ. 2019. Functional status

of xylem through time. Annual Review of Plant Biology 70: 407‐433.

Bromley JR, Busse-Wicher M, Tryfona T, Mortimer JC, Zhang Z, Brown

DM, Dupree P. 2013. GUX1 and GUX2 glucuronyltransferases decorate

distinct domains of glucuronoxylan with different substitution patterns.

The Plant Journal 74: 423-434.

Brown DM, Goubet F, Wong VW, Goodacre R, Stephens E, Dupree P,

Turner SR. 2007. Comparison of five xylan synthesis mutants reveals new

insight into the mechanisms of xylan synthesis. The Plant Journal 52:

1154-1168

Brown D, Wightman R, Zhang Z, Gomez LD, Atanassov I, Bukowski JP,

Tryfona T, McQueen‐Mason SJ, Dupree P, Turner S. 2011.

Arabidopsis genes IRREGULAR XYLEM IRX15 and IRX15L encode

DUF579-containing proteins that are essential for normal xylan deposition

in the secondary cell wall. The Plant Journal 66: 401–413.

Campbell MM, Sederoff RR. 1996. Variation in lignin content and composition

mechanisms of control and implications for the genetic improvement of

plants. Plant Physiology 110: 3–13.

48

Campilho A, Nieminen K, Ragni L. 2020. The development of the periderm:

the final frontier between a plant and its environment. Current Opinion in

Plant Biology 53: 10-14.

Cano-Delgado A, Penfield S, Smith C, Catley M, Bevan M. 2003. Reduced

cellulose synthesis invokes lignification and defense responses in

Arabidopsis thaliana. The Plant Journal 34: 351–362.

Cao Y, Dai Y, Cui S, Ma L. 2008. Histone H2B monoubiquitination in the

chromatin of FLOWERING LOCUS C regulates flowering time in

Arabidopsis. Plant Cell 20: 2586-2602.

Carlquist S. 1988. Axial Parenchyma. In: Comparative Wood Anatomy.

Springer Series in Wood Science. Springer, Berlin, Heidelberg.

Carlquist S. 2013. Comparative Wood Anatomy: Systematic, Ecological, and

Evolutionary Aspects of Dicotyledon Wood. Berlin: Springer-Verlag.

Carmona C, Langan P, Smith JC, Petridis L. 2015. Why genetic modification

of lignin leads to low-recalcitrance biomass. Physical Chemistry Chemical

Physics 17: 358-364.

Carpita N, Tierney M, Campbell M. 2001. Molecular biology of the plant cell

wall: searching for the genes that define structure, architecture and

dynamics. Plant Molecular Biology 47: 1-5.

Cassan-Wang H, Goué N, Saidi MN, Legay S, Sivadon P, Goffner D, Grima-

Pettenati J. 2013. Identification of novel transcription factors regulating

secondary cell wall formation in Arabidopsis. Frontiers in Plant Science 4:

189.

Chaffey N, Cholewa E, Regan S, Sundberg B. 2002. Secondary xylem

development in Arabidopsis: a model for wood formation. Physiologia

Plantarum 114: 594–600.

Chaffey N, Barlow P. 2001. The cytoskeleton facilitates a three-dimensional

symplasmic continuum in the long-lived ray and axial parenchyma cells of

angiosperm trees. Planta 213: 811–823.

Chan J, Coen E. 2020. Interaction between autonomous and microtubule

guidance systems controls cellulose synthase trajectories. Current Biology

30: 941-947.

Chen C, Meyermans H, Burggraeve B, De Rycke RM, Inoue K, De

Vleesschauwer V, Steenackers M, Van Montagu MC, Engler GJ,

Boerjan WA. 2000. Cell-specific and conditional expression of caffeoyl-

coenzyme A-3-O-methyltransferase in poplar. Plant Physiology 123: 853–

868.

Chen HC, Li Q, Shuford CM, Liu J, Muddiman DC, Sederoff RR, Chiang

VL. 2011. Membrane protein complexes catalyze both 4-and 3-

hydroxylation of cinnamic acid derivatives in monolignol biosynthesis.

Proceedings of the National Academy of Sciences, USA 108: 21253-

21258.

Chen HC, Song J, Wang JP, Lin YC, Ducoste J, Shuford CM, Liu J, Li Q,

Shi R, Nepomuceno A, Isik F, Muddiman DC, Williams C, Sederoff

49

RR, Chiang VL. 2014. Systems biology of lignin biosynthesis in Populus

trichocarpa: heteromeric 4-coumaric acid:coenzyme A ligase protein

complex formation, regulation, and numerical modeling. Plant Cell 26:

876-893.

Choat B, Cobb AR, Jansen S. 2008. Structure and function of bordered pits:

new discoveries and impacts on whole-plant hydraulic function. New

Phytologist 177: 608–626.

Chou EY, Schuetz M, Hoffmann N, Watanabe Y, Sibout R, Samuels AL.

2018. Distribution, mobility, and anchoring of lignin-related oxidative

enzymes in Arabidopsis secondary cell walls. Journal of Experimental

Botany 69: 1849-1859.

Courtois-Moreau CL, Pesquet E, Sjodin A, Muniz L, Bollhoner B, Kaneda

M, Samuels L, Jansson S, Tuominen H. 2009. A unique program for cell

death in xylem fibers of Populus stem. The Plant Journal 58: 260-274.

D'Autréaux B, Toledano MB. 2007. ROS as signaling molecules: mechanisms

that generate specificity in ROS homeostasis. Nature Reviews Molecular

Cell Biology 8: 813–824.

Davin LB, Lewis NG. 2000. Dirigent proteins and dirigent sites explain the

mystery of specificity of radical precursor coupling in lignan and lignin

biosynthesis. Plant Physiology 123: 453-461.

Dechend R, Hirano F, Lehmann K, Heissmeyer V, Ansieau S, Wulczyn FG,

Scheidereit C, Leutz A. 1999. The Bcl-3 oncoprotein acts as a bridging

factor between NF-kappaB/Rel and nuclear co-regulators. Oncogene 18:

3316-3323.

Del Pozo JC, Diaz-Trivino S, Cisneros N, Gutierrez C. 2007. The E2FC-DPB

transcription factor controls cell division, endoreplication and lateral root

formation in a SCF-dependent manner. Plant Signaling and Behavior 2:

273–274.

De Meester B, de Vries L, Özparpucu M, Gierlinger N, Corneillie S, Pallidis

A, Goeminne G, Morreel K, De Bruyne M, De Rycke, Vanholme R,

Boerjan W. 2018. Vessel‐specific reintroduction of CINNAMOYL‐COA

REDUCTASE1 (CCR1) in dwarfed ccr1 mutants restores vessel and

xylary fiber integrity and increases biomass. Plant Physiology 176: 611-

633.

Demont-Caulet N, Lapierre C, Jouanin L, Baumberger S, Méchin V. 2010.

Arabidopsis peroxidase-catalyzed copolymerization of coniferyl and

sinapyl alcohols: kinetics of an endwise process. Phytochemistry 71: 1673-

1683.

De Rybel B, Moller B, Yoshida S, Grabowicz I, Barbier de Reuille P, Boeren

S, Smith RS, Borst JW, Weijers D. 2013. A bHLH complex controls

embryonic vascular tissue establishment and indeterminate growth in

Arabidopsis. Developmental Cell 24: 426–437.

Derbyshire P, Ménard D, Green P, Saalbach G, Buschmann H, Lloyd CW,

Pesquet E. 2015. Proteomic analysis of microtubule interacting proteins

50

over the course of xylem tracheary element formation in Arabidopsis. Plant

Cell 27: 2709–2726.

De Souza Lima MM, Borsali R. 2004. Rodlike cellulose microcrystals:

structure, properties, and applications. Macromolecular Rapid

Communications 25: 771–787.

Dolan WL, Dilkes BP, Stout JM, Bonawitz ND, Chapple C. 2017. Mediator

complex subunits MED2, MED5, MED16, and MED23 genetically

interact in the regulation of phenylpropanoid biosynthesis. Plant Cell 29:

3269–3285.

Donaldson LA. 2001. Lignification and lignin topochemistry – an ultrastructural

view. Phytochemistry 57: 859–873.

Edwards KD, Takata N, Johansson M, Jurca M, Novák O, Hényková E,

Liverani S, Kozarewa I, Strnad M, Millar AJ, Ljung K, Eriksson ME.

2018. Circadian clock components control daily growth activities by

modulating cytokinin levels and cell division‐associated gene expression

in Populus trees. Plant, Cell & Environment 41: 1468-1482.

Ehlting J, Mattheus N, Aeschliman DS, Li E, Hamberger B, Cullis IF,

Zhuang J, Kaneda M, Mansfield SD, Samuels L, Ritland K, Ellis BE,

Bohlmann J, Douglas CJ. 2005. Global transcript profiling of primary

stems from Arabidopsis thaliana identifies candidate genes for missing

links in lignin biosynthesis and transcriptional regulators of fiber

differentiation. The Plant Journal 42: 618‐640.

