02.07.2013 Views

The Plant Vascular System: Evolution, Development and FunctionsF

The Plant Vascular System: Evolution, Development and FunctionsF

The Plant Vascular System: Evolution, Development and FunctionsF

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Journal of Integrative <strong>Plant</strong> Biology 2013, 55 (4): 294–388<br />

Invited Expert Review<br />

<strong>The</strong> <strong>Plant</strong> <strong>Vascular</strong> <strong>System</strong>: <strong>Evolution</strong>, <strong>Development</strong><br />

<strong>and</strong> Functions F<br />

William J. Lucas1∗, Andrew Groover2 , Raffael Lichtenberger3 , Kaori Furuta3 , Shri-Ram Yadav3 ,<br />

Ykä Helariutta3 , Xin-Qiang He4 , Hiroo Fukuda5 , Julie Kang6 , Siobhan M. Brady1 ,<br />

John W. Patrick7 , John Sperry8 , Akiko Yoshida1 , Ana-Flor López-Millán9 , Michael A. Grusak10 <strong>and</strong> Pradeep Kachroo11 1Department of <strong>Plant</strong> Biology, College of Biological Sciences, University of California, Davis, CA 95616, USA<br />

2US Forest Service, Pacific SW Res Stn, Institute for Forest Genetics, Davis, CA 95618, USA<br />

3Institute of Biotechnology/Department of Biology <strong>and</strong> Environmental Sciences, University of Helsinki, FIN-00014, Finl<strong>and</strong><br />

4State Key Laboratory of Protein <strong>and</strong> <strong>Plant</strong> Gene Research, College of Life Sciences, Peking University, Beijing 100871, China<br />

5Department of Biological Sciences, Graduate School of Science, <strong>The</strong> University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033,<br />

Japan<br />

6Biology Department, University of Northern Iowa, Cedar Falls, IA, USA<br />

7School of Environmental <strong>and</strong> Life Sciences, <strong>The</strong> University of Newcastle, Callaghan, NSW 2308, Australia<br />

8Biology Department, University of Utah, 257S 1400E Salt Lake City, UT 84112, USA<br />

9<strong>Plant</strong> Nutrition Department, Aula Dei Experimental Station (CSIC), P.O. Box 13034, E-50080, Zaragoza, Spain<br />

10USDA-ARS Children’s Nutrition Research Center, Department of Pediatrics, Baylor College of Medicine, 1100 Bates Street, Houston, TX<br />

77030, USA<br />

11Department of <strong>Plant</strong> Pathology, College of Agriculture, University of Kentucky, Lexington, KY 40546, USA<br />

All authors contributed equally to the preparation of this manuscript.<br />

∗<br />

Corresponding author<br />

Tel: +1 530 752 1093; Fax: +1 530 752 5410; E-mail: wjlucas@ucdavis.edu<br />

F Articles can be viewed online without a subscription.<br />

Available online on 5 March 2013 at www.jipb.net <strong>and</strong> www.wileyonlinelibrary.com/journal/jipb<br />

doi: 10.1111/jipb.12041<br />

Contents<br />

I. Introduction 295<br />

II. <strong>Evolution</strong> of the <strong>Plant</strong> <strong>Vascular</strong> <strong>System</strong> 295<br />

III. Phloem <strong>Development</strong> & Differentiation 300<br />

IV. Molecular Mechanisms Underlying Xylem Cell Differentiation 307<br />

V. Spatial & Temporal Regulation of <strong>Vascular</strong> Patterning 311<br />

VI. Secondary <strong>Vascular</strong> <strong>Development</strong> 318<br />

VII. Physical <strong>and</strong> Physiological Constraints on Phloem Transport Function 321<br />

VIII. Physical & Physiological Constraints on Xylem Function 328<br />

IX. Long-distance Signaling Through the Phloem 339<br />

X. Root-to-shoot Signaling 347<br />

XI. <strong>Vascular</strong> Transport of Microelement Minerals 351<br />

XII. <strong>System</strong>ic Signaling: Pathogen Resistance 356<br />

XIII. Future Perspectives 361<br />

XIV. Acknowledgements 362<br />

XV. References 362<br />

C○ 2013 Institute of Botany, Chinese Academy of Sciences


William J. Lucas<br />

(Corresponding author)<br />

Abstract<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 295<br />

<strong>The</strong> emergence of the tracheophyte-based vascular system of l<strong>and</strong> plants<br />

had major impacts on the evolution of terrestrial biology, in general,<br />

through its role in facilitating the development of plants with increased<br />

stature, photosynthetic output, <strong>and</strong> ability to colonize a greatly exp<strong>and</strong>ed<br />

range of environmental habitats. Recently, considerable progress has<br />

been made in terms of our underst<strong>and</strong>ing of the developmental <strong>and</strong><br />

physiological programs involved in the formation <strong>and</strong> function of the<br />

plant vascular system. In this review, we first examine the evolutionary<br />

events that gave rise to the tracheophytes, followed by analysis of the<br />

genetic <strong>and</strong> hormonal networks that cooperate to orchestrate vascular<br />

development in the gymnosperms <strong>and</strong> angiosperms. <strong>The</strong> two essential<br />

functions performed by the vascular system, namely the delivery of resources (water, essential mineral<br />

nutrients, sugars <strong>and</strong> amino acids) to the various plant organs <strong>and</strong> provision of mechanical support are<br />

next discussed. Here, we focus on critical questions relating to structural <strong>and</strong> physiological properties<br />

controlling the delivery of material through the xylem <strong>and</strong> phloem. Recent discoveries into the role of<br />

the vascular system as an effective long-distance communication system are next assessed in terms<br />

of the coordination of developmental, physiological <strong>and</strong> defense-related processes, at the whole-plant<br />

level. A concerted effort has been made to integrate all these new findings into a comprehensive picture<br />

of the state-of-the-art in the area of plant vascular biology. Finally, areas important for future research<br />

are highlighted in terms of their likely contribution both to basic knowledge <strong>and</strong> applications to primary<br />

industry.<br />

Keywords: <strong>Evolution</strong>; vascular development; phloem; xylem; nutrient delivery; long-distance communication; systemic signaling.<br />

Lucas WJ, Groover A, Lichtenberger R, Furuta K, Yadav SR, Helariutta Y, He XQ, Fukuda H, Kang J, Brady SM, Patrick JW, Sperry J,<br />

Yoshida A, López-Millán AF, Grusak MA, Kachroo P (2013) <strong>The</strong> plant vascular system: <strong>Evolution</strong>, development <strong>and</strong> functions. J. Integr. <strong>Plant</strong><br />

Biol. 55(4), 294–388.<br />

Introduction<br />

<strong>The</strong> plant vascular system carries out two essential functions,<br />

namely the delivery of resources (water, essential mineral<br />

nutrients, sugars <strong>and</strong> amino acids) to the various plant organs,<br />

<strong>and</strong> provision of mechanical support. In addition, the vascular<br />

system serves as an effective long-distance communication<br />

system, with the phloem <strong>and</strong> xylem serving to input information<br />

relating to abiotic <strong>and</strong> biotic conditions above <strong>and</strong> below<br />

ground, respectively. This combination of resource supply <strong>and</strong><br />

delivery of information, including hormones, peptide hormones,<br />

proteins <strong>and</strong> RNA, allows the vascular system to engage in the<br />

coordination of developmental <strong>and</strong> physiological processes at<br />

the whole-plant level.<br />

Over the past decade, considerable progress has been<br />

made in terms of our underst<strong>and</strong>ing of the developmental <strong>and</strong><br />

physiological programs involved in the formation <strong>and</strong> function of<br />

the plant vascular system. In this review, we have made every<br />

effort to integrate these new findings into a comprehensive<br />

picture of the state-of-the-art in this important facet of plant<br />

biology. We also highlight potential areas important for future<br />

research in terms of their likely contribution both to basic<br />

knowledge <strong>and</strong> applications to primary industry.<br />

<strong>Evolution</strong> of the <strong>Plant</strong> <strong>Vascular</strong> <strong>System</strong><br />

Why the need for a vasculature system?<br />

For plants as photosynthetic autotrophs, the evolutionary step<br />

from uni- to multi-cellularity conferred an important selective<br />

advantage in terms of division of labor; i.e., functional specialization<br />

of tissues/organs to more effectively extract, <strong>and</strong><br />

compete for, essential resources in aquatic <strong>and</strong> terrestrial<br />

environments. Successful colonization of terrestrial environments,<br />

by plants, depended upon positioning of organs in both<br />

aerial <strong>and</strong> soil environments to meet their autotrophic requirements.<br />

For example, for photosynthetic efficiency, sufficient<br />

levels of light only co-occur with a supply of CO2 in aerial


296 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

environments, whereas water <strong>and</strong> mineral requirements are<br />

primarily acquired from soil environments. Thus, aerial <strong>and</strong> soil<br />

organs of early l<strong>and</strong> plants were nutritionally interdependent<br />

<strong>and</strong>, consequently, there was intense selection pressure for the<br />

evolution of an inter-organ transport system to allow access to<br />

the complete spectrum of essential resources for cell growth<br />

<strong>and</strong> maintenance.<br />

An important feature of early multicellular plants was the<br />

acquisition of plasmodesmata (PD), whose cytoplasmic channels<br />

established a symplasmic continuum throughout the body<br />

of the plant (Lucas et al. 1993). This symplasm allowed for<br />

the exchange of nutrients between the different plant organs.<br />

However, this symplasmic route, in which intracellular cytoplasmic<br />

streaming is arranged in series with intercellular diffusion<br />

through PD, is effective only over rather short distances. For<br />

example, a PD-mediated sucrose flux of 2 × 10 −4 mol m −2 s −1<br />

into heterotrophic cells would satisfy their metabolic dem<strong>and</strong>.<br />

Maximum reported permeability coefficients for sucrose diffusion<br />

through PD are on the order of 6 × 10 −6 ms −1 (Fisher<br />

<strong>and</strong> Wang 1995). Using this value, diffusion theory predicts that<br />

a significant sucrose concentration drop would be required,<br />

across each adjoining cell wall interface, to sustain this flux<br />

of sucrose from the autotrophic (photosynthetic) cells into the<br />

heterotrophic (water <strong>and</strong> mineral nutrient acquiring) cells. Thus,<br />

the path length would be limited to only a few cells, arranged in<br />

series, <strong>and</strong> the size of the organism would be limited to a few<br />

millimeters.<br />

In order for multicellular autotrophs to overcome these<br />

diffusion-imposed size constraints, a strong selection pressure<br />

existed to evolve an axially-arranged tissue system, located<br />

throughout the plant body, with a greatly increased conductivity<br />

for intercellular transport. <strong>The</strong> solution to this problem began<br />

over 470 Mya <strong>and</strong>, in combination with prevailing global climate<br />

change, including dramatic changes in atmospheric CO2 levels,<br />

gave rise to the development of the cuticle <strong>and</strong> stomata,<br />

important adaptations that both reduced tissue dehydration<br />

<strong>and</strong> increased the capacity for exchange of CO2, thereby<br />

enhancing the rates of photosynthesis (Franks <strong>and</strong> Brodribb<br />

2005; Ruszala et al. 2011; but see Duckett et al. 2009).<br />

Following acquisition of these two traits, early l<strong>and</strong> plants<br />

evolved cells specialized for long-distance transport of food<br />

<strong>and</strong> water (Ligrone et al. 2000, 2012; Raven 2003; van Bel<br />

2003; Pittermann 2010). Irrespective of plant group, these cells<br />

became arranged end-to-end in longitudinal files having a simplified<br />

cytoplasm <strong>and</strong> modified end walls designed to increase<br />

their intra- <strong>and</strong> intercellular conductivities, respectively.<br />

In l<strong>and</strong> plants, the degree of cellular modifications of transport<br />

cells increases from the bryophytes (pretracheophytes—also<br />

termed non-vascular plants—the liverworts, mosses <strong>and</strong> hornworts),<br />

to the early tracheophytes, the vascular cryptogams (lycophytes<br />

<strong>and</strong> pterophytes), on through to seed plants (Ligrone<br />

et al. 2000, 2012; Raven 2003; van Bel 2003). <strong>The</strong>se cell<br />

specializations neatly scale with maximal sizes attained by<br />

each group of l<strong>and</strong> plants. Interestingly, impacts of enhancing<br />

conductivities of cells transporting sugars converges with a<br />

greater influence imposed by evolving water conducting cells<br />

to sustain hydration of aerial photosynthetic tissues.<br />

<strong>Evolution</strong>ary origins <strong>and</strong> diversification of food<br />

<strong>and</strong> water transport systems<br />

Studies based on fossil records <strong>and</strong> extant (living) bryophytes<br />

have established that developmental programs evolved to form<br />

specialized water <strong>and</strong> nutrient conducting tissues. Based on<br />

the fossil record, early pretracheophyte l<strong>and</strong> plants appeared<br />

to have developed simple water-conducting conduits having<br />

smooth walls with small pores, likely derived from the presence<br />

of PD. Similar structures are present, for example, in some of<br />

the mosses, the most ancient being termed water-conducting<br />

cells (WCCs) <strong>and</strong> the more advanced being the hydroids of the<br />

peristomate mosses (Mishler <strong>and</strong> Churchill 1984; Kenrick <strong>and</strong><br />

Crane 1997; Ligrone et al. 2012).<br />

Hydroids often form a central str<strong>and</strong> in the gametophyte<br />

stem/sporophyte seta in the mosses (Figure 1A, B). During<br />

their development, these hydroid cells undergo various structural<br />

modifications to the cell wall <strong>and</strong> are dead at maturity<br />

(Figure 1C, D). Although in some cases the hydroid wall may<br />

become thickened, these are considered to be primary in<br />

nature <strong>and</strong> lack lignin. However, recent studies have indicated<br />

that bryophyte cell walls contain lignin-related compounds,<br />

but these do not impart mechanical strengthening properties<br />

(Ligrone et al. 2012). Although this absence of mechanical<br />

strength served as an impediment to an increase in body size,<br />

it allowed for hydroid collapse during tissue desiccation, <strong>and</strong><br />

rapid rehydration following a resupply of water (Figure 1E, F), a<br />

feature that likely minimized cavitation of these WCCs (Ligrone<br />

et al. 2012) (see also later section). This trait may also have<br />

allowed peristomate mosses to exp<strong>and</strong> into dryer habitats.<br />

<strong>The</strong> evolution of hydroids could have involved modification of<br />

existing WCCs. However, based on the distribution of WCCs<br />

in the early l<strong>and</strong> plants (Figure 2), it seems equally probable<br />

that they arose through an independent developmental<br />

pathway after the loss of perforate WCCs (Ligrone et al.<br />

2012).<br />

<strong>The</strong> fossil record contains less information on the evolution<br />

of specialized food-conducting cells (FCCs), due in large<br />

part to their less robust characteristics that limited effective<br />

preservation. However, insights can be gained from studies<br />

on extant bryophyte species. As with WCCs, early FCCs were<br />

represented by files of aligned elongated cells in which the<br />

cytoplasmic contents underwent a series of positional <strong>and</strong><br />

structural modifications (Figure 3). Here, we will use moss as an<br />

example; in some species (members of the order Polytrichales),


Figure 1. Water-conducting cells of early non-vascular (pretracheophyte)<br />

l<strong>and</strong> plants.<br />

(A) Light micrograph of a leafy stem from the moss Plagiomnium<br />

undulatum illustrating the prominent central str<strong>and</strong> of hydroid cells<br />

(h) surrounded by parenchyma cells (p).<br />

(B–D) Transmission electron micrographs illustrating hydroids having<br />

unevenly thickened walls in the leaf shoot of the moss Polytrichum<br />

juniperinum (B), a differentiating hydroid (C) <strong>and</strong> mature<br />

hydroids (D) in the leafy shoot of Polytrichum formosum.<br />

(E, F) Transverse sections of moss hydroids in the hydrated (E)<br />

<strong>and</strong> dehydrated condition (F); note that in the presence of water,<br />

dehydrated hydroids become rehydrated <strong>and</strong> functional. Images<br />

(A–D) reproduced from Ligrone et al. (2000), with permission of<br />

<strong>The</strong> Royal Society London; (E, F) reproduced from Ligrone et al.<br />

(2012), with permission of Oxford University Press. Scale bars:<br />

50 µm in(A), 5µm in(B–D) <strong>and</strong> 5 µm in(E), common to (F).<br />

the FCCs gave rise to a group of more specialized cells, termed<br />

leptoids <strong>and</strong> associated specialized parenchyma cells. During<br />

development, leptoids undergo a series of cytological changes,<br />

including cytoplasmic polarization <strong>and</strong> microtubule-associated<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 297<br />

alignment of plastids, mitochondria <strong>and</strong> the endoplasmic reticulum<br />

(ER), in a longitudinal pattern. At maturity, the FCCs of<br />

the bryophytes generally lack a large central vacuole <strong>and</strong>, in<br />

some species, there is partial degradation of the nucleus. In<br />

addition, the end walls of cells within these files of aligned FCCs<br />

develop a high density of PD (Figure 4), presumably to optimize<br />

symplasmic continuity for cell-to-cell diffusion of photosynthate<br />

(Ligrone et al. 2000, 2012; Raven 2003). Finally, FCCs/leptoids<br />

often develop in close proximity to WCCs/hydroids; in<br />

some species, the WCCs/hydroids are centrally located<br />

in the tissue/stem being ensheathed by FCCs/leptoids<br />

(Figure 3E).<br />

As with hydroids, the evolutionary events leading to the<br />

development of FCCs, <strong>and</strong> leptoids in particular, appear to<br />

have been driven by the necessity to withst<strong>and</strong> periods in<br />

which the early l<strong>and</strong> plants underwent desiccation. Insight sinto<br />

the presence <strong>and</strong> importance of such traits were offered by<br />

recent physiological <strong>and</strong> anatomical studies performed on the<br />

desiccation-resistant moss Polytrichum formosum. Here, the<br />

unique resilience of the lepoids to an imposed dehydration was<br />

shown to be associated with the unique role of microtubules<br />

in control over the special cytological features of these FCCs<br />

(Pressel et al. 2006). <strong>The</strong> properties of cavitation-resistant<br />

hydroids <strong>and</strong> desiccation-tolerant leptoids would likely have<br />

had an important impact on these mosses in terms of the ability<br />

to penetrate into diverse ecological niches.<br />

Emergence of xylem with lignified tracheids<br />

<strong>and</strong> vessels<br />

As indicated in Figure 2, xylem tissues may well have evolved<br />

independently from WCCs/hydroids. Although hydroids have a<br />

number of similar features to the early tracheary elements,<br />

including functioning after death, there are many important<br />

differences. Perhaps the most critical was the acquisition<br />

of a developmental program for the deposition of patterned<br />

secondary cell wall material. Of equal importance was the<br />

development of lignin <strong>and</strong> its deposition within the secondary<br />

wall of tracheary cells. Collectively, these evolutionary events<br />

imparted biomechanical support <strong>and</strong> compressive strength,<br />

with an ability to withst<strong>and</strong> tracheid collapse when the water column<br />

was placed under tension (see later section). Acquisition of<br />

biomechanical strength afforded the opportunity for an increase<br />

in plant height, with the benefit of enhanced competition for<br />

sunlight.<br />

A further defining feature of the early vascular plants was that<br />

their tracheary (water conducting) elements had pits of varying<br />

architecture that spanned the secondary wall. In contrast to<br />

the WCCs of the bryophytes, the formation of these pits is not<br />

dependent upon the dissolution of PD (Barnett 1982; Lachaud<br />

<strong>and</strong> Maurousset 1996), <strong>and</strong> in the early tracheary element, the<br />

tracheid (Edwards et al. 1992), the primary cell wall remains


298 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Figure 2. Cladogram illustrating the distribution of water-conducting cells (WCCs) in early l<strong>and</strong> plants.<br />

Note that hydroids in the peristomate mosses lack perforate cell walls. Reproduced from Ligrone et al. (2012), with permission of Oxford<br />

University Press.<br />

imperforate. However, in the advanced form, the vessel element<br />

or vessel member, the primary wall is removed in discrete<br />

regions between adjacent members, thereby giving rise to a<br />

perforation plate. This evolutionary adaptation allows water to<br />

flow through many mature vessel members that collectively<br />

form a vessel, unimpeded by the primary cell wall; i.e., the<br />

perforation plate reduces the overall resistance to water flow<br />

through vessels.<br />

<strong>Evolution</strong>ary relationship between FCCs <strong>and</strong> early<br />

tracheophyte sieve elements<br />

<strong>The</strong> cytological features of FCCs are widespread in the<br />

bryophytes <strong>and</strong> many are also present in the phloem sieve<br />

elements of the lycophytes, pterophytes <strong>and</strong> gymnosperms<br />

(Esau et al. 1953) (Table 1). It is also noteworthy that the ER is<br />

present in PD located in the adjoining transverse walls between<br />

FCCs, leptoids <strong>and</strong> the sieve elements of ferns (Evert et al.<br />

1989) <strong>and</strong> conifers (Schulz 1992). Furthermore, both leptoids<br />

<strong>and</strong> early sieve elements, termed sieve cells, have supporting<br />

parenchyma cells. <strong>The</strong>se features, held in common between<br />

the more advanced FCCs <strong>and</strong> the phloem sieve elements of<br />

the early tracheophytes, raise the possibility of a developmental<br />

program having components shared between these nutrient<br />

delivery systems of the plant kingdom.<br />

<strong>Evolution</strong> of molecular mechanisms regulating<br />

vascular development<br />

Significant progress has been made in elucidating the molecular<br />

mechanisms regulating vascular development. In most<br />

cases, a modest number of angiosperm model species have<br />

been the focus of molecular-genetic <strong>and</strong> genomic analysis<br />

of vascular development. At present, individual genes<br />

regulating specific aspects of vascular development have<br />

been characterized in detail. In addition, models of how<br />

vascular tissues are initiated, patterned, balance proliferation<br />

<strong>and</strong> differentiation, <strong>and</strong> acquire polarity have been<br />

developed.<br />

<strong>Vascular</strong> development is currently being modeled at new<br />

levels of complexity in Arabidopsis <strong>and</strong> Populus, using computational<br />

<strong>and</strong> network biology approaches that make use<br />

of extensive genomic gene expression <strong>and</strong> gene regulation<br />

datasets. While incomplete, new models representing important<br />

phylogentic positions in l<strong>and</strong> plant evolution are also being<br />

developed, <strong>and</strong> will provide important insights into the origins<br />

<strong>and</strong> diversification of mechanisms regulating vascular development.<br />

Importantly, many of the key gene families that regulate<br />

vascular development predate tracheophytes. Thus, one major<br />

challenge for underst<strong>and</strong>ing the evolution of vascular development<br />

will be to determine the evolutionary processes by which


Figure 3. Cytological details of moss food-conducting cells.<br />

(A) Cytoplasmic polarity in a leafy stem of Plagiomnium undulatum;<br />

most of the organelles are in the top end of the lower cell.<br />

(B) Sphorophyte seta of Mnium hornum showing the longitudinal<br />

alignment of elongated plastids (p) <strong>and</strong> the highly elongated nucleus<br />

(n).<br />

(C, D) Longitudinal arrays of microtubules associated with tubules<br />

<strong>and</strong> vesicles in a leafy stem of Plagiomnium undulatum (C) <strong>and</strong><br />

Polytrichum juniperinum (D).<br />

(E) Transverse section of leptoids <strong>and</strong> adjacent hydroids (h) in a<br />

stem of Polytrichum commune.<br />

Scale bars: 4 µm in(A), 2µm in(B, E), 0.5µm in(C, D).<br />

Reproduced from Ligrone et al. (2000), with permission of <strong>The</strong> Royal<br />

Society London.<br />

regulatory genes <strong>and</strong> modules were duplicated, modified, or<br />

directly co-opted to function in vascular development (Pires <strong>and</strong><br />

Dolan 2012). Even more challenging will be determining the<br />

evolutionary steps underlying the many biochemical processes<br />

required for the production of vascular tissues <strong>and</strong> lignified<br />

secondary cell walls.<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 299<br />

Figure 4. Abundant plasmodesmata in the trumpet-shaped end<br />

walls between food-conducting cells in the moss Sphagnum<br />

cuspiatum.<br />

Scale bar: 10 µm. Reproduced from Ligrone et al. (2000), with<br />

permission of <strong>The</strong> Royal Society London.<br />

Table 1. Comparison of cytological features present in moss<br />

food-conducting cells <strong>and</strong> sieve cells in ferns <strong>and</strong> conifers<br />

Similaritiesa Differences (in sieve cells)<br />

Absence of vacuoles No cytoplasmic polarization<br />

Nacreous wallsb Apparent lack of polyribosomes<br />

Nuclear degenerationb Presence of endoplasmic<br />

reticulum (ER) within<br />

plasmodesmata (PD)<br />

Callose associated with PDc a Modified after Ligrone et al. (2000).<br />

b Restricted to the Polytrichales in mosses.<br />

c Restricted to the Polytrichales in mosses, absent in some lower<br />

tracheopytes.<br />

Auxin is an evolutionarily ancient regulator of vascular<br />

development<br />

In the following sections we present some examples of the<br />

genes <strong>and</strong> mechanisms regulating specific aspects of vascular<br />

development. This is not a complete review of the literature,<br />

but rather we aim to highlight some of the molecular-genetic<br />

models of vascular development. We begin with the enigmatic<br />

plant hormone auxin, which has been known to play<br />

fundamental roles in vascular development for decades, but<br />

only recently have insights been gleaned at the moleculargenetic<br />

level as to how it exerts its many influences on vascular<br />

development. To underst<strong>and</strong> the myriad of ways that auxin<br />

influences plant development, it is necessary to underst<strong>and</strong><br />

its synthesis, conjugation, transport, perception, <strong>and</strong> effects<br />

on gene expression. Fundamental insights into all of these<br />

processes have been gained, <strong>and</strong> have been summarized


300 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

in recent reviews. Importantly, auxin can apparently be synthesized<br />

by all plants (Johri 2008; Lau et al. 2009) in which<br />

it plays various roles in promoting growth, <strong>and</strong> has thus<br />

been recruited to participate in vascular development in the<br />

tracheopytes. Here, we consider one specific role for auxin<br />

during vascular development: how auxin transport proteins act<br />

during the establishment <strong>and</strong> propagation of interconnected<br />

vascular str<strong>and</strong>s.<br />

<strong>Vascular</strong> str<strong>and</strong>s consist of interconnected files of cells. It<br />

is critical that vascular str<strong>and</strong>s be properly spaced <strong>and</strong> patterned<br />

within tissues, <strong>and</strong> that functional cell types (e.g. water<br />

conducting tracheary elements) coordinate differentiation to<br />

produce interconnected conduits for transport. Auxin has long<br />

been known to play fundamental roles in both the induction of<br />

vascular str<strong>and</strong>s <strong>and</strong> in the differentiation of vascular cell types<br />

(Sachs 1991). How auxin is transported through tissues has<br />

been recognized as a primary factor in determining the location<br />

<strong>and</strong> propagation (or canalization) of vascular str<strong>and</strong>s. Major<br />

insights into how auxin transport is regulated have been gained<br />

through the identification <strong>and</strong> characterization of PIN-FORMED<br />

(PIN) proteins, which include plasma membrane spanning<br />

auxin efflux carriers. PIN proteins act through asymmetric<br />

subcellular localization to direct routes of auxin flow through<br />

cells <strong>and</strong> tissues (Petráˇsek et al. 2006, Petráˇsek <strong>and</strong> Friml<br />

2009).<br />

Importantly, PIN proteins are found in all l<strong>and</strong> plants (Krecek<br />

et al. 2009) including the pretracheophytes, <strong>and</strong> polar auxin<br />

transport (PAT) already existed in the Charophyta (Boot et al.<br />

2012), suggesting that ancestral functions of PIN proteins were<br />

not in vascular development, per se. <strong>The</strong> fossil record also<br />

provides insights into the role of auxin transport in the evolution<br />

of vascular tissues. Routes of auxin transport can be indirectly<br />

inferred from anatomical changes in extant angiosperms <strong>and</strong><br />

gymnosperms, in which circular patterns of tracheary element<br />

develop in secondary xylem above branches, which impede<br />

auxin transport. Amazingly, fossils of wood from 375-millionyear-old<br />

Archaeopteris (a progymnosperm) also show this<br />

pattern (Rothwell <strong>and</strong> Lev-Yadun 2005).<br />

A general expectation is that the auxin-related mechanisms<br />

regulating vascular differentiation are shared (i.e. are homologous)<br />

among vascular plants, but this remains to be verified<br />

through functional studies in all the major vascular plant lineages.<br />

Perhaps more intriguing will be the characterization of<br />

ancestral auxin-related mechanisms in non-vascular plants <strong>and</strong><br />

the determination of the evolutionary steps through which they<br />

were co-opted <strong>and</strong> modified during vascular plant evolution.<br />

CLASS III HD-ZIP transcription factors<br />

Gene transcription is a major mechanism for regulating<br />

vascular development. One increasingly well characterized<br />

transcriptional module is defined by the Class III homeodomain-<br />

leucine zipper (HD-ZIP) transcription factors. <strong>The</strong> genes encoding<br />

these transcription factors are evolutionarily ancient, <strong>and</strong><br />

are found in all l<strong>and</strong> plants (Floyd et al. 2006). Interestingly,<br />

transcript levels of Class III HD ZIPs are negatively regulated by<br />

miRNAs which are also highly conserved (Floyd <strong>and</strong> Bowman<br />

2004). <strong>The</strong> Class III HD-ZIP gene family exp<strong>and</strong>ed <strong>and</strong> diversified<br />

in l<strong>and</strong> plant lineages, acquiring new expression patterns<br />

<strong>and</strong> functions along the way. Indeed, the functions of Class<br />

III HD-ZIPs in regulating vascular tissues are undoubtedly<br />

derived, since Class III HD-ZIPs predate the appearance of<br />

vasculature in l<strong>and</strong> plant evolution.<br />

In Arabidopsis, the Class III HD-ZIP gene family is comprised<br />

of five genes, REVOLUTA (REV), PHABULOSA (PHB),<br />

PHAVOLUTA (PHV), ATHB8, <strong>and</strong> CORONA/ATHB15. Phylogenetic<br />

<strong>and</strong> functional relationships among them support the<br />

conclusion that ATHB8 <strong>and</strong> ATHB15, <strong>and</strong> REV, PHAB, <strong>and</strong><br />

PHAV represent subclades (Prigge et al. 2005, 2006; Floyd<br />

et al. 2006). Functional relationships among the Class III HD-<br />

ZIPs are complex, <strong>and</strong> different family members have been<br />

implicated in shoot apical meristem formation, lateral organ<br />

initiation, embryo patterning, <strong>and</strong> leaf polarity (Floyd et al.<br />

2006). However, all five Arabidopsis Class III HD-ZIPs have<br />

been implicated in some aspect of vascular development. <strong>The</strong><br />

role for the Class III HD-ZIPs in regulating vascular polarity will<br />

be discussed in depth later in the review.<br />

<strong>The</strong> interplay of transcription <strong>and</strong> hormones is just one of<br />

the many areas ripe for exploration in terms of the evolution<br />

<strong>and</strong> development of vascular biology. Powerful new tools are<br />

available or quickly being developed that will result in dramatic<br />

changes to both the scope <strong>and</strong> level of complexity that can<br />

be addressed in future studies. For example, new sequencing<br />

technologies now allow for comprehensive cataloguing of gene<br />

expression in vascular tissues from virtually any species. <strong>The</strong>se<br />

<strong>and</strong> related genomic technologies are also being used to<br />

provide massive datasets for new computational approaches,<br />

including network biology, which can model the complex interactions<br />

of genes that together regulate fundamental features<br />

of vascular development. Importantly, these new technologies<br />

must be integrated within the framework of paleobotany, plant<br />

anatomy, <strong>and</strong> plant physiology to provide meaningful models of<br />

the evolutionary steps that occurred, at the molecular-genetic<br />

level, to provide the diversity of vascular biology that we see in<br />

extant plants.<br />

Phloem <strong>Development</strong> & Differentiation<br />

Recently, several novel regulatory mechanisms that control the<br />

specification of vascular patterning <strong>and</strong> differentiation have<br />

been uncovered. Through the use of novel genomics <strong>and</strong><br />

molecular techniques in several model plant systems, such as<br />

Arabidopsis, Populus <strong>and</strong> Zinnia, new insights have become


available regarding the regulation of vascular development.<br />

<strong>The</strong> importance of signaling in the control of vascular morphogenesis<br />

has become increasingly apparent. <strong>The</strong> primary agents<br />

involved include well-known phytohormones such as auxin,<br />

cytokinin <strong>and</strong> brassinosteroids, as well as other small regulatory<br />

molecules. This level of underst<strong>and</strong>ing also involves the<br />

transporters <strong>and</strong> receptors of these factors. <strong>The</strong>re is increasing<br />

evidence for the notion that xylem <strong>and</strong> phloem development<br />

are highly coordinated. While many of the factors which will be<br />

described in this section of the review act cell-autonomously,<br />

the emerging importance of mobile regulatory factors will also<br />

be highlighted.<br />

Embryonic provascular <strong>and</strong> shoot vascular<br />

development<br />

During embryogenesis, the progenitor cells that will eventually<br />

become the vascular tissues are first established as undifferentiated<br />

procambial tissues. Surrounded by the epidermal <strong>and</strong><br />

ground tissue layers, this procambial tissue forms the innermost<br />

domain of the plant embryo. By the late globular embryonic<br />

stage, the four procambium cells have undergone periclinal<br />

divisions to generate the future pericycle <strong>and</strong> the vascular<br />

primordium. <strong>The</strong> complete root promeristem with all initials <strong>and</strong><br />

derived cell types is contained already in the early torpedo<br />

stage embryo (Scheres et al. 1995). Asymmetric cell divisions<br />

within the vascular primordium go on to establish the number<br />

of vascular initials present in the seedling root meristem.<br />

In the aerial parts of a seedling, the protovascular elements<br />

are first specified <strong>and</strong> start to differentiate in the cotyledons <strong>and</strong>,<br />

only subsequently, in the axis (Bauby et al. 2007). Protophloem<br />

differentiation can first be observed in the midvein by the<br />

typical cell elongation, followed by extension to the distal<br />

loops <strong>and</strong> cotyledonary node. Before vasculature differentiation,<br />

continuous procambial str<strong>and</strong>s can already be observed<br />

(Figure 5). <strong>The</strong> main agent in the establishment of this pattern<br />

is the phytohormone auxin. <strong>The</strong> accumulation of auxin, through<br />

polar transport mechanisms, as shown by the synthetic auxin<br />

reporter DR5 <strong>and</strong> the auxin-induced pre-procambial marker<br />

AtHB8, precedes the formation of vascular str<strong>and</strong>s in leaves<br />

(Scarpella et al. 2004). Facilitating this highly localized auxin<br />

accumulation is the auxin transporter PIN1, which channels<br />

auxin to the provascular regions. PIN1 is already expressed<br />

before procambium formation.<br />

After the initial differentiation, the vasculature develops in<br />

bundles (Esau 1969). <strong>The</strong>se bundles have a very distinct<br />

radial pattern of abaxial phloem <strong>and</strong> adaxial xylem divided by<br />

an active procambium (Figure 6). <strong>The</strong> radial patterning of the<br />

vascular bundles is already established during embryogenesis<br />

by the main factors of radial patterning, KANADI (KAN) <strong>and</strong> the<br />

Class III HD-ZIPs, PHB, PHV, REV <strong>and</strong> CORONA/ATHB15<br />

(McConnell et al. 2001; Emery et al. 2003). Auxin also plays<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 301<br />

a role in this determination of the radial vascular patterning<br />

(Izakhi <strong>and</strong> Bowman 2007). PIN1 localization is affected in kan<br />

mutants, showing the integration of the auxin transport pathway<br />

<strong>and</strong> KAN signaling. <strong>The</strong> triple mutant phb phv rev has radialized<br />

as well as abaxialized leaves <strong>and</strong> vascular bundles. In contrast,<br />

gain-of-function Class III HD-ZIP mutants with faulty microRNA<br />

(miRNA) regulation <strong>and</strong> kan1 kan2 kan3 have radialized <strong>and</strong><br />

adaxialized leaves <strong>and</strong> bundles. In addition, the Class III HD-<br />

ZIP genes also appear to regulate vascular tissue proliferation.<br />

As will be described in more detail later, the phloem translocation<br />

stream, or phloem sap, contains not only photosynthate<br />

but also a wide array of macromolecules, such as mRNA,<br />

small RNAs <strong>and</strong> proteins. Both in animals <strong>and</strong> plants, small<br />

RNAs have already been identified as important regulatory<br />

factors controlling cell fate. A bidirectional cell-to-cell communication<br />

network involving the mobile transcription factor<br />

SHORTROOT (SHR) <strong>and</strong> microRNA165/166 species specifies<br />

the radial position of two types of xylem vessels in Arabidopsis<br />

roots (Carlsbecker et al. 2010; Miyashima et al. 2011). Since<br />

microRNA165/166 is a factor restricting PHB activity, it also<br />

regulates phloem development.<br />

Recent studies have shown that CALLOSE SYNTHASE<br />

3 (CALS3), a membrane-bound enzyme which synthesizes<br />

callose, a β-1,3-glucan (Verma <strong>and</strong> Hong 2001; Colombani<br />

et al. 2004), appears to be involved in regulating the cell-to-cell<br />

movement of microRNA165/166 (Vatén et al. 2011). CALS3 is<br />

expressed both in phloem <strong>and</strong> meristematic tissues. Gain-offunction<br />

mutations in CALS3 result in increased accumulation<br />

of PD callose, a decrease in the PD aperture, multiple defects<br />

in root development, <strong>and</strong> reduced intercellular trafficking of<br />

various molecules. Using an inducible expression system for a<br />

modified version of CALS3 (CALS3m), Vatén et al. (2011) were<br />

able to show that increased callose deposition inhibited SHR<br />

<strong>and</strong> microRNA165 movement between the stele <strong>and</strong> the endodermis.<br />

This interesting result suggested that regulated callose<br />

biosynthesis, at the PD level, may be essential for control over<br />

cell-to-cell communication <strong>and</strong> cell fate determination.<br />

Root vascular development<br />

In the root, protophloem initially differentiates from an independent<br />

differentiation locus in the upper hypocotyl; only later<br />

does protophloem differentiation from the root apical meristem<br />

begin. As the root grows, the cellular pattern is established <strong>and</strong><br />

maintained by the self-renewal of pluripotent root meristem<br />

cells. Different cell identities are initiated from the stem cells<br />

around the quiescent center (QC): the provascular initials of<br />

the stele, the cortex/endodermal initials, the epidermal/lateral<br />

root initials, <strong>and</strong> the columella initials. In the Arabidopsis root,<br />

a central vascular cylinder (consisting of xylem, phloem <strong>and</strong><br />

procambium) is surrounded by radially symmetric layers of<br />

pericycle, endodermis, cortex <strong>and</strong> epidermal cells. In this root


302 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Figure 5. Schematic of the primary phloem organization in Arabidopsis shoot <strong>and</strong> root.<br />

(A) Longitudinal section through the shoot apex, shoot <strong>and</strong> root.<br />

(B) Cross section of an early developing leaf showing the preprocambial bundles which precede the vascular bundles.<br />

(C) Cross section of a leaf showing established vascular bundles where primary phloem <strong>and</strong> xylem differentiate asymmetrically from a<br />

separating layer of procambium.<br />

(D) Cross section of the stem showing primary vascular bundles.<br />

(E) Cross section of the root showing the primary vascular patterning with two phloem poles consisting of sieve elements <strong>and</strong> companion<br />

cells flanking the xylem axis.<br />

(F) Cross section of the root tip showing two poles of protophloem sieve elements flanking the xylem axis.<br />

system, the vascular tissue is comprised of a central axis of<br />

water-conductive xylem tissue that is flanked by two poles of<br />

photoassimilate-conductive phloem tissue.<br />

In the root apical meristem, the phloem cell lineages arise<br />

from two domains of initials through asymmetric cell divisions<br />

(Mähönen et al. 2000). Periclinal divisions establish companion<br />

cells (CCs) <strong>and</strong> tangential divisions establish sieve elements<br />

(SEs). This asymmetry allows these initials to give rise to<br />

multiple cell lineages with different fates; in addition to the<br />

phloem lineage, they also precede undifferented procambial<br />

cell lineages. As opposed to the invariant pattern of cell<br />

lineages in the endodermis <strong>and</strong> outer layers, the number<br />

<strong>and</strong> exact pattern of these procambial divisions vary between<br />

individual seedlings.<br />

Mähönen et al. (2000) have described these initial asymmetric<br />

cell divisions in great detail through sequential<br />

cross-sections of the region immediately above the quiescent<br />

center, allowing the precise determination of the first true<br />

phloem domains. At first, although cell divisions <strong>and</strong> early xylem<br />

specification can be observed (though not yet the complete


Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 303<br />

Figure 6. <strong>Vascular</strong> patterning is regulated by KANADI <strong>and</strong> Class III HD-ZIP genes <strong>and</strong> the phytohormones auxin <strong>and</strong> cytokinin.<br />

(A) In shoot vascular bundles, the default radial pattern has phloem located abaxially <strong>and</strong> xylem adaxially.<br />

(B) In the phb phv rev mutant, phloem surrounds xylem.<br />

(C) Conversely, in Class III HD-ZIP gain-of-function mutants <strong>and</strong> kan1 kan2 kan3, xylem surrounds phloem.<br />

(D) Class III HD-ZIP genes are regulated by miR165/166 <strong>and</strong> interact with auxin <strong>and</strong> brassinosteroids.<br />

(E) In the root, auxin is restricted to the xylem axis by the presence of cytokinin.<br />

(F) In mutants with defective cytokinin signaling such as wol, auxin is abundant throughout the stele, leading to ubiquitous protoxylem<br />

differentiation <strong>and</strong> loss of phloem identity.<br />

xylem axis), phloem identity is not observable directly above<br />

the QC. <strong>The</strong> first newly formed cell walls associated with<br />

phloem development are only visible 27 µm above the QC.<br />

At a distance of 69 µm, the first protophloem SEs are clearly<br />

present.<br />

Hormonal balance determines the development<br />

of vascular poles in the root<br />

Cytokinin, an essential phytohormone for development in the<br />

root, is required for vascular patterning <strong>and</strong> the differentiation<br />

of all cell types except the protoxylem. Recently, it has been<br />

shown that the root vascular pattern is defined by a mutually<br />

inhibitory interaction between cytokinin <strong>and</strong> auxin (Bishopp<br />

et al. 2011a, 2011b). If cytokinin signaling is disturbed, as<br />

in the WOODEN LEG mutant, wol or the triple cytokinin<br />

receptor mutant ahk2 ahk3 ahk4 (ARABIDOPSIS HISTIDINE<br />

KINASE), or if cytokinin levels are reduced, as is found in<br />

transgenic plants overexpressing CYTOKININ OXIDASE, the<br />

effect is always an increased number of protoxylem cell files<br />

<strong>and</strong> the loss of other cell types in the root vasculature. <strong>The</strong><br />

domain of cytokinin activity is restricted by the action of<br />

the cytokinin signaling inhibitor, ARABIDOPSIS HISTIDINE<br />

PHOSPHOTRANSFER PROTEIN 6 (AHP6). Only through this<br />

mechanism does protoxylem differentiation occur in a spatially<br />

specific manner, allowing for the proper development of the<br />

phloem cell types.