Endo S, Pesquet E, Yamaguchi M, Tashiro G, Sato M, Toyooka K,

Nishikubo N, Udagawa-Motose M, Kubo M, Fukuda H, Demura T.

2009. Identifying new components participating in the secondary cell wall

formation of vessel elements in Zinnia and Arabidopsis. Plant Cell 21:

1155–1165.

Etchells JP, Turner SR. 2010. The PXY-CLE41 receptor ligand pair defines a

multifunctional pathway that controls the rate and orientation of vascular

cell division. Development 137: 767‐774.

Etchells JP, Provost CM, Mishra L, Turner SR. 2013. WOX4 and WOX14 act

downstream of the PXY receptor kinase to regulate plant vascular

proliferation independently of any role in vascular organisation.

Development 140: 2224-2234.

Feng H, Li X, Chen H, Deng J, Zhang C, Liu J, Wang T, Zhang X, Dong J.

2018. GhHUB2, a ubiquitin ligase, is involved in cotton fiber development

via the ubiquitin–26S proteasome pathway. Journal of Experimental

Botany 21: 5059–5075.

Fernández-Pérez F, Pomar F, Pedreño MA, Novo-Uzal E. 2015. Suppression

of Arabidopsis peroxidase 72 alters cell wall and phenylpropanoid

metabolism. Plant Science 239: 192-199.

Fischer U, Kucukoglu M, Helariutta Y, Bhalerao RP. 2019. The dynamics of

cambial stem cell activity. Annual Reviews of Plant Biology 70: 293-329.

51

Food and Agriculture Organization of the United Nations. 2014. State of the

World’s Forests 2014: Enhancing the socioeconomic benefits from forests.

Rome. available at http://www.fao.org/3/a-i3710e.pdf.

Food and Agriculture Organization of the United Nations. 2017A.

Sustainable woodfuel for food security. FAO Working Paper. Rome.

available at http://www.fao.org/3/a-i7917e.pdf.

Food and Agriculture Organization of the United Nations. 2017B. Forests and

energy. FAO. Rome. available at http://www.fao.org/3/a-i6928e.pdf.

Food and Agriculture Organization of the United Nations. 2018. The State of

the World’s Forests 2018 - Forest pathways to sustainable development.

Rome.

Fosket DE. 1994. Introduction, Plant growth and development. Academic Press

1-40.

Franke R, McMichael CM, Meyer K, Shirley AM, Cusumano JC, Chapple

C. 2000. Modified lignin in tobacco and poplar plants over-expressing the

Arabidopsis gene encoding ferulate 5-hydroxylase. The Plant Journal 22:

223–234.

Franke R, Hemm MR, Denault JW, Ruegger MO, Humphreys JM, Chapple

C. 2002. Changes in secondary metabolism and deposition of an unusual

lignin in the ref8 mutant of Arabidopsis. The Plant Journal 30: 47-59.

Fry SC. 2004. Primary cell wall metabolism: tracking the careers of wall

polymers in living plant cells. New Phytologist 161: 641-675.

Fujita M, Harada H. 1979. Autoradiographic investigations of cell wall

development. II. Tritiated phenylalanine and ferulic acid assimilation in

relation to lignification. Mokuzai Gakkaishi 25: 89-94.

Fukuda H. 1996. Xylogenesis: initiation, progression and cell death. Annual

Review of Plant Physiology and Plant Molecular Biology 47: 299-325.

Fukushima K, Terashima N. 1990. Heterogeneity in formation of lignin XIII.

Formation of p-hydroxyphenyl lignin in various hardwoods visualized by

microautoradiography. Journal of Wood Chemistry and Technology 10:

413-433.

Fukushima K, Terashima N. 1991. Heterogeneity in formation of lignin XIV.

Formation and structure of lignin in differentiating xylem of Ginkgo

biloba. Holzforschung 45: 87-94.

Gansert D. 2003. Xylem sap flow as a major pathway for oxygen supply to the

sapwood of birch (Betula pubescens Ehr.). Plant, Cell and Environment

26: 1803–1814.

Geng P, Zhang S, Liu J, Zhao C, Wu J, Cao Y, Fu C, Han X, He H, Zhao Q.

2020. MYB20, MYB42, MYB43, and MYB85 regulate phenylalanine and

lignin biosynthesis during secondary cell Wall formation. Plant Physiology

182:1272-1283.

Gómez Ros LV, Paradiso A, Gabaldón C, Pedreño MA, De Gara L, Ros

Barceló A. 2006. Two distinct cell sources of H2O2 in the lignifying Zinnia

elegans cell culture system. Protoplasma 227: 175–183.

52

Gonneau M, Desprez T, Guillot A, Vernhettes S, Höfte H. 2014. Catalytic

subunit stoichiometry within the cellulose synthase complex. Plant

Physiology 166: 1709–1712.

Gorzsás A, Stenlund H, Persson P, Trygg J, Sundberg B. 2011. Cell-specific

chemotyping and multivariate imaging by combined FT-IR

microspectroscopy and orthogonal projections to latent structures OPLS

analysis reveals the chemical landscape of secondary xylem. The Plant

Journal 66: 903–914.

Gorzsás A. 2017. Chemical imaging of xylem by Raman microspectroscopy. In:

M Lucas, JP Etchells, eds. Xylem: methods and protocols. New York, NY,

USA: Springer, 133– 178.

Gou M, Ran X, Martin DW, Liu CJ. 2018. The scaffold proteins of lignin

biosynthetic cytochrome P450 enzymes. Nature Plants 4: 299-310.

Grantham NJ, Wurman-Rodrich J, Terrett OM, Lyczakowski JJ, Stott K,

Iuga D, Simmons TJ, Durand-Tardif M, Brown SP, Dupree R, Busse-

Wicher M, Dupree P. 2017. An even pattern of xylan substitution is

critical for interaction with cellulose in plant cell walls. Nature Plants 3:

859–865.

Gursanscky NR, Jouannet V, Grünwald K, Sanchez P, Laaber-Schwarz M,

Greb T. 2016. MOL1 is required for cambium homeostasis in Arabidopsis.

The Plant Journal 86: 210-220.

Hacke U, Sperry J, Feild T, Sano Y, Sikkema E, Pittermann J. 2007. Water

transport in vesselless angiosperms: conducting efficiency and cavitation

safety. International Journal of Plant Sciences 168: 1113–1126.

Han S, Cho H, Noh J, Qi J, Jung HJ, Nam H, Lee S, Hwang D, Greb T,

Hwang I. 2018. BIL1-mediated MP phosphorylation integrates PXY and

cytokinin signalling in secondary growth. Nature Plants 4: 605‐614.

Harmer SL, Hogenesch JB, Straume M, Chang HS, Han B, Zhu T, Wang X,

Kreps JA, Kay SA. 2000. Orchestrated transcription of key pathways in

Arabidopsis by the circadian clock. Science 290: 2110-2113.

Herrero J, Fernández-Pérez F, Yebra T, Novo-Uzal E, Pomar F, Pedreño

MÁ, Cuello J, Guéra A, Esteban-Carrasco A, Zapata JM. 2013.

Bioinformatic and functional characterization of the basic peroxidase 72

from Arabidopsis thaliana involved in lignin biosynthesis. Planta 237:

1599-1612.

Hill Jr JL, Hammudi MB, Tien M. 2014. The Arabidopsis cellulose synthase

complex: A proposed hexamer of CESA trimers in an equimolar

stoichiometry. Plant Cell 26: 4834–4842.

Himanen K, Woloszynska M, Boccardi TM, De Groeve S, Nelissen H, Bruno

L, Vuylsteke M, Van Lijsebettens M. 2012. Histone H2B

monoubiquitination is required to reach maximal transcript levels of

circadian clock genes in Arabidopsis. The Plant Journal 72: 249-260.

Hirakawa Y, Shinohara H, Kondo Y, Inoue A, Nakanomyo I, Ogawa M,

Sawa S, Ohashi-Ito K, Matsubayashi Y, Fukuda H. 2008. Non-cell-

53

autonomous control of vascular stem cell fate by a CLE peptide/receptor

system. Proceedings of the National Academy of Sciences, USA 105:

15208‐15213.

Hirakawa Y, Kondo Y, Fukuda H. 2010. TDIF peptide signaling regulates

vascular stem cell proliferation via the WOX4 homeobox gene in

Arabidopsis. Plant Cell 22: 2618-2629.

Hoffmann N, Benske A, Betz H, Schuetz M, Samuels AL. 2020. Laccases and

peroxidases co-localize in lignified secondary cell walls throughout stem

development. Plant Physiology 184: 806-822.

Holbrook N. 1995. Stem water storage. In: Gartner BL, ed. Stems and trunks in

plant form and function. San Diego, CA, USA: Academic Press, 151–174.

Höll W. 2000. Distribution, fluctuation and metabolism of food reserves in the

wood of trees. In: Savidge R, Barnett J, Napier R, editors, Cell and

molecular biology of wood formation. Oxford: BIOS Scientific Publishers,

347–362.

Hosmani PS, Kamiya T, Danku J, Naseer S, Geldner N, Guerinot ML, Salt

DE. 2013. Dirigent domain-containing protein is part of the machinery

required for formation of the lignin-based Casparian strip in the root.

Proceedings of the National Academy of Sciences, USA 110: 14498–

14503.

Hosokawa M, Suzuki S, Umezawa T, Sato Y. 2001. Progress of lignification

mediated by intercellular transportation of monolignols during tracheary

element differentiation of isolated Zinnia mesophyll cells. Plant and Cell

Physiology 42: 959–968.

Hu WJ, Kawaoka A, Tsai CJ, Lung J, Osakabe K, Ebinuma H, Chiang VL.

1998. Compartmentalized expression of two structurally and functionally

distinct 4-coumarate:CoA ligase genes in aspen (Populus tremuloides).

Proceedings of the National Academy of Sciences, USA 95: 5407–5412.

Huntley SK, Ellis D, Gilbert M, Chapple C, Mansfield SD. 2003. Significant

increases in pulping efficiency in C4H-F5H-transformed poplars:

Improved chemical savings and reduced environmental toxins. Journal of

Agricultural and Food Chemistry 51: 6178–6183.

Hussey SG, Mizrachi E, Groover A, Berger DK, Myburg AA. 2015. Genome-

wide mapping of histone H3 lysine 4 trimethylation in Eucalyptus grandis

developing xylem. BMC Plant Biology 15: 117.

Hussey SG, Loots MT, van der Merwe K, Mizrachi E, Myburg AA. 2017.

Integrated analysis and transcript abundance modelling of H3K4me3 and

H3K27me3 in developing secondary xylem. Scientific Reports 7: 3370.

IAWA Committee. 1989. IAWA list of microscopic features for hardwood

identification. International Association of Wood Anatomists Bulletin

New Series 10: 219–332.

IAWA Committee. 2004. IAWA list of microscopic features for softwood

identification. International Association of Wood Anatomists Journal 25:

1–70.

54

Jaini R, Wang P, Dudareva N, Chapple C, Morgan JA. 2017. Targeted

metabolomics of the phenylpropanoid pathway in Arabidopsis thaliana

using reversed phase liquid chromatography coupled with tandem mass

spectrometry. Phytochemical Analysis 28: 267–276.

Jansen S, Gortan E, Lens F, Lo Gullo MA, Salleo S, Scholz A, Stein A, Trifilo

P, Nardini A. 2011. Do quantitative vessel and pit characters account for

ion-mediated changes in the hydraulic conductance of angiosperm xylem?

New Phytologist 189: 218–228.

Jensen JK, Kim H, Cocuron JC, Orler R, Ralph J, Wilkerson CG. 2011. The

DUF579 domain containing proteins IRX15 and IRX15-L affect xylan

synthesis in Arabidopsis. The Plant Journal 66: 387–400.

Kaneda M, Rensing KH, Wong JC, Banno B, Mansfield SD, Samuels AL.

2008. Tracking monolignols during wood development in lodgepole pine.

Plant Physiology 147: 1750-1760.

Kaneda M, Schuetz M, Lin BSP, Chanis C, Hamberger B, Western TL,

Ehlting J, Samuels AL. 2011. ABC transporters coordinately expressed

during lignification of Arabidopsis stems include a set of ABCBs

associated with auxin transport. Journal of Experimental Botany 62: 2063–

2077.

Karpinska B, Karlsson M, Schinkel H, Streller S, Süss KH, Melzer M,

Wingsle G. 2001. A novel superoxide dismutase with a high isoelectric

point in higher plants. Expression, regulation, and protein localization.

Plant Physiology 126: 1668–1677.

Katayama H, Iwamoto K, Kariya Y, Asakawa T, Kan T, Fukuda H, Ohashi-

Ito K. 2015. A negative feedback loop controlling bHLH complexes is

involved in vascular cell division and differentiation in the root apical

meristem. Current Biology 25: 3144–3150.

Kim SH, Lee CM, Kafle K. 2013. Characterization of crystalline cellulose in

biomass: basic principles, applications, and limitations of XRD, NMR, IR,

Raman, and SFG. Korean Journal of Chemical Engineering 30: 2127–

2141.

Kitin P, Voelker SL, Meinzer FC, Beeckman H, Strauss SH, Lachenbruch

B. 2010. Tyloses and phenolic deposits in xylem vessels impede water

transport in low-lignin transgenic poplars: a study by cryo-fluorescence

microscopy. Plant Physiology 154: 887-898.

Kondo Y, Ito T, Nakagami H, Hirakawa Y, Saito M, Tamaki T, Shirasu K,

Fukuda H. 2014. Plant GSK3 proteins regulate xylem cell differentiation

downstream of TDIF-TDR signalling. Nature Communications 5: 3504.

Krämer U. 2015. The Natural History of Model Organisms: Planting molecular

functions in an ecological context with Arabidopsis thaliana. eLife 4:

e06100.

Kubo M, Udagawa M, Nishikubo N, Horiguchi G, Yamaguchi M, Ito J,

Mimura T, Fukuda H, Demura T. 2005. Transcription switches for

55

protoxylem and metaxylem vessel formation. Genes & Development 19:

1855–1860.

Kumar M, Campbell L, Turner S. 2016. Secondary cell walls: biosynthesis and

manipulation. Journal of Experimental Botany 67: 515-531.

Lacombe E, Van Doorsselaere J, Boerjan W, Boudet AM, Grima-Pettenati

J. 2000. Characterization of cis-elements required for vascular expression

of the cinnamoyl CoA reductase gene and for protein-DNA complex

formation. The Plant Journal 23: 663‐676.

Laibach, F. 1907. Zur frage nach der individualität der chromosomen im

plfanzenreich. Beih. Bot. Zentralbl. 22: 191–210.

Lan W, Lu F, Regner M, Zhu Y, Rencoret J, Ralph SA, Zakai UI, Morreel

K, Boerjan W, Ralph J. 2015. Tricin, a flavonoid monomer in monocot

lignification. Plant Physiology 167: 1284-1295.

Lauvergeat V, Rech P, Jauneau A, Guez C, Coutos-Thevenot P, Grima-

Pettenati J. 2002. The vascular expression pattern directed by the

Eucalyptus gunnii cinnamyl alcohol dehydrogenase EgCAD2 promoter is

conserved among woody and herbaceous plant species. Plant Molecular

Biology 50: 497‐509.

Lebovka I, Mele BH, Zakieva A, Gursanscky N, Merks R, Greb T. 2020.

Computational modelling of cambium activity provides a regulatory

framework for simulating radial plant growth. Biorxiv preprint 908715.

Lee C, O’Neill MA, Tsumuraya Y, Darvill AG, Ye ZH. 2007. The irregular

xylem9 mutant is deficient in xylan xylosyltransferase activity. Plant and

Cell Physiology 48: 1624-1634.

Lee Y, Rubio MC, Alassimone J, Geldner N. 2013. A mechanism for localized

lignin deposition in the endodermis. Cell 153: 402–412.

Lens F, Tixier A, Cochard H, Sperry JS, Jansen S, Herbette S. 2013.

Embolism resistance as a key mechanism to understand adaptive plant

strategies. Current Opinion in Plant Biology 16: 287-292.

Leplé JC, Dauwe R, Morreel K, Storme V, Lapierre C, Pollet B, Naumann

A, Kang KY, Kim H, Ruel K, Lefèbvre A, Joseleau JP, Grima-

Pettenati J, De Rycke R, Andersson-Gunnerås S, Erban A, Fehrle I,

Petit-Conil M, Kopka J, Polle A, Messens E, Sundberg B, Mansfield

SD, Ralph J, Pilate G, Boerjan W. 2007. Downregulation of cinnamoyl-

coenzyme A reductase in poplar: multiple-level phenotyping reveals

effects on cell wall polymer metabolism and structure. Plant Cell 19: 3669‐

3691.

Lewis AM, Boose ER. 1995. Estimating volume flow rates through xylem

conduits. American Journal of Botany 82: 1112–1116.

Li S, Lei L, Somerville CR, Gu Y. 2012. Cellulose synthase interactive protein

1 CSI1 links microtubules and cellulose synthase complexes. Proceedings

of the National Academy of Sciences, USA 109: 185-190.

56

Liang M, Davis E, Gardner D, Cai X, Wu Y. 2006. Involvement of AtLAC15

in lignin synthesis in seeds and in root elongation of Arabidopsis. Planta

224: 1185-1196.

Liao Y, Koelewijn SF, Van den Bossche G, Van Aelst J, Van den Bosch S,

Renders T, Navare K, Nicolaï T, Van Aelst K, Maesen M, Matsushima

H, Thevelein JM, Van Acker K, Lagrain B, Verboekend D, Sels BF.

2020. A sustainable wood biorefinery for low–carbon footprint chemicals

production. Science 367: 1385-1390.

Liochev SI, Fridovich I. 1994. The role of O2.- in the production of HO.: in vitro

and in vivo. Free Radical Biology and Medicine 16: 29–33.