304 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Phloem differentiation<br />

Sieve elements comprise the main conductive tissue of the<br />

phloem. Like the CCs, they originate from phloem precursor<br />

cells in the procambium. However, very early in primary phloem<br />

development, they undergo dramatic changes in their morphology.<br />

As the SEs mature, they experience extensive degradation<br />

of their organelles. <strong>The</strong> nucleus, vacuoles, rough endoplasmic<br />

reticulum (ER) <strong>and</strong> Golgi are degraded in a process which<br />

has not yet been characterized at the molecular level. This<br />

reduction in cellular contents establishes an effective transport<br />

route through the sieve tubes. However, the SEs still remain<br />

living, as they retain a plasma membrane <strong>and</strong> a reduced<br />

number of other organelles, such as smooth ER, plastids <strong>and</strong><br />

mitochondria. <strong>The</strong> residual ER is localized near the PD which<br />

interconnect the SEs to their neighboring CCs.<br />

<strong>The</strong> cell walls of the SEs also undergo drastic changes in<br />

structure. <strong>The</strong> first observable process is an increase in callose<br />

that is deposited in platelet form around the PD of the SEs,<br />

replacing the already present cellulose. <strong>The</strong> cell walls which<br />

form the interface to adjoining SEs contain a high density<br />

of these callose-ensheathed PD. As these SEs mature, both<br />

these callose deposits <strong>and</strong> the middle lamella in these regions<br />

of the cell wall are removed, thereby forming a sieve plate<br />

with enlarged pores (Lucas et al. 1993). <strong>The</strong> lateral cell walls<br />

of SEs also develop specialized areas of PD-derived pores,<br />

which are called lateral sieve areas. Recent studies have<br />

shown that CALS3 <strong>and</strong> CALS7 are involved in depositing PDcallose<br />

during this developmental process (Vatén et al. 2011;<br />

Xie et al. 2011). <strong>The</strong> newly formed sieve plates, in combination<br />

with the lateral sieve areas, enable each individual SE to<br />

become a component of an integrated sieve tube system that<br />

can facilitate effective fluid transport by bulk flow. It is also<br />

noteworthy that these pores increase considerably in size as<br />

tissues age, thus increasing the transport potential of the more<br />

mature vasculature (Truernit et al. 2008).<br />

<strong>The</strong> survival <strong>and</strong> differentiation of SEs depends on a close<br />

association with their neighboring CCs, a specialized type of<br />

parenchyma cell. <strong>The</strong> cytoplasm of the CC is unusually dense,<br />

due in part to an increased number of plastids, mitochondria<br />

<strong>and</strong> free ribosomes (Cronshaw 1981). <strong>The</strong> CCs are connected<br />

to their adjacent SEs by numerous branched PD. Through<br />

these connections, the enucleate SEs are supplied with energy,<br />

assimilates <strong>and</strong> macromolecular compounds, such as proteins<br />

<strong>and</strong> RNA (Raven 1991; Lough <strong>and</strong> Lucas 2006). <strong>The</strong> size<br />

exclusion limit of these PD connections usually lies between<br />

10 <strong>and</strong> 40 kDa (Kempers <strong>and</strong> van Bel 1997), giving credence<br />

to the concept of protein transport from CCs to SEs.<br />

<strong>The</strong> morphological <strong>and</strong> physiological uniqueness of the<br />

phloem cell types described above is also a result of specific<br />

gene expression patterns, as shown by recent transcriptome<br />

studies (Lee et al. 2006; Brady et al. 2007, 2011). <strong>The</strong>se<br />

transcriptional programs are exquisitely controlled in space<br />

<strong>and</strong> time. To underst<strong>and</strong> how these unique cell identities<br />

are acquired, a deeper underst<strong>and</strong>ing of these programs is<br />

absolutely essential. Microarray analyses of a high-resolution<br />

set of developmental time points, <strong>and</strong> a comprehensive set of<br />

cell types within the root, has resulted in the most detailed root<br />

expression map to date.<br />

Numerous distinct expression patterns have been identified<br />

through these analyses, several of which are specific to SEs<br />

or to SEs <strong>and</strong> CCs together. More than a thous<strong>and</strong> genes<br />

have been identified as having phloem-specific expression,<br />

highlighting the phloem as a highly specialized tissue within<br />

the stele. <strong>The</strong> data from these microarray studies has been<br />

made publicly available in the AREX LITE: <strong>The</strong> Arabidopsis<br />

Gene Expression Database (arexdb.org), providing an invaluable<br />

resource for future studies on phloem function.<br />

Bauby et al. (2007) identified several phloem markers, which<br />

they named Phloem Differentiation 1–5 (PD1-PD5) by screening<br />

the Versailles collection of gene trap mutants for plant lines<br />

expressing the uidA reporter gene in immature vascular tissues.<br />

PD1-4 were restricted to protophloem cells, as determined by<br />

their cell shape. However, PD5 was expressed in both protophloem<br />

<strong>and</strong> metaphloem cells. Using these markers, these<br />

authors could track the onset of phloem development directly<br />

after embryogenesis. PD4 is expressed in the tips of leaf<br />

primordia some 3 days after germination (dag). Spreading from<br />

there, by 7 dag, PD4 has already traced out the entire future<br />

leaf vasculature. <strong>The</strong> first expression of PD1 <strong>and</strong> PD3 was<br />

detected 3 dag in the proximal protophloem of leaf primordia.<br />

<strong>The</strong>se results establish that PD1 <strong>and</strong> PD3 are expressed<br />

during the differentiation of protophloem. PD4 <strong>and</strong> PD5 gene<br />

reporter-based expression was also detected in the location<br />

of the midveins <strong>and</strong> higher order veins before procambium<br />

differentiation, thereby defining the pre-patterning of the future<br />

veins.<br />

Regulation of phloem differentiation<br />

Currently, only two factors are known which specify phloem<br />

identity: ALTERED PHLOEM DEVELOPMENT (APL) <strong>and</strong> OC-<br />

TOPUS (OPS). APL is a MYB coiled-coil transcription factor<br />

essential for the proper differentiation of both SEs <strong>and</strong> CCs<br />

(Bonke et al. 2003). Additionally, APL contributes to the spatial<br />

limiting of xylem differentiation, is expressed both in SEs <strong>and</strong><br />

CCs, <strong>and</strong> has been shown to be nuclear localized. Loss of<br />

APL function has an extraordinary effect on the phloem, as<br />

can be observed in the loss-of-function mutant apl1 (Figure 7).<br />

This mutant is seedling-lethal <strong>and</strong> results in short-rooted plants.<br />

Neither CCs nor SEs can be detected in cross-sections of<br />

apl1 plants. Furthermore, phloem-specific reporters, such as<br />

the CC-specific sucrose transporter, SUC2, or the proto SE<br />

reporter, J0701, cease to be expressed entirely in these mutant


Figure 7. Model of OCTOPUS (OPS) <strong>and</strong> ALTERED PHLOEM DEVELOPMENT (APL) action.<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 305<br />

(A) OPS is located at the apical plasma membrane in the procambium <strong>and</strong> phloem lineage. OPS interprets vascular signals for phloem<br />

differentiation, such as APL.<br />

(B) In the ops mutant, phloem differentiation is delayed. Procambial cell number is increased <strong>and</strong> gaps of undifferentiated cells are visible in<br />

the protophloem str<strong>and</strong>.<br />

(C) In the apl mutant, the initiation of phloem differentiation is largely unperturbed, however proper protophloem fails to emerge. In their<br />

place, protophloem/protoxylem hybrid cells appear.<br />

plants. This dramatic effect is restricted to the phloem poles;<br />

other cell types appear similar to wild-type (WT) plants.<br />

As mentioned above, asymmetrical cell divisions establish<br />

the phloem poles in wild-type plants; periclinal divisions establish<br />

CCs <strong>and</strong> tangential divisions establish SEs. In the apl<br />

mutant, these cell divisions are often delayed, but they still<br />

take place, so APL does not appear to be required for these<br />

asymmetric cell divisions. However, since the subsequent<br />

differentiation of SEs cannot be observed, it can be concluded<br />

that APL is responsible for phloem differentiation, rather than<br />

the establishment of the phloem cell lineage.<br />

It has also been proposed that APL acts as an inhibitor to<br />

xylem differentiation. In the apl mutant, ectopic xylem str<strong>and</strong>s<br />

are seen in the place of the phloem poles. When APL is<br />

expressed ectopically in vascular bundles, xylem formation is<br />

inhibited. Recently, novel imaging techniques were employed<br />

to analyze the apl mutant in more detail (Truernit et al. 2008).<br />

With this increased resolution, it was discovered that protophloem<br />

differentiation proceeds normally in this mutant until<br />

2 dag. At this time, cells in the protophloem position display the<br />

normal characteristic shape <strong>and</strong> cell wall thickening. However,<br />

after this period, the previously described acquisition of xylem<br />

characteristics was observed, although the cells in place of<br />

the SEs still formed sieve plates. This finding suggests that<br />

these cells can be classified as hybrids between phloem <strong>and</strong><br />

xylem (Truernit et al. 2008). Thus, APL is absolutely required<br />

for the later stages of phloem development, although it is<br />

not the earliest factor acting in this process. <strong>The</strong> identity of<br />

such a factor, or factors, has yet to be determined. Additional<br />

support for the presence of such factor(s) was provided by Lee<br />

et al. (2006), who pointed out that there are numerous phloem<br />

markers with earlier SE expression than APL.


306 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

In addition to APL, OPS has also been identified as a gene<br />

related to phloem cell differentiation; this gene is required<br />

for phloem continuity during phloem development (Truernit<br />

et al. 2012). Interestingly, OPS was first reported based on<br />

its vascular expression pattern (Nagawa et al. 2006). Further<br />

detailed analysis revealed that its expression is initially in the<br />

provascular cells at the heart stage of embryo development,<br />

<strong>and</strong> it subsequently becomes restricted to the phloem lineage<br />

cells following phloem cell-type specification (Bauby et al. 2007;<br />

Truernit et al. 2012). Unlike APL, OPS expression can be<br />

observed in the phloem <strong>and</strong> procambium initials near the QC.<br />

<strong>Vascular</strong> patterning of the cotyledons, in the mature embryo<br />

of the ops mutant, is delayed, <strong>and</strong> the number of completed<br />

vascular loops in the developing cotyledon is reduced (Truenit<br />

et al. 2012). In contrast, the progression of vascular patterning<br />

is accelerated by OPS overexpression in the cotyledons of<br />

mature embryos, <strong>and</strong> the number of completed vascular loops<br />

in the developing cotyledon is increased. This suggests that<br />

OPS is involved in promoting the progression of vascular<br />

patterning. Furthermore, Truernit et al. (2012) found that, in ops<br />

hypocotyls <strong>and</strong> roots, the phloem SE cell file was interrupted<br />

by undifferentiated SEs, which failed to undergo an increase<br />

in cell wall thickness, callose deposition, or nuclear degradation;<br />

these cells failed to acquire SE-specific PD1 marker<br />

expression (Truernit et al. 2012). <strong>The</strong>se cellular differentiation<br />

defects caused inefficient phloem transport in the root. <strong>The</strong>se<br />

phenotypes indicated that OPS was required for continuous<br />

phloem development.<br />

OPS encodes a membrane-associated protein (Benschop<br />

et al. 2007) specific to higher plants (Nagawa et al. 2006). <strong>The</strong><br />

function of OPS is currently unknown; no functional domain<br />

has been identified in this protein. However, a functionally<br />

complementing OPS-GFP protein was located at the apical<br />

end of the SEs (Truernit et al. 2012). Inductive cell-to-cell communication<br />

from differentiated to undifferentiated neighboring<br />

cells is known to occur during xylem differentiation; XYLO-<br />

GEN is a polar-localized proteoglycan-like factor required to<br />

direct continuous xylem differentiation in Zinnia elegans L. <strong>and</strong><br />

A. thaliana (Motose et al. 2004). It is thought that, in a similar<br />

manner, OPS contributes to longitudinal signaling, thereby<br />

inducing SE differentiation in undifferentiated SE precursor<br />

cells. Further study of OPS function <strong>and</strong> identification of factors<br />

relating to OPS function will advance our underst<strong>and</strong>ing of how<br />

phloem continuity is organized <strong>and</strong> how the phloem develops.<br />

<strong>The</strong> Arabidopsis LATERAL ROOT DEVELOPMENT 3<br />

(LRD3) is another important gene which has been reported<br />

to regulate early phloem development <strong>and</strong> to control transport<br />

function of phloem (Ingram et al. 2011). <strong>The</strong> LRD3 gene<br />

encodes a LIM-domain protein which is specifically expressed<br />

in the CCs. <strong>The</strong> normal function of LRD3 is to maintain a<br />

balance between primary <strong>and</strong> lateral root growth <strong>and</strong> phloemmediated<br />

resource allocation within the root system. <strong>The</strong> lrd3<br />

loss of function mutant has decreased primary <strong>and</strong> increased<br />

lateral root growth <strong>and</strong> density, without having a significant<br />

effect on sucrose uptake. Additionally, aniline blue staining<br />

of the lrd3 primary root shows an overall reduction in the<br />

callose level in root meristems, developing SEs, <strong>and</strong> the PD<br />

that connect CCs to SEs, suggesting a non-cell-autonomous<br />

role for LRD3 in early phloem development.<br />

A detailed analysis of long-distance transport using different<br />

transport assays, including 14 C-sucrose, the fluorescent tracer<br />

dye carboxyfluorescein diacetate (CFDA) <strong>and</strong> CC-driven GFP<br />

(AtSUC2::GFP), demonstrated that phloem translocation to<br />

the primary root tip is severely limited in young lrd3 plants,<br />

whereas phloem loading <strong>and</strong> export from the shoot appear to<br />

be normal. Notably, these phloem defects were subsequently<br />

rescued, spontaneously, in older plants, along with a subsequent<br />

increase in phloem delivery to <strong>and</strong> growth of the primary<br />

root. Importantly, continuous exogenous auxin treatment could<br />

rescue the early phloem developmental defects <strong>and</strong> transport<br />

function in the primary roots of lrd3. This finding suggested<br />

either that auxin functions downstream of LRD3, or that it may<br />

have an independent key role in early phloem development.<br />

Interestingly, this study of the effects of lrd3 on root system<br />

architecture <strong>and</strong> the pattern of phloem translocation in the root<br />

system suggests that there might be some tightly regulated<br />

mechanism(s) which selectively supports a biased phloemmediated<br />

resource allocation in the lateral roots when the<br />

primary root is compromised.<br />

Phloem: A conduit for delivery of photosynthate <strong>and</strong><br />

information molecules<br />

Phloem is an important organ, vital for more than just the wellestablished<br />

function of photoassimilate transport from the photosynthetic<br />

organs to the sink tissues. New transport functions<br />

continue to be discovered, such as the phloem-based transport<br />

of phytohormones, small RNAs, mRNAs <strong>and</strong> proteins. As will<br />

be discussed later, this transport of macromolecules appears<br />

to play a role in facilitating the coordinated developmental<br />

programs in meristematic regions located at various locations<br />

around the body of the plant.<br />

In addition to its specialized transport functions, phloem also<br />

st<strong>and</strong>s out by the highly distinctive morphology of its cell types.<br />

In the angiosperms, the interaction between the enucleate SEs<br />

<strong>and</strong> the CCs needs to be highly integrated in order to maintain<br />

the operation of the sieve tube system (van Bel 2003). This<br />

process likely involves the cell-to-cell trafficking of a wide range<br />

of molecules via the PD that interconnect the SE-CC complex.<br />

Future studies on the nature of this molecular exchange will be<br />

greatly assisted by the use of the modified CALS3m system<br />

(Vatén et al. 2011) which can serve as an effective tool to<br />

modulate transport between specific cell types.


Molecular Mechanisms Underlying Xylem<br />

Cell Differentiation<br />

In the shoot apical meristem, stem cells differentiate into various<br />

cell types that comprise the shoot, while still proliferating<br />

in order to maintain themselves (Weigel <strong>and</strong> Jürgens 2002).<br />

Similarly, in the vascular meristem, procambial <strong>and</strong> cambial<br />

cells differentiate into specific vascular cells, such as tracheary<br />

elements, xylem fiber cells, xylem parenchyma cells, SEs, CCs,<br />

phloem parenchyma <strong>and</strong> phloem fiber cells, while again maintaining<br />

activity to proliferate (Figure 8). <strong>The</strong>refore, procambial<br />

<strong>and</strong> cambial cells are considered as vascular stem cells (Hirakawa<br />

et al. 2010, 2011; Miyashima et al. 2012). Recent studies<br />

have revealed that local communication between vascular<br />

stem cells <strong>and</strong> differentiated vascular cells directs the wellorganized<br />

formation of vascular tissues (Lehesranta et al. 2010;<br />

Hirakawa et al. 2011). During this vascular formation, plant<br />

hormones, including auxin, cytokinin <strong>and</strong> brassinosteroids, act<br />

as signaling molecules that mediate in this process of cell-cell<br />

communication (Fukuda 2004). In addition, recently, a tracheary<br />

element differentiation inhibitory factor (TDIF), a small<br />

peptide, was found to function as a signaling molecule both inhibiting<br />

xylem cell differentiation from procambial cells <strong>and</strong> promoting<br />

procambial cell proliferation (Ito et al. 2006; Hirakawa<br />

et al. 2008, 2010). TDIF belongs to the CLAVATA3/EMBRYO<br />

SURROUNDING REGION-related (CLE) family, some of<br />

whose members are central players in cell-cell communication<br />

within meristems (Diévart <strong>and</strong> Clark 2004; Matsubayashi <strong>and</strong><br />

Sakagami 2006; Fiers et al. 2007; Fukuda et al. 2007; Jun et al.<br />

2008; Butenko et al. 2009; Betsuyaku et al. 2011).<br />

Further insight into the differentiation of procambial cells<br />

into xylem cells has been gained from recent comprehensive<br />

gene expression <strong>and</strong> function analyses (Kubo et al. 2005;<br />

Zhong et al. 2006; Yoshida et al. 2009; Ohashi-Ito et al. 2010;<br />

Yamaguchi et al. 2011). In particular, the discovery of master<br />

genes that induce differentiation of various xylem cells greatly<br />

enhanced our underst<strong>and</strong>ing of xylem formation (Kubo et al.<br />

2005; Zhong et al. 2006; Mitsuda et al. 2007). Further analysis<br />

of these master genes revealed transcriptional networks that<br />

control xylem cell differentiation, which involves specialized<br />

secondary wall formation. Tracheary element differentiation<br />

also involves programmed cell death (PCD) (Fukuda 2000). In<br />

this section of the review, we will evaluate advances in our underst<strong>and</strong>ing<br />

of xylem cell differentiation from procambial cells,<br />

with a focus on cell-cell signaling, the underlying transcriptional<br />

network <strong>and</strong> the onset of PCD.<br />

Intercellular signaling pathways regulating xylem<br />

differentiation<br />

<strong>The</strong> TDIF-TDR signaling pathway regulates vascular stem cell<br />

maintenance. TDIF is a CLE-family peptide composed of twelve<br />

Figure 8. Xylem cells.<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 307<br />

(A) Poplar vascular tissue in which tracheary elements (TE) <strong>and</strong><br />

xylem fiber (XF) cells are formed.<br />

(B) VND6-induced tracheary elements.<br />

(C) SND1-induced xylem fiber cells.<br />

Scale bars: 25 µm in(B) <strong>and</strong> (C).<br />

amino acids with hydroxylation on two proline residues (Ito<br />

et al. 2006). In the Arabidopsis genome, this TDIF sequence<br />

is encoded by two genes, CLE41 <strong>and</strong> CLE44 (Hirakawa<br />

et al. 2008). <strong>The</strong> TDIF RECEPTOR/PHLOEM INTERCALATED<br />

WITH XYLEM (TDR/PXY) is a receptor for TDIF, which belongs<br />

to the Class XI LEUCINE-RICH REPEAT RECEPTOR-LIKE<br />

KINASE (LRR-RLK) family (Hirakawa et al. 2008).<br />

Genetic <strong>and</strong> physiological analyses have revealed that the<br />

TDIF-TDR signaling pathway is crucial for vascular stem cell<br />

maintenance, by inhibiting xylem differentiation from procambial<br />

cells <strong>and</strong> promoting procambial cell proliferation (Hirakawa<br />

et al. 2008) (Figure 9). TDR is expressed preferentially in the<br />

procambium <strong>and</strong> cambium (Fisher <strong>and</strong> Turner 2007; Hirakawa<br />

et al. 2008), whereas CLE41 <strong>and</strong> CLE44 are expressed<br />

specifically in the phloem <strong>and</strong> more widely in its neighbors,<br />

respectively. Defects in TDR or CLE41 cause the exhaustion<br />

of procambial cells located between the phloem <strong>and</strong> xylem,<br />

resulting in formation of xylem vessels adjacent to phloem<br />

cells in the hypocotyl (Fisher <strong>and</strong> Turner 2007; Hirakawa et al.<br />

2008, 2010). Ectopic expression of CLE41, under either a


308 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Figure 9. Regulation of procambial cell fates by the tracheary<br />

element differentiation inhibitory factor (TDIF) –TDIF receptor<br />

(TDR) signaling pathway.<br />

TDIF is produced in phloem cells, secreted from phloem cells, <strong>and</strong><br />

perceived by TDR in procambial cells. TDR signaling is diverged<br />

into two pathways: one promotes self-renewal via WOX4, <strong>and</strong> the<br />

other inhibits tracheary element (TE) differentiation from procambial<br />

cells, probably indirectly via the suppression of VND6/VND7.<br />

xylem-specific or 35S promoter, disrupts the normal pattern<br />

of vascular tissues, indicating the importance of the synthesis<br />

site for this signaling peptide (Etchells <strong>and</strong> Turner 2010). Thus,<br />

phloem-synthesizing TDIF regulates procambial cell fate in a<br />

non-cell-autonomous fashion.<br />

<strong>The</strong> TDIF peptide signal activates expression of WOX4,<br />

a member of the WUSCHEL-related HOMEOBOX (WOX)<br />

gene family, in procambial <strong>and</strong> cambial cells (Hirakawa et al.<br />

2010; Ji et al. 2010; Suer et al. 2011). Interestingly, WOX4<br />

is required for TDIF-dependent enhancement of procambial<br />

cell proliferation, but not for the TDIF-dependent suppression<br />

of xylem differentiation (Hirakawa et al. 2010). Ethylene/ERF<br />

signaling is reported to be another pathway to regulate procambial/cambial<br />

cell division <strong>and</strong> may function in parallel to the<br />

CLE41-TDR/PXY pathway, <strong>and</strong>, under normal circumstances,<br />

TDR/PXY signaling acts to repress the ethylene/ERF pathway<br />

(Etchells et al. 2012). Hence, at least two intracellular signaling<br />

pathways that diverge after TDIF recognition by TDR<br />

may regulate, independently, the behavior of vascular stem<br />

cells. Lastly, TDIF, which is produced mainly by CLE42, has<br />

also been shown to play a role in axillary bud formation in<br />

Arabidopsis, indicating that it is a multifunctional peptide signal<br />

in plants (Yaginuma et al. 2011).<br />

CLE peptides inhibit protoxylem vessel formation through<br />

activating cytokinin signaling. Cytokinin is a key regulator of<br />

xylem development (Mähönen et al. 2000, 2006; Mok <strong>and</strong> Mok<br />

2001; Matsumoto-Kitano et al. 2008; Bishopp et al. 2011b).<br />

Recent studies have revealed crosstalk between CLE peptide<br />

<strong>and</strong> cytokinin signaling, which regulates xylem differentiation<br />

(Kondo et al. 2011). In roots, TDIF does not significantly<br />

affect vascular development (Kondo et al. 2011). In contrast,<br />

treatment with some CLE peptides, including CLE9/CLE10,<br />

inhibits formation of protoxylem but not of metaxylem vessels in<br />

Arabidopsis roots. CLE9 <strong>and</strong> CLE10, which encode the same<br />

CLE peptide, are preferentially expressed in vascular cells of<br />

roots (Kondo et al. 2011). Microarray analysis revealed that the<br />

CLE9/CLE10 peptide specifically reduces expression of type-A<br />

ARABIDOPSIS RESPONSE REGULATERs (ARRs) which are<br />

known as negative regulators of cytokinin signaling (Kiba et al.<br />

2003; To et al. 2004, 2007).<br />

<strong>The</strong> ARR5 <strong>and</strong> ARR6 are particular CLE9/CLE10 targets<br />

<strong>and</strong>, consistent with this finding, in the root of arr5arr6 double<br />

mutant plants, protoxylem vessel formation is often inhibited<br />

(Kondo et al. 2011). Conversely, arr10arr12, a double mutant<br />

for two type-B ARRs, which function positively in cytokinin<br />

signaling, displayed ectopic protoxylem vessel formation. Furthermore,<br />

arr10arr12 was resistant to the CLE9/CLE10 peptide<br />

in terms of protoxylem vessel formation. Interestingly,<br />

other combinations of type-B ARR mutants, such as arr1arr12<br />

<strong>and</strong> arr1arr10, showed much weaker resistance against the<br />

CLE9/CLE10 peptide compared with arr10arr12. This result<br />

implies that ARR10 <strong>and</strong> ARR12 act as major Type-B ARRs.<br />

Thus, the CLE9/CLE10 peptide activates cytokinin signaling<br />

through the repression of ARR5 <strong>and</strong> ARR6, resulting in the<br />

inhibition of protoxylem vessel formation. Genetic analysis<br />

suggests that the CLV2 membrane receptor <strong>and</strong> its partner<br />

CRN/SOL2 kinase (Miwa et al. 2008; Müller et al. 2008)<br />

may act in protoxylem vessel formation, downstream of the<br />

CLE9/CLE10 peptide signaling (Kondo et al. 2011).<br />

For cell-to-cell communication, plant cells send signaling<br />

molecules via the symplasmic pathway. A GRAS-family transcription<br />

factor, SHR, is a signal that moves cell to cell selectively<br />

through PD. SHR proteins are known to move from the<br />

stele into the endodermis to induce another GRAS-family transcription<br />

factor, SCARECROW (SCR), <strong>and</strong> then, together with<br />

SCR, they up-regulate expression of target genes, including the<br />

miR165/166 genes (Levesque et al. 2006; Cui et al. 2007; Gallagher<br />

<strong>and</strong> Benfy 2009). <strong>The</strong> mature miR165/166 moves back<br />

from the endodermis into the pericyle <strong>and</strong> protoxylem vessel<br />

poles in the stele, most likely through PD. Here, miR165/166<br />

degrades the transcripts of PHB <strong>and</strong> its family of Class III HD-<br />

ZIP genes (Carlsbecker et al. 2010). <strong>The</strong>se transcripts within


xylem precursors specify central metaxylem vessels, at high<br />

levels, <strong>and</strong> peripheral protoxylem vessels, at low levels. This<br />

reciprocal signaling between the inner vascular tissues <strong>and</strong> the<br />

surrounding cell layers allows the domain of response to be<br />

confined to one of two tissue compartments (Scheres 2010).<br />

Genome-wide analysis revealed the presence of direct targets<br />

of SHR not only in the endodermis but also in the xylem <strong>and</strong><br />

pericycle, suggesting a complex function of this transcription<br />

factor in vascular development (Cui et al. 2012).<br />

<strong>The</strong>rmospermine, a structural isomer of spermine (Ohsima<br />

1979), has been shown to act as a suppressor of xylem development.<br />

ACAULIS 5 (ACL5) encodes a thermospermine synthase<br />

(Kakehi et al. 2008) that is expressed specifically in early<br />

developing vessel elements (Muñiz et al. 2008). ACL5 lossof-function<br />

mutants cause excessive differentiation of xylem<br />

cells (Hanzawa et al. 1997; Clay <strong>and</strong> Nelson 2005; Muñiz et al.<br />

2008). Exogenously applied themospermine suppresses xylem<br />

vessel differentiation in both Arabidopsis plants <strong>and</strong> a Zinnia<br />

xylogenic culture (Kakehi et al. 2010). Genetic analysis of acl5<br />

identified a suppressor of the acl5 phenotype, sac51, whose<br />

causal gene encodes a basic helix-loop-helix (bHLH) transcription<br />

factor (Imai et al. 2006). <strong>The</strong>rmospermine is considered<br />

to regulate translational activity of SAC51 mRNA, resulting<br />

in the suppression of xylem development (Imai et al. 2008).<br />

A recent chemical biology approach also indicated that the<br />

SAC51-mediated thermospermine signaling pathway can limit<br />

auxin mediated promotion of xylem differentiation (Yoshimoto<br />

et al. 2012). Thus, the possibility exists that ACL5 may control<br />

xylem specification through the prevention of premature cell<br />

death (Muñiz et al. 2008; Vera-Sirera et al. 2010).<br />

Transcriptional regulation of xylem cell differentiation<br />

<strong>The</strong> Class III HD-ZIP genes have been shown to regulate<br />

xylem differentiation. In the phb phv rev cna athb8 mutant<br />

background, procambial cells fail to differentiate into xylem<br />

cells, but proliferate actively to produce many procambium<br />

cells. However, every quadruple loss-of-function mutant of the<br />

five Class III HD-ZIP genes exhibits ectopic xylem formation<br />

in the roots (Carlsbecker et al. 2010). In contrast, a gain-offunction<br />

mutant of PHB induces ectopic metaxylem vessel<br />

formation (Carlsbecker et al. 2010), <strong>and</strong> overproduction of<br />

ATHB8 promotes xylem differentiation (Baima et al. 1995).<br />

<strong>The</strong>se findings indicate that the Class III HD-ZIP members function<br />

positively in xylem specification. However, the regulation<br />

of xylem differentiation by these genes is more complicated.<br />

In roots, miR165/166, which degrades Class III HD-ZIP transcripts,<br />

promotes protoxylem vessel differentiation. <strong>The</strong>refore,<br />

it is proposed that high transcript levels of these genes inhibit<br />

protoxylem vessel formation, but promote metaxylem vessel<br />

formation. Because exogenously applied brassinosteroids promote<br />

the expression of Class III HD-ZIP genes (Ohashi-Ito <strong>and</strong><br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 309<br />

Fukuda 2003) <strong>and</strong> xylem cell differentiation (Yamamoto et al.<br />

1997), brassinosteroids may promote xylem differentiation, at<br />

least partly, through activation of these genes. <strong>The</strong>se Class III<br />

HD-ZIP <strong>and</strong> KANADI transcription factors were also reported to<br />

regulate cambium cell differentiation, in which KANADI might<br />

act by inhibiting auxin transport <strong>and</strong> Class III HD-ZIPs by<br />

promoting xylem differentiation (Ilegems et al. 2010; Robischon<br />

et al. 2011).<br />

Members of a subgroup of NAM/ATAF/CUC (NAC) domain<br />

proteins, namely the VASCULAR-RELATED NAC-DOMAINs<br />

(VNDs) <strong>and</strong> NAC SECONDARY WALL THICKENING PRO-<br />

MOTING FACTORs/SECONDARY WALL-ASSOCIATED NAC<br />

DOMAIN PROTEINs (NSTs/SNDs), function as master transcription<br />

factors that can induce xylem cell differentiation by<br />

their ectopic expression (Demura <strong>and</strong> Fukuda 2007; Zhong <strong>and</strong><br />

Ye 2007). VND6 <strong>and</strong> VND7 initiate metaxylem <strong>and</strong> protoxylem<br />

vessel differentiation, respectively (Kubo et al. 2005). Similarly,<br />

SND1/NST3 <strong>and</strong> NST1 induce xylem fiber differentiation<br />

(Mitsuda et al. 2005, 2007; Zhong et al. 2006). However, a<br />

single loss-of-function mutant of each gene shows no morphological<br />

defects, suggesting that other family members may<br />

have redundant functions to induce xylem differentiation, although<br />

each does not induce xylem cell differentiation when<br />

overexpressed (Kubo et al. 2005).<br />

<strong>The</strong> activity of these master transcription factors appears to<br />

be regulated by the following three mechanisms. (1) Expressions<br />

of VND7 <strong>and</strong> two genes for AS2/LBD domain-containing<br />

proteins, ASL20/LBD18 <strong>and</strong> ASL19/LBD30, form a positive<br />

feedback loop to amplify their expression (Soyano et al. 2008).<br />

This rapid amplification of the master transcription factor may<br />

drive xylem cell differentiation promptly <strong>and</strong> irreversibly. (2)<br />

VND7 activity is also regulated at the protein level by its<br />

proteasome-mediated degradation (Yamaguchi et al. 2008). (3)<br />

A NAC domain transcription repressor, VND-INTERACTING2<br />

(VNI2), represses VND activity by protein-protein interaction<br />

(Yamaguchi et al. 2010). VNI2 has an unstable property because<br />

of the PEST proteolysis target motif in its C-terminal<br />

region, which may allow VND7 to exert its function promptly<br />

when it is required.<br />

Xylem cells form characteristic secondary walls. <strong>The</strong>se morphological<br />

events are regulated by master regulators such as<br />

SND1/NST3, VND6 <strong>and</strong> VND7. Microarray experiments revealed<br />

that VND6, VND7 <strong>and</strong> SND1 induce a hierarchical gene<br />

expression network (Zhong et al. 2008; Ohashi-Ito et al. 2010;<br />

Yamaguchi et al. 2010). <strong>The</strong>se master transcription factors<br />

induce, probably directly, the expression of genes for other<br />

transcription factors, such as MYB46, MYB83, <strong>and</strong> MYB103.<br />

MYB46 <strong>and</strong> MYB83 regulate redundant biosynthetic pathways<br />

for all three major secondary wall components, namely cellulose,<br />

lignin, <strong>and</strong> xylan (Zhong et al. 2007; McCarthy et al.<br />

2009). Here, two NACs <strong>and</strong> 10 MYBs appear to act downstream<br />

of SND1 (Zhong et al. 2008; Zhou et al. 2009). Of them,


310 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

MYB58, MYB63 <strong>and</strong> MYB85, which might be target genes of<br />

MYB46 <strong>and</strong>/or MYB83, specifically upregulate genes related<br />

to the lignin biosynthetic pathway (Zhou et al. 2009). Some of<br />

these key transcription factors are also induced by VND6 <strong>and</strong><br />