Lion C, Simon C, Huss B, Blervacq AS, Tirot L, Toybou D, Spriet C,

Slomianny C, Guerardel Y, Hawkins S, Biot C. 2017. BLISS: a

bioorthogonal dual-labeling strategy to unravel lignification dynamics in

plants. Cell Chemical Biology 24: 326-338.

Liu L, Shang-Guan K, Zhang B, Liu X, Yan M, Zhang L, Shi Y, Zhang M,

Qian Q, Li J, Zhou Y. 2013A. Brittle culm1, a COBRA-like protein,

functions in cellulose assembly through binding cellulose microfibrils.

PLoS Genetics 9: e1003704.

Liu F, Rehmani I, Esaki S, Fu R, Chen L, de Serrano V, Liu A. 2013B. Pirin

is an iron‐dependent redox regulator of NF‐κB. Proceedings of the

National Academy of Sciences, USA 110: 9722-9727.

Liu Y, Li X, Li K, Liu H, Lin C. 2013C. Multiple bHLH proteins form

heterodimers to mediate CRY2-dependent regulation of flowering-time in

Arabidopsis. PLOS Genetics 9: e1003861.

Lyczakowski JJ, Wicher KB, Terrett OM, Blanc NF, Yu X, Brown D, Krogh

KBRM, Dupree P, Wicher MB. 2017. Removal of glucuronic acid from

xylan is a strategy to improve the conversion of plant biomass to sugars for

bioenergy. Biotechnology for Biofuels 10: 224.

Mahboubi A, Linden P, Hedenström M, Moritz T, Niittylä T. 2015. 13C

tracking after 13CO2 supply revealed diurnal patterns of wood formation in

aspen. Plant Physiology 168: 478-489.

Mahon EL, Mansfield S. 2019. Tailor-made trees: engineering lignin for ease

of processing and tomorrow’s bioeconomy. Current Opinion in

Biotechnology 56: 147-155.

Mähönen AP, Bishopp A, Higuchi M, Nieminen KM, Kinoshita K,

Tormakangas K, Ikeda Y, Oka A, Kakimoto T, Helariutta Y. 2006.

Cytokinin signaling and its inhibitor AHP6 regulate cell fate during

vascular development. Science 311: 94–98.

Mähönen AP. 2019. High levels of auxin signaling define the stem-cell organizer

of the vascular cambium. Nature 565: 485-489.

Maloney VJ, Mansfield SD. 2010. Characterization and varied expression of a

membrane-bound endo-β-1,4-glucanase in hybrid poplar. Plant

Biotechnology Journal 8: 294–307.

57

Manabe Y, Verhertbruggen Y, Gille S, Harholt J, Chong SL, Mohan-

Anupama Pawar P, Mellerowicz EJ, Tenkanen M, Cheng K, Pauly M,

Scheller HV. 2013. Reduced wall acetylation proteins play vital and

distinct roles in cell wall O-acetylation in Arabidopsis. Plant Physiology

163: 1107-1117.

Mansfield SD. 2009. Solutions for dissolution – engineering cell walls for

deconstruction. Current Opinion in Biotechnology 20: 286-294.

Mao X, Kim JI, Wheeler MT, Heintzelman AK, Weake VM, Chapple C.

2019A. Mutation of Mediator subunit CDK8 counteracts the stunted

growth and salicylic acid hyper‐accumulation phenotypes of an

Arabidopsis MED5 mutant. New Phytologist 223: 233–245.

Mao X, Weake VM, Chapple C. 2019B. Mediator function in plant metabolism

revealed by large-scale biology. Journal of Experimental Botany 70: 5995-

6003.

Mazur E, Kurczynska EU. 2012. Rays, intrusive growth, and storied cambium

in the inflorescence stems of Arabidopsis thaliana (L.) Heynh.

Protoplasma 249: 217-220.

Mayer KF, Schoof H, Haecker A, Lenhard M, Jürgens G, Laux T. 1998. Role

of WUSCHEL in regulating stem cell fate in the Arabidopsis shoot

meristem. Cell 95: 805‐815.

Meents MJ, Watanabe Y, Samuels AL. 2018. The cell biology of secondary

cell wall biosynthesis. Annals of Botany 121: 1107-1125.

Mele G, Ori N, Sato Y, Hake S. 2003. The knotted1-like homeobox gene

BREVIPEDICELLUS regulates cell differentiation by modulating

metabolic pathways. Genes & Development 17: 2088–2093.

Mellor N, Adibi M, El-Showk S, De Rybel B, King J, Mahonen AP, Weijers

D, Bishopp A. 2017. Theoretical approaches to understanding root

vascular patterning: a consensus between recent models. Journal of

Experimental Botany 68: 5-16.

Ménard D, Pesquet E. 2015. Cellular interactions during tracheary elements

formation and function. Current Opinion in Plant Biology 23: 109-115.

Miao YC, Liu CJ. 2010. ATP-binding cassette-like transporters are involved in

the transport of lignin precursors across plasma and vacuolar membranes.

Proceedings of the National Academy of Sciences, USA 107: 22728-

22733.

Miedes E, Vanholme R, Boerjan W, Molina A. 2014. The role of the secondary

cell wall in plant resistance to pathogens. Frontiers in Plant Science 5: 358.

Mitsuda N, Seki M, Shinozaki K, Ohme-Takagi M. 2005. The NAC

transcription factors NST1 and NST2 of Arabidopsis regulate secondary

wall thickenings and are required for anther dehiscence. Plant Cell 17:

2993-3006.

Mitsuda N, Iwase A, Yamamoto H, Yoshida M, Seki M, Shinozaki K, Ohme-

Takagi M. 2007. NAC transcription factors, NST1 and NST3, are key

58

regulators of the formation of secondary walls in woody tissues of

Arabidopsis. Plant Cell 19: 270-280.

Miyashima S, Roszak P, Sevilem I, Toyokura K, Blob B, Heo JO, Mellor N,

Help-Rinta-Rahko H, Otero S, Smet W, Boekschoten M, Hooiveld G,

Hashimoto K, Smetana O, Siligato R, Wallner ES, Mähönen AP,

Kondo Y, Melnyk CW, Greb T, Nakajima K, Sozzani R, Bishopp A,

De Rybel B, Helariutta Y. 2019. Mobile PEAR transcription factors

integrate positional cues to prime cambial growth. Nature 565: 490-494.

Morris H, Plavcová L, Cvecko P, Fichtler E, Gillingham MAF, Martínez-

Cabrera HI, McGlinn DJ, Wheeler E, Zheng J, Ziemińska K, Jansen

S. 2016. A global analysis of parenchyma tissue fractions in secondary

xylem of seed plants. New Phytologist 209: 1553-1565.

Moura JCMS, Bonine CAV, De Oliveira Fernandes Viana J, Dornelas MC,

Mazzafera P. 2010. Abiotic and biotic stresses and changes in the lignin

content and composition in plants. Journal of Integrative Plant Biology 52:

360–376.

Myburg AA, Lev-Yadun S, Sederoff RR. 2013. Xylem structure and function.

In: eLS. John Wiley & Sons Ltd, Chichester.

Nardini A, Salleo S, Jansen S. 2011. More than just a vulnerable pipeline: xylem

physiology in the light of ion-mediated regulation of plant water transport.

Journal of Experimental Botany 62: 4701-4718.

Nguyen HTK, Hyoung S, Him HJ, Cho KM, Shin JS. 2019. The transcription

factor γMYB2 acts as a negative regulator of secondary cell wall thickening

in anther and stem. Gene 702: 158–165.

Novaes E, Kirst M, Chiang V, Winter-Sederoff H, Sederoff R. 2010. Lignin

and biomass: a negative correlation for wood formation and lignin content

in trees. Plant Physiology 154: 555-561.

Ogawa KI, Kanematsu S, Asada K. 1997. Generation of superoxide anion and

localization of CuZn-superoxide dismutase in the vascular tissue of

spinach hypocotyls: their association with lignification. Plant and Cell

Physiology 38: 1118–1126.

Ohashi-Ito K, Saegusa M, Iwamoto K, Oda Y, Katayama H, Kojima M,

Sakakibara H, Fukuda H. 2014. A bHLH complex activates vascular cell

division via cytokinin action in root apical meristem. Current Biology 24:

2053–2058.

Ohashi-Ito K, Iwamoto K, Nagashima Y, Kojima M, Sakakibara H, Fukuda

H. 2019. A positive feedback loop comprising LHW–TMO5 and local

auxin biosynthesis regulates initial vascular development in Arabidopsis

roots. Plant and Cell Physiology 60: 2684–2691.

Öhman D, Demedts B, Kumar M, Gerber L, Gorzsás A, Goeminne G,

Hedenström M, Ellis B, Boerjan W, Sundberg B. 2013. MYB103 is

required for FERULATE-5-HYDROXYLASE expression and syringyl

lignin biosynthesis in Arabidopsis stems. The Plant Journal 73: 63‐76.

59

Ohtani M, Demura T. 2019. The quest for transcriptional hubs of lignin

biosynthesis: beyond the NAC-MYB-gene regulatory network model.

Current Opinion in Biotechnology 56: 82-87.

Oikawa A, Joshi HJ, Rennie EA, Ebert B, Manisseri C, Heazlewood JL,

Scheller HV. 2010. An integrative approach to the identification of

Arabidopsis and rice genes involved in xylan and secondary wall

development. PLoS One 5: e15481.

Önnerud H, Zhang L, Gellerstedt G, Henriksson G. 2002. Polymerization of

monolignols by redox shuttle-mediated enzymatic oxidation: a new model

in lignin biosynthesis I. Plant Cell 14: 1953-1962.

Østergaard L, Teilum K, Mirza O, Mattsson O, Petersen M, Welinder KG,

Mundy J, Gajhede M, Henriksen A. 2000. Arabidopsis ATP A2

peroxidase. Expression and high-resolution structure of a plant peroxidase

with implications for lignification. Plant Molecular Biology 44: 231-243.

Oyarce P, De Meester B, Fonseca F, De Vries L, Goeminne G, Pallidis A, De

Rycke R, Tsuji Y, Li Y, Van den Bosch S, Sels B, Ralph J, Vanholme

R, Boerjan W. 2019. Introducing curcumin biosynthesis in Arabidopsis

enhances lignocellulosic biomass processing. Nature Plants 5: 225–237.

Panshin AJ, de Zeeuw C. 1970. Textbook of wood technology, Vol. I: Structure,

identification, uses, and properties of the commercial woods of the United

States and Canada. New York: McGraw-Hill. 3rd ed.

Panshin AJ, de Zeeuw C. 1980. Textbook of wood technology. Part 1.

Formation, anatomy, and properties of wood. New York: McGraw-Hill.

Paredez AR, Somerville CR, Ehrhardt DW. 2006. Visualization of cellulose

synthase demonstrates functional association with microtubules. Science

312: 1491-1495.

Paris Agreement to the United Nations Framework Convention on Climate

Change. 2015. T.I.A.S. No. 16-1104.

Perkins M, Smith RA, Samuels L. 2019. The transport of monomers during

lignification in plants: anything goes but how? Current Opinion in

Biotechnology 56: 69-74.

Persson S, Wei H, Milne J, Page GP, Somerville CR. 2005. Identification of

genes required for cellulose synthesis by regression analysis of public

microarray data sets. Proceedings of the National Academy of Sciences,

USA 102: 8633–8638.

Pesquet E, Zhang B, Gorzsás A, Puhakainen T, Serk H, Escamez S, Barbier

O, Gerber L, Courtois-Moreau C, Alatalo E, Paulin L, Kangasjärvi J,

Sundberg B, Goffner D, Tuominen H. 2013. Non-cell-autonomous

postmortem lignification of tracheary elements in Zinnia elegans. Plant

Cell 25: 1314-1328.

Pickett-Heaps JD. 1968. Xylem wall deposition: Radioautographic

investigations using lignin precursors. Protoplasma 65: 181-205.

Piquemal J, Lapierre C, Myton K, O’Connell A, Schuch W, Grima-Pettenati

J, Boudet AM. 1998. Down-regulation of cinnamoyl-CoA reductase

60

induces significant changes of lignin profiles in transgenic tobacco plants.

The Plant Journal 13: 71-83.

Pittermann J. 2010. The evolution of water transport in plants: an integrated

approach. Geobiology 8: 112–139.

Plavcová L, Jansen S. 2015. The role of xylem parenchyma in the storage and

utilization of non-structural carbohydrates. In: Hacke UG, ed. Functional

and ecological xylem anatomy. Heidelberg, Germany: Springer

International, 209–234.

Ployet R, Soler M, Carocha V, Ladouce N, Alves A, Rodrigues JC, Harvengt

L, Marque C, Teulières C, Grima-Pettenati J, Mounet F. 2018. Long

cold exposure induces transcriptional and biochemical remodelling of

xylem secondary cell wall in Eucalyptus. Tree Physiology 38: 409–422.

Polko JK, Kieber JJ. 2019. The regulation of cellulose biosynthesis in plants.

Plant Cell 31: 282-296.

Pott DM, Osorio S, Vallarino JG. 2019. From central to specialized

metabolism: an overview of some secondary compounds derived from the

primary metabolism for their role in conferring nutritional and organoleptic

characteristics to fruit. Frontiers in Plant Science 10: 835.

Ragauskas AJ, Beckham GT, Biddy MJ, Chandra R, Chen F, Davis MF,

Davison BH, Dixon RA, Gilna P, Keller M, Langan P, Naskar AK,

Saddler JN, Tschaplinski TJ, Tuskan GA, Wyman CE. 2014. Lignin

valorization: improving lignin processing in the biorefinery. Science 344:

1246843.

Ragni L, Greb T. 2018. Secondary growth as a determinant of plant shape and

form. Seminars in Cell & Developmental Biology 79: 58-67.

Ralph J, Grabber JH, Harfield RD. 1995. Lignin-ferulate cross-links in

grasses: active incorporation of ferulate polysaccharide esters into ryegrass

lignins. Carbohydrate Research 275: 167-178.

Ralph J, Lundquist K, Brunow G, Lu F, Kim H, Schatz PF, Marita JM,

Hatfield RD, Ralph SA, Christensen JH, Boerjan W. 2004. Lignins:

Natural polymers from oxidative coupling of 4-hydroxyphenyl-

propanoids. Phytochemistry Reviews 3: 29-60.

Ralph J, Brunow G, Harris PJ, Dixon RA, Schatz PF, Boerjan W. 2008.

Lignification: are lignins biosynthesized via simple combinatorial

chemistry or via proteinaceous control and template replication? Recent

Advances in Polyphenol Research 1: 36–66.

Ralph J, Lapierre C, Boerjan W. 2019. Lignin structure and its engineering.

Current Opinion in Biotechnology 56: 240-249.

Rejab NA, Nakano Y, Yoneda A, Ohtani M, Demura T. 2015. Possible

contribution of TED6 and TED7, secondary cell wall-related membrane

proteins, to evolution of tracheary element in angiosperm lineage. Plant

Biotechnology 32: 343–347.

Ren Y, Hansen SF, Ebert B, Lau J, Scheller HV. 2014. Site-directed

mutagenesis of IRX9, IRX9L and IRX14 proteins involved in xylan

61

biosynthesis: glycosyltransferase activity is not required for IRX9 function

in Arabidopsis. PLoS One 9: e105014.

Rennie EA, Hansen SF, Baidoo EE, Hadi MZ, Keasling JD, Scheller HV.

2012. Three members of the Arabidopsis glycosyltransferase family 8 are

xylan glucuronosyltransferases. Plant Physiology 159: 1408-1417.

Rogers LA, Dubos C, Cullis IF, Surman C, Poole M, Willment J, Mansfield

SD, Campbell MM. 2005. Light, the circadian clock, and sugar perception

in the control of lignin biosynthesis. Journal of Experimental Botany 56:

1651-1663.

Rojas-Murcia N, Hématy K, Lee Y, Emonet A, Ursache R, Fujita S, De Bellis

D, Geldner N. 2020. High-order mutants reveal an essential requirement

for peroxidases but not laccases in Casparian strip lignification. bioRxiv

154617.

Ros Barceló A. 2005. Xylem parenchyma cells deliver the H2O2 necessary for

lignification in differentiating xylem vessels. Planta 220: 747–756.

Ruelland E, Campalans A, Selman-Housein G, Puigdomenecha P, Rigaua J.

2003. Cellular and subcellular localization of the lignin biosynthetic

enzymes caffeic acid-O-methyltransferase, cinnamyl alcohol

dehydrogenase and cinnamoyl-coenzyme A reductase in two monocots,

sugarcane and maize. Physiologia Plantarum 117: 93-99.

Ruonala R, Ko D, Helariutta Y. 2017. Genetic networks in plant vascular

development. Annual Review of Genetics 51: 335–359.

Sagi M, Fluhr R. 2006. Production of reactive oxygen species by plant NADPH

oxidases. Plant Physiology 141: 336–340.

Salleo S, Lo Gullo MA, Trifilo P, Nardini A. 2004. New evidence for a role of

vessel-associated cells and phloem in the rapid xylem refilling of cavitated

stems of Lauris nobilis L. Plant, Cell and Environment 27: 1065–1076.

Salleo S, Trifilo P, Esposito S, Nardini A, Lo Gullo MA. 2009. Starch-to-sugar

conversion in wood parenchyma of field-growing Laurus nobilis plants: a

component of the signal pathway for embolism repair? Functional Plant

Biology 36: 815–825.

Sanio K. 1873. Anatomie der gemeinen Kiefer Pinus sylvestris L. Jahrbücher für

Wissenschaftliche Botanik 9: 50-126.

Sarkar AK, Luijten M, Miyashima S, Lenhard M, Hashimoto T, Nakajima

K, Scheres B, Heidstra R, Laux T. 2007. Conserved factors regulate

signalling in Arabidopsis thaliana shoot and root stem cell organizers.

Nature 446: 811‐814.

Sauter JJ, Kloth S. 1986. Plasmodesmatal frequency and radial translocation

rates in ray cells of poplar (Populus × canadensis Moench ‘robusta’).