VND7, suggesting that this hierarchal structure is also true in<br />

the case of VND6 <strong>and</strong> VND7 (Ohashi et al. 2010; Yamaguchi<br />

et al. 2010). However, each master regulator also induces the<br />

expression of distinct genes, including transcription factors.<br />

Thus, SND1/NST3, VND6 <strong>and</strong> VND7, as master regulators,<br />

switch at the top of the hierarchy to upregulate transcription<br />

factors such as MYBs, which, in turn, functioning as the second<br />

<strong>and</strong> third regulators, upregulate expression of genes encoding<br />

enzymes catalyzing secondary wall thickening during specific<br />

stages of xylem cell differentiation.<br />

Interestingly, VND6 <strong>and</strong> VND7 can directly upregulate the<br />

expression of genes for enzymes such as XCP1 <strong>and</strong> CESA4,<br />

which are ranked lowest, as well as genes for transcription<br />

factors, such as MYB46, which are ranked higher in the<br />

gene expression hierarchy (Ohashi-Ito et al. 2010; Yamaguchi<br />

et al. 2011). Similarly, genes for enzymes such as 4CL1 are<br />

direct targets of SND1 (Zhong et al. 2008; McCarthy et al.<br />

2009). <strong>The</strong>se findings indicate a sophisticated transcriptional<br />

regulatory network, by the master regulator, over a hierarchy.<br />

Tracheary elements <strong>and</strong> xylem fibers possess different characteristics,<br />

such as cell wall structure <strong>and</strong> PCD. In accordance<br />

with these characters, VND6, but not SND1, induces the<br />

expression of genes related to rapid PCD, such as XCP1<br />

<strong>and</strong> XCPs, while SND1, but VND6 preferentially, upregulates<br />

genes for lignin monomer synthesis, such as PAL1, 4CL3, <strong>and</strong><br />

CCoAOMT (Ohashi-Ito et al. 2010).<br />

It is well-established that an 11-bp cis-element named<br />

the tracheary-element-regulating cis-element (TERE), which<br />

is found in upstream sequences of many genes expressed<br />

in xylem vessel cells, is responsible for xylem vessel cellspecific<br />

expression (Pyo et al. 2007). VND6 binds the TERE<br />

sequence <strong>and</strong> activates the TERE-containing promoter, in<br />

planta, but not a mutated promoter having substitutions in the<br />

TERE sequence (Ohashi-Ito et al. 2010). VND7 also binds<br />

TERE (Yamaguchi et al. 2011). <strong>The</strong>se results demonstrate<br />

that TERE is one of the target sequences contained within<br />

the VND6 promoter. In contrast, SND1 specifically binds to a<br />

19-bp sequence named SECONDARY WALL NAC BINDING<br />

ELEMENT (SNBE), to activate its target genes (Zhong et al.<br />

2010).<br />

Cellular events underlying xylem cell formation<br />

Xylem cell differentiation involves temporal <strong>and</strong> spatial regulation<br />

of secondary cell wall deposition. A number of xylem<br />

cell types exist, such as those with annular <strong>and</strong> spiral patterns<br />

in protoxylem vessels, reticulate <strong>and</strong> pitted patterns in<br />

metaxylem vessels, <strong>and</strong> a smeared pattern in xylem fibers.<br />

<strong>The</strong> cortical microtubules regulate the spatial pattern of the<br />

secondary cell wall by orientating cellulose deposition. By<br />

using cultures expressing GFP-tubulin it was discovered that<br />

cortical microtubules became gradually bundled, which, in turn,<br />

was followed by secondary wall deposition (Oda et al. 2005,<br />

2006, 2010). It is important to find microtubule associated<br />

proteins (MAPs) involved in secondary wall formation <strong>and</strong> to<br />

know their function. In this context, some important secondary<br />

wall-related MAPs that regulate cortical microtubule orientation<br />

have been discovered. For example, AtMAP70 family proteins<br />

appear to be involved in the formation of the secondary wall<br />

boundary (Pesquet et al. 2010). <strong>The</strong> plant-specific microtubule<br />

binding protein MIDD1/RIP3 promoted microtubule depolymerization<br />

in the future secondary wall pit area, resulting in a<br />

secondary wall-depletion domain (Oda et al. 2010). Further<br />

analysis revealed that ROPGEF4 <strong>and</strong> ROPGAP3 mediate local<br />

activation of the plant Rho GTPase ROP11, <strong>and</strong> this activated<br />

ROP11 then recruits MIDD1 to induce local disassembly of<br />

cortical microtubules (Oda <strong>and</strong> Fukuda 2012b). Interestingly,<br />

<strong>and</strong> conversely, cortical microtubules eliminate active ROP11<br />

from the plasma membrane through MIDD1. Such a mutual<br />

inhibitory interaction between active domains of ROP <strong>and</strong><br />

cortical microtubules gives rise to distinct patterns of secondary<br />

cell walls. <strong>The</strong>se findings shed new insights into the microtubule<br />

organizing mechanism regulating secondary wall patterning<br />

(Oda <strong>and</strong> Fukuda 2012a).<br />

PCD is a genetically regulated cell suicide process involved<br />

in many aspects of plant growth, such as seed germination,<br />

vascular differentiation, aerechyma tissue formation, reproductive<br />

organ development <strong>and</strong> leaf senescence (Kuriyama <strong>and</strong><br />

Fukuda 2002). During xylem development, rapid <strong>and</strong> slow PCD<br />

occurs in tracheary elements <strong>and</strong> xylem fiber cells, respectively,<br />

in order to facilitate the removal of cellular content for the<br />

formation of dead cells with secondary walls (Bollhoner et al.<br />

2012). PCD during tracheary element differentiation has long<br />

been recognized as an example of developmental PCD in<br />

plants (Fukuda 2004; Turner et al. 2007). This process includes<br />

cell death signal induction, accumulation of autolytic enzymes<br />

in the vacuole, vacuole swelling <strong>and</strong> collapse, <strong>and</strong> degradation<br />

of cell contents followed by mature tracheary element formation<br />

(Fukuda 2000).<br />

It has been suggested that the signals for xylem cell death are<br />

produced early during xylem differentiation, <strong>and</strong> that cell death<br />

is prevented through the action of inhibitors <strong>and</strong> the storage<br />

of hydrolytic enzymes in the vacuole (Bollhoner et al. 2012).<br />

According to the morphological process, the death of xylem<br />

tracheary elements is defined as a vacuolar type of cell death<br />

(Kuriyama <strong>and</strong> Fukuda 2002; Van Doorn et al. 2011a). Vacuolar<br />

membrane breakdown is the crucial event in tracheary element<br />

PCD, <strong>and</strong> bursting of the central vacuole triggers autolytic<br />

hydrolysis of the cell contents, thereby leading to cell death<br />

(Bollhoner et al. 2012).


Microarray analyses of gene expression have revealed a<br />

simultaneous expression of many genes involved in both secondary<br />

wall formation <strong>and</strong> PCD (Demura et al. 2002; Milioni<br />

et al. 2002; Kubo et al. 2005; Pesquet et al. 2005; Ohashi<br />

et al. 2010). As mentioned above, recent results have demonstrated<br />

that a transcriptional regulatory system, composed of<br />

transcription factors such as VND6 <strong>and</strong> a TERE-cis sequence,<br />

regulates the simultaneous expression of genes related to<br />

both secondary wall formation <strong>and</strong> PCD in tracheary elements<br />

(Ohashi-Ito et al. 2010). <strong>The</strong>se findings indicate that tracheary<br />

element-differentiation-inducing master genes initiate at least<br />

a part of PCD directly by activating PCD-related genes through<br />

binding the TERE sequence in their promoters. This suggests<br />

that, in contrast to apoptosis in animals, in which a common<br />

intracellular signaling system induces PCD, in plants, various<br />

developmental processes involving PCD may be regulated independently<br />

involving their own specific developmental steps.<br />

Nitric oxide (NO) <strong>and</strong> polyamine have been suggested as<br />

signals involved in the cell death induction in xylem development.<br />

NO production is largely confined to xylem cells; removal<br />

of NO from the cultured Zinnia cells, with its scavenger PTIO,<br />

results in dramatic reductions in both PCD <strong>and</strong> in the formation<br />

of tracheary elements (Gabaldon et al. 2005). Thus, NO might<br />

well be an important factor mediating PCD during tracheary<br />

element differentiation.<br />

Execution of PCD in developing tracheary elements involves<br />

expression <strong>and</strong> vacuole-accumulation of several hydrolytic<br />

enzymes, such as the cysteine proteases XCP1 <strong>and</strong> XCP2<br />

(Zhao et al. 2000; Funk et al. 2002; Avci et al. 2008), the Zn 2+ -<br />

dependent nuclease ZEN1 (Ito <strong>and</strong> Fukuda 2002) <strong>and</strong> RNases<br />

(Lehmann et al. 2001). Ca 2+ -dependent DNases were also<br />

detected in secondary xylem cells, <strong>and</strong> their activity dynamics<br />

were closely correlated with secondary xylem development<br />

(Chen et al. 2012a). ATMC9 encodes a caspase-like protein,<br />

which does not function as a caspase but as an arginine/lysinespecific<br />

cysteine protease (Vercammen et al. 2004). ATMC9<br />

was specifically expressed in differentiating vessels but not<br />

in fully-differentiated vessels (Ohashi-Ito et al. 2010). This<br />

result suggested the involvement of ATMC9 in PCD. However,<br />

application of caspase inhibitors significantly delays the time of<br />

tracheary element formation <strong>and</strong> inhibits DNA breakdown <strong>and</strong><br />

appearance of TUNEL-positive nuclei in Zinnia xylogenic cell<br />

culture (Twumasi et al. 2010).<br />

Recently, the protease responsible for developing xylemrelated<br />

caspase-3-like activity was purified <strong>and</strong> identified to<br />

be 20S proteasome (Han et al. 2012). <strong>The</strong> fact that treatment<br />

with a caspase-3 inhibitor Ac-DEVD-CHO causes a defect in<br />

veins in Arabidopsis cotyledons, <strong>and</strong> the proteasome inhibitor<br />

clasto-lactacystin β-lactone delays tracheary element PCD in<br />

VND6-induced Arabidopsis xylogenic culture, strongly suggest<br />

that the proteasome is involved in PCD during this process of<br />

differentiation (Han et al. 2012). Consistent with this notion, the<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 311<br />

26S proteasome inhibitors lactacystin <strong>and</strong> MG132 also delay<br />

or block the differentiation of suspension-cultured tracheary<br />

elements (Woffenden et al. 1998; Endo et al. 2001). Autophagy<br />

has also been suggested to be involved in tracheary element<br />

PCD (Weir et al. 2005). A recent finding that a small GTP<br />

binding protein RabG3b plays a positive role in PCD during tracheary<br />

element differentiation by activating autophagy (Kwon<br />

et al. 2010) provided support for this notion.<br />

<strong>The</strong> PCD of xylem fibers is less well characterized compared<br />

to that of xylem tracheary elements, likely due to the fact that<br />

this process proceeds slowly in these cell types. Microarray<br />

analyses revealed that a large number of genes encoding<br />

previously-uncharacterized transcription factors, as well as<br />

genes involved in ethylene, sphingolipids, light signaling <strong>and</strong><br />

autophagy-related factors, are expressed preferentially during<br />

xylem fiber development (Courtois-Moreau et al. 2009). Further<br />

comparison of genes related to PCD between xylem fibers<br />

<strong>and</strong> tracheary elements, in a model species like poplar, may<br />

help in advancing our underst<strong>and</strong>ing of PCD as it occurs in<br />

plants.<br />

Control over master transcription factors <strong>and</strong> crosstalk<br />

between signaling pathways<br />

It is now well established that xylem cell differentiation is<br />

regulated by various factors, both at the cell-autonomous <strong>and</strong><br />

non-cell-autonomous level. Auxin, cytokinin, brassinosteroids<br />

<strong>and</strong> CLE peptides act, cooperatively, at different stages of<br />

xylem cell differentiation. Importantly, an as-yet-unidentified<br />

intracellular signaling system initiates the expression of genes<br />

for master transcription factors such as VND6, VND7 <strong>and</strong><br />

SND1/NST1, each of which in turn induces distinctive xylem<br />

cell-specific gene expression. Further advances in our underst<strong>and</strong>ing<br />

of the events underlying xylem differentiation will<br />

be gained by studies on the related intracellular signaling<br />

pathways <strong>and</strong> the nature of the crosstalk that occurs between<br />

these specific signaling pathways.<br />

Spatial & Temporal Regulation<br />

of <strong>Vascular</strong> Patterning<br />

<strong>Vascular</strong> organization in leaves<br />

<strong>The</strong> leaf vascular system is a network of interconnecting<br />

veins, or vascular str<strong>and</strong>s, consisting of two main conducting<br />

tissue types: the xylem <strong>and</strong> the phloem. While the specialized<br />

conducting elements are composed of tracheary or vessel<br />

elements in the xylem <strong>and</strong> SEs in the phloem, the vascular<br />

system also contains non-conducting supporting cells, such as<br />

parenchyma, sclerenchyma <strong>and</strong> fibers. Thus, the development<br />

of cell types within the radial arrangement of the vascular<br />

bundle must be precisely spatially coordinated along with the


312 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

temporal longitudinal vein pattern in order to efficiently carry<br />

out their function as the long-distance transport system of the<br />

plant (Dengler <strong>and</strong> Kang 2001).<br />

<strong>The</strong> spatial organization of the leaf vascular system is both<br />

species- <strong>and</strong> organ-specific. Despite the diverse vein patterns<br />

found within leaves, the one commonality that is present during<br />

the ontogeny of the vascular system is the organization of the<br />

vascular bundles into a hierarchical system. Veins are organized<br />

into distinct size classes, based on their width at the most<br />

proximal point of attachment to the parent vein (Nelson <strong>and</strong><br />

Dengler 1997). Primary <strong>and</strong> secondary veins are considered<br />

to be major veins, not only due to their width, but because<br />

they are typically embedded in rib parenchyma, whereas higher<br />

order, or minor, veins such as tertiary <strong>and</strong> quaternary veins are<br />

embedded in mesophyll (Esau 1965a). <strong>The</strong> highest order veins,<br />

the freely ending veinlets, are the smallest in diameter <strong>and</strong> end<br />

blindly in surrounding mesophyll (Figure 10A).<br />

<strong>The</strong> presence of this hierarchical system in leaves reflects the<br />

function of the veins such that larger diameter veins function<br />

in bulk transport of water <strong>and</strong> metabolites, whereas smaller diameter<br />

veins function in phloem loading (Haritatos et al. 2000).<br />

In both the juvenile <strong>and</strong> adult phase leaves of Arabidopsis,<br />

the vein pattern is characterized by the major secondary veins<br />

that loop in opposite pairs in a series of conspicuous arches<br />

along the length of the leaf (Hickey 1973). This looping pattern,<br />

termed brochidodromous, is present in both juvenile <strong>and</strong> adult<br />

phase leaves. However, the hierarchical pattern is well defined<br />

in the adult leaves; there is a higher vein density <strong>and</strong> vein order<br />

(up to the 6 th order) when compared with the juvenile leaves<br />

(Kang <strong>and</strong> Dengler 2004). Despite this increasing vascular<br />

complexity, the overall vein pattern within a given species is<br />

highly conserved <strong>and</strong> reproducible, yet the vasculature itself<br />

is highly amenable to changes <strong>and</strong> re-modification during leaf<br />

development (Kang et al. 2007).<br />

Longitudinal vein pattern—procambium<br />

As indicated above, the procambium is a primary meristematic<br />

tissue that develops de novo from ground meristem cells to<br />

form differentiated xylem <strong>and</strong> phloem. In a temporal sense, the<br />

longitudinal vein pattern in Arabidopsis develops basipetally.<br />

However, the individual differentiating str<strong>and</strong>s of the preprocambium,<br />

procambium, <strong>and</strong> xylem develop in various directions<br />

(basipetally, acropetally or perpendicular to/or from the<br />

leaf midvein) depending on the stage of vascular development,<br />

as well as the local auxin levels (Figure 10B). Based strictly<br />

on its anatomical appearance, procambium is first identifiable<br />

by its cytoplasmically dense narrow cell shape <strong>and</strong> continuous<br />

cell files that seemingly appear either simultaneously or<br />

progressively along the length of the vascular str<strong>and</strong> (Esau<br />

1965b; Nelson <strong>and</strong> Dengler 1997).<br />

Figure 10. Longitudinal <strong>and</strong> radial vein patterning in leaves.<br />

(A) Venation pattern in the lamina of a mature Arabidopsis leaf.<br />

Vein size hierarchy is based on diameter of the veins at their<br />

most proximal insertion point. Vein size classes are color coded<br />

as follows: Orange, mid (primary) vein; purple, secondary/marginal<br />

veins; blue, tertiary veins; red, quaternary/freely ending veinlets.<br />

(B) <strong>Development</strong> of vein pattern in young leaves, as indicated by<br />

AtHB-8 (Kang <strong>and</strong> Dengler 2004; Scarpella et al. 2004). Establishment<br />

of the overall vein pattern in Arabidopsis is basipetal (black<br />

arrow). Secondary pre-procambium of the first pair of loops develop<br />

out from the midvein (dotted pink arrows, arrow indicates direction<br />

of pre-procambial str<strong>and</strong> progression). Pre-procambium of the<br />

second pair of secondary vein loops progresses either basipetally or<br />

acropetally. Third <strong>and</strong> higher secondary vein loop pairs progress out<br />

from the midvein towards the leaf margin <strong>and</strong> reconnect with other<br />

extending str<strong>and</strong>s (dotted black arrows). Procambium differentiates<br />

simultaneously along the procambial str<strong>and</strong> (blue solid lines). Xylem<br />

differentiation occurs approximately 4 d later <strong>and</strong> can develop<br />

either continuously, or as discontinuous isl<strong>and</strong>s, along the vascular<br />

str<strong>and</strong> (purple lines, arrow indicates direction of xylem str<strong>and</strong><br />

progression).<br />

(C) Differentiation of procambial cells. Pre-procambium is isodiametric<br />

in cell shape <strong>and</strong> is anatomically indistinguishable<br />

from ground meristem cells (maroon cell). Cell divisions of<br />

the pre-procambium are parallel to the direction of growth<br />

(light blue cells) of the vascular str<strong>and</strong>, resulting in elongated<br />

shaped cells characteristic of the procambium (dark blue<br />

cell).<br />

(D) Radial vein pattern in leaves. (Left to right): Procambial cells<br />

(as indicated by AtHB-8) are present within the vascular bundle.<br />

In a typical angiosperm leaf, xylem cells are dorsal to phloem<br />

cells (collateral vein pattern). In severely radialized leaf mutants,<br />

vein cell arrangement also becomes radialized. In adaxialized<br />

mutants such as phabulosa (phb), phavoluta (phv), <strong>and</strong> revolute<br />

(rev), xylem cells surround phloem cells (amphivasal), whereas in<br />

abaxialized mutants, such as those in the KANADI gene family,<br />

phloem cells surround the xylem cells (amphicribral) (Eshed et al.<br />

2001; McConnell et al. 2001; Emery et al. 2003).


<strong>The</strong> elongated procambium cells develop through distinct<br />

cell division patterns in which they divide parallel to the vascular<br />

str<strong>and</strong> (Kang et al. 2007) (Figure 10C). Although the<br />

anatomical distinction of procambium is clearly evident by its<br />

elongated shape, the precursor cells, pre-procambium, are<br />

isodiametric <strong>and</strong> are anatomically indistinguishable from surrounding<br />

ground meristem cells. Due to the difficulty of clearly<br />

identifying pre-procambium <strong>and</strong> ground meristem through<br />

anatomy alone, the use of molecular markers such as Arabidopsis<br />

thaliana HOMEOBOX 8 (AtHB-8), MONOPTEROS<br />

(MP), <strong>and</strong> PINFORMED 1 (PIN1) to identify procambium <strong>and</strong><br />

pre-procambium has facilitated the visualization of these early<br />

stages of vascular development (Mattsson et al. 2003; Kang<br />

<strong>and</strong> Dengler 2004; Scarpella et al. 2004; Wenzel et al. 2007).<br />

Auxin has been shown to regulate many aspects of plant<br />

development <strong>and</strong> play a critical role during vascular patterning,<br />

specifically vascular cell differentiation <strong>and</strong> vascular str<strong>and</strong><br />

formation (Aloni 1987; Aloni et al. 2003; Berleth et al. 2000;<br />

Mattsson et al. 2003). Early classical experiments showed<br />

that auxin is capable of inducing new str<strong>and</strong>s in response to<br />

wounding by promoting the transdifferentiation of parenchyma<br />

cells into continuous cell files towards the basal parts of the<br />

plant (Sachs 1981). <strong>The</strong>se early observations led to the “auxin<br />

canalization hypothesis” which suggested that auxin is transported<br />

directionally through a cell, progressively narrowing into<br />

discrete canals <strong>and</strong> operating through self-reinforcing positive<br />

feedback (Sachs 1981).<br />

In recent years, it has been shown that auxin is synthesized<br />

predominantly in leaf primordia <strong>and</strong> transported unidirectionally<br />

from apical to basal regions of the plant. Polar auxin transport<br />

(PAT) is accomplished by translocating auxin in a targeted<br />

manner through the cell, via auxin influx <strong>and</strong> efflux carriers<br />

(Lomax <strong>and</strong> Hicks 1992). Several gene families are known<br />

to affect vascular str<strong>and</strong> formation by modulating auxin levels<br />

during leaf development. <strong>The</strong> Arabidopsis family of efflux<br />

carriers, PIN1, regulates the polarity <strong>and</strong> elevated auxin levels<br />

from the shoot apical meristem into developing leaf primordia<br />

(Reinhardt et al. 2000; Benkova et al. 2003). <strong>The</strong> subcellular<br />

epidermal localization <strong>and</strong> convergence of auxin flow to the<br />

tip of the leaf primordia <strong>and</strong> subsequent basal transport of<br />

auxin directs the location of the future midvein <strong>and</strong> the sites of<br />

vascular str<strong>and</strong> formation (Reinhardt et al. 2003; Petráˇsek et al.<br />

2006).<br />

In young leaf primordia, the lateral marginal convergence<br />

points of PIN1 are required for vascular str<strong>and</strong> positioning <strong>and</strong><br />

arrangement (Sieburth 1999; Wenzel et al. 2007). Specifically,<br />

the initial broad expression domain of auxin in the developing<br />

leaf converges <strong>and</strong> tapers to narrow cell files (presumptive vascular<br />

str<strong>and</strong>s) that are dependent on auxin transport (Scarpella<br />

et al. 2006; Sawchuk et al. 2008). <strong>The</strong> large family of auxin<br />

response factors, which include transcription factors such as<br />

MP/AUXIN RESPONSE FACTOR 5 (ARF5), plays a key role<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 313<br />

in vascular str<strong>and</strong> formation (Wenzel et al. 2007). It is now<br />

well documented that mp loss-of-function mutants have reduced<br />

vasculature, discontinuous veins, <strong>and</strong> also affect embryo<br />

polarity <strong>and</strong> root meristem patterning (Hardtke <strong>and</strong> Berleth<br />

1998; Hardtke et al. 2004; Wenzel et al. 2007; Schuetz et al.<br />

2008).<br />

MP regulates vascular formation by inducing PIN1 expression,<br />

<strong>and</strong> recently, MP has been shown to directly target AtHB-<br />

8 through an activator that binds to the TGTCTG element<br />

in the AtHB-8 promoter to induce pre-procambial expression<br />

(Donner et al. 2009). Expression of AtHB-8 is simultaneously<br />

present along with the expression of SHR to demarcate presumptive<br />

vascular cells (Wenzel et al. 2007; Donner et al.<br />

2009; Gardiner et al. 2011). Although AtHB-8 expression is<br />

specifically localized to <strong>and</strong> remains in pre-procambial <strong>and</strong><br />

procambial cells, the expression domain of SHR is localized<br />

beyond the vascular str<strong>and</strong>, suggesting an alternative function<br />

beyond vascular development in leaves (Gardiner et al. 2011).<br />

<strong>The</strong> Class III HD-ZIP family of transcription factors, which<br />

includes AtHB-8, act as known regulators of both longitudinal<br />

<strong>and</strong> radial vascular patterning. <strong>The</strong> AtHB-8 gene is one of the<br />

earliest expressed in pre-procambial <strong>and</strong> procambial str<strong>and</strong>s<br />

to set up vascular patterning (Baima et al. 1995; Kang <strong>and</strong><br />

Dengler 2004; Scarpella et al. 2004). In Arabidopsis leaves,<br />

longitudinal vein pattern is initiated early in development<br />

through the acquisition of pre-procambial cells along a presumptive<br />

procambial str<strong>and</strong> (Kang <strong>and</strong> Dengler 2004; Scarpella<br />

et al. 2004; Sawchuck et al. 2007). Here, AtHB-8 is expressed<br />

in pre-procambial cells that are genetically identifiable from<br />

surrounding ground meristem cells. <strong>The</strong> distinct spatial organization<br />

of the secondary loops, characteristic of Arabidopsis<br />

vein patterning, develop uniquely based on the position of the<br />

secondary loop.<br />

<strong>The</strong> pre-procambium of the first secondary loop pair develops<br />

progressively away from the point of origin, the central midvein,<br />

to form a continuous loop (Figure 10B) (Kang <strong>and</strong> Dengler<br />

2004; Scarpella et al. 2004; Sawchuck et al. 2007). Later<br />

formed secondary pre-procambial str<strong>and</strong>s (i.e. second or third<br />

secondary loop pairs) also develop away from the point of<br />

origin; however, the direction of the extending str<strong>and</strong> can<br />

develop either acropetally or basipetally (Kang <strong>and</strong> Dengler<br />

2004). Procambial str<strong>and</strong>s develop simultaneously along the<br />

entire length of the vascular str<strong>and</strong> (Sawchuck et al. 2007). This<br />

simultaneous occurrence of the procambial str<strong>and</strong> is commonly<br />

seen in the first secondary loop pair. Importantly, procambial<br />

str<strong>and</strong>s of later formed secondary loop pairs can differentiate<br />

towards the leaf margin <strong>and</strong> reconnect with other str<strong>and</strong>s<br />

(Figure 10B).<br />

Approximately four d after procambium differentiation, xylem<br />

begins to develop both continuously from a point of origin<br />

or as discontinuous isl<strong>and</strong>s that connect either acropetally<br />

or basipetally with other str<strong>and</strong>s (Figure 10B). To date, many


314 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

vascular pattern mutants have been identified (Scarpella <strong>and</strong><br />

Meijer 2004; Scarpella <strong>and</strong> Helariutta 2010), <strong>and</strong> invariably,<br />

these mutants have disrupted <strong>and</strong>/or discontinuous vascular<br />

str<strong>and</strong>s, suggesting that proper formation or continuity of vascular<br />

str<strong>and</strong>s occurs early during the pre-procambial stages of<br />

development (Scarpella et al. 2010).<br />

Radial vein pattern—polarity <strong>and</strong> cell proliferation<br />

<strong>The</strong> spatial <strong>and</strong> temporal coordination of both longitudinal <strong>and</strong><br />

radial vein pattern is essential for proper functioning of the vascular<br />

system. As leaves arise from the shoot apical meristem,<br />

the incipient leaf primordia are initially radialized, but internal<br />

tissues quickly become polarized acquiring adaxial (dorsal)<br />

<strong>and</strong> abaxial (ventral) cellular identities. <strong>The</strong> juxtaposition of<br />

adaxial <strong>and</strong> abaxial characteristics allows the leaf to grow<br />

out into a flattened lateral organ. In a typical eudicot leaf, the<br />

vascular tissues are arranged in a collateral pattern with xylem<br />

adaxial to the phloem. This differentiation must occur from<br />

a uniform procambium cell population (Figure 10D). Much of<br />

what we currently know concerning vascular polarity is derived<br />

from work conducted on leaf polarity mutants, as alterations in<br />

leaf polarity often result in vascular bundle defects (Scarpella<br />

<strong>and</strong> Meijer 2004; Husb<strong>and</strong>s et al. 2009). In a completely<br />

radialized polarity mutant, vascular tissue is also radialized in<br />

that either xylem tissue surrounds a central cylinder of phloem<br />

(amphivasal) or phloem tissue surrounds a central cylinder of<br />

xylem (amphicribral) (Figure 10D).<br />

<strong>The</strong> establishment of adaxial-abaxial polarity is temporally<br />

regulated in the shoot apical meristem. Early experiments<br />

showed that meristem-derived signals may act to promote<br />

adaxial cell fate, as leaf primordia that were altered surgically<br />

became abaxialized (Sussex 1954). However, it is unlikely that<br />

meristem-derived factors alone are sufficient in establishing<br />

organ (<strong>and</strong> vascular) polarity, <strong>and</strong> that patterning of adaxialabaxial<br />

cell fate requires a number of genetic inputs. Transcription<br />

factors, such as those in the Class III HD-ZIP family, were<br />

identified as adaxial determinants based on radialized mutant<br />

phenotypes in Arabidopsis. Of these, the gain-of-function mutants<br />

in PHB, PHV, <strong>and</strong> REV display radialized leaves with<br />

amphivasal vascular bundles (McConnell et al. 2001; Emery<br />

et al. 2003). <strong>The</strong>ir abaxial counterparts, such as the KAN genes<br />

(MYB-like GARP transcription factors) are expressed in abaxial<br />

tissues, promoting abaxial identity. Gain-of-function mutations<br />

in these genes can result in amphicribral (phloem cells outside<br />

of a ring of xylem cells) vascular bundles (Kerstetter et al. 2001;<br />

Emery et al. 2003), or in the most extreme case, result in the<br />

complete elimination of vascular tissue within the radialized<br />

organ (McConnell <strong>and</strong> Barton 1998; Sawa et al. 1999).<br />

<strong>The</strong> intimate connection between the development of the<br />

procambium <strong>and</strong> the surrounding ground meristem (mesophyll)<br />

during tissue histogenesis has yet to be deciphered. <strong>The</strong><br />

genetic mechanisms coupling vascular cell proliferation with<br />

organ formation during tissue histogenesis are also largely<br />

unknown. However, it is known that cell proliferation is a<br />

critical developmental process that is required during tissue<br />

histogenesis. Attaining proper cell numbers within the radial<br />

vascular bundle is essential in order for vein size hierarchy<br />

to be properly established during vascular development. <strong>The</strong><br />

organization of this vein hierarchy is controlled, at least in part,<br />

by cell cycle regulators such as CyclinB1;1 (Kang <strong>and</strong> Dengler<br />

2002). Expression of CyclinB1;1::GUS is modulated within the<br />

vein orders so that cell cycling is prolonged in larger vein size<br />

classes, such as the midvein, but ceases first in smaller order<br />

veins. Modification of cell proliferation in leaves established<br />

that vein patterning is tightly coordinated with maintenance of<br />

meristematic competency of ground meristem cells to regulate<br />

(higher order) vein architecture (Kang et al. 2007). Although<br />

the direct association between the cell cycle <strong>and</strong> vascular<br />

patterning has yet to be determined, genes known to play a<br />

role in cell proliferation <strong>and</strong>/or stem cell maintenance may aid<br />

in elucidating this mechanistic pathway (Ji et al. 2010; Vanneste<br />

et al. 2011).<br />

Spatio-temporal regulation of root vascular<br />

development<br />

A number of comprehensive reviews exist that cover different<br />

aspects of root xylem development (Cano-Delgado et al. 2010;<br />

Scarpella <strong>and</strong> Helariutta 2010). In this section of the review,<br />

we will focus first on the morphological evidence for the timing<br />

of events that control vascular specification <strong>and</strong> differentiation<br />

within the root. We will then assess progress in elucidating<br />

the molecular markers <strong>and</strong> regulatory factors that govern<br />

spatio-temporal aspects of root vascular development. Spatial<br />

regulation involves cellular mechanisms that determine the<br />

arrangement of vascular cell types relative to each other.<br />

<strong>The</strong> temporal regulation of vascular development comprises<br />

mechanisms that determine cell specification of vascular cell<br />

type lineages beyond the quiescent center (QC), as well as differences<br />

in the timing of these differentiation events (Mähönen<br />

et al. 2000). <strong>The</strong> developmental time at which various cell types<br />

differentiate can be read according to the cell’s distance from<br />

the QC, or position relative to the root meristematic, elongation<br />

or maturation zone (Figure 11A). As the majority of research<br />

into the regulation of the spatial or temporal aspects of vascular<br />

development has been performed in the Arabidopsis root, we<br />

will use this model system to highlight progress in this area.<br />

Morphological markers of root vascular development<br />

Vasculature in the Arabidopsis root, as discussed earlier, is<br />

composed of the radially symmetric pericycle cell layer that<br />

surrounds the diarch vasculature. <strong>The</strong> pericycle is differentiated


Figure 11. Spatio-temporal regulation of root vascular patterning.<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 315<br />

(A) Spatial markers of vascular differentiation. <strong>Development</strong>al time points at which morphological markers consistent with differentiation of<br />

vascular cell types are indicated relative to the position along the longitudinal axis of the root. Changes in differentiation are highlighted<br />

in a change in cell color. MeZ, meristematic zone; El, elongation zone; MaZ, maturation zone; PP, protophloem; MP, metaphloem; CC,<br />

companion cells; PX, protoxylem; MX, metaxylem.<br />

(B) Temporal regulation of vascular regulator gene expression. <strong>The</strong> distinct temporal patterns of different vascular regulators are demonstrated<br />

along with the cell type with which these markers are associated. If a gene has a much higher peak of gene expression, then only this peak<br />

is shown.<br />

(C) Examples of genes whose expression shows fluctuating peaks in developmental time (first row), or dynamic expression between roots<br />

(compare first <strong>and</strong> second rows).<br />

into two cell types, the xylem <strong>and</strong> phloem pole pericycle cells.<br />

<strong>The</strong> former are located at the poles of the xylem axis <strong>and</strong><br />

are the only cells competent to become lateral root primordia,<br />

whereas the latter occupy the position between the xylem<br />

poles. <strong>The</strong>re are no morphological markers for phloem pole<br />

pericycle differentiation, other than their position relative to<br />

xylem pole pericycle cells, <strong>and</strong> the function of these cells<br />

remains to be elucidated.<br />

Phloem tissue is positioned interior to the pericycle cell<br />

layer <strong>and</strong> is located at the opposing poles of the vascular<br />

cylinder, whereas the central xylem axis cells form a median<br />

line transecting the vascular cylinder, perpendicular to the two<br />

phloem poles. Procambial cells are positioned between the<br />

xylem <strong>and</strong> phloem tissues. Xylem tissue is composed of two<br />

different cell types: protoxylem <strong>and</strong> metaxylem vessels. In the<br />

Arabidopsis root, there are two outer protoxylem cells <strong>and</strong> three<br />

inner metaxylem cells that can be distinguished based on their<br />

secondary cell wall characteristics. Protoxylem cells have a<br />

helical or annular pattern of secondary cell wall deposition,<br />

whereas metaxylem cells have a pitted deposition pattern.


316 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Protoxylem vessels in the root mature before the surrounding<br />

tissues elongate; during cell expansion of these surrounding<br />

cells, these protoxylem vessels are often destroyed. Thus, the<br />

metaxylem vessels act as the primary water conducting tissue<br />

throughout the main body of the plant (Esau 1965b). Metaxylem<br />

cell differentiation is temporally separated from protoxylem<br />

differentiation in that the outer metaxylem cells differentiate<br />

only after protoxylem cells differentiate <strong>and</strong> the surrounding<br />

tissues have completed their expansion. <strong>The</strong> inner metaxylem<br />

vessel differentiates later than the outer two metaxylem cells.<br />

Phloem tissue is composed of three cell types: protophloem<br />

SEs to the outside, <strong>and</strong> metaphloem SEs to the interior of<br />

the vascular cylinder, with CCs flanking the SEs. Protophloem<br />

SEs differentiate earlier than the metaphloem SEs <strong>and</strong> their<br />

associated CCs.<br />

Detailed anatomical studies of the Arabidopsis root tip have<br />

elucidated the earliest events in the timing <strong>and</strong> patterning of<br />

vascular initial cell divisions that give rise to all vascular cell<br />

types in the primary root (Mähönen et al. 2000). Just above<br />

the QC, asymmetric cell divisions of vascular initial cells give<br />

rise to the presumptive pericycle layer <strong>and</strong> protoxylem cells.<br />

At a position close to the QC (∼9 µm), five xylem cells are<br />

visible, <strong>and</strong> these will eventually differentiate into protoxylem<br />

<strong>and</strong> metaxylem vessels (Figure 11A). Two domains of vascular<br />

initials give rise to the phloem <strong>and</strong> procambial cell lineages,<br />

<strong>and</strong> they are located between 3 µm <strong>and</strong> 6 µm above the<br />

QC (Mähönen et al. 2000; Bonke et al. 2003). <strong>The</strong> number<br />

<strong>and</strong> exact pattern of future procambial cell divisions is variable<br />

between individual plants of the same species.<br />

<strong>The</strong> full set of phloem cells (protophloem, metaphloem<br />

<strong>and</strong> CC) can be observed at a distance above the QC<br />

(∼27 µm) (Mähönen et al. 2000) (Figure 11A). Protophloem<br />

<strong>and</strong> metaphloem SEs result from one tangential division of<br />

precursor cells, whereas CCs arise from one periclinal division<br />

of precursor cells (Bonke et al. 2003). At a further distance<br />

above the QC (∼70 µm), the first histological evidence of<br />

differentiation can be observed in protophloem SEs, as determined<br />

by staining with toluidine blue (Mähönen et al. 2000).<br />

Thus, protophloem SE differentiation occurs much earlier in<br />

developmental time compared to protoxylem vessel formation<br />

(Figure 11A). Metaphloem SEs <strong>and</strong> CCs differentiate at an<br />

approximately similar time to the outer metaxylem SEs. However,<br />

morphological analyses have determined that the spatial<br />

patterning of xylem cells occurs temporally prior to the spatial<br />

patterning of the phloem cells within the root.<br />

<strong>Vascular</strong> proliferation—cytokinin signaling<br />

<strong>Vascular</strong> initial cells or stem cells are the progenitor cell type for<br />

all vascular cells within the primary root. Regulation of vascular<br />

initial cell division is the first step in vascular development<br />

<strong>and</strong> is accomplished, in part, by the two-component cytokinin<br />

receptor WOL (Mähönen et al. 2000). WOL is expressed early<br />

in the Arabidopsis embryo during the globular stage <strong>and</strong> is<br />

present throughout the vascular cylinder during all subsequent<br />

stages of embryo <strong>and</strong> primary root development (Figure 11B).<br />

Interestingly, vascular defects within the embryonic root have<br />

not yet been reported. In the primary root of a wol mutant,<br />

there are fewer vascular initial cells, <strong>and</strong> the entire vascular<br />

bundle differentiates as protoxylem. Although this suggests<br />

that wol is deficient in procambial, metaxylem vessel <strong>and</strong><br />

phloem cell specification, a double mutant between wol <strong>and</strong><br />

fass (which results in supernumerary cell layers) produces<br />

phenotypically normal procambial <strong>and</strong> phloem cells, as well as<br />

both protoxylem <strong>and</strong> metaxylem vessels. This demonstrates<br />

that the role of WOL is in vascular initial cell proliferation, <strong>and</strong><br />

that any influence on cell specification is secondary to this<br />

defect.<br />

Transcriptional master regulators<br />

<strong>and</strong> xylem development<br />

Xylem cell differentiation, as marked by secondary cell<br />

wall synthesis <strong>and</strong> deposition, occurs much later in root<br />

developmental time relative to protophloem cell differentiation<br />

(Figure 11A). However, cells destined to become xylem cells<br />

are morphologically identifiable immediately after division of<br />

vascular initial cells. Based on gene expression data, a downstream<br />

regulator of cytokinin signaling, the AHP6, an inhibitory<br />

pseudophosphotransfer protein, is likely one of the earliest<br />

regulators of protoxylem cell specification (Mähönen et al.<br />

2006), but is unlikely to be the sole regulator (Figure 11B). AHP6<br />

functions to negatively regulate cytokinin signaling through<br />

spatial restriction of signaling within protoxylem cells. In a<br />

wol mutant, therefore, there is a lack of cytokinin signaling,<br />

a decrease in the asymmetric division of vascular initial cells<br />

<strong>and</strong> ectopic protoxylem cell differentiation in the few remaining<br />

vascular cells. AHP6 acts in a negative feedback loop with<br />

cytokinin signaling – cytokinin represses AHP6 expression,<br />

while AHP6 represses <strong>and</strong> spatially restricts cytokinin signaling<br />

(Mähönen et al. 2006). Cytokinin regulates the spatial domain<br />

of AHP6 expression in embryogenesis prior to when primary<br />

root protoxylem differentiation occurs. Thus, it appears that<br />

this negative regulatory feedback between cytokinin <strong>and</strong> AHP6<br />

occurs upstream of protoxylem specification in the primary root<br />

(Mähönen et al. 2006).<br />

<strong>The</strong> earliest marker of protoxylem cell specification in the primary<br />

root is achieved through a TARGET OF MONOPTEROS<br />

5 (TMO5) promoter:GFP fusion, named S4 (Lee et al. 2006;<br />

Schlereth et al. 2010). TMO5 is required for embryonic root<br />

initiation, <strong>and</strong> expression of this bHLH transcription factor is<br />

turned on shortly after division of vascular initial cells in the<br />

primary root <strong>and</strong> is turned off prior to secondary cell wall<br />

differentiation in protoxylem cells. This marker then turns on


in metaxylem cells <strong>and</strong> subsequently turns off again prior<br />

to secondary cell wall differentiation. MYB46 is one of the<br />

transcription factors partially required for synthesis of various<br />

components of the secondary cell wall in the protoxylem <strong>and</strong><br />

subsequent metaxylem (Lee et al. 2006; Zhong et al. 2008).<br />

This gene is expressed towards the end of the elongation<br />

zone in protoxylem cells <strong>and</strong> then later in metaxylem cells.<br />

Together, these findings suggest that, although there are no<br />

morphological markers of protoxylem cell specification early<br />

in developmental time, there are indeed molecular markers<br />

<strong>and</strong> two distinct developmental states for protoxylem <strong>and</strong><br />

metaxylem cells: an “early” state <strong>and</strong> a “late” state. Gene<br />

expression data support this observation, as there are distinct<br />

gene expression profiles in cells marked by TMO5 as compared<br />

to MYB46 (Brady et al. 2007).<br />

As mentioned earlier, PHB, PHV, REV, COR <strong>and</strong> ATHB-<br />

8 act redundantly to regulate xylem cell fate differentiation<br />

<strong>and</strong> are sufficient to regulate xylem patterning. In a dominant<br />

mutant background of PHB, phb-1d, which its mRNA is resistant<br />

to miRNA degradation, ectopic protoxylem is observed<br />

(Carlsbecker et al. 2010). PHB is therefore sufficient to specify<br />

protoxylem differentiation. <strong>The</strong> developmental time point at<br />

which the patterning of xylem cells is first regulated by these<br />

transcription factors remains unknown. Upstream regulators<br />

of SHR have yet to be identified, although SHR is expressed<br />

throughout the vascular cylinder during vascular development.<br />

Class III HD-ZIP transcription factors are also expressed in<br />

various cell types of the vasculature, primarily early in root<br />

developmental time in the meristematic zone; miR165a <strong>and</strong><br />

miR166b are also expressed with a peak in endodermis cells in<br />

this same zone (Carlsbecker et al. 2010). Based solely on the<br />

temporal nature of these expression patterns, the patterning<br />

of protoxylem <strong>and</strong> metaxylem cells appears to occur prior to<br />

the action of VND6 <strong>and</strong> VND7, although validation of this<br />

hypothesis requires further experimentation.<br />

Phloem cell patterning <strong>and</strong> differentiation<br />

Only a few factors are known to regulate phloem cell patterning<br />

<strong>and</strong> differentiation, despite protophloem being histologically<br />

evident quite early in development relative to xylem cell types<br />

(Mähönen et al. 2000). Mutations in OPS, ops-1 <strong>and</strong> ops-2,<br />

result in irregular phloem differentiation within the root. In<br />

WT roots, SE <strong>and</strong> CC differentiation is first marked by cell<br />

elongation followed by callose deposition <strong>and</strong> subsequent cell<br />

wall thickening in the longitudinal dimension. In the ops-1<br />

mutant, cell elongation, callose deposition <strong>and</strong> cell wall thickening<br />

fail to occur within the phloem cell lineage (Truernit<br />

et al. 2012). In addition, the ops-1 mutant has impaired longdistance<br />

phloem transport, likely because of gaps in phloem SE<br />

continuity. However, OPS is not sufficient to specify phloem cell<br />

differentiation. Overexpression of OPS results in precocious<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 317<br />

phloem cell differentiation within the root, but only within already<br />

specified protophloem <strong>and</strong> metaphloem lineages (Truernit et al.<br />

2012).<br />

In roots, APL is expressed later than OPS, that is, at a<br />

distance from the QC (Bonke et al. 2003; Truernit et al.<br />

2012) (Figure 11B). In an apl mutant, protophloem cells are<br />

misspecified as protoxylem cells, <strong>and</strong> there is a short root<br />

phenotype (Bonke et al. 2003). Contrary to an ops mutant<br />

phenotype, in an apl mutant background, no protophloem,<br />

metaphloem, or CCs are present anywhere in the phloem pole<br />

position within the vascular cylinder. APL plays a multifaceted<br />

role in phloem development. First, APL regulates the timing<br />

of the asymmetric cell divisions that would normally give rise<br />

to the SE <strong>and</strong> CC lineages. Second, APL is required for<br />

protophloem <strong>and</strong> metaphloem SE differentiation. Third, APL<br />

represses protoxylem differentiation within cells in the phloem<br />

pole position. However, despite all these roles, APL is not<br />

sufficient to specify phloem cell specification <strong>and</strong> differentiation.<br />

Clearly, additional factors remain to be identified in phloem<br />

cell development. <strong>The</strong> presence of a master regulator for<br />

phloem development like VND6/7 <strong>and</strong> the Class III HD-ZIP<br />

family in xylem development that is both necessary <strong>and</strong><br />

sufficient has not been identified. Protophloem cells differentiate<br />

earlier than metaphloem cells <strong>and</strong> CCs. APL protein<br />

localization reflects this temporal difference in differentiation,<br />

but the factor(s) that determines this temporal delay has yet to<br />

be isolated. Finally, factors that determine the spatial patterning<br />

of protophloem, metaphloem <strong>and</strong> CCs are similarly unknown.<br />

Pericycle cell specification <strong>and</strong> differentiation<br />

Pericycle cells have been divided into two populations based<br />

on gene expression <strong>and</strong> function. Only xylem pole pericyle cells<br />

are competent to become lateral root primordia. One marker of<br />

xylem pole pericycle cell differentiation is the J0121 enhancer<br />

trap that marks xylem pole pericycle cells after they exit from the<br />

meristematic zone <strong>and</strong> pass through the elongation zone (Parizot<br />

et al. 2008). A marker of intervening cells within the pericycle<br />

tissue layer helped identify phloem pole pericycle cells. <strong>The</strong>se<br />

cells are marked by expression of S17, a basic leucine zipper<br />

transcription factor. <strong>The</strong> function of phloem pole pericycle cells<br />

has not been determined, nor are there histological markers of<br />

phloem pole pericycle differentiation. However, phloem pole<br />

pericycle cells have a very distinct expression pattern <strong>and</strong><br />

underlying transcriptional signature from that of xylem pole<br />

pericycle cells, as determined by expression profiling of marked<br />

populations of both of these cells relative to other cell types<br />

in the Arabidopsis root (Brady et al. 2007). Interestingly, the<br />

expression pattern of phloem pole pericycle cells more closely<br />

reflects that of cells in the developing phloem cell lineage.<br />

No early markers of either phloem or xylem pole pericycle<br />

cells have been identified, nor have regulatory factors been


318 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

found that are both necessary <strong>and</strong> sufficient for pericycle<br />

cell specification <strong>and</strong> differentiation. Cytokinin is sufficient to<br />

suppress xylem pole pericycle differentiation, as marked by<br />

the J0121 enhancer trap line, <strong>and</strong> this, in part, requires AHP6<br />

(Mähönen et al. 2006). MiR165/166 repression of PHB is also<br />

required for pericycle differentiation (Miyashima <strong>and</strong> Nakajima<br />

2011). In WT roots, AHP6 is expressed in protoxylem cells<br />

<strong>and</strong> the two abutting xylem pole pericycle cells (Figure 11B).<br />

In the scr-3 <strong>and</strong> phb-1d mutants, AHP6 expression was either<br />

completely lost or detected in only one of the aforementioned<br />

three cells. In addition, expression of an additional xylem pole<br />

pericycle marker gene, STELAR K + OUTWARD RECTIFIER<br />

(SKOR), was greatly reduced in these phb-1d <strong>and</strong> scr-3 mutant<br />

lines. Finally, in response to external auxin, which is sufficient to<br />

induce periclinal divisions in xylem pole pericyle cells, reduced<br />

periclinal divisions were observed in the phb-1d mutant. Expression<br />

of a mutant form of miR165, which is able to target the<br />

phb-1d miRNA-resistant PHB transcripts, was able to rescue<br />

the AHP6 <strong>and</strong> SKOR expression patterns in phb-1d. However,<br />

lateral root primordium development was somewhat delayed,<br />

suggesting that SHR/miR165-dependent regulation of PHB is<br />

required for pericycle function.<br />

Dynamic regulation of gene expression within the root<br />

vasculature<br />

Dynamic gene expression patterns have also been identified<br />

in the root vasculature, <strong>and</strong> thus far, oscillatory expression<br />

patterns have been linked to lateral root initiation. First, oscillations<br />

in auxin responsiveness, as measured by the DR5:GUS<br />

synthetic auxin response reporter in xylem pole pericycle cells<br />

within the root meristematic zone, were shown to temporally<br />

correlate with lateral root initiation at regular intervals in the maturation<br />

zone (De Smet et al. 2007). Further work demonstrated<br />

that this oscillatory auxin responsiveness likely determines<br />

competence of xylem pole pericycle cells to become lateral<br />

root primordia (Moreno-Risueno et al. 2010).<br />

To identify additional factors that play a role in lateral root development<br />

<strong>and</strong> which may act at the same time or downstream<br />

of this fluctuating auxin responsiveness, microarray analysis<br />

was used to identify genes whose expression oscillates in<br />

phase with DR5 auxin responsiveness as well as antiphase with<br />

auxin responsiveness. Many of these oscillating genes were<br />

shown to regulate prebranch initiation site formation <strong>and</strong> lateral<br />

root number (Moreno-Risueno et al. 2010). Numerous other<br />

genes have been identified that regulate dynamic expression in<br />

root developmental time. <strong>The</strong>se were obtained by sectioning an<br />

individual Arabidopsis root into 12 successive sections, each<br />

representing a specific point in a developmental time (Brady<br />

et al. 2007). Sections of an independent root served as a<br />

biological replicate.<br />

Based on these data, genes were identified whose expression<br />

fluctuates over root developmental time; e.g., they showed<br />

peaks of expression in the meristematic zone <strong>and</strong> maturation<br />

zone, with downregulation of expression in the meristematic<br />

zone (Figure 11C). A rigorous statistical method was developed<br />

to identify cases of dynamic expression between roots<br />

(Figure 11C), <strong>and</strong> many of these were expressed specifically in<br />

root phloem cell types or in xylem pole pericycle cells. <strong>The</strong>ir<br />

function was inferred to be associated with energy capture<br />

<strong>and</strong> lateral root initiation, respectively (Orl<strong>and</strong>o et al. 2010).<br />

Together, these data indicate that oscillatory, rhythmic <strong>and</strong><br />

fluctuating gene expression within roots <strong>and</strong> between roots in<br />

the root vasculature serve to regulate patterning of vascular<br />

cells, <strong>and</strong> likely other vascular biological functions.<br />

Spatio-temporal regulation of root <strong>and</strong> shoot vascular<br />

development <strong>and</strong> connectivity<br />

Studies on Arabidopsis root mutants defective in cell proliferation<br />

<strong>and</strong> cell differentiation may provide insight into possible<br />

common genetic regulatory mechanisms controlling radial<br />

vascular development in shoots <strong>and</strong> leaves. Mutants such as<br />

wol show decreased cell proliferation in root procambium <strong>and</strong><br />

differentiate completely into protoxylem (Mähönen et al. 2000),<br />

whereas the apl mutant shows defects in phloem development<br />

(Bonke et al. 2003). Although these genes appear to affect<br />

root cells exclusively, recent studies revealed that SHR is also<br />

involved in regulating cell proliferation in leaves (Dhondt et al.<br />

2010). Furthermore, expression of SHR is tightly correlated with<br />

AtHB-8 expression during vascular str<strong>and</strong> formation (Gardiner<br />

et al. 2011). While the spatial <strong>and</strong> temporal patterns of cell<br />

proliferation during root vascular development are becoming<br />

clearer, underst<strong>and</strong>ing this mechanism in leaves has proved<br />

to be more challenging. It is likely that combinatorial control<br />

of both hormones <strong>and</strong> genetics are in effect, <strong>and</strong> that these<br />

components are tightly integrated in both tissue <strong>and</strong> organ<br />

(leaf) morphogenesis. <strong>The</strong>refore, isolating these developmental<br />

components will be critical in order to thoroughly underst<strong>and</strong><br />

the spatial <strong>and</strong> temporal control of vascular development in<br />

leaves.<br />

Secondary <strong>Vascular</strong> <strong>Development</strong><br />

<strong>The</strong> term “secondary growth” refers to the radial growth of<br />

stems, <strong>and</strong> is ultimately the result of cell division within a<br />

lateral meristem, the vascular cambium (Larson 1994). <strong>The</strong><br />

vascular cambium produces daughter cells towards the center<br />

of the stem which become part of the secondary xylem, or<br />

wood (Figure 12). <strong>The</strong> cambium also produces daughter cells<br />

towards the outside of the stem which become part of the inner<br />

bark. <strong>The</strong>re are two types of cambial initials: fusiform <strong>and</strong> ray.


Figure 12. Internal structure of a woody plant stem.<br />

<strong>The</strong> vascular cambium consists of a centrifugal layer of fusiform<br />

secondary phloem <strong>and</strong> a centripetal layer of secondary xylem cells<br />

surrounding a central zone comprising phloem <strong>and</strong> xylem transit<br />

amplifying cells with a central uniseriate layer of cambial stem cells.<br />

Most angiosperms <strong>and</strong> gymnosperm trees species also contain<br />

radial files of near isodiametric ray cells that play a role in nutrient<br />

transport <strong>and</strong> storage (reproduced from Matte Risopatron et al.<br />

(2010) with permission).<br />

Fusiform initials give rise to the vertically-oriented cells, including<br />

water-conducting tracheary elements of the secondary<br />

xylem, <strong>and</strong> nutrient- <strong>and</strong> molecular signaling-conducting SEs of<br />

the secondary phloem. Ray initials produce procumbent cells<br />

that serve to transport materials radially in the stem, <strong>and</strong> likely<br />

serve storage <strong>and</strong> other functions that are currently poorly<br />

defined.<br />

To produce a functional, woody stem, these <strong>and</strong> many<br />

other developmental processes must be coordinated (Du <strong>and</strong><br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 319<br />

Groover 2010). Importantly, these developmental processes<br />

are also highly influenced by environmental cues. This is<br />

evident when observing annual rings in a tree stump, where<br />

favorable environmental conditions in the spring can lead to<br />

rapid growth <strong>and</strong> production of wood with anatomical <strong>and</strong><br />

chemical differences from wood produced under draught <strong>and</strong><br />

less favorable conditions later in the growing season. Another<br />

notable example of how environmental cues influence secondary<br />

growth is the formation of reaction wood in response<br />

to gravity <strong>and</strong> mechanical stresses, with reaction wood serving<br />

to right bent stems or to support horizontal branches (Du <strong>and</strong><br />

Yamamoto 2007).<br />

In this section of the review, we will highlight some of the more<br />

recent advances in the underst<strong>and</strong>ing of how secondary growth<br />

is regulated. This is an exciting period in the study of secondary<br />

growth, as genomic approaches applied to a number of species<br />

have yielded comprehensive lists of the genes expressed in<br />

the cambium <strong>and</strong> secondary vascular tissues. Additionally, the<br />

model forest tree genus Populus now has a fully sequenced<br />

genome (Tuskan et al. 2006), which, when paired with relatively<br />

efficient transformation systems for some Populus genotypes,<br />

has allowed detailed functional characterization of a modest<br />

number of regulatory genes.<br />

One emerging theme from these studies is that at least some<br />

of the major regulatory genes <strong>and</strong> mechanisms that regulate<br />

the cambium <strong>and</strong> secondary vascular development have been<br />

either directly co-opted from the shoot apical meristem, or else<br />

represent genes derived from duplication of an ancestral shoot<br />

apical meristem regulator (Spicer <strong>and</strong> Groover 2010). Thus,<br />

the study of the cambium <strong>and</strong> secondary growth also presents<br />

opportunities to underst<strong>and</strong> the evolution of meristems <strong>and</strong><br />

details as to how regulatory modules of genes can be reused<br />

<strong>and</strong> repurposed during plant evolution. Furthermore, since<br />

secondary growth in angiosperms, <strong>and</strong> perhaps in both angiosperms<br />

<strong>and</strong> gymnosperms, is likely homologous, advances<br />

in our underst<strong>and</strong>ing within model species like Populus can<br />

potentially greatly accelerate our underst<strong>and</strong>ing of secondary<br />

growth in less tractable species.<br />

<strong>The</strong> many values of secondary woody growth<br />

To better underst<strong>and</strong> the importance of the fundamental processes<br />

involved in secondary growth, it is worthwhile to first<br />

gain an underst<strong>and</strong>ing of the practical reasons of how research<br />

of secondary growth is important to ecosystems, societies <strong>and</strong><br />

industries. <strong>The</strong> wood produced by forest trees during secondary<br />

growth represents durable sequestration of the greenhouse gas<br />

CO2, which is ultimately incorporated into wood as byproducts<br />

of photosynthesis. Wood is primarily composed of fiber <strong>and</strong><br />

tracheary element cells which undergo complex processes of<br />

differentiation in which they synthesize thickened secondary<br />

cell walls before undergoing PCD to produce cell corpses.


320 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Fibers impart mechanical strength to wood, whereas tracheary<br />

elements provide both mechanical strength as well as water<br />

conduction. Lignin <strong>and</strong> cellulose are primary components of<br />

secondary cell walls, <strong>and</strong> thus of wood, <strong>and</strong> impart mechanical<br />

strength <strong>and</strong> resistance to degradation. Together, cellulose <strong>and</strong><br />

lignin are the most abundant biopolymers on the planet.<br />

<strong>The</strong> secondary cell walls in wood ultimately reflect storage<br />

of energy <strong>and</strong> CO2 derived from photosynthesis. With regards<br />

to carbon sequestration, forests are second only to the oceans<br />

in the biological sequestration of carbon, <strong>and</strong> are thus central<br />

to the carbon cycle <strong>and</strong> to mediating the levels of atmospheric<br />

CO2. <strong>The</strong> energy stored in wood has played central roles in<br />

history by providing heating <strong>and</strong> cooking fuels, <strong>and</strong> continues<br />

to play these vital roles in developing countries today (Salim<br />

<strong>and</strong> Ullsten 1999; FAO 2008). Looking to the future, the woody<br />

tissues of trees are increasingly of interest as a net-carbon<br />

neutral source of bioenergy (FAO 2008). While wood wastes<br />

have long played roles in cogeneration plants to supplement<br />

coal or to provide energy to forest industry mills, more recently<br />

woody biomass is being utilized as a “next generation” biofuel.<br />

Wood from trees can be utilized as a feedstock to produce<br />

ethanol, syngas, or other biofuels. In addition, similar to a<br />

petroleum refinery, the utility <strong>and</strong> economics of biorefineries<br />

will be bolstered by production of value added products such<br />

as acetic acid <strong>and</strong> other chemicals (Naik et al. 2010). Fast<br />

growing woody perennial crops, including forest trees such<br />

Populus, are beginning to be utilized as a source of biofuels<br />

on industrial scales (Sannigrahi et al. 2010).<br />

Wood is also used to produce pulp, paper, lumber, <strong>and</strong><br />

countless derived forest products. Forest products have been<br />

estimated to represent three percent of the total world trade<br />

(FAO 2009). Harvesting <strong>and</strong> processing of wood products, as<br />

well as related follow-on industries, represent vital components<br />

of rural economies in forested regions worldwide, where few<br />

alternative industries exist. Importantly, forests <strong>and</strong> the woody<br />

bodies of trees provide numerous ecosystem services to which<br />

it is difficult to ascribe an economic value (Salim <strong>and</strong> Ullsten<br />

1999; FAO 2009). Forests provide unique habitats, underpin<br />

crucial ecosystems, provide clean water, <strong>and</strong> are the focus of<br />

tourism industries worldwide. Perhaps the most difficult to value<br />

are the aesthetic, cultural, <strong>and</strong> spiritual aspects of forests, trees,<br />

<strong>and</strong> wood. In short, wood produced by forests is central to the<br />

health of our planet <strong>and</strong> society.<br />

<strong>The</strong> biology <strong>and</strong> regulation of secondary growth<br />

Secondary growth represents the culmination of a number<br />

of fascinating developmental processes. Currently, there are<br />

mechanisms identified <strong>and</strong> partially characterized that regulate<br />

specific developmental processes, including cambium initiation<br />

<strong>and</strong> maintenance, tissue patterning, <strong>and</strong> the balance of cell<br />

division <strong>and</strong> cell differentiation. While our underst<strong>and</strong>ing of<br />

secondary growth is far from complete, past <strong>and</strong> ongoing<br />

genomic studies have provided exhaustive lists of all the genes<br />

expressed during secondary growth. Molecular genetics studies<br />

have provided insights into the function of a modest number<br />

of regulatory genes, primarily encoding transcription factors,<br />

signaling peptides, <strong>and</strong> receptors. <strong>Plant</strong> growth regulators,<br />

including cytokinin, ethylene, gibberellic acid <strong>and</strong> most notably<br />

auxin, have all been implicated in influencing some aspect of<br />

secondary growth.<br />

Here, we provide a brief analysis of some of the mechanisms<br />

identified that regulate secondary growth. First, we will focus<br />

on auxin, as it has profound influences on secondary growth.<br />

Auxin has long been known to be a critical regulator of cambium<br />

functions <strong>and</strong> secondary growth. For example, exogenous<br />

auxin applied to decapitated shoots can stimulate cambial<br />

formation <strong>and</strong> activity (Snow 1935). It is generally assumed<br />

that auxin is produced in leaves <strong>and</strong> apical meristems, <strong>and</strong><br />

transported down the cambium to stimulate growth. However,<br />

direct determinations of the actual routes of auxin transport in<br />

the stem <strong>and</strong> the relative amount of auxin synthesized in the<br />

stem are currently lacking. Indeed, looking at trees like the giant<br />

sequoia in which the canopy foliage can be a hundred feet from<br />

the active cambium at the base of the stem, bring into question<br />

models for auxin synthesis <strong>and</strong> transport that were developed<br />

in smaller <strong>and</strong> more tractable plant species.<br />

<strong>The</strong> role of auxin transport <strong>and</strong> auxin gradients during secondary<br />

growth have been researched directly in forest trees<br />

(Uggla et al. 1996; Schrader et al. 2003; Kramer et al. 2008;<br />

Nilsson et al. 2008), but remain inconclusive. A radial auxin gradient<br />

is present across secondary vascular tissues <strong>and</strong> peaks<br />

in the region of the cambium <strong>and</strong> nascent secondary xylem in<br />

both angiosperm <strong>and</strong> gymnosperm trees (Uggla et al. 1996;<br />

Tuominen et al. 1997). This observation spurred speculation<br />

that auxin could create a radial morphogen gradient, but that<br />

concept has been brought into question by studies showing<br />

that few genes that are auxin-responsive actually show peaks<br />

of expression in the cambial zone (Nilsson et al. 2008).<br />

Auxin has also been shown to be transported basipitally<br />

in the stem, <strong>and</strong> to be involved with polarity determination<br />

in secondary vascular tissues (Kramer et al. 2008). Genes<br />

encoding PIN-type auxin efflux carriers are preferentially expressed<br />

in the cambial zone <strong>and</strong> developing xylem in the<br />

radial gradient, <strong>and</strong> in the apical-basal gradient they show a<br />

sharp peak of expression in internodes that are transitioning to<br />

secondary growth (Schrader et al. 2003). Currently, the subcellular<br />

localization <strong>and</strong> function of PIN transporters is largely<br />

uncharacterized in stems, <strong>and</strong> the functional significance of<br />

the auxin gradients remains contentious (Nilsson et al. 2008).<br />

Thus, auxin is a central point for important new advances in the<br />

study of secondary growth.<br />

Indirect observations suggest that cell-cell communication<br />

might play important roles in secondary growth, <strong>and</strong> studies


are beginning to reveal important mechanisms in how the<br />

balance of cell differentiation <strong>and</strong> cell division is regulated, <strong>and</strong><br />

how tissue identity is maintained across layers of secondary<br />

vascular tissues. Recent studies in Arabidopsis <strong>and</strong> Populus<br />

have identified mechanisms by which the balance of cell differentiation<br />

<strong>and</strong> tissue identity are established through cell-cell<br />

signaling. As previously discussed, in Arabidopsis, TDIFisa<br />

small peptide product of the phloem-expressed CLE41/44 gene<br />

(Ito et al. 2006). TDIF is secreted from the phloem, <strong>and</strong> acts to<br />

inhibit tracheary element differentiation <strong>and</strong> stimulate cambial<br />

activity (Ito et al. 2006; Hirakawa et al. 2008; Etchells <strong>and</strong><br />

Turner 2010), as well as to regulate the orientation of cambial<br />

divisions (Etchells <strong>and</strong> Turner 2010). This peptide is perceived<br />

by the LRR-Receptor kinase Phloem Intercalated with Xylem<br />

(PXY), which is expressed in the procambium (Hirakawa et al.<br />

2008; Hirakawa et al. 2010). Loss-of-function pxy mutants<br />

show phloem cells intermixed in the xylem (Fisher <strong>and</strong> Turner<br />

2007). TRIF/PXY signaling activates WOX4 (Hirakawa et al.<br />

2010), which encodes a transcription factor that presumably<br />

influences gene expression associated with meristematic cell<br />

fate. Interestingly, stimulation of cambial activity by auxin<br />

requires functional WOX4 <strong>and</strong> PXY (Suer et al. 2011), providing<br />

insight into how auxin integrates into transcriptional regulation<br />

in secondary vascular tissues.<br />

Patterning <strong>and</strong> polarity in secondary vascular tissues<br />

Cross sections of a typical woody stem show that secondary<br />

vascular tissues are highly patterned (Figure 12), <strong>and</strong> that the<br />

proper position <strong>and</strong> patterning of the cambium, secondary<br />

xylem, <strong>and</strong> secondary phloem are crucial to the function of<br />

these tissues. Additionally, secondary vascular tissues can<br />

be described in terms of polarity, analogous to vasculature in<br />

leaves. Take for example the vasculature of a typical dicot tree,<br />

poplar, in which the vascular bundles in leaves always have<br />

xylem towards the adaxial <strong>and</strong> phloem towards the abaxial<br />

surface of the leaf. By following those vascular bundles through<br />

the leaf trace <strong>and</strong> into the stem, it becomes apparent that<br />

the same polarity relationships are found in both primary <strong>and</strong><br />

secondary vascular tissues, with xylem towards the center <strong>and</strong><br />

phloem towards the outside of the stem.<br />

Insights into how polarity is established <strong>and</strong> maintained<br />

in vascular tissues has been provided by pioneering studies<br />

of the Class III HD-ZIP <strong>and</strong> KANADI transcription factors in<br />

Arabidopsis. <strong>The</strong>se Class III HD-ZIPs are highly conserved<br />

in plants, act antagonistically with KANADIs, <strong>and</strong> have been<br />

shown to regulate fundamental aspects of meristem function,<br />

polarity, <strong>and</strong> vascular development (Emery et al. 2003; Izhaki<br />

<strong>and</strong> Bowman 2007; Bowman <strong>and</strong> Floyd 2008). In Arabidopsis,<br />

REV is implicated in various developmental processes,<br />

including patterning of primary vascular bundles (Emery et al.<br />

2003; Bowman <strong>and</strong> Floyd 2008). A recent study showed that<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 321<br />

misexpression of a Populus REV ortholog results in formation<br />

of ectopic cambia in the cortex of the stem, <strong>and</strong> that these<br />

cambia can produce secondary xylem with reversed polarity<br />

(Robischon et al. 2011), indicating that the Class III HD-ZIPs<br />

also affect patterning <strong>and</strong> polarity in secondary vasculature.<br />

Genetics <strong>and</strong> genomics: critical tools for advancing<br />

knowledge on woody growth<br />

Research on secondary growth is at an exciting point, as genomic<br />

tools are now allowing the characterization of the genetic<br />

variation within species that is responsible for wood quality <strong>and</strong><br />

growth traits. Association mapping is being taken to a wholegenome<br />

scale for some tree species, <strong>and</strong> will undoubtedly<br />

provide fascinating insights into macro- <strong>and</strong> micro-evolution<br />

of wood formation (Neale <strong>and</strong> Kremer 2011). Genomic tools<br />

are also now enabling the first generation of network biology<br />

approaches in the model tree genus Populus (Street et al.<br />

2011) that can be used in underst<strong>and</strong>ing woody growth. Such<br />

approaches utilize a variety of genomic data types to model<br />

the genetic networks that regulate specific aspects of woody<br />

growth, <strong>and</strong> can ultimately produce a “wiring diagram” of regulatory<br />

networks. This will be important for both better directing<br />

future research <strong>and</strong> for providing predictive models that can<br />

potentially be used to better direct breeding programs, identify<br />

regulatory genes for biotechnology, <strong>and</strong> provide insights into<br />

the complexities of biological processes fundamental to the<br />

future of forests worldwide.<br />

Physical <strong>and</strong> Physiological Constraints<br />

on Phloem Transport Function<br />

We now turn our attention to an examination of the constraints<br />

of photoassimilate transport in the most evolutionarily<br />

advanced plants, the angiosperms. Here, photoassimilate<br />

conducting units are comprised of SEs arranged end-to-end to<br />

form conduits that are referred to as sieve tubes. At maturity,<br />

SEs lack nuclei <strong>and</strong> vacuoles, <strong>and</strong> their parietal cytoplasm<br />

has a greatly reduced number of organelles. In contrast to<br />

xylem tracheary elements, SEs retain their semi-permeable<br />

plasma membrane. <strong>The</strong>ir shared end walls contain interconnecting<br />

pores (sieve plate pores) formed from PD coalescing<br />

within pit fields (Evert 2006). In addition, each SE is highly<br />

interconnected symplasmically with a metabolically active CC,<br />

through specialized PD, to form a functional unit referred to as<br />

the SE-CC complex.<br />

In order to provide a framework on which to identify the<br />

physical <strong>and</strong> physiological constraints regulating phloem transport,<br />

we must first examine the physical mechanisms responsible<br />

for resource transport through the sieve tube system.<br />

As photoassimilate flow is polarized from source leaves (net<br />

exporters of resources) to heterotrophic sinks (net importers


322 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

of resources), phloem transport can be envisaged in terms<br />

of three key physiological components arranged in series.<br />

<strong>The</strong>se components are: (a) loading of photosynthate (most<br />

commonly sucrose) in collection phloem (minor veins) within<br />

source leaves, (b) long-distance delivery in transport phloem,<br />

<strong>and</strong> (c) unloading from release phloem (Figure 13A). Principles<br />

of flux control analysis dictate that each component will confer<br />

an influence (constraint) on overall flow, from source to sink;<br />

thus, the question of constraint becomes one of degree.<br />

Bulk flow identifies the regulatory elements<br />

<strong>The</strong> phloem is generally buried deep within plant tissues, <strong>and</strong><br />

this location, along with its sensitivity to mechanical perturbation,<br />

has made it technically challenging to observe flows<br />

through sieve tubes, non-invasively <strong>and</strong> in real-time. However,<br />

there is a critical body of experimental evidence showing that<br />

solutes <strong>and</strong> water move at similar velocities through sieve tubes<br />

(van Bel <strong>and</strong> Hafke 2005; Windt et al. 2006). <strong>The</strong>se studies<br />

suggest that the phloem translocation stream moves by bulk<br />

flow, consistent with the now widely accepted pressure flow<br />

hypothesis put forward originally by Münch (1930).<br />

On the premise that transport through sieve tubes conforms<br />

to bulk flow, then the transport rate (Rf ) of a nutrient species is<br />

given by the product of transport velocity (V), sieve tube crosssectional<br />

area (A) <strong>and</strong> phloem sap concentration (C), whereby:<br />

Rf = V · A · C (1)<br />

Both A <strong>and</strong> C are finite elements, whereas, given that flow<br />

through sieve tubes approximates laminar flow through capillaries,<br />

the factors determining transport velocity (V), or solvent<br />

volume flux (Jv, having units of m 3 m −2 s −1 ,orms −1 ), are<br />

identified by the Hagen-Poiseuille Law as:<br />

Jv = Pπr 2 /8ηl (2)<br />

Here, P is the hydrostatic pressure difference between either<br />

end of each sieve tube of length (l), the translocation stream<br />

has a viscosity (η), <strong>and</strong> it moves though sieve tubes of known<br />

radii (r) that have sub-structural elements that serve to impede<br />

transport; i.e., sieve plate pores <strong>and</strong> parietal cytoplasm<br />

(Mullendore et al. 2010).<br />

<strong>The</strong> water potential (ψw) of any cell (e.g., a SE-CC complex)<br />

is given by:<br />

ψw = ψP + ψπ<br />

where ψP is the pressure potential <strong>and</strong> ψπ is the osmotic or<br />

solute potential within the cell of interest. <strong>The</strong> value of ψP within<br />

a cell is determined by the magnitude of ψπ <strong>and</strong> the ψw of the<br />

cell wall (apoplasmic potential). As ψπ in the wall is generally<br />

close to zero, <strong>and</strong> given that the cell is in quasi water potential<br />

equilibrium with its wall (i.e., the values of ψw for the cell <strong>and</strong><br />

its surrounding wall [apoplasm] are close to being equal), then<br />

(3)<br />

the value of ψP in the cell (CC-SE) is given by:<br />

ψP = ψw − ψπ<br />

Armed with the above information, we can identify key elements<br />

that may constrain rates of phloem transport.<br />

Sieve element osmotic potentials are determined<br />

by phloem loading/retrieval<br />

<strong>The</strong> ψπ of the SE content (generally termed phloem sap) is<br />

primarily (60% to 75%) determined by one of several sugar<br />

species (sucrose, polyol or a raffinose family oligosaccharides<br />

– RFOs) with K + <strong>and</strong> the accompanying anions accounting for<br />

most of the remaining osmotic potential (Turgeon <strong>and</strong> Wolf<br />

2009). Thus, loading/retrieval of sugars plays a key role in<br />

setting the SE hydrostatic pressure (Equation 4), <strong>and</strong>, for this<br />

reason, we shall primarily focus attention on sugar transport,<br />

but where appropriate, we will comment on the involvement of<br />

other solutes.<br />

Proposed phloem-loading mechanisms are based on thermodynamic<br />

considerations <strong>and</strong> cellular pathways of loading.<br />

<strong>The</strong> most intensively studied is the apoplasmic loading mechanism<br />

in which sucrose (or polyol) loading requires the direct<br />

input of metabolic energy (Figure 13B, I), <strong>and</strong> is widespread<br />

amongst monocot <strong>and</strong> herbaceous eudicots species. An<br />

energy-dependent symplasmic loading mechanism also has<br />

been described. Here, sugar (sucrose or polyol) diffuses down<br />

its concentration gradient through a symplasmic route from<br />

mesophyll cells to specialized CCs, termed intermediary cells<br />

(ICs), where biochemical energy is required for sucrose/polyol<br />

conversion to large RFOs. This loading mechanism is referred<br />

to as the polymer trap mechanism (Figure 13B, II) <strong>and</strong> is thought<br />

to operate predominantly in herbaceous eudicots.<br />

Another loading system has been suggested to operate in<br />

woody plants (Davidson et al. 2011; Liesche <strong>and</strong> Schulz 2012)<br />

in which sugars are passively loaded through diffusion, driven<br />

by high sugar concentrations maintained in the mesophyll<br />

cell cytosol (Rennie <strong>and</strong> Turgeon 2009; Turgeon 2010a). This<br />

pathway is considered to be symplasmic, based on observed<br />

high PD densities between each cellular interface from mesophyll<br />

to SE-CC complexes (Figure 13B, III). It has also been<br />

suggested that delivery of sugars into SE-CC complexes could<br />

be achieved by bulk flow operating through interconnecting PD<br />

(Voitsekhovskaja et al. 2006).<br />

Several caveats must be considered for both the symplasmic<br />

diffusion <strong>and</strong> bulk flow models for passive phloem loading.<br />