Planta 168: 337–380.

Scheller HV, Ulvskov P. 2010. Hemicelluloses. Annual Review of Plant Biology

61: 263–289.

Schlereth A, Möller B, Liu W, Kientz M, Flipse J, Rademacher EH, Schmid

M, Jürgens G, Weijers D. 2010. MONOPTEROS controls embryonic

62

root initiation by regulating a mobile transcription factor. Nature 464: 913–

916.

Schmitt U, Liese W. 1993. Response of xylem parenchyma to suberisation in

some hardwoods after mechanical injury. Trees – Structure and Function

8: 23–30.

Schneider R, Tang L, Lampugnani ER, Barkwill S, Lathe R, Zhang Y,

McFarlane HE, Pesquet E, Niittyla T, Mansfield SD, Zhou Y, Persson

S. 2017. Two complementary mechanisms underpin cell wall patterning

during xylem vessel development. Plant Cell 29: 2433–2449.

Schoof H, Lenhard M, Haecker A, Mayer KF, Jürgens G, Laux T. 2000. The

stem cell population of Arabidopsis shoot meristems in maintained by a

regulatory loop between the CLAVATA and WUSCHEL genes. Cell 100:

635–644.

Schuetz M, Benske A, Smith RA, Watanabe Y, Tobimatsu Y, Ralph J,

Demura T, Ellis B, Samuels AL. 2014. Laccases direct lignification in

the discrete secondary cell wall domains of protoxylem. Plant Physiology

166: 798-807.

Secchi F, Zwieniecki MA. 2011. Sensing embolism in xylem vessels: the role of

sucrose as a trigger for refilling. Plant, Cell and Environment 34: 514–524.

Secchi F, Zwieniecki MA. 2012. Analysis of xylem sap from functional

(nonembolized) and nonfunctional (embolized) vessels of Populus nigra:

chemistry of refilling. Plant Physiology 160: 955-964.

Shi D, Lebovka I, Lopez-Salmeron V, Sanchez P, Greb T. 2019. Bifacial

cambium stem cells generate xylem and phloem during radial plant growth.

Development 146: dev171355.

Shigeto J, Kiyonaga Y, Fujita K, Kondo R, Tsutsumi Y. 2013. Putative

cationic cell-wall-bound peroxidase homologues in Arabidopsis, AtPrx2,

AtPrx25, and AtPrx71, are involved in lignification. Journal of

Agricultural and Food Chemistry 6: 3781-3788.

Shigeto J, Itoh Y, Hirao S, Ohira K, Fujita K, Tsutsumi Y. 2015.

Simultaneously disrupting AtPrx2, AtPrx25 and AtPrx71 alters lignin

content and structure in Arabidopsis stem. Journal of Integrative Plant

Biology 57: 349–356.

Showalter AM. 1993. Structure and function of plant-cell wall proteins. Plant

Cell 5: 9–23.

Sibout R, Baucher M, Gatineau M, Doorsselaere JV, Mila I, Pollet B, Maba

B, Pilate G, Lapierre C, Boerjan W, Jouanin L. 2002. Expression of a

poplar cDNA encoding a ferulate-5-hydroxylase/ coniferaldehyde 5-

hydroxylase increases S lignin deposition in Arabidopsis thaliana. Plant

Physiology and Biochemistry 40: 1087–1096.

Sjöström E. 1993. Wood Chemistry, Fundamentals and Applications, 2nd

Edition. Academic Press 293.

Smet W, Sevilem I, de Luis Balaguer MA, Wybouw B, Mor E, Miyashima S,

Blob B, Roszak P, Jacobs TB, Boekschoten M, Hooiveld G, Sozzani R,

63

Helariutta Y, De Rybel B. 2019. DOF2.1 controls cytokine-independent

vascular cell proliferation downstream of TMO5/LHW. Current Biology

29: 520–529.

Smetana O, Makila R, Lyu M, Amiryousefi A, Sanchez Rodriguez F, Wu

MF, Sole-Gil A, Leal Gavarron M, Siligato R, Miyashima S, Roszak P,

Blomster T, Reed JW, Broholm S, Mähönen AP. 2019. High levels of

auxin signalling define the stem-cell organizer of the vascular cambium.

Nature 565: 485‐489.

Smith RA, Schuetz M, Roach M, Mansfield SD, Ellis B, Samuels L. 2013.

Neighboring parenchyma cells contribute to Arabidopsis xylem

lignification, while lignification of interfascicular fibers is cell

autonomous. Plant Cell 25: 1923-1935.

Smith RA, Schuetz M, Karlen SD, Bird D, Tokunaga N, Sato Y, Mansfield

SD, Ralph J, Samuels L. 2017. Defining the diverse cell populations

contributing to lignification in Arabidopsis stems. Plant Physiology 174:

1028-1036.

Sperry JS, Hacke U, Feild T, Sano Y, Sikkema E. 2007. Hydraulic

consequences of vessel evolution in angiosperms. International Journal of

Plant Sciences 168: 1127–1139.

Spicer R. 2014. Symplasmic networks in secondary vascular tissues:

parenchyma distribution and activity supporting long-distance transport.

Journal of Experimental Botany 65: 1829‐1848.

Stahl Y, Wink RH, Ingram GC, Simon R. 2009. A signaling module

controlling the stem cell niche in Arabidopsis root meristems. Current

Biology 19: 909–914.

Sterjiades R, Dean JFD, Gamble G, Himmelsbach DS, Eriksson KEL. 1993.

Extracellular laccases and peroxidases from sycamore maple (Acer

pseudoplatanus) cell-suspension cultures. Planta 190: 75–87.

Stewart CM. 1966. Excretion and heartwood formation in living trees. Science

153: 1068-1074.

Stewart JJ, Akiyama T, Chapple C, Ralph J, Mansfield SD. 2009. The effects

on lignin structure of overexpression of ferulate 5-hydroxylase in hybrid

poplar. Plant Physiology 150: 621–635.

Studer MH, DeMartini JD, Davis MF, Sykes RW, Davison B, Keller M,

Tuskan GA, Wyman CE. 2011. Lignin content in natural Populus

variants affects sugar release. Proceedings of the National Academy of

Sciences, USA 108: 6300-6305.

Sun Q, Rost TL, Matthews MA. 2008. Wound-induced vascular occlusions in

Vitis vinifera (Vitaceae): Tyloses in summer and gels in winter. American

Journal of Botany 95: 1498-1505.

Szyjanowicz PMJ, McKinnon I, Taylor NG, Gardiner J, Jarvis MC, Turner

SR. 2004. The irregular xylem 2 mutant is an allele of korrigan that affects

the secondary cell wall of Arabidopsis thaliana. The Plant Journal 37: 730–

740.

64

Takabe K, Fujita M, Harada H, Saiki H. 1985. Autoradiographic

investigations of lignification in the cell walls of Cryptomeria

(Cryptomeria japonica D. Don). Mokuzai Gakkaishi 3: 613-619.

Takabe K, Takeuchi M, Sato T, Ito M, Fujita M. 2001. Immunocytochemical

localization of enzymes involved in lignification of the cell wall. Journal

of Plant Research 114: 509-515.

Takeuchi M, Kegasa T, Watanabe A, Tamura M, Tsutsumi Y. 2018.

Expression analysis of transporter genes for screening candidate

monolignol transporters using Arabidopsis thaliana cell suspensions

during tracheary element differentiation. Journal of Plant Research 131:

297-305.

Taylor GN, Howells RM, Huttly AK, Vickers K, Turner SR. 2003.

Interactions among three distinct CesA proteins essential for cellulose

synthesis. Proceedings of the National Academy of Sciences, USA 100:

1450-1455.

Taylor-Teeples M, Lin L, De Lucas M, Turco G, Toal TW, Gaudinier A,

Young NF, Trabucco GM, Veiling MT, Lamothe R, Handakumbura

PP, Xiong G, Wang C, Corwin J, Tsoukalas A, Zhang L, Ware D,

Pauly M, Kliebenstein DJ, Dehesh K, Tagkopoulos I, Breton G,

Pruneda-Paz JL, Ahnert SE, Kay SA, Hazen SP, Brady SM. 2015. An

Arabidopsis gene regulatory network for secondary cell wall synthesis.

Nature 517: 571-575.

Terashima N, Fukushima K. 1988. Heterogeneity in formation of lignin. XI.

An autographic study of heterogeneous formation and structure of pine

lignin. Wood Science and Technology 22: 259–270.

Terashima N, Fukushima K, Sano Y, Takabe K. 1988. Heterogeneity in

formation of lignin. X. Visualization of lignification process in

differentiating xylem of pine by microautoradiography. Holzforschung 42:

347-350.

Timell TE. 1967. Recent progress in the chemistry of wood hemicelluloses.

Wood Science and Technology 1: 45– 70.

Tobimatsu Y, Schuetz M. 2019. Lignin polymerization: how do plants manage

the chemistry so well? Current Opinion in Biotechnology 56: 75-81.