If PD were to allow diffusion of sugars from mesophyll cells<br />

into sieve tubes, then all similarly-sized metabolites <strong>and</strong> ions<br />

should also pass into SE-CC complexes; i.e., the system would<br />

lack specificity. For situations in which bulk flow might transport<br />

sucrose, again all other soluble constituents present within the<br />

cytosol of cells forming the loading pathway should also gain<br />

(4)


Figure 13. Diagrammatic representation of resource flow through the phloem pathway.<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 323<br />

(A) Overall flow of resources through the phloem pathway. Within source leaves, resources (nutrients – green arrows <strong>and</strong> water – blue<br />

arrows) are loaded into sieve tubes (ST) of the collection phloem (chalk) formed by a network of minor veins. <strong>The</strong>se loaded solutes lower ψπ<br />

in the ST that then causes water to move, by osmosis, across the ST semi-permeable plasma membrane resulting in a high ψPsl condition.<br />

This osmotically-driven increase in ψP serves as the thermodynamic driving force to drive bulk flow (green highlighted blue arrows) of ST sap<br />

throughout the phloem system. In this context, loaded resources flow from collection phloem STs into STs of lower order veins, functioning as<br />

transport phloem (light blue) for export from source leaves. Transport phloem supports long-distance axial transport of resources from source<br />

leaves to sinks through STs of exceptionally high hydraulic conductivities that homeostatically sustain their ψP by resource exchange with<br />

surrounding tissues (curved green superimposed on blue arrows). Upon reaching the release phloem (light mauve), resources are unloaded<br />

from STs by bulk flow through plasmodesmata (PD) that interconnect the surrounding cells. <strong>The</strong> difference in pressure potential between the<br />

source <strong>and</strong> sink (ψPsl – ψPs) represent the hydrostatic pressure differential that drives bulk flow through the phloem pathway from source to<br />

sink.<br />

(B) Phloem loading pathways <strong>and</strong> mechanisms within source leaves. Photosynthetically reduced carbon, generated in chloroplasts, is used<br />

to drive sucrose (Suc) or polyol (Poly) biosynthesis (green broken arrows) within the cytosol of mesophyll cells (MC). Excess Suc/Poly is<br />

transiently stored in vacuoles (V) of mesophyl cells <strong>and</strong>, along with carbon from remobilized chloroplastic starch grains (SG), buffers their<br />

cytosolic pool sizes available for phloem loading. Suc/poly (green arrows) moves from mesophyll cells along a phloem-loading pathway that<br />

includes bundle sheath cells (BSC), phloem parenchyma cells (PPC), companion cells (CC) or intermediary cells (IC) to finally enter sieve<br />

elements (SE) of the collection phloem. Three loading mechanisms are considered to function in different species. (I) Active apoplasmic<br />

loading: Suc (<strong>and</strong>/or Poly) is first released from phloem parenchyma cells (PPCs) by the action of a permease (Chen et al. 2012b), <strong>and</strong><br />

subsequently retrieved into SE-CC complexes by symporters located along SE-CC plasma membranes. (II) Symplasmic loading: Raffinose<br />

oligosaccharides (RFOs) are synthesised in specialized CCs, termed ICs, from Suc delivered symplasmically from MCs. <strong>The</strong> larger molecular


324 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

entry into the phloem. When operating over a prolonged period,<br />

either of these proposed passive-loading systems would be<br />

anticipated to cause a perturbation to metabolism within the<br />

mesophyll cells.<br />

Rates of phloem loading depend upon the pool size of each<br />

transported solute available for loading, as well as the loading<br />

<strong>and</strong> retrieval mechanisms. For sugars, sucrose (Grodzinski<br />

et al. 1998) <strong>and</strong> polyol (Teo et al. 2006) pools are generated<br />

in mesophyll cells, whereas RFOs are synthesized from<br />

sucrose that enters the specialized ICs (Turgeon <strong>and</strong> Wolf<br />

2009)(Figure 13B, II). Irrespective of sugar species <strong>and</strong> phloem<br />

loading mechanism, during the photoperiod, transported sugars<br />

arise from current photosynthesis <strong>and</strong> export rates are<br />

linked positively with the sugar pool size (Grodzinski et al.<br />

1998; Leonardos et al. 2006; Lundmark et al. 2006). During<br />

the night, sugar pools are fed by starch reserves remobilized<br />

from chloroplasts (Smith <strong>and</strong> Stitt 2007) <strong>and</strong> sugars released<br />

from vacuolar storage in mesophyll cells (Eom et al. 2011)<br />

(Figure 13B). Depending upon carbon gain by leaf storage pools<br />

during the preceding photoperiod, remobilizing reserves during<br />

the night can sustain sugar pool sizes <strong>and</strong>, hence, export rates<br />

(Grimmer <strong>and</strong> Komor 1999).<br />

In situations where source leaves are operating at suboptimal<br />

photosynthetic activity, analyses of metabolic control<br />

have provided estimates that source leaf metabolism exercised<br />

approximately 80% of the control exerted over photoassimilates<br />

transported into developing potato tubers (Sweetlove et al.<br />

1998). However, the relationship between leaf metabolism <strong>and</strong><br />

export rates also depends upon prevailing source/sink ratios.<br />

This can be illustrated by studies aimed at investigating effects<br />

associated with CO2 enrichment. Under conditions of source<br />

limitation, leaf photosynthetic rates are increased substantially<br />

by CO2 enrichment, <strong>and</strong> are matched proportionately by those<br />

of photoassimilate export (Farrar <strong>and</strong> Jones 2000). In contrast,<br />

more attenuated responses of leaf photosynthetic rates are<br />

elicited by CO2 enrichment under sink limitation, <strong>and</strong> these are<br />

not proportionately matched by export (Grodzinski et al. 1998;<br />

Grimmer <strong>and</strong> Komor 1999). <strong>The</strong> latter response suggests that,<br />

under sink limitation, predominant control of photoassimilate<br />

transport shifts to processes downstream of source leaf sugar<br />

metabolism.<br />

Estimates of membrane fluxes of sucrose loaded into SE-<br />

CC complexes in sugar beet leaves fall into the maximal<br />

range for plasma membrane transporter activity (Giaquinta<br />

1983). <strong>The</strong>refore, if sucrose transporters are indeed operating<br />

at maximum capacity, then their overexpression might<br />

be expected to result in enhanced rates of phloem loading<br />

<strong>and</strong> photoassimilate export. However, overexpression of the<br />

spinach sucrose transporter (SoSUT1) in potato, while altering<br />

leaf metabolism, exerted no impact on biomass gain by the<br />

tubers (Leggewie et al. 2003). This finding indicates an absence<br />

of any constraint imposed by endogenous sucrose transporters<br />

on phloem loading. Indeed, phloem loading can respond quite<br />

rapidly (within minutes) to changes in sink dem<strong>and</strong> (Lalonde<br />

et al. 2003).<br />

A striking example of the dynamic range available to the<br />

phloem loading system is shown by studies performed on<br />

Ricinus, a plant whose phloem sap will exude (bleed) from<br />

severed SE-CC complexes. Here, excisions made in Ricinus<br />

stems reduced ψP to zero in this region of the sieve tube<br />

system. This treatment resulted in exudation of phloem sap<br />

from the severed sieve tubes <strong>and</strong> an increase in translocation<br />

<strong>and</strong>, hence, phloem-loading rates of sucrose, by an order of<br />

magnitude (Smith <strong>and</strong> Milburn 1980a). This observed rapid<br />

response is envisaged to reflect signaling from the sink region<br />

(in this case, the site of SE-CC stem excision) to source leaves.<br />

Here, pressure-concentration waves transmitted through interconnecting<br />

sieve tubes (Mencuccini <strong>and</strong> Hölttä 2010) could act<br />

to regulate transporter activity mediating phloem loading (Smith<br />

<strong>and</strong> Milburn 1980b; Ransom-Hodgkins et al. 2003).<br />

←−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−<br />

sizes of RFOs are thought to prevent their backward diffusion through PD interconnecting ICs with PPCs; however, the dilated PD that<br />

interconnect the IC-SE complexes permit forward diffusion of RFOs into the SEs (polymer trap model). (III) Passive – symplasmic loading:<br />

Suc or Poly is proposed to move by diffusion through the symplasm, via PD, down their concentration gradients from MCs to SEs. For all<br />

phloem-loading mechanisms, water enters (curved blue arrows) SE/CC or IC complexes through aquaporins (paired khaki ovals) (Fraysse<br />

et al. 2005).<br />

(C) Phloem unloading pathways from release phloem in sink organs. In all sinks, it is highly probable that imported resources are unloaded<br />

symplasmically by bulk flow from release phloem SE-CC complexes (green-highlighted blue arrows) into adjacent PPCs. Onward resource<br />

movement through the phloem-unloading pathway may occur by the following pathways. (I) Continuous symplasmic unloading: here,<br />

resources (green-highlighted blue arrows) likely continue to move by diffusion through PD into surrounding phloem parenchyma (PC)<br />

<strong>and</strong> ultimately sink cells (SC). (II) Apoplasmic step unloading: here, the phloem-unloading route involves resource transit through the sink<br />

apoplasm due to a symplasmic discontinuity in unloading pathways at either the PPC/SC or PC/SC interface. Membrane exchange of nutrients<br />

to, <strong>and</strong> from, the sink apoplasm occurs by transporter-mediated (khaki circles) membrane efflux <strong>and</strong> influx mechanisms, respectively. In both<br />

cases, water exiting SE-CC complexes can enter SCs (in the case of growing sinks) or, for non-exp<strong>and</strong>ing storage sinks, water returns to<br />

the xylem transpiration stream by exiting PPC/PCs (blue curved arrows) to the sink apoplasm through aquaporins (paired khaki ovals).


Other layers of post-translational control of sucrose transporter<br />

activity include protein-protein interactions, e.g., SUT1-<br />

SUT4 regulation of phloem loading (Chincinska et al. 2008),<br />

redox-induced dimerization of SUT1 (Krügel et al. 2008), <strong>and</strong><br />

cytochrome b5 interaction with MdSUT1 <strong>and</strong> MdSOT6 (Fan<br />

et al. 2009). Interestingly, the question as to whether symplasmic<br />

loading species also have the capacity for short-term<br />

adjustments in phloem loading capacity is less certain (Amiard<br />

et al. 2005).<br />

Magnetic resonance imaging studies, conducted on longdistance<br />

transport of water through the vascular system in<br />

a range of species, have established that phloem transport<br />

remains unaffected by diurnal variations in transpiration-driven<br />

changes in apoplasmic leaf water potential (Windt et al. 2006).<br />

Thus, a regulatory mechanism must operate to maintain a<br />

constant pressure gradient (ψP) to drive bulk flow through<br />

sieve tubes. This likely involves osmoregulatory activities<br />

at the level of the SE-CC complex (Pommerrenig et al.<br />

2007).<br />

In general, it would appear that phloem loading of major<br />

osmotic species (sugars <strong>and</strong> K + ) does not constrain phloem<br />

transport under optimal growth conditions. During periods of<br />

abiotic stress, phloem loading can minimize the impact of<br />

water/salt stress through osmoregulatory activities of cells<br />

comprising the phloem-loading pathway(s) (Koroleva et al.<br />

2002; Pommerrenig et al. 2007). Interestingly, <strong>and</strong> perhaps<br />

surprisingly, both apoplasmic (Wardlaw <strong>and</strong> Bagnall 1981) <strong>and</strong><br />

symplasmic (Hoffman-Thom et al. 2001) loaders undergo maintenance<br />

of phloem loading activities in cold-adapted plants.<br />

This indicates that changes in the viscosity of the phloem<br />

translocation stream may have little impact on bulk flow through<br />

sieve tubes. In contrast, elevated temperatures can slow<br />

translocation by callose occlusion of sieve pores (Milburn <strong>and</strong><br />

Kallarackal 1989). In addition, deficiencies of K + <strong>and</strong> Mg 2+ can<br />

impact apoplasmic loading of sucrose into SE-CC complexes<br />

(Hermans et al. 2006). In the case of K + , this is thought to reflect<br />

a limitation in charge compensation across the SE-CC plasma<br />

membrane which could impede the operation of the sucrose-H +<br />

symport system (Deeken et al. 2002), whereas Mg 2+ deficiency<br />

could lower the availability of Mg 2+ -ATP which serves as<br />

substrate for the H + -ATPase that generates the proton motive<br />

force to power the sucrose H + symporter (Cakmak <strong>and</strong> Kirkby<br />

2008).<br />

In terms of minor osmotic species, direct control of their<br />

phloem translocation rates is determined entirely by the concentrations<br />

to which they accumulate in SE-CC complexes.<br />

This situation is nicely illustrated by studies performed on transgenic<br />

peas expressing a yeast S-methylmethione transporter<br />

under the control of the phloem-specific AtAAP1 promoter.<br />

Here, S-methylmethione levels in developing seeds were found<br />

to be proportional to their concentrations detected in phloem<br />

exudates (Tan et al. 2010).<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 325<br />

Mechanisms of phloem unloading<br />

<strong>The</strong> cellular pathway of phloem unloading may extend, functionally,<br />

from SE lumens of the release phloem to sites of<br />

nutrient utilization/storage in the particular sink organ/tissue<br />

(Lalonde et al. 2003; Figure 13C). Within these bounds, the<br />

cellular pathways followed circumscribe the physical conditions<br />

under which an unloading mechanism operates. Most sink<br />

systems investigated to date have PD interconnecting the<br />

SE-CC complex to cells of the surrounding ground tissues,<br />

<strong>and</strong>, thus, confer the potential for universal symplasmic unloading<br />

(Figure 13C, I). In general, such routes for symplasmic<br />

unloading have low densities of PD that interconnect<br />

SE-CCs with adjacent phloem parenchyma cells. Thus, a<br />

marked bottleneck for symplasmic nutrient delivery may exist<br />

at this cellular interface.<br />

Unloading routes in a variety of sink systems have<br />

been mapped by using membrane-impermeant fluorochromes<br />

loaded into phloem of source leaves. Upon import into the<br />

release phloem zone, fluorochrome movement can be retained<br />

within the vascular system of fleshy fruit during their major<br />

phase of sugar accumulation (e.g., apple, a sorbitol transporter<br />

(Zhang et al. 2004), grape berry, a sucrose transporter (Zhang<br />

et al. 2006), <strong>and</strong> cucumber, an RFO transporter (Hu et al.<br />

2011). However, more commonly, the fluorochrome moves<br />

symplasmically out from the phloem into surrounding ground<br />

tissues (Figure 13C, I) as found for root <strong>and</strong> shoot meristems<br />

(Stadler et al. 2005), exp<strong>and</strong>ing leaves (Stadler et al. 2005),<br />

young fruit prior to their major phase of sugar accumulation<br />

(Zhang et al. 2006), <strong>and</strong> developing seeds in which movement<br />

is restricted to maternal tissues (Zhang et al. 2007).<br />

In developing fruits, during the phase of sugar accumulation,<br />

SE-CC complexes are thought to be the site of sucrose release<br />

into the fruit apoplasm (Zhang et al. 2004, 2006; Hu et al. 2011).<br />

Studies conducted on tomato fruit have indicated that the<br />

cumulative membrane surface area of SE-CC complexes would<br />

be barely adequate to support sucrose unloading at maximal<br />

fluxes known to be associated with membrane transport. In<br />

contrast, using the range of reported PD-associated fluxes<br />

(Fisher 2000), it can be shown that PD densities could readily<br />

accommodate unloading of sucrose into surrounding phloem<br />

parenchyma cells (Figure 13C, II). Clearly, further studies are<br />

required to resolve whether or not phloem unloading universally<br />

includes a symplasmic passage from SE-CC complexes to<br />

phloem parenchyma cells in the release phloem zone, as found<br />

for developing seeds (Zhang et al. 2007) (Figure 13C, II).<br />

An obvious constraint on facilitated apoplasmic unloading<br />

from SE-CC complexes is a co-requirement for a hydrolysable<br />

transported sugar (e.g., sucrose or RFO) <strong>and</strong> an invertase<br />

present within the cell walls of the release phloem zone. This<br />

combination ensures maintenance of an outwardly directed diffusion<br />

gradient for the transported sugar, due to its conversion


326 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

into a different chemical species by the cell wall invertase.<br />

Some fleshy fruits (Zhang et al. 2006; Hu et al. 2011) <strong>and</strong><br />

pre-storage phase seeds (Zhang et al. 2007) satisfy these<br />

requirements. However, these features do not apply to plants<br />

that transport polyol through their phloem, as exemplified by<br />

apple <strong>and</strong> temperate seeds during their storage phase; these<br />

systems may well rely on energy-coupled sugar release to the<br />

seed apoplasmic space (Zhang et al. 2004, 2007) (Figure 13C,<br />

II). In this context, maximal activities of symporters retrieving<br />

sucrose from seed apoplasmic spaces into pea cotyledons or<br />

wheat endosperm cells clearly constrain phloem unloading,<br />

as shown by increases in seed dry weight of transgenic<br />

plants overexpressing these transporters (Rosche et al. 2002;<br />

Weichert et al. 2010) (Figure 13C, II).<br />

A less obvious constraint, but an essential component of an<br />

unloading mechanism, is that unloading rates of all solutes (<strong>and</strong><br />

particularly the major osmotic species) <strong>and</strong> water must match<br />

their rates of phloem import. Any mismatch will impact water relations<br />

of release phloem sieve tubes <strong>and</strong>, hence, translocation<br />

rates. This requirement is essentially irreconcilable if significant<br />

leakage were to occur by diffusion. To ensure matching rates<br />

of phloem import <strong>and</strong> membrane efflux, it is important to stress<br />

that membrane transport of all solutes <strong>and</strong> water needs to be<br />

facilitated. This allows activities of membrane transporters to be<br />

potentially coordinated to ensure that rates of resource import<br />

through, <strong>and</strong> unloading from, sieve tubes in the release phloem<br />

are matched (see Zhang et al. 2007).<br />

A simple resolution to this problem is for unloading from the<br />

SE-CC complex to follow a symplasmic route <strong>and</strong> occur by<br />

bulk flow. Currently, experimental support for bulk flow, as an<br />

unloading mechanism, is limited. Such evidence is derived from<br />

experiments in which hydrostatic pressure differences between<br />

SE-CC complexes <strong>and</strong> surrounding cells were manipulated at<br />

root tips (e.g., Gould et al. 2004b; Pritchard et al. 2004). However,<br />

given that PD may represent a low hydraulic conductivity<br />

pathway, a significant pressure differential would be required to<br />

ensure the operation of an effective bulk flow delivery system.<br />

Consistent with this prediction, large osmotic potential differences<br />

(−0.7 MPa to −1.3 MPa) between SEs of release<br />

phloem <strong>and</strong> downstream cells have been measured in symplasmic<br />

unloading pathways of root tips (Warmbrodt 1987;<br />

Pritchard 1996; Gould et al. 2004a) <strong>and</strong> developing seeds<br />

(Fisher <strong>and</strong> Cash-Clark 2000b). <strong>The</strong>se differences translate<br />

into equally large differences in hydrostatic pressures, since<br />

sink apoplasmic ψP values approach zero (see Lalonde et al.<br />

2003). Expulsion of sieve tube contents by bulk flow into the<br />

much larger cell volumes of phloem parenchyma cells will<br />

dissipate the hydrostatic pressure of the expelled phloem sap<br />

within these cells. This, together with frictional drag imposed<br />

by a low hydraulic conductivity PD pathway, dictates that<br />

the large differentials in hydrostatic pressure between SE-<br />

CC complexes <strong>and</strong> phloem parenchyma cells are the result<br />

rather than the cause of bulk flow (Fisher <strong>and</strong> Cash-Clark<br />

2000b).<br />

<strong>The</strong> above considerations draw attention to the possibility<br />

that hydraulic conductivities of PD linking SE-CC complexes<br />

with phloem parenchyma cells play a significant role in regulating<br />

phloem unloading. Interestingly, size exclusion limits of<br />

PD at these cellular boundaries appear to be unusually high<br />

in roots (60 kDa; Stadler et al. 2005), sink leaves (50 kDa;<br />

Stadler et al. 2005) <strong>and</strong> developing seeds (400 kDa; Fisher <strong>and</strong><br />

Cash-Clarke 2000a), compared to the frequently reported value<br />

of 0.8 kDa −1.0 kDa for PD linking various cell types in ground<br />

tissues (Fisher 2000). However, size exclusion limits based on<br />

molecular weight can be misleading, <strong>and</strong> Stokes radius is a<br />

preferable measure (Fisher <strong>and</strong> Cash-Clarke 2000a). Taking<br />

this into account, PD hydraulic conductivities computed on this<br />

basis appear to be sufficient to accommodate the required<br />

rates of bulk flow out from the SE-CC complex (Fisher <strong>and</strong><br />

Cash-Clark 2000b).<br />

Large PD conductivities offer scope for considerable control<br />

over bulk flow across the SE-CC complex <strong>and</strong> adjoining phloem<br />

parenchyma cellular interfaces. A hint that such a system may<br />

operate is illustrated by studies on developing wheat grains.<br />

Here, imposition of a pharmacological block on sucrose uptake<br />

into the endosperm of an attached wheat grain was not accompanied<br />

by a change in sucrose concentration in cells forming<br />

the unloading pathway within maternal grain tissue (Fisher<br />

<strong>and</strong> Wang 1995). This finding points to a direct link between<br />

sucrose uptake by the endosperm <strong>and</strong> PD conductance at SE-<br />

CC-phloem parenchyma cell interfaces. How PD gating could<br />

be linked with sink dem<strong>and</strong> has yet to be determined, but this<br />

would have significant implications for phloem translocation.<br />

Phloem-imported water drives cell expansion in growing<br />

sinks (Walter et al. 2009). However, in non-exp<strong>and</strong>ing storage<br />

sinks, phloem-imported water is recycled by the xylem back to<br />

the parent plant body (Choat et al. 2009). This necessitates<br />

water exit across plasma membranes, irrespective of the unloading<br />

pathway, <strong>and</strong> likely depends upon movement facilitated<br />

by aquaporins (Zhou et al. 2007). Thus, except for symplasmic<br />

unloading into growing sinks, aquaporins could play a vital role<br />

in constraining rates of phloem unloading <strong>and</strong>, hence, overall<br />

phloem transport (Figure 13C).<br />

Transport phloem – Far from being a passive conduit<br />

interconnecting sources <strong>and</strong> sinks<br />

Compared to collection <strong>and</strong> release phloem, transport phloem<br />

(Figure 13A) extends over considerable distances of up to 100 m<br />

in tall trees. Axial flows along sieve tubes occur at astonishingly<br />

high rates, as demonstrated by estimates of specific mass<br />

transfers of approximately 500 g biomass m −2 sieve tube crosssectional<br />

area s −1 (Canny 1975). For a time, these observations<br />

supported the notion that sieve tube cross-sectional areas


(Equation 1) were a major constraint over rates of phloem<br />

translocation (Canny 1975), a notion later dispelled by the<br />

finding that specific mass transfer rates could be elevated by<br />

an order of magnitude in modified plant systems (Passioura<br />

<strong>and</strong> Ashford 1974; Smith <strong>and</strong> Milburn 1980a; Kallarackal <strong>and</strong><br />

Milburn 1984). More recently, a quest to obtain greater insights<br />

into sieve tube transport function (Thompson 2006) has reignited<br />

an interest in obtaining measures of sieve tube geometries<br />

to obtain meaningful estimates of sieve tube hydraulic<br />

conductivities (Thompson <strong>and</strong> Wolniak 2008; Mullendore et al.<br />

2010; Froelich et al. 2011).<br />

On the assumption that flows through sieve tube lumens<br />

<strong>and</strong> sieve pores are laminar, hydraulic conductivity (L) can be<br />

derived from the Hagen-Poiseulli Law as:<br />

L = πr 2 /8ηl (5)<br />

A technically innovative <strong>and</strong> thorough quantitative plant<br />

anatomical study, undertaken across a range of eudicot herbaceous<br />

life forms, yielded estimates of sieve tube hydraulic<br />

conductivities (Mullendore et al. 2010). Values of L were<br />

found to be dominated by sieve pore radii, as predicted<br />

by Equation 5, <strong>and</strong> were inversely related with independent<br />

measures of phloem transport velocities. Such an outcome<br />

is in contradiction to the Hagen-Poiseulli Law. Together<br />

with other phloem transport anomalies, such as gradients<br />

of sieve tube hydrostatic pressures not scaling with plant<br />

size (Turgeon 2010b), these studies point to a key feature<br />

in phloem translocation likely being overlooked. As outlined<br />

below, we contend that sieve tube properties that establish conduits<br />

of exceptionally high hydraulic conductivities, combined<br />

with their ability to osmoregulate, can account for transport<br />

phloem being capable of supporting high fluxes over long (m)<br />

distances.<br />

As indicated by Equation 5, sieve tube length (l), <strong>and</strong> most<br />

importantly, sieve pore radius (r), have a direct influence on<br />

hydraulic conductivity, along with sap viscosity (η). Sieve-tube<br />

sap viscosity varies approximately 2.5-fold across the range<br />

of measured sieve-tube sucrose concentrations (300 mM to<br />

1000 mM sucrose) <strong>and</strong>, hence, could influence sieve tube<br />

hydraulic conductivity. <strong>The</strong> viscosity of a 600 mM sucrose<br />

solution increases approximately 2-fold from 25 ◦ C to 0 ◦ C<br />

(Misra <strong>and</strong> Varshin 1961). A controlled experiment, in which an<br />

approximate doubling of phloem sap viscosity can be achieved<br />

without impacting source or sink activities, is to gradually (to<br />

avoid shock) cool a stem zone (from 25 ◦ C to just above 0 ◦ C).<br />

Absence of any slowing of transport rates through the cooled<br />

zone (Wardlaw 1974; Minchin <strong>and</strong> Thorpe 1983; Peuke et al.<br />

2006) argues that this range of phloem sap viscosities exerts<br />

little influence over phloem transport.<br />

To date, studies of hydraulic conductivities deduced from<br />

sieve tube geometries have yielded ambiguous results (Mullendore<br />

et al. 2010; Froelich et al. 2011). However, indirect<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 327<br />

observations suggest that sieve tube hydraulic conductivity<br />

is unlikely to constrain phloem transport. For instance, removal<br />

of substantial proportions of transport phloem crosssectional<br />

area from the stem had little impact on rates of<br />

translocation through the narrowed phloem zone (Wardlaw<br />

<strong>and</strong> Moncur 1976), thus indicating a considerable spare capacity<br />

for phloem transport. A spectacular example illustrating<br />

excess transport capacity is provided by a study of translocation<br />

rates through pedicels supporting developing apical<br />

fruits in racemes of Ricinus. Upon removal of apical fruits,<br />

<strong>and</strong> allowing exudation from their severed petiole stumps to<br />

proceed, translocation rates increased from 166 g to 3111<br />

g biomass m −2 sieve-tube area s −1 , a response suggesting<br />

that phloem transport was sink controlled, not phloem pathway<br />

controlled (Smith <strong>and</strong> Milburn 1980a; Kallarackal <strong>and</strong> Milburn<br />

1984).<br />

Experimental measurements conducted using a microfluidic<br />

system simulating phloem pressure flow as well as transport<br />

properties of ‘real’ plants (including tall trees) also yielded<br />

results that conformed with predictions of the Münch model<br />

(Jensen et al. 2011, 2012). Studies performed on an Arabidopsis<br />

mutant lacking P-protein agglomerations in sieve tubes, <strong>and</strong><br />

hence conferring higher sieve tube conductivity, established<br />

that these plants had similar transport velocities (or volume<br />

flux – see Equation 2) to WT plants (Froelich et al. 2011).<br />

Collectively, these studies support the notion that sieve-tube<br />

hydraulic conductivities do not impose a significant limitation<br />

on transport fluxes along phloem pathways, even over considerable<br />

lengths of sieve tubes. Rather, as discussed above, the<br />

majority of control may well be exercised by bulk flow through<br />

PD linking SE-CC complexes of release phloem with adjacent<br />

phloem parenchyma cells (Figure 13C).<br />

Pressure-concentration waves generated by phloem unloading<br />

are transmitted over considerable distances (m) at velocities<br />

an order of magnitude higher than those of phloem translocation<br />

(Smith <strong>and</strong> Milburn 1980a; Mencuccini <strong>and</strong> Hölttä 2010).<br />

Such a signaling system is envisioned to underpin unified<br />

responses by all SE-CC complexes, comprising phloem paths<br />

from release to collection phloem, to altered resource dem<strong>and</strong>s<br />

by the various sinks (Thompson 2006). <strong>The</strong>se responses are<br />

mediated by turgor-regulated membrane transport of sugars<br />

into SE-CC complexes; these sugars are supplied from mesophyll<br />

<strong>and</strong> axial pools for compensation within collection <strong>and</strong><br />

transport phloem, respectively (Figure 13A). This mechanism<br />

results in homeostasis of hydrostatic pressure in sieve tubes<br />

along the phloem pathway (Gould et al. 2004a). <strong>The</strong> action<br />

of this pressure-concentration signaling system could account<br />

for differentials in hydrostatic pressures between collection <strong>and</strong><br />

release phloem not scaling with transport distance, particularly<br />

in tall trees (Turgeon 2010b). In addition, such a mechanism<br />

could maintain sieve-tube sap concentrations of all solutes<br />

(Gould et al 2004a) <strong>and</strong>, hence, their rates of phloem transport.


328 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

<strong>The</strong> interconnected nature of the plant’s vascular system<br />

does not appear to exert a constraint over patterns of resource<br />

flow, as demonstrated by their plasticity in response to changes<br />

in source/sink ratios (Wardlaw 1990). This leads one to the<br />

notion that the elements of the phloem function, kinetically, as<br />

a single pool of resources. A ‘common’ kinetic pool of phloem<br />

sap depends upon hydraulic connectivity between all functional<br />

phloem conduits. Lateral sieve areas, phloem anastomoses<br />

(Evert 2006) <strong>and</strong> intervening phloem parenchyma cells (Oross<br />

<strong>and</strong> Lucas 1985) provide conduits for resource flows to function<br />

as a common kinetic pool.<br />

<strong>The</strong> above described transport behaviors led Don Fisher<br />

to propose a variant of the Münch pressure flow model in<br />

which he envisaged phloem systems functioning as highpressure<br />

manifolds (Fisher 2000). High hydrostatic pressures<br />

generated by phloem loading in source leaves (Gould et al.<br />

2005) are maintained throughout the transport phloem by<br />

osmoregulated loading in collection or re-loading by transport<br />

phloem (Figure 13). In the region of the release phloem,<br />

gradients in hydrostatic pressure across PD connecting SE-<br />

CC complexes to adjoining phloem parenchyma cells, in<br />

combination with PD hydraulic conductivity, control overall<br />

flows from source regions to each sink. As a corollary,<br />

relative magnitudes of PD conductivities between various<br />

sinks could control resource partitioning at a whole plant<br />

level.<br />

<strong>The</strong> high-pressure manifold model (Fisher 2000) accounts for<br />

all known aspects of phloem transport, except direct unloading<br />

across SE-CC plasma membranes. However, as we mentioned<br />

above, conclusive evidence for this pathway forming a major<br />

phloem-unloading route has not yet been established. Indeed,<br />

symplasmic flow into surrounding phloem parenchyma cells<br />

remains a real possibility in all sinks (Figure 13C) <strong>and</strong>, hence,<br />

the high-pressure manifold model appears to be universally<br />

applicable.<br />

Directing future studies to testing the phloem<br />

high-pressure manifold model<br />

In broad terms, there is strong evidence that the high-pressure<br />

manifold model (Fisher 2000) accounts for key elements underpinning<br />

phloem transport <strong>and</strong> resource partitioning at the<br />

whole plant level. <strong>The</strong> model highlights hydraulic conductivities<br />

of PD linking release phloem SE-CC complexes with<br />

phloem parenchyma cells as the pivotal point at which phloem<br />

transport is constrained both physically <strong>and</strong> physiologically.<br />

We consider the evidence sufficiently compelling to invest<br />

significant effort in future investigations to further test the<br />

general applicability of this model. Resolving the underpinning<br />

regulatory mechanisms could open up substantial biotechnological<br />

opportunities to divert biomass flows to enhance crop<br />

yields.<br />

Physical & Physiological Constraints on<br />

Xylem Function<br />

<strong>The</strong> xylem of the plant vascular system transports more fluid<br />

longer distances than any other vascular tissue. <strong>The</strong> collective<br />

flow of xylem sap summed over all the plants on a watershed<br />

can exceed the total runoff in streams (Schlesinger 1997).<br />

Typically, less than 5% of the xylem water is consumed<br />

by osmotically-driven cell expansion, <strong>and</strong> less than 1% is<br />

consumed by photosynthesis. <strong>The</strong> bulk of the transported<br />

water is lost to transpiration: the water evaporates from cell<br />

wall surfaces into the intercellular air spaces of the leaves,<br />

<strong>and</strong> diffuses out into the atmosphere through open stomata.<br />

Hence, the term “transpiration stream” is used to refer to xylem<br />

sap flow. Although the transpiration stream carries nutrients,<br />

molecular signals, <strong>and</strong> other compounds from roots to leaves,<br />

<strong>and</strong> evaporative cooling can minimize overheating of larger<br />

leaves, these benefits are usually regarded as secondary to the<br />

cost of having to lose such large quantities of water in exchange<br />

for stomatal CO2 uptake (Holtta et al. 2011). Under typical<br />

diffusion gradients, plants transpire hundreds of molecules of<br />

water for every CO2 molecule fixed by photosynthesis. If plants<br />

could evolve a way of obtaining CO2 without simultaneously<br />

losing water, their water consumption would be substantially<br />

reduced <strong>and</strong> water would presumably be much less of a limiting<br />

factor for their productivity.<br />

As expected for such a poor water-for-carbon exchange<br />

rate, plants have evolved a metabolically cheap mechanism for<br />

driving the transpiration stream; otherwise, the cost of moving<br />

water could easily exceed the meager energy return. According<br />

to the well-substantiated cohesion-tension mechanism summarized<br />

in Figure 14, water is pulled to the site of evaporation in<br />

the leaves by the tension established within the surface of the<br />

water at the top of the water column (capillary) (Pickard 1981).<br />

<strong>The</strong> plant functions more or less as a ‘water wick’. Once the<br />

‘wick’ is grown, the driving force for the transpiration stream<br />

is free of charge from the plant’s perspective. Most directly,<br />

the energy to drive the transpiration stream comes from the<br />

sun. However, despite its energetic efficiency, the cohesiontension<br />

mechanism has important limitations that constrain the<br />

productivity <strong>and</strong> survival of plants. Current research questions<br />

include the evolution, physiology, <strong>and</strong> ecology of these water<br />

transport constraints.<br />

<strong>The</strong> problem of frictional resistance to flow<br />

<strong>The</strong> basic wicking process (Figure 14A) presents a physical<br />

paradox. A narrower tube is better for generating capillary at<br />

the evaporating meniscus for pulling water up, but it is worse<br />

for creating high frictional resistance to the upward flow. <strong>The</strong><br />

maximum drop in pressure (Pmin) created by an air-waterinterface<br />

across a cylindrical pore is inversely proportional to


Figure 14. <strong>The</strong> cohesion-tension mechanism for transpirationdriven<br />

xylem flow.<br />

(A) <strong>The</strong> basic hydraulic lift process: evaporation from a meniscus<br />

coupled to bulk liquid flow by capillary action.<br />

(B) In plants, surface tension created at the surface of narrow pores<br />

within the cell walls (w) acts as the energy gradient to drive longdistance<br />

bulk flow through a low-resistance network of dead xylem<br />

conduits (c). Conduits are connected by pits (p), which also protect<br />

against air-entry <strong>and</strong> embolism in the inevitable event of damage<br />

(see Figures 15, 16). Water is filtered through living cell membranes<br />

at the root endodermis (en) by reverse osmosis.<br />

(C) Living cells (rounded rectangles enclosed by hatched cell<br />

walls) do not generate transpirational flow (ep, leaf epidermis; m,<br />

mesophyll; cs, Casparian strip of the endodermis; rc, root cortex).<br />

Water flows through them to the site of evaporation via symplastic (s<br />

arrows) <strong>and</strong> apoplastic routes (a arrows). In the root, the apoplastic<br />

route (a’) is interrupted by the Casparian strip. Transpiration is<br />

actively regulated by stomatal opening through which the waterfor-carbon<br />

exchange occurs (from Sperry 2011).<br />

the cylinder radius (Pmin proportional to 1/radius), whereas the<br />

hydraulic resistivity (resistance per unit length) through the<br />

cylinder increases with 1/radius 4 . <strong>The</strong> higher the resistivity,<br />

the lower the flow rate at which P drops to Pmin, pulling<br />

the meniscus down the tube <strong>and</strong> drying out the wick. <strong>The</strong><br />

evaporating menisci of the plant are held in the nanometerscale<br />

pores of primary walls facing internal (intercellular) air<br />

spaces. Although ideal for generating a potentially substantial<br />

driving force for pulling up the transpiration stream, the high<br />

resistivity of these pores to bulk flow limits the distance <strong>and</strong> rate<br />

of water flow. Hence, high flow resistivity of parenchymatous<br />

plant ground tissue was a major factor limiting the size of nonvascular<br />

(pre-tracheophyte) plants.<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 329<br />

With the evolution of xylem conduits, the wick paradox<br />

was solved: maximum capillary action at nano-scale cell wall<br />

pores is coupled to minimum bulk flow resistivity in micro-scale<br />

xylem conduit lumens for carrying the transpiration stream over<br />

most of the soil-to-leaf distance (Figure 14B). <strong>The</strong> conduits are<br />

dead cell wall structures that resist collapse by having lignified<br />

secondary walls. <strong>The</strong>y form a continuous apoplasmic pipeline<br />

from terminal leaf vein to root tip, propagating the tensional<br />

component to the supply of water in the soil. As water is pulled<br />

into the stele of the absorbing root tissues, it is forced across<br />

the endodermal membrane by reverse osmosis, providing a<br />

mechanism for filtration <strong>and</strong> selective uptake (Figure 14B, en).<br />

Like cell wall water, the soil water is held by capillary <strong>and</strong><br />

absorptive forces. In short, the cohesion-tension mechanism is<br />

a “tug-of-war” on a rope of liquid water between capillary forces<br />

in soil vs. plant apoplasm. Living protoplasts do not participate<br />

in driving the transpiration stream, but draw from it by osmosis<br />

during cell expansion growth <strong>and</strong> to stay hydrated <strong>and</strong> turgid<br />

(Figure 14C).<br />

Low resistivity xylem facilitates larger plants <strong>and</strong> higher flow<br />

rates (equals greater photosynthetic productivity), <strong>and</strong> xylem<br />

evolution is a history of innovations for presumably moving<br />

water more efficiently. <strong>The</strong> increasing literature on the topic<br />

is beyond the more physiological emphasis of this section<br />

of the current review, but one example will suffice. <strong>The</strong> rise<br />

of high-productivity angiosperms appears to coincide with the<br />

evolution of greater vein density within leaves, which minimizes<br />

the distance the transpiration stream flows in high-resistance<br />

parenchymatous ground tissue (Boyce et al. 2009). Presumably,<br />

the evolutionary pressure for maximizing hydraulic efficiency<br />

varied over geological time scales, being less during<br />

periods of higher atmospheric CO2 <strong>and</strong> arid (dry) conditions,<br />

<strong>and</strong> increasing during periods of low CO2 <strong>and</strong> mesic (wet)<br />

conditions (Boyce <strong>and</strong> Zwieniecki 2012).<br />

<strong>The</strong> reason that lower frictional resistance correlates with<br />

greater photosynthetic rate is because it also correlates with<br />

greater diffusive conductance of the stomatal pores (Meinzer<br />

et al. 1995; Hubbard et al. 2001). <strong>The</strong> physics of the xylem<br />

conduit does not account for this coupling between low flow<br />

resistance <strong>and</strong> high diffusive conductance (equates to high<br />

evaporation rate). As long as the integrity of the xylem conduit<br />

remains constant, its evaporation rate is essentially independent<br />

of its internal liquid phase flow resistance. <strong>The</strong> observed<br />

coupling must result from a physiological response of the plant.<br />

<strong>The</strong> simplest explanation for the coordination of low hydraulic<br />

resistance <strong>and</strong> high diffusive conductances is that stomata are<br />

responding in a feedback manner to some measure of leaf<br />

or plant water status (Sperry 2000; Brodribb 2009). Increased<br />

plant hydraulic resistances result in a greater tensional component<br />

(i.e., more negative values of ψP) at a given transpiration<br />

rate <strong>and</strong> soil water status. If the ψP falls below some regulatory<br />

set-point (which need not be constant), hydraulic or chemical


330 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

signals are sent to the stomatal complex, reducing the stomatal<br />

aperture <strong>and</strong> transpiration rate, which causes the ψP to rise<br />

back above the set-point. This ψP feedback is at least broadly<br />

consistent with most observed stomatal behavior in response to<br />

hydraulic resistance, as well as to evaporative dem<strong>and</strong> <strong>and</strong> soil<br />

moisture stress (Oren et al. 1999; Brodribb 2009; Pieruschka<br />

et al. 2010).<br />

Recent evidence indicates that the ancestral feedback was<br />

entirely passive, as indeed it appears to still be in many seedless<br />

vascular plants (Brodribb <strong>and</strong> McAdam 2011). Accordingly,<br />

as ψP becomes more negative, guard cell turgor drops without<br />

any chemical signaling or active osmotic adjustment. Ensuing<br />

stomatal closure reduces the transpiration rate, which causes<br />

ψP to stop falling <strong>and</strong> rise back up. Only with the divergence<br />

of the seed plants 360 Mya did apparently active signaling<br />

evolve, which involves ψP sensing mechanisms, triggering of<br />

abscisic acid <strong>and</strong> other chemical signaling molecules, <strong>and</strong><br />

active osmotic adjustment via ion pumps at the stomatal<br />

complex (McAdam <strong>and</strong> Brodribb 2012). <strong>The</strong> details of how this<br />

active feedback is achieved, the extent of active vs. passive<br />

mechanisms, the actual sites of evaporation within the leaf,<br />

the sites of water potential sensing, <strong>and</strong> the extent of hydraulic<br />

coupling between the stomatal complex <strong>and</strong> xylem are basic<br />

questions that are still poorly understood, <strong>and</strong> are the subject<br />

of considerable research <strong>and</strong> debate (Pieruschka et al. 2010;<br />

Mott <strong>and</strong> Peak 2011).<br />

Regardless of the feedback details, plants appear to have<br />

evolved to regulate ψP at the expense of sacrificing photosynthesis<br />

via reduced stomatal diffusive conductance. Presumably,<br />

this response avoids deleterious consequences of<br />

excessively negative ψP. Certainly most physiological processes<br />

are more energy-dem<strong>and</strong>ing at more negative ψP: the<br />

xylem conduit has to be stronger, <strong>and</strong> protoplasmic osmotic<br />

concentrations have to be greater. <strong>The</strong> implication is that<br />

there is some optimal midday ψP, in so far as it maximizes<br />

the cost/benefit margin of water transport vs. CO2 processing<br />

(Holtta et al. 2011). A priori, the optimal ψP would differ across<br />

habitats. For example, it would need to be more negative in drier<br />

habitats where plants have to pull harder to extract soil water as<br />

compared to wetter habitats. A full underst<strong>and</strong>ing of the costs<br />

associated with xylem transport requires detailed consideration<br />

of how xylem structure relates to its role in the water conducting<br />

process.<br />

Confining embolism with inter-conduit pitting<br />

While the evolution of xylem conduits solved the so called<br />

“wick” paradox, a new problem was created: the low-resistivity<br />

xylem conduits are necessarily too wide to generate, in <strong>and</strong><br />

of themselves, much of a tension, i.e., a negative ψP value.<br />

In the inevitable event that a conduit becomes damaged <strong>and</strong><br />

exposed to air, the surface tension present in the new meniscus<br />

spanning the conduit lumen is too weak to resist retreat (Pmin is<br />

not negative enough), <strong>and</strong> so water is pulled from this specific<br />

conduit into neighboring conduits <strong>and</strong> becomes embolized; i.e.,<br />

it eventually is filled with N2, O2, CO2 <strong>and</strong> water vapor gases.<br />

Thus, capillary rise within the xylem conduit lumen cannot do<br />

the job of pulling up the transpiration stream. For the system<br />

to function, the xylem conduits must be primed by being water<br />

filled from inception, as they are, having developed from living<br />

cells. Nevertheless, there must be a means of limiting the<br />

spread of embolism when inevitably a conduit is damaged,<br />

even if by normal developmental events such as abscission of<br />

parts or protoxylem rupture.<br />

<strong>The</strong> problem of embolism is mitigated most fundamentally<br />

by dividing the fluid conducting space into thous<strong>and</strong>s<br />

of overlapping <strong>and</strong> inter-connected conduits (Figure 14B, C).<br />

Each one embolizes as a unit because the inter-connections<br />

consist of porous partitions (inter-conduit pits) fine enough<br />

to trap <strong>and</strong> hold an air-water meniscus against a sufficiently<br />

negative ψP to minimize further gas propagation (Figure 14B,p).<br />

This multi-conduit system necessarily compromises hydraulic<br />

conductance because of the added resistance to flow through<br />

the inter-conduit pitting. Presumably, the lowest hydraulic resistance<br />

would be achieved by a single branching tube akin<br />

to the animal positive-pressure cardiovascular system. But a<br />

tensional system of such design would fail completely from a<br />

single point of air entry without any partitions to check the influx<br />

of gas.<br />

<strong>The</strong> presence of inter-conduit pitting is of great consequence<br />

for xylem functioning. <strong>The</strong> distribution of pitting <strong>and</strong> the structure<br />

<strong>and</strong> chemistry of individual pits influence both the flow<br />

resistance through the xylem <strong>and</strong> the Pmin for the xylem system<br />

(Choat et al. 2008). Although there is tremendous variation in<br />

inter-conduit pit structure across lineages, their basic structure<br />

has three elements held in common (Figure 15). As water flows<br />

from one conduit lumen to another, it passes through a pit<br />

aperture in the secondary wall, which opens into a usually<br />

wider pit chamber. Spanning the chamber is a porous pit<br />

membrane through which the water filters before passing out<br />

the downstream aperture. <strong>The</strong> pit membrane is the modified<br />

primary cell walls plus middle lamella of the adjacent conduits.<br />

<strong>The</strong>re is no cell membrane or protoplasm in the dead but<br />

functioning conduit.<br />

Hydrolysis during xylem cell death is thought to remove all<br />

hemicelluloses <strong>and</strong> a debatable portion of pectins from the<br />

pit membrane, leaving a porous cellulosic mesh of microfibrils<br />

(Butterfield 1995). However, atomic force microscopy suggests<br />

that the microfibrillar structure lies beneath a non-porous<br />

coating of amorphous non-fibrillar material (Pesacreta et al.<br />

2005). Pit membrane chemistry, structure, <strong>and</strong> development<br />

are crucial to underst<strong>and</strong>ing the frictional resistance to flow as<br />

well as the ability of the xylem to sustain significant tensions in<br />

the water column (Choat et al. 2008; Lee et al. 2012).


Figure 15. Inter-vessel pit structure in angiosperm wood.<br />

Upper-right insert shows brightfield image of two overlapping vessels.<br />

<strong>The</strong>ir common wall is studded with inter-vessel pits as shown in<br />

the main scanning electron microscopy (SEM) image where an intervessel<br />

wall has been sectioned to show individual pit structure. Pits<br />

consist of openings in the secondary wall (apertures) leading to pit<br />

chambers that are spanned by a pit membrane which is the modified<br />

primary cell wall <strong>and</strong> middle lamella of the adjacent vessel elements.<br />

<strong>The</strong> lower-right insert shows an SEM face view of the micro-porous<br />

pit membrane. Scale bars: 5 µm <strong>and</strong> 30 µm for the upper inset.<br />

Micrographs courtesy of Fredrick Lens, Jarmila Pittermann, <strong>and</strong><br />

Brendan Choat.<br />

Inter-conduit pits add substantial flow resistance to the xylem<br />

conduit lumen. An unobstructed lumen conducts water as<br />

efficiently as an ideal cylindrical capillary tube of the same<br />

diameter (Zwieniecki et al. 2001a; Christman <strong>and</strong> Sperry 2010).<br />

<strong>The</strong> most extensive survey indicates that adding inter-conduit<br />

pits increases flow resistivity over that of an unobstructed lumen<br />

by an average factor of 2.8 in conifers with unicellular tracheids<br />

<strong>and</strong> 2.3 in angiosperms with multicellular vessels (Hacke et al.<br />

2006; Pittermann et al. 2006a). <strong>The</strong> lower number for vessels<br />

is not surprising given that they are roughly 10 times longer<br />

than a tracheid of the same diameter (Pittermann et al. 2005),<br />

thereby spacing high resistance pits further apart <strong>and</strong> reducing<br />

the length-normalized resistance (resistivity). What is surprising<br />

is that the greater length of vessels does not have more of<br />

an effect: if inter-vessel pits have the same area-specific pit<br />

resistance as inter-tracheid pits, placing the end-walls 10 times<br />

further apart should increase lumen resistivity by less than a<br />

factor of 1.18. <strong>The</strong> higher observed factor of 2.3 indicates that<br />

inter-vessel pits have higher flow resistance than inter-tracheid<br />

ones, a difference consistent with anatomy <strong>and</strong> estimations<br />

based on modeling (Pittermann et al. 2005).<br />

Inter-vessel pits have nano-porous “homogeneous” pit membranes<br />

(pores usually < 100 nm) (Choat et al. 2008), or<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 331<br />

Figure 16. Structure <strong>and</strong> function of inter-conduit pits in<br />

conifer tracheids (left) <strong>and</strong> angiosperm vessels (right).<br />

(A) Pit membranes, face view.<br />

(B) Side view schematic of membranes within the pit chamber<br />

formed by the secondary walls. Pits open <strong>and</strong> functioning in water<br />

transport.<br />

(C) Schematic of pit location within conduit network.<br />

(D) Side view of pits in sealed position showing proposed airseeding<br />

process (from Pittermann et al. 2005).<br />

often with no pores detectable, whereas inter-tracheid pits<br />

of conifers have a highly porous margo (pores ≫ 500 nm)<br />

(Petty <strong>and</strong> Preston 1969) peripheral to a central thickened torus<br />

(Figure 16A, left). <strong>The</strong> greater porosity of the margo decreases<br />

the area-specific pit resistance by an estimated 59-fold relative<br />

to inter-vessel pits of angiosperms (Pittermann et al. 2005).<br />

<strong>The</strong> homogenous-type pit membrane is presumably ancestral,<br />

<strong>and</strong> the implication is that the evolution of efficient torus-margo<br />

pitting, within in the gymnosperm lineage, was as hydraulically<br />

advantageous as the evolution of vessels in angiosperms.


332 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Pit membrane chemistry interacts with xylem sap chemistry<br />

to influence xylem flow resistance in a very complex <strong>and</strong> poorly<br />

understood manner. According to the “ionic effect”, increasing<br />

the concentration of KCl up to 50 mM (encompassing the<br />

physiological range) can decrease resistivity anywhere from<br />

2% to 37% relative to pure water, depending on the angiosperm<br />

species <strong>and</strong> even the season (Nardini et al. 2011b). <strong>The</strong><br />

KCl effect is much less or even negative in the already<br />

low-resistance torus-margo pits of conifer species (Cochard<br />

et al. 2010b). Furthermore, this KCl effect can be reduced or<br />

eliminated in the presence of as little as 1 mM Ca 2+ in some<br />

species <strong>and</strong> conditions (Van Iepeeren <strong>and</strong> Van Gelder 2006),<br />

but not in others (Nardini et al. 2011b). Skepticism about the<br />

importance of the phenomenon in planta (Van Iepeeren 2007)<br />

has been answered with observations indicating KCl-mediated<br />

decreases in resistivity associated with embolism <strong>and</strong> exposure<br />

of branches to sunlight (Nardini et al. 2011b). Interestingly,<br />

adaptive adjustments in KCl concentration may be mediated<br />

by xylem-phloem exchange (Zwieniecki et al. 2004).<br />

<strong>The</strong> ionic effect has been localized to the pit membranes,<br />

but the mechanism remains unknown. A “hydrogel” model implicates<br />

ionic shrinkage of pit membrane pectins or equivalent<br />

hydrogel polymers, <strong>and</strong>, hence, a widening of membrane pores<br />

(Zwieniecki et al. 2001b). However, recent observations with<br />

atomic force microscopy do not support a pore-widening effect.<br />

Although KCl was observed to thin the membrane, pores were<br />

not observed, suggesting that the decrease in resistance resulted<br />

from membrane thinning <strong>and</strong> perhaps increased permeability<br />

of non-porous gel material (Lee et al. 2012). <strong>The</strong> extent of<br />

pectins or similar gel materials in pit membranes appears to be<br />

highly variable across species, perhaps underlying the extreme<br />

variation in the ionic effect (Nardini et al. 2011b). Uncertainty<br />

about the extent of hydrogel components of pit membranes has<br />

led to an alternative (<strong>and</strong> perhaps complementary) hypothesis<br />

that ions increase permeability by reducing the diffuse-double<br />

layer of cations lining negatively charged nano-scale pores<br />

in the membrane (Van Doorn et al. 2011b). All of these<br />

hypotheses are consistent with a minor effect in torus-margo<br />

pit membranes, with their large micro-scale pores between<br />

cellulosic str<strong>and</strong>s having presumably minimal pectin content.<br />

Although inter-conduit pits have the disadvantage of adding<br />

substantial flow resistance, they perform the highly advantageous<br />

function of trapping an air-water meniscus <strong>and</strong> minimizing<br />

the embolism event such that it does not compromise<br />

the conducting system (Figure 16B, C). <strong>The</strong> homogenous pits<br />

of angiosperm vessels have pores narrow enough to trap<br />

themeniscuswithaPmin negative enough to hold against a<br />

substantial range of negative ψP values (Figure 16D). <strong>The</strong> torusmargo<br />

pits function somewhat differently. <strong>The</strong> wider margo<br />

pores cannot sustain a very negative Pmin, but they can<br />

generate just enough pressure difference to aspirate the solid<br />

torus against the pit aperture on the water-filled side (Petty<br />

1972). In this way, the torus can seal off the pit with a sufficiently<br />

negative Pmin to minimize air passage (Figure 16D).<br />

While inter-conduit pits minimize the propagation of embolism,<br />

as the next section indicates, they nevertheless play<br />

a major role in limiting the tensional gradient that can be<br />

generated by the cohesion-tension mechanism.<br />

Limits to negative ψ P values: the problem of cavitation<br />

Periodically, the cohesion-tension mechanism comes under<br />

question for its prediction of liquid pressures that fall below<br />

the vapor pressure of water, <strong>and</strong> also below pure vacuum for<br />

a gas (Canny 1998; Zimmermann et al. 2004). A tree 30 m tall<br />

requires a ψP of −0.3 MPa on its stationary water column just<br />

to balance the gravity component. To this we need to add,<br />

say, −0.3 MPa to balance a favorable soil water potential<br />

of −0.3 MPa. Finally, we need to add the typical ψP of<br />

−1 MPa needed to overcome frictional resistance under midday<br />

transpiration rates. <strong>The</strong> required ψP totals −1.6 MPa. At sea<br />

level <strong>and</strong> 20 ◦ C, a vapor pressure of only −0.098 MPa will<br />

bring water to its boiling point, <strong>and</strong> −0.1013 MPa corresponds<br />

to pure vacuum for a gas. Clearly, for the cohesion-tension<br />

mechanism to operate, transition from the liquid phase to the<br />

vapor phase must be suppressed, <strong>and</strong> the xylem sap must<br />

remain in a metastable liquid state. <strong>The</strong> xylem sap is in effect<br />

super-heated, although “super-tensioned” is more descriptive.<br />

<strong>The</strong> liquid water column becomes analogous to a solid whose<br />

strong atomic <strong>and</strong> intermolecular bonds allow it to be placed<br />

under tension; i.e., water is a tensile liquid!<br />

<strong>The</strong> concept of metastable water is foreign to the macroscopic<br />

world of normal human experience, hence the cohesiontension<br />

skeptics. Water boils at 100 ◦ C, <strong>and</strong> vacuum pumps become<br />

gas-locked at or above −0.098 MPa. But in these familiar<br />

cases, the phase change to vapor (cavitation) is nucleated by<br />

contact with foreign agents that destabilize the inter-molecular<br />

hydrogen-bonding of liquid water (Pickard 1981). Such “heterogeneous<br />

nucleation” of cavitation is typically triggered by<br />

minute <strong>and</strong> ubiquitous gas bubbles in the system. When care<br />

is taken to minimize such heterogeneous nucleation, liquid<br />

water can develop substantially metastable negative ψP values.<br />

<strong>The</strong>oretical calculations, based on equations of state for water,<br />

put the limiting ψP at homogeneous cavitation (where energy<br />

of the water molecules themselves is sufficient to trigger the<br />

phase change) below −200 MPa at 20 ◦ C(Mercury <strong>and</strong> Tardy<br />

2001). Experiments with a variety of systems ranging from<br />

centrifuged capillary tubes to water-filled quartz crystals have<br />

reached values well below −25 MPa, with some as low as −180<br />

MPa <strong>and</strong> approaching the theoretical limit (Briggs 1950; Zheng<br />

et al. 1991). Such values dwarf even the most negative ψP in<br />

plants, which is about −13 MPa (Jacobson et al. 2007); more<br />

typical plant ψP values are less negative than −3 MPa.


<strong>The</strong>re is abundant evidence that cavitation “pressures” in<br />

plants are negative enough for the cohesion-tension mechanism<br />

to operate. Such pressures are determined from a “vulnerability<br />

curve” which usually plots the hydraulic conductivity<br />

(reciprocal of resistivity) of the xylem (often as a percentage<br />

loss from maximum) as a function of the ψP value in the xylem<br />

sap. Curves are generated in several ways, but the centrifuge<br />

method is commonly used because of its rapidity (Alder et al.<br />

1997; Cochard et al. 2005). Stems or roots are spun in a custom<br />

centrifuge rotor that places their xylem under a known tension<br />

at the center of rotation. <strong>The</strong> conductivity is measured either<br />

during or between spinning, <strong>and</strong> the experiment is continued<br />

until the conductivity has dropped to negligible values, thus<br />

indicating complete blockage of flow by cavitation. Typical<br />

species-specific variation is shown in Figure 17A, <strong>and</strong> Figure 17B<br />

compares the ψP value causing complete loss of xylem conductivity<br />

with the minimum ψP values measured in nature for<br />

102 species (Sperry 2000). Clearly, the xylem of some species<br />

cavitates much more readily than others, but these vulnerable<br />

species also do not develop very negative ψP values in nature.<br />

Across the board, the minimum ψP is generally less negative<br />

than the value of ψP at zero conductivity, as is required by the<br />

cohesion-tension mechanism.<br />

Although the centrifuge method is widely accepted for<br />

conifers <strong>and</strong> for short-vesseled angiosperms, there is some<br />

controversy about its ability to measure the vulnerability of longvesseled<br />

taxa where many conduits can exceed the length of<br />

the spinning conductivity segment. Response curves in these<br />

taxa indicate that a significant number of these large vessels<br />

cavitate at very modest negative ψP values. This pattern is<br />

seen in two of the curves shown in Figure 17A (open symbols).<br />

Comparisons with other methods have verified this type of<br />

curve in many (Christman et al. 2012; Jacobson <strong>and</strong> Pratt<br />

2012; Sperry et al. 2012), but not in all cases (Choat et al.<br />

2010; Cochard et al. 2010a), <strong>and</strong> resolving the matter requires<br />

further research.<br />

A major cause of the cavitation in plant xylem is air-seeding<br />

through the inter-conduit pits that normally are responsible<br />

for confining gas embolism (Crombie et al. 1985; Sperry <strong>and</strong><br />

Tyree 1988; Sperry <strong>and</strong> Tyree 1990; Jarbeau et al. 1995).<br />

<strong>The</strong> Pmin of inter-conduit connections is easily measured from<br />

the positively applied gas pressure that just breaches their<br />

seal (Christman et al. 2012). From a variety of techniques<br />

employed across a wide range of species, the Pmin range<br />

of inter-conduit pitting is generally indistinguishable from the<br />

range of ψP values that cause cavitation. In consequence,<br />

vulnerability curves can usually be reproduced using positive<br />

gas pressures rather than placing the xylem sap under tension<br />

(Cochard et al. 1992). <strong>The</strong> prevailing model of cavitation<br />

nucleation is that when the ψP in the transpiration stream<br />

drops to the Pmin of the inter-conduit seal against adjoining<br />

embolized vessels, a gas bubble is pulled into the transpi-<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 333<br />

ration stream which then nucleates cavitation. <strong>The</strong> sap ψP<br />

in the cavitated conduit immediately rises to 0, <strong>and</strong> the sap<br />

is very quickly drained out of the conduit by the surrounding<br />

transpiration stream until the entire conduit becomes filled with<br />

water vapor <strong>and</strong> air, the exact composition of the embolism<br />

depending on diffusive exchange with the surrounding tissue<br />

(Tyree <strong>and</strong> Sperry 1989). According to this model, cavitation<br />

<strong>and</strong> embolism can propagate from conduit-to-conduit via pit<br />

failure.<br />

Considerable attention has been given to how the structure<br />

of inter-conduit pitting relates to its function in cavitation resistance.<br />

In conifer tracheids, where torus aspiration seals the<br />

pits, it has been proposed that air-seeding occurs when the pit<br />

membrane is stretched sufficiently to displace the torus from<br />

the aperture, exposing part of the margo through which air can<br />

readily pass (Figure 16D). Displacement has been observed microscopically,<br />

<strong>and</strong> conifers that are more resistant to cavitation<br />

generally have less flexible pit membranes, <strong>and</strong> can have a<br />

torus that covers the aperture with greater overlap (Sperry <strong>and</strong><br />

Tyree 1990; Domec et al. 2006; Hacke <strong>and</strong> Jansen 2009). An<br />

alternative model is that air seeding occurs by displacement of<br />

the air-water-interface between the torus <strong>and</strong> the pit chamber<br />

wall. In some conifers, air-seeding could also occur through<br />

pores in the torus. In support of these mechanisms is a dependence<br />

of Pmin on the sap surface tension (Cochard et al. 2009;<br />

Holtta et al. 2012; Jansen et al. 2012). However, it is not clear<br />

whether sap solutions that lower the surface tension also alter<br />

pit membrane flexibility <strong>and</strong>, hence, the ease of torus displacement.<br />

It is also unclear whether the torus-margo membrane can<br />

de-aspirate to function a second time if the embolized tracheids<br />

refill.<br />

In angiosperm inter-vessel pits, air seeding probably occurs<br />

by displacement of the air-water meniscus from pores in<br />

their homogenously nano-porous pit membranes (Figure 16D),<br />

but beyond this, the details are surprisingly complex <strong>and</strong><br />

ambiguous. Air seeding at pores is implied by the predicted<br />

sensitivity to surfactants <strong>and</strong> correspondence between pore<br />

dimensions, sized by particle penetration, <strong>and</strong> air-seeding<br />

pressures (Crombie et al. 1985; Sperry <strong>and</strong> Tyree 1988;<br />

Jarbeau et al. 1995). <strong>The</strong> pores may be pre-existing in the<br />

membrane, or perhaps created or enlarged by the partial<br />

or complete aspiration of the membrane that likely precedes<br />

the air-seeding (Thomas 1972). A major role of membrane<br />

strength is indicated by the effects of removing Ca 2+ from<br />

the membrane using various chelators such as oxalic acid.<br />

<strong>The</strong>se treatments do not alter surface tension or hydraulic<br />

conductivity, but they can dramatically increase membrane<br />

flexibility <strong>and</strong> vulnerability to cavitation (Sperry <strong>and</strong> Tyree<br />

1988; Sperry <strong>and</strong> Tyree 1990; Herbette <strong>and</strong> Cochard 2010).<br />

Further support for the importance of membrane mechanics is<br />

the cavitation “fatigue” phenomenon wherein xylem becomes<br />

more vulnerable to cavitation after having been cavitated <strong>and</strong>


334 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Figure 17. Vulnerability of xylem to cavitation by negative pressure.<br />

(A) Vulnerability curves for five species showing the drop in xylem hydraulic conductivity (normalized by stem cross-sectional area) as the<br />

xylem pressure becomes more negative. Curves were generated using the centrifugal force method. Species can differ considerably in their<br />

maximum hydraulic conductivity (x axis intercept) <strong>and</strong> how readily they lose it to cavitation (from Hacke et al. 2006).<br />

(B) Xylem pressure at zero hydraulic conductivity from cavitation (the y intercept of a, above) vs. the minimum pressure observed in nature<br />

for 102 species (from Sperry 2000).<br />

re-filled. <strong>The</strong> implication is that mechanical stress has plastically<br />

deformed the membrane to make it more porous.<br />

Cavitation fatigue is potentially reversible in vivo <strong>and</strong> with<br />

artificial xylem saps, suggesting the stressed <strong>and</strong> air-seeded<br />

pit membrane can be restored back to its normal form (Hacke<br />

et al. 2001b; Stiller <strong>and</strong> Sperry 2002).<br />

<strong>The</strong> possibility that pores are created by mechanical stress<br />

on the pit membrane is also consistent with the typical rarity of


observable pores of predicted air-seeding size in non-stressed<br />

membranes (Choat et al. 2008), as well as the tendency for<br />

pit membranes to be thicker, pit chambers shallower (less<br />

deflection), <strong>and</strong> apertures proportionally smaller in cavitationresistant<br />

species (Choat et al. 2008; Jansen et al. 2009;<br />

Lens et al. 2010). Furthermore, the KCl-induced decrease in<br />

membrane flow resistance does not translate into increased<br />

vulnerability to cavitation (Cochard et al. 2010b). This could<br />

mean there are few to no pores in the non-stressed water conducting<br />

pit membrane, with water penetrating the hydrated gel<br />

phase. Alternatively, this could support the diffuse-double layer<br />

hypothesis for the KCl effect which does not require changes<br />

in pore dimensions (Van Doorn et al. 2011b). <strong>The</strong> apparent<br />

interaction between membrane stress <strong>and</strong> air-seeding would<br />

be largely eliminated for “vestured” pits where, in some plant<br />

lineages, the membrane is supported by outgrowths from the pit<br />

chamber wall, a factor that must lie behind the elusive adaptive<br />

significance of these structures (Jansen et al. 2001; Choat et al.<br />

2004).<br />

An additional complexity in linking cavitation resistance to<br />

pit structure is a seemingly inescapable role for probability. A<br />

single inter-conduit seal consists of hundreds if not thous<strong>and</strong>s<br />

of individual pits. Not all of these pits will be created equal, <strong>and</strong><br />

it only takes one leaky pit to air-seed the cavitation. According<br />

to the “rare pit” hypothesis, the more pits that constitute the<br />

seal, the weaker the seal will be, because by chance the<br />

leakier will be the leakiest pit (Wheeler et al. 2005). This will<br />

be true whether the air-seeding pores are pre-existing, or are<br />

created by mechanical stress. This hypothesis is supported<br />

by the general trend for larger vessels to be more vulnerable<br />

to cavitation (because they would have more pits), <strong>and</strong> by a<br />

correlation between the extent of inter-conduit pit area <strong>and</strong><br />

increasing vulnerability (Hacke et al. 2006; Christman et al.<br />

2009; Christman et al. 2012). It seems safe to conclude that<br />

both pit quantity <strong>and</strong> pit quality interact to set the ψP value at<br />

which air-seeding occurs <strong>and</strong>, hence, the cavitation resistance<br />

(Lens et al. 2010).<br />

Because cavitation appears to spread from conduit to conduit,<br />

the three dimensional connectivity of the conduit network<br />

should have an additional impact on cavitation resistance.<br />

Woody species differ considerably in the extent of overlap<br />

or connectivity between individual conduits (Carlquist 1984).<br />

Modeling studies suggest that greater connectivity would tend<br />

to facilitate the spread of embolism, whereas more limited connectivity<br />

would tend to confine it (Loepfe et al. 2007). However,<br />

limited comparative data do not always support this prediction.<br />

A greater vessel grouping index (higher connectivity) in some<br />

lineages correlates with higher cavitation resistance rather than<br />

enhanced vulnerability (Lens et al. 2010). This trend supports<br />

the earlier notion that the greater redundancy provided by high<br />

connectivity would be advantageous for minimizing the effect<br />

of embolism in xeric habitats (Carlquist 1984).<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 335<br />

<strong>The</strong> evident complexity that links xylem conduit structure to<br />

cavitation resistance comes from the multitude of variables<br />

involved: features at the individual pit level (e.g. membrane<br />

porosity, thickness, Ca 2+ content, mechanical properties) can<br />

be balanced by features at the inter-conduit wall scale (e.g.,<br />

pit number), which in turn can be balanced by features at<br />

the conduit network level (e.g., conduit connectivity). Thus, the<br />

same cavitation resistance is likely to be achieved by multiple<br />

structural combinations.<br />

Freeze-thaw associated cavitation<br />

Cavitation can also be induced by freeze-thaw cycles <strong>and</strong> likely<br />

by a “thaw expansion” nucleating mechanism (Pittermann <strong>and</strong><br />

Sperry 2006) (Figure 18A). Freezing of the xylem sap in nature<br />

usually occurs under conditions where transpiration is minimal.<br />

Thus, xylem blockage by ice formation would normally not<br />

result in ψP becoming more negative. Instead, ψP would likely<br />

become less negative, or even become positive because of<br />

the expansion of ice. However, dissolved gases in the sap are<br />

insoluble in ice, <strong>and</strong> under typical freezing conditions will form<br />

bubbles in the middle of the conduit. On thawing, if these bubbles<br />

do not dissolve fast enough they can nucleate cavitation<br />

if the thawed sap ψP is sufficiently negative. Hence, cavitation<br />

would occur during the thawing rather than the freezing phase,<br />

a prediction supported by experimental observation (Mayr <strong>and</strong><br />

Sperry 2010).<br />

Just as for cavitation by water stress, there is considerable<br />

variation between species in their vulnerability to freeze-thaw<br />

cavitation. It is scarcely detectible in some species regardless<br />

of the negative sap ψP values, <strong>and</strong> others are blocked<br />

completely by a single freeze-thaw event at ψP values close<br />

to zero (Davis et al. 1999) (Figure 18B). However, unlike the<br />

water stress situation, there appears to be a single structural<br />

variable of over-riding importance for determining vulnerability<br />

to freeze-thaw cavitation. That variable is the xylem conduit<br />

lumen diameter: wider conduits are uniformly more susceptible<br />

to cavitation than narrower ones, regardless of whether the<br />

conduit is a conifer tracheid or angiosperm vessel (Figure 18B).<br />

<strong>The</strong> simplest explanation is that wider vessels form larger<br />

bubbles during freezing because of their greater water volume,<br />

<strong>and</strong> large bubbles take longer to re-dissolve <strong>and</strong>, hence, are<br />

more likely to nucleate cavitation, post-thaw. As the thawexpansion<br />

model predicts, a more negative ψP post-thaw, or<br />

a more rapid thawing rate, will also induce more cavitation at<br />

a given conduit diameter (Langan et al. 1997; Pittermann <strong>and</strong><br />

Sperry 2003, 2006). Similarly, the amount of embolism should<br />

decrease with a greater rate of freezing, which reduces gas<br />

bubble size (Sevanto et al. 2012).<br />

In some species, the amount of embolism increases with the<br />

minimum freezing temperature, an observation not necessarily<br />

predicted from the thaw-expansion mechanism (Pockman <strong>and</strong>


336 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Figure 18. Vulnerability of xylem to cavitation by freeze-thaw<br />

events.<br />

(A) <strong>The</strong> “thaw expansion” mechanism for cavitation by freezing<br />

<strong>and</strong> thawing. Freezing of a sap-filled functional vessel creates<br />

gas bubbles in the ice-filled frozen conduit. If bubbles persist long<br />

enough after thawing, <strong>and</strong> negative pressures are low enough, they<br />

will trigger cavitation <strong>and</strong> result in an embolized conduit (from Sperry<br />

1993).<br />

(B) Loss of hydraulic conductivity caused by a single freeze-thaw<br />

cycle at -0.5 MPa versus the average conduit lumen diameter. Data<br />

are species averages for angiosperm vessels (black) <strong>and</strong> conifer<br />

tracheids (blue) (from Pittermann <strong>and</strong> Sperry 2003).<br />

Sperry 1997; Ball et al. 2006; Pittermann <strong>and</strong> Sperry 2006). It is<br />

possible that the lower ice temperature <strong>and</strong> consequent tissue<br />

dehydration creates, locally, a more negative sap ψP, postthaw<br />

(Ball et al. 2006), or perhaps causes tissue damage that<br />

nucleates cavitation, post-thaw. Acoustic emissions are often<br />

detected during the freezing phase, which could indicate that at<br />

least some cavitation occurs prior to the thawing of tissue (Mayr<br />

<strong>and</strong> Zublasing 2010). However, experiments indicate no loss of<br />

conductivity when stems frozen under negative ψP conditions<br />

are thawed at atmospheric pressure; the conductivity only<br />

drops when the thaw occurs under negative ψP values (Mayr<br />

<strong>and</strong> Sperry 2010). It is possible that other phenomena besides<br />

conduit cavitation are causing freezing-associated acoustic<br />

emissions. Importantly, not all embolism events during winter<br />

are necessarily caused by freeze-thaw cycles. Sublimation <strong>and</strong><br />

cavitation by water stress in thawed <strong>and</strong> transpiring crowns<br />

with frozen boles or soil represent other potential causes<br />

(Peguero-Pina et al. 2011).<br />

Negative xylem sap ψ P <strong>and</strong> conduit collapse<br />

<strong>The</strong> cohesion-tension mechanism requires conduit walls that<br />

are sufficiently rigid to withst<strong>and</strong> collapse by the required<br />

negative ψP values. Hence, the evolution of secondary walls<br />

<strong>and</strong> lignification necessarily paralleled the evolution of xylem<br />

tissues. While many factors contribute to the strength of conduit<br />

walls to prevent implosion, a dominant variable is the ratio<br />

of wall thickness to conduit lumen radius. This ratio tends to<br />

increase with cavitation resistance, as expected from concomitantly<br />

more negative sap ψP values. A higher thickness-to-span<br />

ratio also increases wood density, consistent with the tendency<br />

for greater wood density in more cavitation-resistant trees that<br />

generally experience more negative ψP values (Hacke et al.<br />

2001a; Domec et al. 2009).<br />

Estimates of wall strength give an average safety factor<br />

from implosion of 1.9 in woody angiosperm vessels <strong>and</strong> 6.8<br />

in conifer stem tracheids (Hacke et al. 2001a). <strong>The</strong> lower value<br />

for vessels presumably reflects their minimal role in mechanical<br />

support of the tree, a function performed by wood fibers.<br />

However, conifer tracheids must be additionally reinforced<br />

because they not only have to hold up against negative sap<br />

ψP, but they also support the tree itself. Interestingly, not all<br />

conduits avoid implosion, as it has been observed in the axial<br />

tracheids of pine needles (Cochard et al. 2004), transfusion<br />

tracheids of podocarps (Brodribb <strong>and</strong> Holbrook 2005), <strong>and</strong><br />

metaxylem vessels in maize (Kaufman et al. 2009). In each<br />

case, the collapse was apparently reversible. Not unexpectedly,<br />

implosion is also observed in conduits of lignin-deficient<br />

mutants (Piquemal et al. 1998).<br />

Trade-offs between efficiency <strong>and</strong> safety<br />

<strong>The</strong> cohesion-tension mechanism, <strong>and</strong> its limitation by cavitation<br />

<strong>and</strong> conduit collapse, suggest potential trade-offs in<br />

the xylem conduit structure for minimizing flow resistance on<br />

the one h<strong>and</strong> (efficiency), <strong>and</strong> sustaining greater negative ψP<br />

without cavitation or conduit collapse on the other h<strong>and</strong> (safety).<br />