Tokunaga N, Kaneta T, Sato S, Sato Y. 2009. Analysis of expression profiles

of three peroxidase genes associated with lignification in Arabidopsis

thaliana. Physiologia Plantarum 136: 237-249.

Tronchet M, Balagué C, Kroj T, Jouanin L, Roby D. 2010. Cinnamyl alcohol

dehy-drogenases C and D, key enzymes in lignin biosynthesis, play an

essential role in disease resistance in Arabidopsis. Molecular Plant

Pathology 11: 83–92.

Tsai KL, Tomomori-Sato C, Sato S, Conaway RC, Conaway JW, Asturias

FJ. 2014. Subunit architecture and functional modular rearrangements of

the transcriptional Mediator complex. Cell 157: 1430–1444.

65

Tsuyama T, Kawai R, Shitan N, Matoh T, Sugiyama J, Yoshinaga A, Takabe

K, Fujita M, Yazaki, K. 2013. Proton-dependent coniferin transport, a

common major transport event in differentiating xylem tissue of woody

plants. Plant Physiology 162: 918-926.

Tsuyama T, Matsushita Y, Fukushima K, Takabe K, Yazaki K, Kamei I.

2019. Proton gradient-dependent transport of p-glucocoumaryl alcohol in

differentiating xylem of woody plants. Scientific Reports 9: 8900.

Turner SR, Somerville CR. 1997. Collapsed xylem phenotype of Arabidopsis

identifies mutants deficient in cellulose deposition in the secondary cell

wall. Plant Cell 9: 689–701.

Turner S, Gallois P, Brown D. 2007. Tracheary element differentiation. Annual

Review of Plant Biology 58: 407‐433.

Tyree MT, Zimmermann MH. 2002. Hydraulic architecture of whole plants

and plant performance. In Xylem Structure and the Ascent of Sap, pp. 175–

214. New York: Springer. 2nd ed.

United Nations. 2017. Sustainable Development Goals Report 2017 available at

https://unstats.un.org/sdgs/files/report/2017/thesustainabledevelopmentgo

alsreport2017.pdf.

Urbanowicz BR, Peña MJ, Ratnaparkhe S, Avci U, Backe J, Steet HF,

Foston M, Li H, O’Neill MA, Ragauskas AJ, Darvill AG, Wyman C,

Gilbert HJ, York WS. 2012. 4-O-methylation of glucuronic acid in

Arabidopsis glucuronoxylan is catalyzed by a domain of unknown function

family 579 protein. Proceedings of the National Academy of Sciences,

USA 109: 14253-14258.

Urbanowicz BR, Peña MJ, Moniz HA, Moremen KW, York WS. 2014. Two

Arabidopsis proteins synthesize acetylated xylan in vitro. The Plant

Journal 80: 197–206.

Van Acker R, Déjardin A, Desmet S, Hoengenaert L, Vanholme R, Morreel

K, Laurans F, Kim H, Santoro N, Foster C, Goeminne G, Légée F,

Lapierre C, Pilate G, Ralph J, Boerjan W. 2017. Different metabolic

routes for coniferaldehyde and sinapaldehyde with CINNAMYL

ALCOHOL DEHYDROGENASE1 deficiency. Plant Physiology 175:

1018-1039.

Vanholme R, Demedts B, Morreel K, Ralph J, Boerjan W. 2010. Lignin

biosynthesis and structure. Plant Physiology 153: 895-905.

Vanholme R, De Meester B, Ralph J, Boerjan W. 2019. Lignin biosynthesis

and its integration into metabolism. Current Opinion in Biotechnology 56:

230‐239.

Vera-Sirera F, De Rybel B, Urbez C, Kouklas E, Pesquera M, Alvarez-

Mahecha JC, Minguet EG, Tuominen H, Carbonell J, Borst JW,

Weijers D, Blázquez MA. 2015. A bHLH-based feedback loop restricts

vascular cell proliferation in plants. Developmental Cell 35: 432–443.

Vermaas JV, Dixon RA, Chen F, Mansfield SD, Boerjan W, Ralph J,

Crowley MF, Beckham GT. 2019. Passive membrane transport of lignin-

66

related compounds. Proceedings of the National Academy of Sciences,

USA 116: 23117–23123.

Voelker SL, Lachenbruch B, Meinzer FC, Jourdes M, Ki C, Patten AM,

Davin LB, Lewis NG, Tuskan GA, Gunter L, Decker SR, Selig MJ,

Sykes R, Himmel ME, Kitin P, Shevchenko O, Strauss SH. 2010.

Antisense down-regulation of 4CL expression alters lignification, tree

growth, and saccharification potential of field-grown poplar. Plant

Physiology 154: 874‐886.

Voxeur A, Wang Y, Sibout R. 2015. Lignification: different mechanisms for a

versatile polymer. Current Opinion in Plant Biology 23: 83-90.

Wagner A, Donaldson L, Kim H, Phillips L, Flint H, Steward D, Torr K,

Koch G, Schmitt U, Ralph J. 2009. Suppression of 4-coumarate-CoA

ligase in the coniferous gymnosperm Pinus radiata. Plant Physiology 149:

370-383.

Wagner A, Tobimatsu Y, Phillips L, Flint H, Geddes B, Lu F, Ralph J. 2015.

Syringyl lignin production in conifers: Proof of concept in a Pine tracheary

element system. Proceedings of the National Academy of Sciences, USA

112: 6218-6223.

Wang H, Avci U, Nakashima J, Hahn MG, Chen F, Dixon RA. 2010. Mutation

of WRKY transcription factors initiates pith secondary wall formation and

increases stem biomass in dicotyledonous plants. Proceedings of the

National Academy of Sciences, USA 107: 22338–22343.

Wang Y, Frei M. 2011. Stressed food – the impact of abiotic environmental

stresses on crop quality. Agriculture, Ecosystems & Environment 141:

271–286.

Wang CY, Zhang S, Yu Y, Luo YC, Liu Q, Lu C, Zhang YC, Qu LH, Lucas

WJ, Wang X, Chen YQ. 2014. MiR397b regulates both lignin content

and seed number in Arabidopsis via modulating a laccase involved in

lignin biosynthesis. Plant Biotechnology Journal 12: 1132–1142.

Wang JP, Chuang L, Loziuk PL, Chen H, Lin Y-C, Shi R, Qu G-Z,

Muddiman DC, Sederoff RR, Chiang VL. 2015. Phosphorylation is an

on/off switch for 5-hydroxyconiferaldehyde O-methyltransferase activity

in poplar monolignol biosynthesis. Proceedings of the National Academy

of Sciences, USA 112: 8481-8486.

Wang YY, Meng X, Pu Y, Ragauskas AJ. 2020. Recent advances in the

application of functionalized lignin in value-added polymeric materials.

Polymers (Basel) 12: E2277.

Watanabe Y, Meents MJ, McDonnell LM, Barkwill S, Sampathkumar A,

Cartwright HN, Demura T, Ehrhardt DW, Samuels AL, Mansfield

SD. 2015. Visualization of cellulose synthases in Arabidopsis secondary

cell walls. Science 350: 198–203.

Wendler WMF, Kremmer E, Forster R, Winnacker EL. 1997. Identification

of Pirin, a novel highly conserved nuclear protein. Journal of Biological

Chemistry 272: 8482-8489.

67

Weng JK, Li X, Bonawitz ND, Chapple C. 2008. Emerging strategies of lignin

engineering and degradation for cellulosic biofuel production. Current

Opinion in Biotechnology 19: 166–172.

Weng JK, Chapple C. 2010. The origin and evolution of lignin biosynthesis.

New Phytologist 187: 273-285.

Whitford R, Fernandez A, De Groodt R, Ortega E, Hilson P. 2008. Plant CLE

peptides from two distinct functional classes synergistically induce

division of vascular cells. Proceedings of the National Academy of

Sciences, USA 105: 18625–18630.

Wiedenhoeft AC. 2013. Structure and function of wood. In Handbook of Wood

Chemistry and Wood Composites, 2nd ed p 9-32.

Wilkerson CG, Mansfield SD, Lu F, Withers S, Park JY, Karlen SD,

Gonzales-Vigil E, Padmakshan D, Unda F, Rencoret J, Ralph J. 2014.

Monolignol ferulate transferase introduces chemically labile linkages into

the lignin backbone. Science 344: 90-93.

Wu AM, Hörnblad E, Voxeur A, Gerber L, Rihouey C, Lerouge P, Marchant

A. 2010. Analysis of the Arabidopsis IRX9/IRX9-L and IRX14/IRX14-L

pairs of glycosyltransferase genes reveals critical contributions to

biosynthesis of the hemicellulose glucuronoxylan. Plant Physiology 153:

542-554.

Xiong G, Cheng K, Pauly M. 2013. Xylan O-acetylation impacts xylem

development and enzymatic recalcitrance as indicated by the Arabidopsis

mutant tbl29. Molecular Plant 6: 1373-1375.

Xu L, Ménard R, Berr A, Fuchs J, Cognat V, Meyer D, Shen W‐H. 2009. The

E2 ubiquitin‐conjugating enzymes, AtUBC1 and AtUBC2, play redundant

roles and are involved in activation of FLC expression and repression of

flowering in Arabidopsis thaliana. The Plant Journal 57: 279– 288.