With respect to greater resistance to collapse, large increases


in the thickness-to-span ratio would be more readily achieved<br />

by narrowing the lumen, because walls arguably have a limited<br />

maximum thickness (Pittermann et al. 2006b). Tradeoffs arise<br />

because narrower lumens would have exponentially greater<br />

flow resistance, <strong>and</strong> higher thickness-to-span increases construction<br />

costs. Greater resistance to cavitation by freeze-thaw<br />

cycles also comes at the expense of narrower lumens, <strong>and</strong><br />

consequently, higher flow resistance per conduit (Davis et al.<br />

1999). Resistance to freeze-thaw embolism can, in turn, tradeoff<br />

with photosynthetic capacity in evergreen species (Choat<br />

et al. 2011).<br />

Enhanced resistance to cavitation by water stress also tends<br />

to be associated with higher flow resistances, as seen in the<br />

collection of vulnerability curves shown in Figure 17A. <strong>The</strong><br />

more vulnerable species tend to have higher initial hydraulic<br />

conductivities than the more cavitation-resistant ones. <strong>The</strong><br />

basis for this association is not straightforward because so<br />

many variables interact to determine a species’ vulnerability<br />

curve. While intuitively one might expect that pits which are<br />

more resistant to air-seeding would also have a greater flow<br />

resistance (Holtta et al. 2011), the data suggest otherwise:<br />

membrane flow resistance is uncoupled from membrane airseeding<br />

resistance. Estimates of pit flow resistances across<br />

several angiosperm <strong>and</strong> conifer species showed no relationship<br />

with cavitation resistance (Pittermann et al. 2005; Hacke et al.<br />

2006). Furthermore, the drop in flow resistance via the ionic<br />

effect on pit membranes had no effect on cavitation resistance<br />

(Cochard et al. 2010b). Instead, the efficiency-safety trade-off<br />

may arise at the whole-conduit <strong>and</strong> conduit-network scale. If the<br />

rare pit hypothesis were correct, greater cavitation resistance<br />

would require fewer pits per conduit, which would generally<br />

correspond to narrower <strong>and</strong> shorter conduits of consequently<br />

greater flow resistance (Wheeler et al. 2005). And if reducing<br />

the connectivity of a conduit (the number of conduits it contacts)<br />

limits the spread of embolism, as expected, the lower<br />

connectivity may translate to lower conductivity at the network<br />

scale (Loepfe et al. 2007).<br />

<strong>The</strong> flow resistance penalty of narrower <strong>and</strong> shorter conduits<br />

can be compensated by increasing their number per wood<br />

area (Hacke et al. 2006). This “packing” strategy is exemplified<br />

by conifers which devote over 95% of their wood volume<br />

to tracheids <strong>and</strong>, thus, achieve similar whole stem hydraulic<br />

conductivity as angiosperms whose generally wider <strong>and</strong> longer<br />

vessels occupy only a small fraction of wood space, most of<br />

which is devoted to structural fibers <strong>and</strong> storage parenchyma.<br />

<strong>The</strong> packing strategy exemplifies how trade-offs at one level<br />

of structure can be compensated for at another scale, vastly<br />

complicating the adaptive interpretation of wood structure <strong>and</strong><br />

function.<br />

Trade-offs of one sort or another presumably underlie the<br />

observed correlation between cavitation resistance <strong>and</strong> the<br />

range of native sap ψP values (Figure 17B). Accordingly, mesic-<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 337<br />

adapted species that experience less negative xylem ψP are<br />

vulnerable to cavitation because being overly resistant would<br />

cost them in terms of greater flow resistance <strong>and</strong> vascular<br />

construction costs. Conversely, xeric-adapted species that<br />

periodically endure more negative ψP values must be more<br />

resistant to cavitation, <strong>and</strong> their consequently greater flow<br />

resistance <strong>and</strong> construction costs competitively exclude them<br />

from more mesic habitats. <strong>The</strong> result of this adaptive scenario<br />

is that species tend to be only as resistant to cavitation as they<br />

have to be for the native ψP range they experience over their<br />

lifespan (Maherali et al. 2003; Markesteijn et al. 2011).<br />

<strong>The</strong> limiting process of cavitation naturally constrains the<br />

xylem ψP over which productivity can be sustained. Indeed,<br />

an important adaptive advantage of stomatal regulation of ψP<br />

is to keep it from reaching such negative values that would<br />

induce excessive cavitation (Tyree <strong>and</strong> Sperry 1988). “Runaway<br />

cavitation,” which is the loss of all hydraulic conductivity<br />

caused by unregulated ψP, can be induced experimentally, <strong>and</strong><br />

it is a dramatic <strong>and</strong> quick cause of mortality (Holtta et al.<br />

2012). Not surprisingly, plants have evolved the necessary<br />

cavitation resistance <strong>and</strong> stomatal control mechanisms to avoid<br />

such an efficient suicidal scenario. But stomatal control cannot<br />

directly prevent the gradual accumulation of cavitation as the<br />

xylem ψP becomes more negative due to limited soil water<br />

availability. In consequence, extreme stress events or climatic<br />

shifts push plants towards excessive cavitation, resulting in<br />

partial or complete dieback from chronic reductions (70% or<br />

greater) in hydraulic conductivity from cavitation (Anderegg<br />

et al. 2012; Plaut et al. 2012). Soil-plant-atmosphere models<br />

that incorporate cavitation resistance can be successful in<br />

predicting responses of vegetation to drought (Sperry et al.<br />

2002), allowing effects of climate change on plant water <strong>and</strong><br />

carbon flux to be anticipated.<br />

Refilling of embolized conduits<br />

<strong>The</strong> refilling of embolized xylem conduits has been documented<br />

in numerous studies. Embolisms accumulating over the winter<br />

from freeze-thaw cycles <strong>and</strong> other causes can be reversed in<br />

the spring in many species (Sperry 1993; Hacke <strong>and</strong> Sauter<br />

1996). Diurnal embolism <strong>and</strong> refilling cycles have also been<br />

documented (Stiller et al. 2005; Yang et al. 2012), as well as<br />

refilling after relief from prolonged drought (West et al. 2008).<br />

Two kinds of refilling have been observed. In “bulk” refilling,<br />

the sap ψP of the entire bulk xylem stream rises close to or<br />

above zero to force sap back into the embolized conduits. For<br />

this to happen, at a minimum, transpiration must be negligible<br />

<strong>and</strong> the soil ψw must be close to zero. Xylem ψP would thus<br />

decrease only to what is necessary to counter-act gravity,<br />

dropping by approximately 0.01 MPa per meter height. Less<br />

negative ψP values, even positive values, could develop from<br />

foliar uptake of rain or dew, <strong>and</strong> especially from osmotically


338 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

generated root pressures. A few species, Acer in particular,<br />

also develop positive stem ψP values in response to freezethaw<br />

cycles in early spring (Tyree <strong>and</strong> Zimmermann 2002).<br />

Root pressures can exceed 0.5 MPa <strong>and</strong> are strongly associated<br />

with bulk refilling in a variety of woody <strong>and</strong> herbaceous<br />

plants (Sperry 1993; Stiller et al. 2005; Yang et al. 2012).<br />

Experimental suppression of root pressure has been shown to<br />

block refilling in some species (Sperry 1993). <strong>The</strong> natural failure<br />

of root pressure <strong>and</strong> spring refilling, owing to freezing-related<br />

mortality of shallow roots, has been linked to birch dieback<br />

episodes (Cox <strong>and</strong> Malcolm 1997). Diminishing root pressure<br />

with plant height has also been invoked as a limit to the stature<br />

of refilling bamboos (Yang et al. 2012).<br />

In a second type of “novel refilling,” the bulk xylem ψP is much<br />

too negative to allow sap to move back into the embolized conduits<br />

(Salleo et al. 1996). <strong>The</strong>re must be a pumping mechanism<br />

that brings sap into the embolized conduit <strong>and</strong> keeps it there<br />

until the gas is dissolved or escapes. <strong>The</strong> pumping mechanism<br />

is unknown, but several hypotheses have been proposed (see<br />

Nardini et al. 2011a for a recent review). Two basic driving<br />

forces are suggested: forward osmosis associated with solute<br />

accumulation in the thin water film along the embolized conduit<br />

wall, or reverse osmosis driven by either tissue pressure, or<br />

perhaps more likely, Münch pressure flow redirected from the<br />

phloem to the embolism, via ray tissue. In the latter case,<br />

refilling becomes a special case of phloem unloading. <strong>The</strong> data<br />

suggest the mechanism is: (a) triggered by the presence of a<br />

gas-filled conduit (rather than a particular plant water potential<br />

(Salleo et al. 1996)), (b) associated with starch hydrolysis<br />

(Bucci et al. 2003; Salleo et al. 2009), (c) upregulation of<br />

certain aquaporins (Secchi <strong>and</strong> Zwieniecki 2010), <strong>and</strong> (d) active<br />

phloem transport in the vicinity of the embolism (Salleo et al.<br />

2006).<br />

Xylem conduit wall sculpturing <strong>and</strong> chemistry may also be<br />

important (Kohonen <strong>and</strong> Hell<strong>and</strong> 2009) with wettable areas<br />

assisting water uptake <strong>and</strong> gas dissolution, <strong>and</strong> hydrophobic<br />

areas perhaps allowing gas escape through minute wall pores<br />

(Zwieniecki <strong>and</strong> Holbrook 2009). Two very different roles have<br />

been proposed for inter-conduit pits in the refilling process.<br />

In one model, air pockets in the pit chamber <strong>and</strong> “wicking”<br />

forces at the aperture serve to isolate the pressurized sap in<br />

the embolized vessel from the negative ψP in the adjacent transpiration<br />

stream (Zwieniecki <strong>and</strong> Holbrook 2000). Alternatively,<br />

it has been proposed that pit membranes can act as osmotic<br />

membranes, with sap being pulled from the transpiration stream<br />

into the refilling conduit by an osmotic gradient, analogous<br />

to the generation of positive turgor pressures in protoplasts<br />

(Hacke <strong>and</strong> Sperry 2003).<br />

A particularly informative study is the imaging work of Brodersen<br />

et al. (2010). Embolized vessels in grapevine were observed<br />

to refill while the ψP of the surrounding xylem was more<br />

negative than −0.7 MPa, confirming the need for a pumping<br />

process. Water entered empty vessels from the direction of<br />

the rays, a pattern consistent with phloem-directed water influx<br />

rather than either pit membrane osmosis from the transpiration<br />

stream or root pressure. <strong>The</strong>re was no obvious mechanism<br />

to prevent drainage of the accumulating water back into the<br />

surrounding sap stream, contradicting a role of inter-vessel<br />

pits in hydraulic isolation. Whether the vessel refilled or stayed<br />

partially embolized depended on the difference between the<br />

rate of water influx from the rays, versus drainage to the<br />

surrounding transpiration stream. <strong>The</strong>re was considerable variation<br />

in the onset, rate, <strong>and</strong> eventual success or failure, in the<br />

plants imaged. Unfortunately, the ψπ of the refilling sap was not<br />

determined, so forward- versus reverse-osmosis mechanisms<br />

could not be distinguished. Nevertheless, the results lend<br />

strong support to a phloem-coupled refilling mechanism that<br />

refills by pumping water into the embolized vessels faster than<br />

it is withdrawn.<br />

Engineering xylem properties: A path to increased<br />

plant productivity<br />

<strong>The</strong> cohesion-tension mechanism constrains the productivity<br />

<strong>and</strong> survival of plants, arguably constituting the “functional<br />

backbone of terrestrial plant productivity” (Brodribb 2009). Because<br />

of the stomatal regulation of canopy xylem ψP, frictional<br />

resistance to water flow through the plant is coupled to the<br />

maximum potential photosynthetic rate <strong>and</strong>, hence, to productivity<br />

in general. <strong>The</strong> coupling in turn is necessary for avoiding<br />

hydraulic failure by cavitation, which limits plant survival<br />

in extremis. <strong>The</strong> cavitation limit presumably evolved in response<br />

to complex trade-offs with frictional resistance, with<br />

competition selecting for minimal flow resistance at the expense<br />

of excessive cavitation safety margins. Although the driving<br />

force for the transpiration stream is passive, flow resistance<br />

(via the ionic effect) <strong>and</strong> conduit refilling is modulated by<br />

active metabolic processes. Probably the single most important<br />

structures in the pipeline are the inter-conduit pits: their distribution,<br />

chemistry, structure, <strong>and</strong> mechanical properties greatly<br />

influence both frictional resistance to flow <strong>and</strong> vulnerability to<br />

cavitation by water stress.<br />

<strong>The</strong> tools of molecular biology have the potential to greatly<br />

advance our knowledge of the flow resistance, cavitation, <strong>and</strong><br />

refilling phenotypes, as well as the nature of trade-offs among<br />

them. As the genetic <strong>and</strong> developmental controls of xylem<br />

anatomical traits become better understood (Demari-Weissler<br />

et al. 2009), they can be manipulated to untangle structurefunction<br />

relationships that can otherwise only be inferred from<br />

comparative studies. Crucial to advancement in this area<br />

are model organisms in which the hydraulic physiology can<br />

be phenotyped <strong>and</strong> manipulated. Among woody plants, the<br />

Populus system is perhaps most promising, <strong>and</strong> much<br />

has already been learned from it (Secchi <strong>and</strong> Zwieniecki


2010; Schreiber et al. 2011). <strong>The</strong> next decade should<br />

bring rapid progress as molecular biology continues to<br />

merge with comparative <strong>and</strong> evolutionary whole plant<br />

physiology.<br />

Long-distance Signaling Through<br />

the Phloem<br />

Over the past several decades, considerable attention has<br />

been paid to unraveling the mechanics of phloem loading.<br />

Genetic <strong>and</strong> molecular studies have identified the major players<br />

that mediate in the loading of sugars, predominantly sucrose,<br />

into the CC-SE complex. Interestingly, in terms of the apoplasmic<br />

loaders, the recent identification of the permease that<br />

controls release of sucrose from the phloem parenchyma cells<br />

into the CC apoplasm (Figure 13B, I) served to complete the<br />

molecular characterization of this important pathway (Chen<br />

et al. 2012b). Based on such studies <strong>and</strong> extensive physiological<br />

experiments, the nature of the photosynthates (sugars<br />

<strong>and</strong> amino acids) loaded into the phloem translocation stream<br />

is well established.<br />

<strong>The</strong> phloem has also been shown to carry additional cargo,<br />

including the phytohormones auxin, gibberellins, cytokines <strong>and</strong><br />

abscisic acid (Hoad 1995), signaling agents involved in plant<br />

defense (discussed later in the review), as well as certain<br />

proteins <strong>and</strong> various forms of RNA (Lough <strong>and</strong> Lucas 2006;<br />

Buhtz et al. 2010). That specific proteins are present in the<br />

phloem has been recognized for some time (Fisher et al.<br />

1992; Bostwick et al. 1992), <strong>and</strong> furthermore, some such<br />

proteins have been shown to move within the translocation<br />

stream (Golecki et al. 1998, 1999; Xoconostle-Cázares et al.<br />

1999).<br />

Phloem proteins: Potential roles in enucleate SE<br />

maintenance <strong>and</strong> long-distance signaling<br />

Phloem exudate can be collected from a number of plant<br />

species, <strong>and</strong> this feature has been used to develop proteomic<br />

databases for these species (Barnes et al. 2004; Giavalisco<br />

et al. 2006; Lin et al. 2009; Rodriguez-Medina et al. 2011).<br />

This collection process requires that an incision be made<br />

into the petiole or stem in order to allow the phloem to<br />

“bleed.” Thus, due care is required to minimize the level<br />

of protein contamination from surrounding (CCs <strong>and</strong> phloem<br />

parenchyma) tissues. As excision results in an abrupt pressure<br />

drop between the sieve tube system <strong>and</strong> the surrounding cells,<br />

it is generally appreciated that some level of contamination<br />

is unavoidable (Atkins et al. 2011). Here, use of molecular<br />

markers such as Rubisco (Doering-Saad et al. 2006; Giavalisco<br />

et al. 2006; Lin et al. 2009), can help in assessing the extent<br />

to which contamination may have occurred. Generally,<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 339<br />

contamination does not appear to be an important issue,<br />

especially for the prominent proteins, but proteins present in<br />

very low abundance need to be viewed with a degree of<br />

caution.<br />

Other methods, including cutting aphid stylets (Aki et al.<br />

2008; Gaupels et al. 2008a), EDTA-induced phloem exudation<br />

(Gaupels et al. 2008b; Batailler et al. 2012) <strong>and</strong> laser microdissection<br />

of phloem tissues (Deeken et al. 2008) have also<br />

been employed to develop phloem databases. Collectively,<br />

these studies have established phloem proteome databases<br />

containing more than 1,000 proteins, with activities encompassing<br />

a very broad range of activities, including enzymes involved<br />

in metabolic networks, amino acid synthesis, protein turnover,<br />

RNA binding, transcriptional regulation, stress responses, defense,<br />

<strong>and</strong> more.<br />

<strong>The</strong> next step will be to partition these proteins into those<br />

involved in local maintenance of the functional enucleate sieve<br />

tube system <strong>and</strong> long-distance signaling. For these studies,<br />

a combination of hetero-grafting experiments conducted between<br />

species from different genera or families, <strong>and</strong> advanced<br />

mass spectroscopy methods, will prove most useful. <strong>The</strong> cucurbits,<br />

such as pumpkin, cucumber, melon <strong>and</strong> watermelon,<br />

from which analytical quantities of phloem exudate can generally<br />

be collected, may prove ideal for this purpose. <strong>The</strong><br />

recent completion of annotated genomes for three of these<br />

cucurbits (Huang et al. 2009; Garcia-Mas et al. 2012; Guo<br />

et al. 2012) adds to the utility of these species for such critical<br />

experiments.<br />

<strong>The</strong> complexity of the phloem proteome raises the question<br />

as to the stability of these proteins <strong>and</strong> the mechanism by<br />

which they might be turned over within the sieve tube system.<br />

<strong>The</strong> large population of proteinase inhibitors probably<br />

prevents turnover by simple proteolysis (Dinant <strong>and</strong> Lucas<br />

2012). However, identification in the phloem sap of ubiquitin<br />

<strong>and</strong> numerous enzymes involved in protein ubiquitination <strong>and</strong><br />

turnover, including all the components of the 26S proteasome<br />

(Figure 19), indicates that enucleate SEs likely have retained the<br />

ability to engage in protein sorting <strong>and</strong> turnover (Lin et al. 2009).<br />

Thus, once they have performed their function(s), phloem<br />

proteins can be degraded either through export into neighboring<br />

CCs, or in loco via the ubiquitin-26S proteasome pathway.<br />

<strong>The</strong> mature, enucleate sieve tube system also has been<br />

shown to contain all the enzymes <strong>and</strong> associated activities<br />

required for a complete antioxidant defense system (Walz et al.<br />

2002; Lin et al. 2009; Batailler et al. 2012). Interestingly, these<br />

enzyme activities appear to increase in response to imposed<br />

drought stress (Walz et al. 2002). This complement of enzymes<br />

would appear to function, locally, to afford protection against<br />

oxidative stresses, thereby preventing damage to essential<br />

components of the SEs. Such local maintenance functions will<br />

likely be performed by a specific subset of the proteins detected<br />

in phloem exudates.


340 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Figure 19. Pumpkin phloem sap contains the machinery for ubiquitin-mediated proteolysis.<br />

(A) Schematic representation of the 26S proteasome indicating that all components except for Rpn4 were identified from the pumpkin phloem<br />

sap. Orange boxes indicate components identified by Lin et al. (2009).<br />

(B) Identification of phloem proteins associated with ubiquitin-mediated proteolysis. Note that green boxes represent proteins present in the<br />

Arabidopsis genome, <strong>and</strong> red lettering indicates identification in pumpkin phloem proteins. White boxes represent Saccharomyces cerevisiaespecific<br />

proteins. Ub, ubiquitin; CHIP, carboxyl terminus of Hsc 70-interacting protein; APC/C, anaphase promoting complex/cyclosome;<br />

DCAF, DD B1-CUL4-associated factor; SCF, Skp1-Cul1-F-box protein; ECS, Elongin C-Cul2-SOCS box; ECV, SCF-like E3 ubiquitin ligase<br />

complex (from Lin et al. 2009).


FLOWERING LOCUS T (FT) as the phloem-mobile<br />

florigenic signal<br />

It has long been known that, for plants whose flowering is<br />

controlled by day length, the phloem is involved in the transmission<br />

of a photoperiod-induced signal that moves from the<br />

mature/source leaves to the shoot apex where it induces the<br />

onset of flowering (Zeevaart 1976). <strong>The</strong> nature of this grafttransmissible<br />

signal, termed florigen (Zeevaart 2006), was<br />

recently identified as FT, a member of the CETS protein<br />

family (consisting of CENTRORADIALIS [CEN], TERMINAL<br />

FLOWER 1 [TFL1] <strong>and</strong> FT). FT expression is confined to CCs<br />

in source leaves, <strong>and</strong> this small protein enters the sieve tube<br />

system by passage through the CC-SE PD.<br />

Direct evidence for the presence of FT in the phloem translocation<br />

stream was provided by studies performed on a pumpkin<br />

(Cucurbita moschata) accession in which flowering occurs<br />

only under short-day (SD) conditions (Lin et al. 2007). Mass<br />

spectroscopy studies conducted on phloem sap collected from<br />

plants grown under long-day (LD) <strong>and</strong> SD conditions provided<br />

unequivocal evidence that the C. moschata FT orthologue<br />

was present only in exudate collected from SD-grown plants<br />

in which flowering was induced. Supporting evidence for FT<br />

as a component of the long-distance florigenic signal was<br />

provided by studies on Arabidopsis (Corbesier et al. 2007) <strong>and</strong><br />

rice (Tamaki et al. 2007). Here, FT <strong>and</strong> Hd3a, the rice FT<br />

orthologue, were expressed as GFP fusions driven by a CCspecific<br />

promoter. Detection of a GFP signal in the meristem<br />

of these transgenic plants was consistent with movement from<br />

the phloem into the meristem where floral induction was taking<br />

place.<br />

An absolute quantitation peptide approach was used to<br />

determine the level of FT in the pumpkin phloem translocation<br />

stream. Recorded values were in the low femtomolar range (Lin<br />

et al. 2007), clearly placing FT in a protein hormone category<br />

(Shalit et al. 2009). Here, it is noteworthy that FT peptides<br />

could not be detected in these pumpkin phloem exudates<br />

when analyzed using the proteomics approach reported by Lin<br />

et al. (2009). This is important, as it indicates that not all low<br />

abundance proteins detected in phloem exudates represent<br />

contaminants from surrounding tissues.<br />

As a number of examples exist in which both mRNA <strong>and</strong><br />

the encoded protein have been detected in phloem exudates<br />

(Lough <strong>and</strong> Lucas 2006), Lin et al. (2007) carried out extensive<br />

tests for the presence of the pumpkin FT transcripts using the<br />

same phloem exudates in which FT peptides were identified.<br />

As FT transcripts could not be amplified, it appeared that,<br />

in these plants, FT protein, but not its mRNA, serves as<br />

the florigenic signal. This finding was consistent with earlier<br />

studies conducted on the tomato FT orthologue, SINGLE-<br />

FLOWER TRUSS (SFT). SFT-dependent graft-transmissible<br />

signals were found to induce flowering in day length neutral<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 341<br />

tomato <strong>and</strong> tobacco plants; however, no evidence was obtained<br />

for the presence of SFT transcripts within the scion meristem<br />

tissues (Lifschitz et al. 2006). Extensive mutagenesis of the<br />

FT gene, in which sequence <strong>and</strong> structural modifications<br />

were made to the mRNA whilst leaving the FT protein unaltered,<br />

had little or no effect on long-distance floral induction<br />

(Notaguchi et al. 2008). Again, these findings are consistent<br />

with FT protein, not its mRNA serving as the phloem-mobile<br />

signal.<br />

Experimental support for FT mRNA as a component of<br />

the florigenic signaling mechanism has been suggested from<br />

studies based on movement-defective viral expression systems<br />

(Li et al. 2009). Here, the first 100 nucleotides of the FT RNA<br />

sequence were shown to function in cis to allow systemic<br />

movement of heterologous viral sequences. Similar findings<br />

with Arabidopsis have been reported in terms of FT sequences<br />

acting in cis to mediate in the long-distance transport of<br />

otherwise cell-autonomous transcripts (Lu et al. 2012). Further<br />

support for the role of endogenous FT mRNA, as a component<br />

of the florigenic signal, was claimed from studies with transgenic<br />

Arabidopsis lines in which flowering was delayed through<br />

expression, in the apex, of RNAi/artificial miRNA-FT (Lu et al.<br />

2012). Unfortunately, these results are contradictory to findings<br />

from a similar FT silencing experiment in which expression of<br />

amiRNA-FT in the phloem caused a significant delay in floral<br />

induction, whereas its expression in the apex failed to delay<br />

flowering (Mathieu et al. 2007). Although the controversy over<br />

whether or not FT mRNA contributes to floral induction remains<br />

to be resolved, there is no a priori reason why, for any gene, its<br />

mRNA <strong>and</strong> protein cannot both serve as signaling agents via<br />

the phloem.<br />

ATC as a phloem-mobile anti-florigenic signal<br />

<strong>System</strong>ic floral inhibitors or anti-florigens have been proposed<br />

to participate in down regulating floral induction under noninducing<br />

conditions (Zeevaart 2006). Evidence in support of<br />

this concept was recently provided by studies performed on<br />

Arabidopsis plants carrying mutations in ATC, aCEN/TFL1<br />

homologue. Flowering in Arabidopsis is promoted under LD<strong>and</strong><br />

inhibited under SD-conditions. Expression of ATC is enhanced<br />

during short days <strong>and</strong>, based on the effect of an atc-2<br />

mutant, the WT gene appears to contribute to the suppression<br />

of flowering (Huang et al. 2012).<br />

At the tissue level, ATC was found to be expressed in the<br />

vascular system, <strong>and</strong> the phloem in particular, but not in the<br />

apex. This suggested a non-cell-autonomous function for either<br />

the ATC transcripts or protein. A range of grafting experiments<br />

were performed with Arabidopsis stock <strong>and</strong> scions connected<br />

above the hypocotyl region (with the stock containing several<br />

mature source leaves). Analysis of RNA extracted from atc-2


342 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

scions grafted onto WT stocks provided clear evidence for<br />

the movement of ATC transcripts across the graft union, <strong>and</strong><br />

compared to WT:WT grafts, flowering time was delayed (Huang<br />

et al. 2012). Parallel experiments were performed to address<br />

whether ATC protein moves across the graft union. In these<br />

Western blot experiments, a clear ATC signal was detected in<br />

atc-2 scions grafted onto WT stocks. <strong>The</strong>se findings support the<br />

possibility that, in Arabidopsis, both ATC transcripts <strong>and</strong> protein<br />

are phloem-mobile; i.e., together, they may enter the shoot<br />

apex to compete with FT for FD, thereby inhibiting the transition<br />

to flowering. However, it is also possible that the phloemmobile<br />

ATC transcripts enter CCs located in the atc-2 scion<br />

tissues where they then produce ATC protein. Irrespective of<br />

this potential complication, identification of ATC as a negative<br />

regulator of flowering time in Arabidopsis constitutes an important<br />

step forward in underst<strong>and</strong>ing the role of the phloem in the<br />

overall regulation of plant growth <strong>and</strong> development.<br />

Phloem-mediated long-distance lipid-based signaling?<br />

Lipids <strong>and</strong> lipid-binding proteins have been detected in phloem<br />

exudates collected from a number of plant species. Some 14<br />

putative lipid-binding proteins were detected in Arabidopsis<br />

phloem exudates collected from excised petioles that were<br />

incubated in EDTA to facilitate the bleeding process (Guelette<br />

et al. 2012). Bioinformatics analysis of these proteins indicated<br />

potential roles in membrane synthesis <strong>and</strong>/or turnover, prevention<br />

of lipid aggregation, participation in synthesis of the<br />

glycosyphosphatidylinositol (GPI) anchor, <strong>and</strong> biotic <strong>and</strong> abiotic<br />

stress. A range of lipids have also been reported in phloem<br />

exudates, including simple lipids to complex glycolipids <strong>and</strong><br />

phytosterols such as cholesterol (Behmer et al. 2011; Guelette<br />

et al. 2012).<br />

An interesting study recently conducted on an Arabidopsis<br />

small (20 kDa) phloem lipid-associated family protein (PLAFP)<br />

revealed that it displayed specific bind properties for phosphatidic<br />

acid (PA) (Benning et al. 2012). As both PA <strong>and</strong><br />

PLAFP were detected in Arabidopsis exudate, these results<br />

suggest that PA may well be either trafficked into or translocated<br />

through the sieve tube system by PLAFP. In any event,<br />

detection of lipids <strong>and</strong> lipid-binding proteins within phloem<br />

exudates certainly raises the question as to whether they function<br />

in membrane maintenance <strong>and</strong>/or long-distance signaling<br />

events.<br />

Messenger RNA: A smart way to send a “message”!<br />

A number of recent studies have identified specific mRNA<br />

populations within the phloem sap of various plant species<br />

(Sasaki et al. 1998; Doering-Saad et al. 2006; Lough <strong>and</strong><br />

Lucas 2006; Omid et al. 2007; Deeken et al. 2008; Gaupels<br />

et al. 2008a; Rodriguez-Medina et al. 2011; Guo et al. 2012).<br />

<strong>The</strong>se databases indicate that the phloem translocation stream<br />

of the angiosperms likely contains in excess of 1,000 mRNA<br />

species that encode for proteins involved in a very wide range<br />

of processes. While many of these transcripts are held in<br />

common between plant species, specific differences have been<br />

reported. For example, a comprehensive analysis carried out<br />

using the phloem transcriptomes prepared from cucumber<br />

(1,012 transcripts) <strong>and</strong> watermelon (1,519 transcripts) phloem<br />

exudate indicated that 55% were held in common (Guo et al.<br />

2012). In contrast, the vascular transcriptomes (13,775 <strong>and</strong><br />

14,242 mRNA species in watermelon <strong>and</strong> cucumber, respectively)<br />

were 97% identical. Thus, differences in phloem transcripts<br />

most likely reflect unique functions specific to these<br />

species.<br />

A comparative analysis of the vascular <strong>and</strong> phloem transcriptomes<br />

for cucumber <strong>and</strong> watermelon identified populations of<br />

transcripts that are highly enriched in phloem exudates over<br />

the level detected in excised vascular bundles. <strong>The</strong> numbers<br />

given above represent the transcripts that were present at<br />

≥2-fold higher than the level detected in vascular bundles.<br />

Concerning cucumber, more than 30% of the phloem transcripts<br />

were enriched >10-fold above the level in the vascular<br />

bundles. Importantly, some transcripts were enriched above<br />

500-fold, with another 210 displaying >20-fold enrichment. A<br />

similar situation was observed for watermelon, with some 120<br />

transcripts displaying >10-fold enrichment <strong>and</strong> 320 having 5fold<br />

or greater enrichment in the phloem sap. <strong>The</strong>se data indicate<br />

that, following transcription in the CCs, many transcripts<br />

must undergo sequestration in the sieve tube system through<br />

trafficking mediated by the CC-SE PD.<br />

To date, only a limited number of these phloem mRNAs have<br />

been characterized in terms of whether they act locally or traffic<br />

long-distance to specific target sites. Excellent examples where<br />

translocation through the phloem has been established include<br />

NACP (Ruiz-Medrano et al. 1999), PP16 (Xoconostle-Cázares<br />

et al. 1999), the PFP-LeT6 fusion gene (Kim et al. 2001), GAIP<br />

(Haywood et al. 2005), BEL5 (Benerjee et al. 2006; Hannapel<br />

2010), POTH1 (a KNOTTED1-Like transcription factor) (Mahajan<br />

et al. 2012) <strong>and</strong> Aux/IAA18 <strong>and</strong> Aux/IAA28 (Notaguchi et al.<br />

2012). <strong>The</strong> stability of these phloem-mobile transcripts is made<br />

possible by the fact that phloem exudates have been shown<br />

to lack RNase activity (Xoconostle-Cázares et al. 1999), <strong>and</strong><br />

thus, by extension, the phloem translocation stream is likely<br />

also devoid of this activity.<br />

Phloem delivery of GAIP transcripts modifies<br />

development in tomato sink organs<br />

<strong>The</strong> pumpkin phloem sap was found to contain transcripts for<br />

two members of the DELLA subfamily of GRAS transcription<br />

factors, CmGAIP <strong>and</strong> CmGAIPB, known to function in<br />

the GA signaling pathway (Ruiz-Medrano et al. 1999). <strong>The</strong>


function of CmGAIP <strong>and</strong> GAI from Arabidopsis was investigated<br />

using transgenic Arabidopsis <strong>and</strong> tomato lines expressing<br />

engineered dominant gain-of-function Cmgaip <strong>and</strong><br />

DELLA-gai genes. Importantly, these transgenic plants exhibit<br />

clear morphological changes in leaf development, <strong>and</strong> this<br />

characteristic was used to test whether phloem delivery of the<br />

Cmgaip/DELLA-gai transcripts into sink tissues could induce<br />

this mutant phenotype. <strong>The</strong>se engineered gai transcripts were<br />

found to move long-distance through the phloem, <strong>and</strong> could<br />

then exit from the terminal phloem <strong>and</strong> subsequently traffic<br />

into the apex, where they accumulated in developing leaf<br />

primordial (Haywood et al. 2005). Parallel studies conducted<br />

with transgenic plants expressing EGFP revealed the inability<br />

of these transcripts to enter the phloem. This finding suggested<br />

that phloem entry of Cmgaip/DELLA-gai transcripts must<br />

occur by a selective process.<br />

Analysis of WT tomato scions grafted onto Cmgaip <strong>and</strong><br />

DELLA-gai stocks indicated that import of these transcripts<br />

caused highly reproducible morphological changes in leaf phenotype<br />

(Figure 20). Unexpectedly, tomato leaflet morphology<br />

was found to be influenced quite late in development. Of<br />

equal importance, the presence of Cmgaip <strong>and</strong> DELLAgai<br />

transcripts, within the various tissues of the scion, was<br />

not correlated with overall sink strength. Strong signals were<br />

detected in young developing flowers <strong>and</strong> the apex, but signal<br />

could not be amplified from fruit stalks or rapidly exp<strong>and</strong>ing<br />

fruit (Figure 20C). <strong>The</strong>se findings indicated an unexpected<br />

complexity in the events underlying phloem delivery<br />

of these transcripts, suggesting a high degree of regulation<br />

over such trafficking of macromolecules. Furthermore, these<br />

studies revealed that phloem long-distance delivery of RNA<br />

can afford flexibility in adjusting developmental programs to<br />

ensure that newly emerging leaves are optimized for performance<br />

under existing environmental conditions (Haywood et al.<br />

2005).<br />

A model has been proposed that selective entry of transcripts<br />

from the CC into the sieve tube system involves specific<br />

sequences within the RNA (Lucas et al. 2001). As both Cmgaip<br />

<strong>and</strong> DELLA-gai transcripts were able to move within the<br />

heterologous plant, tomato, this finding suggested that such<br />

sequence motifs, termed “zip codes,” must be conserved <strong>and</strong>,<br />

further, the molecular machinery required for this recognition<br />

<strong>and</strong> trafficking must similarly be conserved between pumpkin,<br />

tomato <strong>and</strong> Arabidopsis.<br />

Support for this hypothesis was provided by mutational<br />

analysis of GAI in which it was clearly established that mRNA<br />

entry into the phloem is facilitated by a motif located within the 3 ′<br />

region of the transcript (Huang <strong>and</strong> Yu 2009). Furthermore, this<br />

motif was specific to GAI, as parallel experiments conducted<br />

with the four additional members of the DELLA family failed<br />

to detect their transcripts in heterografting assays. By testing<br />

an extensive series of GAI mutants, it was found that two zip<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 343<br />

Figure 20. Phloem delivery of Cmgaip/DELLA-gai transcripts<br />

modifies leaf development.<br />

(A) Schematic illustrating the V-grafting method used to test the<br />

influence of phloem-mobile transcripts, derived from the stock <strong>and</strong><br />

imported across the graft union (gu) into the scion, on developmental<br />

processes occurring in the scion.<br />

(B) Schematic showing that import of Cmgaip/DELLA-gai transcripts<br />

into wild-type tomato scions results in the development of<br />

scion leaves with characteristic morphological features associated<br />

with transgenic stock plants carrying the dominant gain-of-function<br />

Cmgaip or DELLA-gai gene (Haywood et al. 2005). Numbers<br />

represent tomato leaflet position along the leaf axis. Leaflet L0 does<br />

not exhibit this change in morphology as it had passed through this<br />

developmental stage prior to formation of the graft union.<br />

(C) Detection of Cmgaip/DELLA-gai transcripts in wild-type tomato<br />

scions grafted onto Cmgaip/DELLA-gai transgenic tomato stocks;<br />

presence (+), absence (-). Absence of phloem-mobile transcripts<br />

from the rapidly growing fruit indicates the operation of a regulatory<br />

system controlling delivery of phloem mobile transcripts to specific<br />

tissues/organs (from Haywood et al. 2005).<br />

codes appear to be required for phloem entry <strong>and</strong> translocation,<br />

one being located in the coding sequence <strong>and</strong> the other in the<br />

3 ′ untranslated region. Finally, GFP transcripts carrying these<br />

two zip codes were detected in the scion, confirming that these<br />

sequence motifs are necessary <strong>and</strong> sufficient for targeting GAI<br />

transcripts to the phloem (Huang <strong>and</strong> Yu 2009).