Xu B, Ohtani M, Yamaguchi M, Toyooka K, Wakazaki M, Sato M, Kubo M,

Nakano Y, Sano R, Hiwatashi Y, Murata T, Kurata T, Yoneda A, Kato

K, Hasebe M, Demura T. 2014. Contribution of NAC transcription

factors to plant adaptation to land. Science 343: 1505-1508.

Yamaguchi M, Kubo M, Fukuda H, Demura T. 2008. VASCULAR-

RELATED NAC-DOMAIN7 is involved in the differentiation of all types

of xylem vessels in Arabidopsis roots and shoots. The Plant Journal 55:

652-664.

Yamaguchi M, Ohtani M, Mitsuda N, Kubo M, Ohme-Takagi M, Fukuda H,

Demura T. 2010. VND-INTERACTING2, a NAC domain transcription

factor, negatively regulates xylem vessel formation in Arabidopsis. Plant

Cell 22: 1249–1263.

Yan X, Liu J, Kim H, Liu B, Huang X, Yang Z, Lin YCJ, Chen H, Yang C,

Wang JP, Muddiman DC, Ralph J, Sederoff RR, Li Q, Chaing VL.

2019. CAD1 and CCR2 protein complex formation in monolignol

biosynthesis in Populus trichocarpa. New Phytologist 222: 244-260.

68

Yang F, Mitra P, Zhang L, Prak L, Verhertbruggen Y, Kim JS, Sun L,

Zheng K, Tang K, Auer M, Scheller HV, Loqué D. 2013. Engineering

secondary cell wall deposition in plants. Plant Biotechnology Journal 11:

325‐335.

Yuan YX, Teng Q, Zhong RQ, Ye ZH. 2013. The Arabidopsis DUF231

domain-containing protein ESK1 mediates 2-O- and 3-O-acetylation of

xylosyl residues in xylan. Plant and Cell Physiology 54: 1186–1199.

Yuan Y, Teng Q, Zhong R, Ye ZH. 2016A. TBL3 and TBL31, two Arabidopsis

DUF231 domain proteins, are required for 3-O-monoacetylation of xylan.

Plant and Cell Physiology 57: 35-45.

Yuan Y, Teng Q, Zhong R, Haghighat M, Richardson EA, Ye ZH. 2016B.

Mutations of Arabidopsis TBL32 and TBL33 affect xylan acetylation and

secondary wall deposition. PLOS ONE 11: e0146460.

Yuan Y, Teng Q, Zhong R, Ye ZH. 2016C. Roles of Arabidopsis TBL34 and

TBL35 in xylan acetylation and plant growth. Plant Science 243: 120-130.

Zeng W, Jiang N, Nadella R, Killen TL, Nadella V, Faik A. 2010. A

glucuronoarabinoxylan synthase complex from wheat contains members

of the GT43, GT47, and GT75 families and functions cooperatively. Plant

Physiology 154: 78–97.

Zeng W, Lampugnani ER, Picard KL, Song L, Wu AM, Farion IM, Zhao J,

Ford K, Doblin MS, Bacic A. 2016. Asparagus IRX9, IRX10, and

IRX14A are components of an active xylan backbone synthase complex

that forms in the Golgi apparatus. Plant Physiology 171: 93–109.

Zeng Y, Zhao S, Yang S, Ding SY. 2014. Lignin plays a negative role in the

biochemical process for producing lignocellulosic biofuels. Current

Opinion in Biotechnology 27: 38-45.

Zhang B, Tremousaygue D, Denancé N, van Esse HP, Hörger AC, Dabos P,

Goffner D, Thomma BPHJ, van der Hoorn RAL, Tuominen H. 2014.

PIRIN2 stabilizes cysteine protease XCP2 and increases susceptibility to

the vascular pathogen Ralstonia solanacearum in Arabidopsis. The Plant

Journal 79: 1009-1019.

Zhang H, Lin X, Han Z,Wang J, Qu LJ, Chai J. 2016. SERK family receptor-

like kinases function as co-receptors with PXY for plant vascular

development. Molecular Plant 9: 1406–1414.

Zhang J, Xie M, Tuskan GA, Muchero W, Chen JG. 2018. Recent advances

in the transcriptional regulation of secondary cell wall biosynthesis in the

woody plants. Frontiers in Plant Science 9: 1535.

Zhang X, Gou M, Liu CJ. 2013. Arabidopsis Kelch repeat F-box proteins

regulate phenylpropanoid biosynthesis via controlling the turnover of

phenylalanine ammonia-lyase. Plant Cell 25: 4994–5010.

Zhang X, Dominguez PG, Kumar M, Bygdell J, Miroshnichenko S,

Sundberg B, Wingsle G, Niittylä T. 2018. Cellulose synthase

stoichiometry in aspen differs from Arabidopsis and Norway spruce. Plant

Physiology 177: 1096-1107.

69

Zhang B, Sztojka B, Escamez S, Vanholme R, Hedenström M, Wang Y,

Turumtay H, Gorzsás A, Boerjan W, Tuominen H. 2020A. PIRIN2

suppresses S-type lignin accumulation in a noncell-autonomous manner in

Arabidopsis xylem elements. New Phytologist 225: 1923‐1935.

Zhang Q, Luo F, Zhong Y, He J, Li L. 2020B. Modulation of NAC transcription

factor NST1 activity by XYLEM NAC DOMAIN1 regulates secondary

cell wall formation in Arabidopsis. Journal of Experimental Botany 71:

1449‐1458.

Zhang Y, Li D, Zhang H, Hong Y, Huang L, Liu S, Li X, Ouyang Z, Song F.

2015. Tomato histone H2B monoubiquitination enzymes SlHUB1 and

SlHUB2 contribute to disease resistance against Botrytis cinerea through

modulating the balance between SA‐ and JA/ET‐mediated signaling

pathways. BMC Plant Biology 15: 252.

Zhao C, Avci U, Grant EH, Haigler CH, Beers EP. 2008. XND1, a member of

the NAC domain family in Arabidopsis thaliana, negatively regulates

lignocellulose synthesis and programmed cell death in xylem. The Plant

Journal 53: 425–436.

Zhao C, Lasses T, Bakó L, Kong D, Zhao B, Chanda B, Bombarely A, Cruz-

Ramírez A, Scheres B, Brunner AM, Beers EP. 2017. XYLEM NAC

DOMAIN1, an angiosperm NAC transcription factor, inhibits xylem

differentiation through conserved motifs that interact with

RETINOBLASTOMA-RELATED. New Phytologist 216: 76-89.

Zhao J, Chen QH, Zhou S, Sun YH, Li X, Li YZ. 2020. H2Bub1 regulates

RbohD-dependent H2O2 signal pathway in the defense responses to

Verticillium dahliae toxins. Plant Physiology 182: 640-657.

Zhao Q, Nakashima J, Chen F, Yin Y, Fu C, Yun J, Shao H, Wang X, Wang

ZY, Dixon RA. 2013A. LACCASE is necessary and nonredundant with

PEROXIDASE for lignin polymerization during vascular development in

Arabidopsis. Plant Cell 25: 3976–3987.

Zhao Q, Tobimatsu Y, Zhou R, Pattathil S, Gallego-giraldo L, Fu C. 2013B.

Loss of function of cinnamyl alcohol dehydrogenase 1 leads to

unconventional lignin and a temperature-sensitive growth defect in

Medicago truncatula. Proceedings of the National Academy of Sciences,

USA 110: 13660-13665.

Zhao Y, Lin S, Qiu Z, Cao D, Wen J, Deng X, Wang X, Lin J, Li X. 2015.

MicroRNA857 is involved in the regulation of secondary growth of

vascular tissues in Arabidopsis. Plant Physiology 169: 2539–2552.

Zhong R, Demura T, Ye ZH. 2006. SND1, a NAC domain transcription factor,

is a key regulator of secondary wall synthesis in fibers of Arabidopsis.

Plant Cell 18: 3158-3170.

Zhong R, Lee C, Ye ZH. 2010. Global analysis of direct targets of secondary

wall NAC master switches in Arabidopsis. Molecular Plant 3: 1087-1103.

70

Zhong R, Ye ZH. 2012. MYB46 and MYB83 bind to the SMRE sites and

directly activate a suite of transcription factors and secondary wall

biosynthetic genes. Plant and Cell Physiology 53: 368-380.

Zhou J, Lee C, Zhong R, Ye ZH. 2009. MYB58 and MYB63 are transcriptional

activators of the lignin biosynthetic pathway during secondary cell wall

formation in Arabidopsis. Plant Cell 21: 248-266.

Zobel BJ, Van Buijtenen JP. 1989. Wood variation: its causes and control.

Springer-Verlag.

Zou B, Yang D‐L, Shi Z, Dong H, Hua J. 2014. Monoubiquitination of histone

2B at the disease resistance gene locus regulates its expression and impacts

immune responses in Arabidopsis. Plant Physiology 165: 309-318.