344 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Role of phloem RNP complexes in RNA delivery<br />

to target tissues<br />

Considering that the phloem translocation stream contains in<br />

excess of 1,000 transcripts, it is perhaps not surprising that the<br />

pumpkin phloem proteome was found to contain in excess of<br />

80 recognized RNA binding proteins (RBPs) (Lin et al. 2009).<br />

Several of these RBPs have been characterized, with the first<br />

being CmPP16-1 <strong>and</strong> CmPP16-2 from pumpkin (Xoconostle-Cázares<br />

et al. 1999). <strong>The</strong>se two proteins display properties<br />

equivalent to those of viral movement proteins (Lucas 2006)<br />

in that they bind RNA in a sequence non-specific manner <strong>and</strong><br />

mediate the cell-to-cell trafficking of transcripts through PD.<br />

Entry of CmPP16-1/2 into the sieve tube system appears to<br />

be controlled by post-translational modifications. Interestingly,<br />

both CmPP16 <strong>and</strong> its PD receptor, NCAPP1 (Lee et al. 2003),<br />

require serine residues to be phosphorylated <strong>and</strong> glycosylated<br />

for effective interaction <strong>and</strong> delivery of CmPP16 to <strong>and</strong> through<br />

PD (Taoka et al. 2007). In an elegant experiment, Aoki et al.<br />

(2005) used severed brown leafhopper stylets to introduce<br />

CmPP16-1 <strong>and</strong> CmPP16-2 directly into the sieve tube system<br />

of rice. Analysis of the long-distance movement of these two<br />

pumpkin proteins, within the rice plant, clearly revealed that<br />

they did not simply follow the direction of bulk flow. Destinationselective<br />

movement was shown to be controlled by proteins<br />

from the pumpkin phloem sap that interact with CmPP16-<br />

1/2. Collectively, these studies on CmPP16 provide important<br />

insights into the complexity of the processes that underlie<br />

macromolecular trafficking within the phloem translocation<br />

system.<br />

<strong>The</strong> most extensively characterized phloem RBP is Cm-<br />

RBP50, a polypyrimidine tract-binding protein that accumulates<br />

to high levels in pumpkin phloem sap (Ham et al. 2009). Pull<br />

down assays, using a polyclonal antibody directed against<br />

CmRBP50 <strong>and</strong> pumpkin phloem exudates, led to the identification<br />

of the proteins <strong>and</strong> mRNA species contained within<br />

a CmRBP50-associated ribonucleoprotein (RNP) complex<br />

(Figure 21A). Interestingly, CmGAIP transcripts were contained<br />

within this CmRBP50 RNP complex. Binding specificity between<br />

CmRBP50 <strong>and</strong> these CmGAIP transcripts is imparted<br />

by a series of polypyrimidine tracts located within the mRNA.<br />

As these sites differ from those involved in mediating CmGAIP<br />

transcript entry into the phloem (Huang <strong>and</strong> Yu 2009), it is likely<br />

that assembly of the CmRBP50 RNP complex occurs within the<br />

sieve tube system.<br />

Heterografting studies conducted between pumpkin (stock)<br />

<strong>and</strong> cucumber (scion) established that this RNP complex is<br />

engaged in the long-distance delivery of CmGAIP mRNA to<br />

developing tissues. Important insights into the basis for the stability<br />

of this CmRBP50-CmGAIP mRNA complex were provided<br />

by reconstitution experiments. <strong>The</strong>se studies identified a series<br />

of serine residues within the CmRBP50 C-terminus that, when<br />

phosphoryated, allow for the assembly of the RNP complex.<br />

Sequential binding of CmPP16 <strong>and</strong> the other proteins that form<br />

the complex results in an increase in its overall stability (Li et al.<br />

2011) (Figure 21B).<br />

In addition to CmGAIP, CmSCARECROW-LIKE, CmSHOOT<br />

MERISTEMLESS, CmETHYLENE RESPONSE FACTOR <strong>and</strong><br />

CmMybP transcripts were also isolated from CmRBP50 RNP<br />

complexes. Given that the watermelon phloem exudate was<br />

found to contain transcripts for some 118 transcription factors,<br />

there remains much to be done in terms of identifying <strong>and</strong> characterizing<br />

the associated RNP complexes involved in mediating<br />

their entry into, <strong>and</strong> presumed long-distance transport through,<br />

the phloem.<br />

Phloem transcripts <strong>and</strong> protein synthesis<br />

in the enucleate sieve tube system<br />

Analysis of cucumber phloem proteome <strong>and</strong> transcriptome<br />

databases identified some 169 proteins for which transcripts<br />

were also present in phloem exudates. This represents around<br />

15% of the phloem transcripts <strong>and</strong> raises the question as to why<br />

there would be the need for such transcripts when, presumably,<br />

the proteins can enter the sieve tube system by trafficking<br />

through CC-SE PD. <strong>The</strong> possibility exists that some proteins<br />

required for SE maintenance are cell-autonomous. If this were<br />

the case, synthesis within the enucleate sieve tube system<br />

would be required. It has long been assumed that the mature<br />

SE does not have the capacity for protein synthesis. However,<br />

the pumpkin phloem proteome contains numerous proteins<br />

involved in translation (Lin et al. 2009). Furthermore, gel<br />

filtration chromatography experiments performed on pumpkin<br />

phloem exudates identified complexes of proteins containing<br />

CmeIF5A <strong>and</strong> elongation factor 2, both known to be involved<br />

in protein synthesis (Ma et al. 2010). Thus, synthesis of a<br />

discrete set of essential proteins may well occur within mature<br />

SEs.<br />

Phloem-based delivery of small RNA <strong>and</strong> systemic<br />

gene silencing<br />

In recent years, post-transcriptional gene silencing has<br />

emerged as an important component of the regulatory networks<br />

that control a broad array of developmental <strong>and</strong> physiological<br />

processes (Brodersen <strong>and</strong> Voinnet 2006). <strong>The</strong>se events can<br />

occur in local tissues, <strong>and</strong> the phloem also functions as a<br />

conduit for the systemic spread of gene silencing (Melnyk et al.<br />

2011). <strong>The</strong> pioneering work of Palauqui <strong>and</strong> coworkers laid a<br />

solid foundation for this concept. Transgenic tobacco plants<br />

expressing additional copies of a nitrate reducase gene (Nia)<br />

were found to undergo a perplexing process in which small<br />

clusters of cells within mature source leaves were observed to


Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 345<br />

Figure 21. Delivery of phloem-mobile transcripts to sink tissues requires formation of stable ribonucleoprotein complexes.<br />

(A) Model illustrating the protein composition of the pumpkin CmRBP50-based ribonucleoprotein (RNP) complex that binds specifically to<br />

five phloem transcripts that encode transcription factors which are delivered into sink tissues (from Ham et al. 2009).<br />

(B) Phosphorylation of four serine residues at the C-terminus of CmRBP50 is essential for assembly of a stabilized RNP complex (left image).<br />

Mutating these serine residues prevents RNP complex assembly in the sieve tube system (right image) (from Li et al. 2011).<br />

turn white (Palauqui et al. 1996). Subsequently, this process<br />

moved up the body of the plant in a source-to-sink pattern<br />

reflective of phloem translocation.<br />

An insightful follow-up series of experiments revealed that a<br />

graft-transmissible signal moved into the scion where it caused<br />

the turnover of Nia transcripts. A deficiency in fixed nitrogen<br />

then caused the leaves of the scion to turn white (Palauqui<br />

et al. 1997) (Figure 22A). Cell-to-cell movement of the Nia<br />

silencing signal was tested by placement of a WT stem seg-<br />

ment between the silenced stock <strong>and</strong> the non-silenced scion.<br />

That white leaves still developed in these scions confirmed<br />

the involvement of the phloem (Figure 22B). This conclusion<br />

was further supported by grafting Nia-silenced scions onto<br />

non-silenced root stocks. As the direction of the phloem is<br />

from the stock to the scion, this graft combination did not<br />

result in the generation of white leaves (Figure 22C), again<br />

consistent with transmission of the silencing signal through the<br />

phloem.


346 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Figure 22. Grafting experiments illustrating that transmission<br />

of a phloem long-distance signal can induce posttranscriptional<br />

gene silencing within developing scion tissues.<br />

(A–C) Transgenic tobacco plants expressing a nitrate reductase<br />

gene (Nia) segregated into silenced (S) <strong>and</strong> non-silenced (NS)<br />

phenotypes were employed in grafting studies to test for the longdistance<br />

propagation of the silencing condition through the phloem.<br />

(A) Grafting of an NS scion onto an S stock resulted in silencing<br />

of Nia within the scion leaves. (B) Placement of a wild-type (WT)<br />

stem segment between the NS scion <strong>and</strong> S stock did not prevent<br />

the transmission of the silencing signal. (C) Grafting of a silenced<br />

scion (S) onto a NS stock does not activate silencing in the NS stock<br />

tissues.<br />

(D) Analysis of RNA samples collected from specific grafted tissues<br />

( ∗ ) confirmed that the observed silencing phenotype reflected<br />

sequence-specific targeting of the Nia transcripts (redrawn from<br />

Palauqui et al. 1997).<br />

Sequence-specificity of the silencing signal was established<br />

by grafting studies performed with nitrite reductase (Nii) transgenic<br />

tobacco stocks <strong>and</strong> non-silenced Nia scions (Figure 22D).<br />

A hypothesis was advanced that phloem-mobile silencing<br />

signals involved the translocation of antisense RNA, whose<br />

entry into the developing scion tissues caused an enzymemediated<br />

cleavage of the double-str<strong>and</strong>ed form of the target<br />

RNA (Jorgensen et al. 1998). Detection, in silenced tissues,<br />

of small (20–25 nucleotide [nt]) antisense RNA complementary<br />

to the silenced gene (sRNA) (Hamilton <strong>and</strong> Baulcombe 1999)<br />

provided strong support for the general features of this model.<br />

Analysis of RNA extracted from pumpkin phloem sap identified<br />

a population of 21 nt – 24 nt sRNA. Sequencing <strong>and</strong><br />

bioinformatics analysis indicated that these sRNAs belong to<br />

both the micro(mi)RNA <strong>and</strong> small interfering (si)RNA silencing<br />

pathways (Yoo et al. 2004). Interestingly, although equal signal<br />

strength was detected for sense <strong>and</strong> antisense sRNA probes,<br />

they did not appear to exist in the phloem sap as duplexes.<br />

<strong>The</strong> involvement of these phloem sRNAs in systemic silencing<br />

was explored using silencing (stock) <strong>and</strong> non-silencing (scion)<br />

transgenic squash (Cucurbita pepo) plants expressing a viral<br />

coat protein gene. Phloem sap from both the stock <strong>and</strong> scion<br />

tested positive for CP siRNA, <strong>and</strong> analysis of apical tissues<br />

from these scions confirmed that the level of CP mRNA had<br />

been greatly reduced. Interestingly, a low level signal for<br />

the antisense CP transcript could also be detected in the<br />

phloem sap of both stock <strong>and</strong> scion plants. Collectively, these<br />

findings offered support for the hypothesis that both siRNA <strong>and</strong><br />

antisense RNA are likely components of the systemic silencing<br />

machinery.<br />

Limited information is available concerning the mechanism<br />

by which these sRNA molecules enter <strong>and</strong> move longdistance<br />

through the phloem. Biochemical studies performed<br />

on pumpkin <strong>and</strong> cucumber phloem exudate identified a 20 kDa<br />

PHLOEM SMALL RNA BINDING PROTEIN1 (PSRP1) that<br />

bound specifically to sRNA (Yoo et al. 2004). This protein has<br />

the capacity to traffic its sRNA cargo through PD, <strong>and</strong> in the<br />

cucurbits, PSRP1 may be involved in shuttling sRNA from CCs<br />

into the sieve tube system. Interestingly, PSRP1 homologues<br />

have yet to be identified in the genomes of other plant species.<br />

This raises the possibility that additional proteins have evolved<br />

to carry out these same functions.<br />

Role of phloem-mobile sRNA in directing<br />

transcriptional gene silencing in target tissues<br />

<strong>The</strong> phloem sap collected from pumpkin <strong>and</strong> oilseed rape<br />

contains a significant population of 24-nt sRNA (Yoo et al.<br />

2004; Buhtz et al. 2008), indicating a likely involvement in<br />

transcriptional gene silencing (TGS) within sink tissues (Mosher<br />

et al. 2008). Grafting experiments performed with various<br />

combinations of GFP transgenic <strong>and</strong> DICER-LIKE mutant<br />

Arabidopsis lines provided further confirmation that a significant<br />

population of exogenous/endogenous 23 nt – 24 nt sRNA can<br />

cross the graft union (Molnar et al. 2010). Methylation analysis<br />

of DNA extracted from these grafted target tissues provided


compelling evidence for the hypothesis that phloem-mobile 24nt<br />

sRNA can mediate in epigenetic TGS of specific genomic<br />

loci.<br />

Similar findings were reported for studies on endogenous<br />

inverted repeats that generate double-str<strong>and</strong>ed RNA molecules<br />

(Dunoyer et al. 2010), as well as for transgenic plants expressing<br />

a hairpin-structured gene under the control of a viral<br />

companion-cell-specific promoter (Bai et al. 2011). In this latter<br />

case, this phloem-transmissible TGS event was shown to be<br />

inherited by subsequent progeny. Collectively, these findings<br />

support the notion that phloem-mobile sRNA can serve to regulate<br />

gene expression within developing tissues epigenetically<br />

to allow for adaptation to environmental inputs.<br />

Phloem-mobile miRNA<br />

Although generally considered to act cell-autonomously<br />

(Voinnet 2009), as mentioned above, numerous miRNAs have<br />

been detected in phloem exudates from various plant species,<br />

<strong>and</strong> the roles played by some of these have recently been<br />

established. In plants, adaptation to changing nutrient availability<br />

in the soil involves both root-to-shoot (see next section)<br />

<strong>and</strong> shoot-to-root signaling. In the case of phosphate (Pi),<br />

changes in availability within leaves leads to an upregulation in<br />

miR399 production <strong>and</strong>, subsequently, its entry into the phloem<br />

translocation stream (Lin et al. 2008; Pant et al. 2008). Delivery<br />

of miR399 into the roots results in the cleavage of the target<br />

mRNA encoding for PHO2, a ubiquitin-conjugating E2 enzyme<br />

(UBC24). This gives rise to increased uptake of Pi into these<br />

roots <strong>and</strong> restoration of Pi levels within the body of the plant.<br />

Loss of PHO2 activity probably allows for an increase in Pi<br />

transporter capacity; i.e., of influx carriers located in the outer<br />

region of the root <strong>and</strong> xylem parenchyma-located efflux carriers<br />

that function in Pi loading into the transpiration stream (Chiou<br />

<strong>and</strong> Lin 2011).<br />

Tuber induction in potato is regulated by phloem delivery<br />

of BEL5 transcripts <strong>and</strong> miR172 (Martin et al. 2009). <strong>Vascular</strong><br />

expression of miR172 <strong>and</strong> its upregulation under tuber-inducing<br />

SD conditions suggested that this miRNA may act as a longdistance<br />

signaling component in the control of potato tuber<br />

induction. Support for this notion was provided by grafting studies<br />

involving P35S:MIR172 stocks grafted to WT potato scions.<br />

Here, tuberization occurred as early as in P35S:MIR172 potato<br />

lines. In contrast, when P35S:MIR172 scions were grafted to<br />

WT stocks, early tuber induction did not occur. Although these<br />

findings are consistent with miR172 serving as a phloem-mobile<br />

signal, it is also possible that it might act through regulation, in<br />

the CCs, of an independent mobile signal.<br />

Phloem-mobile sRNA control over host infection<br />

by parasitic plants<br />

Parasitic plants cause major losses in some regions of the<br />

world (Ejeta 2007). Recent studies have established that host<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 347<br />

transcripts can enter into parasitic plants (Roney et al. 2007;<br />

David-Schwartz et al. 2008). <strong>The</strong> pathway for this trafficking<br />

is through the haustoria of the parasite that interconnects its<br />

vascular system to that of the host. This suggested that host<br />

invasion by parasitic plants might be controlled by phloem<br />

delivery of sRNA species designed specifically to target critical<br />

genes involved in the physiology or development of the<br />

parasitic weed (Yoder et al. 2009).<br />

Based on the observation that parasitic broomrape<br />

(Orobanche aegyptiaca) accumulates large quantities of mannitol,<br />

Aly et al. (2009) engineered transgenic tomatoes to<br />

express a hairpin construct to target the mannose 6-phosphate<br />

reductase (M6PR) that functions as a key enzyme in mannitol<br />

biosynthesis. Analysis of tissue from broomrape growing on<br />

these transgenic tomato plants indicated a significant reduction<br />

in both M6PR transcript <strong>and</strong> mannitol levels. This strategy<br />

gave rise to a level of tomato protection against this plant<br />

parasite.<br />

An alternate control approach based on targeting a parasitic<br />

developmental program involved the development of transgenic<br />

tobacco plants expressing hairpin constructs for two<br />

dodder (Cuscuta pentagona) haustoria-expressed KNOTTEDlike<br />

HOMEOBOX1 (KNOX1) genes. <strong>The</strong>se constructs were<br />

driven by the CC-specific SUC2 promoter <strong>and</strong> were based<br />

on 3 ′ UTRs that did not display sequence homology to the<br />

related tobacco orthologues, STM <strong>and</strong> KNAT1–3 (Alakonya<br />

et al. 2012). Defects in haustoria development <strong>and</strong> connection<br />

to the transgenic tobacco plants were highly correlated with<br />

the presence of KNOX1 siRNA, delivered most likely through<br />

the vascular system, <strong>and</strong> down-regulation of the C. pentagona<br />

KNOX1 transcript levels. Importantly, dodder plants growing<br />

on these transgenic tobacco plants exhibited greatly reduced<br />

vigor. Collectively, these studies indicate that an effective<br />

control of plant parasitism may be achieved by targeting a<br />

pyramided combination of parasite genes involved in various<br />

aspects of growth <strong>and</strong> development.<br />

Root-to-shoot Signaling<br />

Response to abiotic stress<br />

Signals arising within the root system can provide shoots with<br />

an early warning of root conditions, such as water deficiency,<br />

nutrient availability/deficiency, <strong>and</strong> so forth (Figure 23). <strong>The</strong><br />

xylem transports hormones, such as abscisic acid (ABA)<br />

(Bahrun et al. 2002; Jiang <strong>and</strong> Hartung 2008), ethylene <strong>and</strong><br />

cytokinin (CK) (Takei et al. 2002; Hirose et al. 2008; Kudo<br />

et al. 2010; Ghanem et al. 2011), as well as strigolactones<br />

(SLs) (Gomez-Roldan et al. 2008; Umehara et al. 2008; Brewer<br />

et al. 2013; Ruyter-Spira et al. 2012) from roots to aboveground<br />

tissues. In this section of the review, we will address the role of<br />

these xylem-borne signaling agents.


348 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Figure 23. <strong>The</strong> plant vascular system serves as an effective<br />

inter-organ communication system.<br />

In response to a wide range of environmental <strong>and</strong> endogenous<br />

inputs, the xylem (blue lines) transmits root-to-shoot signals (blue<br />

circles), including hormones such as abscisic acid, ACC (ethylene<br />

precursor) <strong>and</strong> cytokinin, as well as strigolactones (SLs). <strong>The</strong>se<br />

xylem-borne signaling agents serve to communicate the prevailing<br />

conditions within the soil. <strong>The</strong> phloem (pink lines) transports a wide<br />

array of shoot-to-root signaling molecules (pink circles), including<br />

auxin, cytokinin, proteins <strong>and</strong> RNA species, including mRNA <strong>and</strong><br />

sRNA. <strong>The</strong>se phloem-borne signaling agents complete the longdistance<br />

communication circuit that serves to integrate developmental<br />

<strong>and</strong> physiological events, occurring within shoot <strong>and</strong> root<br />

tissues, in order to optimize the plant performance under the existing<br />

growth/environmental conditions.<br />

When soils become dry, root-derived signals are transported<br />

through the xylem to leaves in order to effect a reduction in both<br />

leaf transpiration <strong>and</strong> vegetative growth. Tight control between<br />

water uptake by the root system <strong>and</strong> the xylem transpiration<br />

stream is achieved through regulation of leaf stomatal aperture.<br />

Production of ABA within roots <strong>and</strong> its transport to the leaves<br />

could contribute to preventing excess water loss, as it is has<br />

long been known that ABA is a key regulator of stomatal conductance<br />

(Mittelheuser <strong>and</strong> Van Steveninck 1969; Schachtman<br />

<strong>and</strong> Goodger 2008).<br />

<strong>The</strong> ABA content in roots is well correlated with both soil<br />

moisture <strong>and</strong> root relative water content (Davis <strong>and</strong> Zhang<br />

1991; Thompson et al. 2007). Although large increases in ABA<br />

are detected in the xylem sap, when plants are exposed to<br />

drought conditions (Christmann et al. 2007), grafting studies<br />

have indicated that root-derived ABA is not necessary for<br />

drought-induced stomatal closure (Holbrook et al. 2002). Furthermore,<br />

recent studies have shown that leaf-derived synthesis<br />

of ABA contributes to water-stress-induced down-regulation<br />

of stomatal conductance (Holbrook et al. 2002; Thompson<br />

et al. 2007). Thus, further studies are required to evaluate the<br />

relative contribution of root-derived versus shoot-synthesized<br />

ABA in terms of the overall efficacy of stomatal control over the<br />

transpiration stream.<br />

<strong>The</strong>re is some indication that ethylene-based signaling may<br />

also contribute to root-to-shoot communication under abiotic<br />

stress conditions. For example, the anaerobic environment<br />

caused by soil flooding can increase the level of aminocyclopropane<br />

carboxylic acid (ACC, the immediate precursor of<br />

ethylene) in plant roots. ACC has been detected in the xylem<br />

from both flooded <strong>and</strong> drought-stressed plants (Tudela <strong>and</strong><br />

Primo-Millo 1992; Belimov et al. 2009). This root-derived ACC<br />

is transported to the shoot where it then gives rise to increased<br />

ethylene production which can play a role in regulating shoot<br />

growth <strong>and</strong> development under these stress conditions (Voesenek<br />

et al. 2003; Pérez-Alfocea et al. 2011).<br />

Changes in xylem sap pH have also been reported for<br />

plants exposed to drought conditions. Alkalinization of the<br />

xylem sap appears to be correlated with enhanced stomatal<br />

closure (Jia <strong>and</strong> Davies 2008; Sharp <strong>and</strong> Davies 2009). <strong>The</strong>se<br />

pH changes may act synergistically with ABA <strong>and</strong> ACC to<br />

generate an effective root-to-shoot signaling system for water<br />

stress. <strong>The</strong> involvement of other known xylem-based root-toshoot<br />

signals, such as CK, etc., remains to be established in<br />

terms of contributing to water stress signaling. In any event,<br />

advancing our underst<strong>and</strong>ing of the mechanisms of root-toshoot<br />

signaling associated with water stress should lead the<br />

way for the development of crops with improved water use<br />

efficiencies.<br />

Xylem signals associated with nutrient stress<br />

<strong>The</strong> phenotypic plasticity that plants display in response to<br />

changes in their nutrient supply requires the operation of rootto-shoot<br />

signaling. Such signals from roots can provide shoots<br />

with an early warning of decreases in nutrient supply, while<br />

signals from shoots can ensure that the nutrient acquisition<br />

by roots is integrated to match the nutrient dem<strong>and</strong> of shoots<br />

(Lough <strong>and</strong> Lucas 2006; Liu et al. 2009).<br />

CK plays an important role in plant growth <strong>and</strong> development<br />

<strong>and</strong> its involvement as a xylem-mobile signal in regulating the<br />

nutrient starvation response, such as occurs under nitrogen <strong>and</strong><br />

phosphorus deficiency conditions, is well established (Takei<br />

et al. 2002; Hirose et al. 2008; Ghanem et al. 2011). Nitrate<br />

deprivation leads to a reduction in the level of mobile CK in the<br />

xylem sap, whereas upon resupply of nitrate to these stress


oots, CK again increases in the xylem transpiration stream<br />

(Rahayu et al. 2005; Ruffel et al. 2011). Interestingly, transzeatin-type<br />

CK moves in the xylem, <strong>and</strong> isopentenyl-type CK<br />

is present in the phloem translocation stream. This suggests<br />

that these structural variations carry specific information from<br />

the root-to-shoot <strong>and</strong> shoot-to-root, respectively (Hirose et al.<br />

2008; Werner <strong>and</strong> Schmülling 2009).<br />

As discussed above, phosphate acquisition by the root system<br />

involves phloem-mobile signals from the shoot (Figure 24).<br />

In terms of the root-to-shoot component of this phosphate<br />

signaling network, it has been suggested that the level of<br />

phosphate in the xylem transpiration stream may serve as one<br />

component in this signaling pathway (Bieleski 1973, Poirier<br />

et al. 1991; Burleigh <strong>and</strong> Harrison 1999; Hamburger et al. 2002;<br />

Lai et al. 2007; Stefanovic et al. 2007; Chiou <strong>and</strong> Lin 2011;<br />

Thibaud et al. 2010). Studies on the growth of Arabidopsis roots<br />

being exposed to low phosphate conditions identified the tip of<br />

the primary root, including the meristem region <strong>and</strong> root cap,<br />

as the site that may sense local phosphate availability (Linkohr<br />

et al. 2002; Svistoonoff 2007). However, currently, there is no<br />

evidence for the existence of a phosphate sensor or receptor.<br />

Both CK <strong>and</strong> SLs have also been considered to function in<br />

xylem transmission of root phosphate status (Martin et al. 2000;<br />

Franco-Zorrilla et al. 2005; Kohlen et al. 2011). <strong>Plant</strong>s grown<br />

under limiting phosphate conditions have repressed levels of<br />

trans-zeatin-type CK in their xylem sap (Martin et al. 2000) <strong>and</strong>,<br />

under these conditions, expression of the CK receptor CRE1<br />

is similarly decreased (Franco-Zorrilla et al. 2002, 2005). In<br />

many plant species, the SLs are up-regulated upon exposure<br />

to phosphate deficiency conditions. Grafting studies have indicated<br />

that SLs produced in the root can move to the shoot<br />

(Beveridge et al. 1994; Napoli 1996; Turnbull et al. 2002). In<br />

such studies, WT rootstocks grafted to mutant scions lacking<br />

the ability to produce SLs were able to restore WT branching<br />

patterns in these scions. Thus, xylem-transported SLs can<br />

contribute to the regulation of shoot architectural responses<br />

to phosphate-limiting conditions (Kohlen et al. 2011). Collectively,<br />

these findings suggest that the levels of phosphate, CK<br />

<strong>and</strong> SLs in the xylem transpiration stream play an important<br />

role in coordinating vegetative growth with phosphate nutrient<br />

availability (Rouached et al. 2011).<br />

Xylem signaling in plant-symbiotic associations<br />

<strong>The</strong> interaction of nitrogen-fixing bacteria (Rhizobia) is generally<br />

confined to legumes, whereas most flowering plants<br />

establish symbiotic associations with arbuscular mycorrhizal<br />

(AM) fungi for phosphate acquisition. In both types of plantsymbiont<br />

association, there is a significant metabolic cost to<br />

the plant host. Thus, there is a need for the plant to ensure<br />

that the cost-benefit ratio remains favorable. To this end,<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 349<br />

Figure 24. Long-distance signaling in response to phosphate<br />

deficiency conditions.<br />

Phosphate (Pi) availability in the soil solution is transmitted through<br />

the xylem transpiration stream (blue lines) which passes predominantly<br />

to the mature source leaves. Pi per se, <strong>and</strong>/or other root-toshoot<br />

signals (1), including cytokinin <strong>and</strong> strigolactones, are thought<br />

to be involved in this nutrient signal transduction pathway. When<br />

roots encounter low levels of available Pi, changes in these xylemborne<br />

signaling components are decoded in the leaves (2), resulting<br />

in the activation of Pi deficiency responsive pathways. Outputs from<br />

this Pi-stress response pathway, including miR399 21-nucleotide<br />

silencing agents <strong>and</strong> other potential signaling components (3) are<br />

loaded into the phloem (red lines). Phloem-mobile signals move<br />

down to the root where they enter different target receiver cells<br />

to mediate an increase in Pi uptake (4) <strong>and</strong> alter root architecture<br />

(4’). <strong>The</strong> miR399 signal targets PHOSPHATE2 (PHO2) transcripts<br />

to derepress Pi transporter activity. A different set of phloemmobile<br />

signals are likely delivered to developing leaves (5) <strong>and</strong><br />

the shoot apex (6) to regulate growth <strong>and</strong> development in order to<br />

survive under the Pi-stress condition. This long-distance signaling<br />

network operates to ensure that the root system integrates its<br />

physiological activities to optimize growth conditions within the<br />

shoot.


350 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Figure 25. Autoregulation of symbiosis between plants <strong>and</strong><br />

bacteria/fungi involves long-distance signaling through the<br />

vascular system.<br />

During plant-symbiont associations a significant metabolic cost is<br />

incurred by the plant host. To ensure that the cost-benefit ratio<br />

remains favorable to the plant, systemic autoregulation systems<br />

evolved to control the level of root nodulation associated with<br />

nitrogen-fixing Rhizobium <strong>and</strong> growth on the roots of arbuscular<br />

mycorrhizal (AM) fungi. In this system, root-derived signals (blue<br />

circles) which are still unknown but may involve CLE peptides in<br />

the case of nodulation, are transported through the xylem (blue<br />

lines) to the shoot. In mature source leaves, these autoregulation<br />

of nodulation signals appear to be recognized by a LRR-RLK.<br />

Although grafting studies have established that signals (pink circles)<br />

enter <strong>and</strong> move through the phloem (pink lines) to the roots, their<br />

identities remain to be elucidated. <strong>The</strong>se shoot-to-root signals are<br />

involved in down-regulating the root – Rhizobium/AM fungi symbiotic<br />

association.<br />

plants have evolved a systemic feedback regulatory system<br />

termed “autoregulation” in which nodule formation <strong>and</strong> the<br />

ongoing development of the association with the AM fungus is<br />

negatively controlled by long-distance signaling (Catford et al.<br />

2003; Staehelin et al. 2011) (Figure 25).<br />

<strong>The</strong> process of autoregulation in legumes in which the<br />

number of symbiotic root nodules is controlled by long-distance<br />

communication between the root <strong>and</strong> shoot has been well<br />

studied. Two component pathways are thought to operate<br />

involving root-derived signals through the xylem <strong>and</strong> shootderived<br />

signals through the phloem. Following rhizobial infection,<br />

a root-derived signal is generated that is then translocated<br />

to the shoot. This xylem-borne signal is perceived in the<br />

shoot <strong>and</strong> subsequently leads to the production of a shootderived<br />

signal whose movement through the phloem to the<br />

roots causes a block to further nodulation. Consistent with this<br />

notion, mutant legumes have been identified that are defective<br />

in autoregulation of nodulation, <strong>and</strong> grafting experiments established<br />

that these mutants were not capable of producing the<br />

requisite shoot-derived signals.<br />

Genetic studies suggest that CLE peptides, induced in<br />

response to rhizibial nodulation signals in roots, serve as<br />

signaling agents that travel through the xylem to the shoots.<br />

A leucine-rich repeat (LRR) CLAVATA-like receptor kinase,<br />

located in the leaves, appears to function in this signaling<br />

pathway (Searle et al. 2003). Although not yet proven, the<br />

CLE peptides imported from roots are likely perceived by the<br />

LRR autoregulation receptor kinase in the shoots (Okamoto<br />

et al. 2009; Miyazawa et al. 2010; Osipova et al. 2012). In any<br />

event, it will be of considerable interest to identify the feedback<br />

signal(s) that enters the phloem to down-regulate nodulation<br />

back in the roots (Oka-Kira et al. 2005).<br />

In addition to CLE – LRR-RLK signals from the root, soil available<br />

nitrogen also appears to participate in this long-distance<br />

signaling system. Consistent with this notion, in legumes, high<br />

levels of soil nitrate cause strong up-regulation of root CLE<br />

gene expression (Okamoto et al. 2009; Reid et al. 2011).<br />

Furthermore, high nitrate or ammonia levels abolish nodulation,<br />

<strong>and</strong> autoregulation-defective mutants exhibit more or fewer<br />

nitrate-insensitive phenotypes. Thus, a combination of CLE<br />

peptides <strong>and</strong> nitrate/ammonia levels in the xylem transpiration<br />

stream could function as the root-to-shoot signals that allow<br />

legumes to integrate autoregulation of nodulation with environmental<br />

nitrogen conditions.<br />

With respect to phosphatesignaling, studies using cucumber<br />

revealed that application of root extracts from mycorrhizal<br />

plants reduced the degree of root colonization by AM fungi<br />

(Vierheilig et al. 2003). In contrast, root extracts from noninfected<br />

plants stimulated successful AM fungal colonization.<br />

This study supports the hypothesis that an equivalent autoregulation<br />

system operates to control the plant-AM fungal<br />

association involved in plant phosphate homeostasis. Given<br />

the attributes of the cucurbits for analysis of phloem <strong>and</strong> xylem<br />

sap, this system might prove invaluable for studies aimed<br />

at identifying the agents that serve to coordinate phosphate<br />

acquisition by the roots with utilization in the shoots.


Role of xylem signals in coordination of shoot<br />

architecture<br />

A number of studies have indicated that root-derived signals<br />

play a role in the regulation of vegetative growth (Van Norman<br />

et al. 2004; Van Norman <strong>and</strong> Sieburth 2007; Sieburth <strong>and</strong><br />

Lee 2010). <strong>The</strong>se signals contribute to water <strong>and</strong> nutrient use<br />

efficiency through the control over shoot branching <strong>and</strong> growth.<br />

<strong>The</strong> BYPASS1, 2, 3 (BPS1, 2, 3) genes are required to prevent<br />

the synthesis of a novel substance, a bps signal that moves<br />

from root-to-shoot, where it modifies shoot growth (Van Norman<br />

et al. 2004; Van Norman <strong>and</strong> Sieburth 2007; Lee et al. 2012;<br />

Lee <strong>and</strong> Sieburth 2012).<br />

Although the functions of the BPS proteins remain to be<br />

elucidated, studies based on bps mutants clearly established<br />

that the roots of these mutants cause arrested shoot growth,<br />

likely due to the over-production of an inhibitor of shoot growth.<br />

Grafting experiments established that the root system of the<br />

bps mutants was both necessary <strong>and</strong> sufficient to induce shoot<br />

arrest, <strong>and</strong> revealed that BPS proteins can work to generate<br />

mobile root-to-shoot signals that can inhibit shoot growth (Van<br />

Norman et al. 2004; Lee et al. 2012). It will be of great interest<br />

to unravel the underlying mechanism by which shoot growth is<br />

controlled by this signaling pathway.<br />

Recent studies reported that bps mutants show normal<br />

responses to both exogenous auxin <strong>and</strong> polar auxin transport<br />

inhibitors, suggesting that the primary target of the bps signal<br />

is independent of auxin. Furthermore, this root-to-shoot signal<br />

appears to act in parallel with auxin to regulate patterning<br />

<strong>and</strong> growth in various tissues <strong>and</strong> at multiple developmental<br />

stages (Lee et al. 2012). Clearly, further characterization of<br />

the BPS signaling pathway could well open the door to novel<br />

approaches towards controlling shoot architecture in specialty<br />

crops.<br />

<strong>The</strong> SLs have been referred to as rhizosphere signaling<br />

molecules (Nagahashi <strong>and</strong> Douds 2000; Akiyama et al. 2005)<br />

that also participate in the regulation of shoot architecture by<br />

suppressing lateral shoot development. Biosynthetic SL mutants<br />

exhibit a highly branching phenotype <strong>and</strong>, interestingly,<br />

phosphate starvation in these plants causes a reduction in<br />

shoot branching (Umehara et al. 2008, 2010). As previously<br />

reported, plants grown under phosphate deficiency conditions<br />

have fewer shoots <strong>and</strong> an increase in lateral roots (Umehara<br />

et al. 2010; Kohlen et al. 2011). Several plant hormones,<br />

such as auxin <strong>and</strong> ethylene, also appear to be involved in<br />

linking phosphate signaling with plant growth responses (Chiou<br />

<strong>and</strong> Lin 2011). Auxin signaling was shown to be associated<br />

with changes in root system architecture under phosphate<br />

deficiency conditions (López-Bucio et al. 2002). Recent studies<br />

suggest that an auxin receptor TRI <strong>and</strong> the auxin signaling<br />

pathway are involved in this SL-regulated root-sensing of low<br />

phosphate conditions (Mayzlish-Gati et al. 2012). It is likely<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 351<br />

that auxin <strong>and</strong> SLs function, cooperatively, to control shoot<br />

branching (Brewer et al. 2009; Domagalska <strong>and</strong> Leyser 2011).<br />

<strong>The</strong>se hormones move through the xylem <strong>and</strong> phloem, respectively,<br />

to form a network of systemic signals to orchestrate plant<br />

architecture at the whole plant level.<br />

<strong>Vascular</strong> Transport of Microelement<br />

Minerals<br />

In the last decade, significant advances have been made in the<br />

underst<strong>and</strong>ing of the mechanisms that control the intracellular<br />

homeostasis of microelement minerals (Takano et al. 2008;<br />

Curie et al. 2009; Williams <strong>and</strong> Pittman 2010; Conte <strong>and</strong><br />

Walker 2011; Waters <strong>and</strong> Sankaran 2011; Ivanov et al. 2012;<br />

Sperotto et al. 2012). However, relatively little is known about<br />

the processes governing their long-distance transport. Major<br />

questions remaining relate to the mechanisms of vascular<br />

loading/unloading, as well as the chemical speciation of these<br />

elements during their transport. Furthermore, transport is not<br />

a static process <strong>and</strong>, therefore, may differ not only with the<br />

nutrient <strong>and</strong> plant species but also with other factors, such as<br />

developmental stage, circadian cycle, <strong>and</strong> nutritional status. In<br />

this section of the review, we assess the current knowledge<br />

on microelement vascular transport focusing on these open<br />

questions.<br />

Microelement trafficking <strong>and</strong> speciation in xylem sap<br />

In the xylem sap, the non-proteinogenic amino acid nicotianamine<br />

(NA), histidine, <strong>and</strong> organic acids are usually associated<br />

with cationic microelements (Figure 26, Table 2). NA<br />

binds several transition metals with very high affinity, including,<br />

in order of stability, Fe(III), Cu(II), Ni(II), Co(II), Zn(II), Fe(II)<br />

<strong>and</strong> Mn(II) (von Wirén et al. 1999; Rellán-Álvarez et al. 2008;<br />

Curie et al. 2009). Insights into the role of NA complexation<br />

<strong>and</strong> trafficking have been provided by studies of NA-deficient<br />

mutants. Here, the chloronerva tomato mutant is interesting<br />

as it has high root Cu concentrations, but the concentration<br />

in xylem sap is low, indicating a failure in Cu transport into<br />

mature leaves. This finding indicates that Cu(II)-NA likely<br />

serves as a key complex in the xylem sap (Herbik et al.<br />

1996; Pich <strong>and</strong> Scholz 1996). With regard to other cationic<br />

microelements, analysis of an A. thaliana nicotianamine synthase<br />

(NAS) quadruple mutant (which has low levels of NA)<br />

showed that long-distance transport of Fe through the xylem<br />

was not affected, <strong>and</strong> Fe accumulated in the leaves (Klatte<br />

et al. 2009). Other studies performed on NAS-overexpressing<br />

tobacco <strong>and</strong> A. thaliana plants reported elevated Ni tolerance<br />

<strong>and</strong> high Zn levels in young leaves. Finally, studies conducted<br />

on several metal hyperaccumulator species (Krämer 2010)<br />

identified Cu(II)-NA, Zn(II)-NA <strong>and</strong> Ni(II)-NA complexes in<br />

the roots, xylem sap <strong>and</strong> leaves (Schaumlöffel et al. 2003;


352 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Figure 26. Schematic depicting membrane transporters involved<br />

in loading <strong>and</strong> unloading of micronutrient elements in<br />

the vascular systems.<br />

Homologues from different plant species (At, Arabidopsis thaliana;<br />

Tc, Thlaspi caerulescens; Os,Oryza sativa) are given as examples.<br />

P-type ATPase (HMA), ferroportin (IREG) <strong>and</strong> <strong>and</strong> MATE.<br />

(FRD) families are involved in loading Zn <strong>and</strong> Cu, Fe, <strong>and</strong> citrate,<br />

respectively, into the xylem. Borate is loaded into the xylem by<br />

the anion efflux system, AtBOR1. <strong>The</strong> chemical species present<br />

in the xylem <strong>and</strong> phloem sap are indicated; several micronutrient<br />

species may occur in xylem sap. Histidine (His), Nicotianamine<br />

(NA), <strong>and</strong> organic acids are the most likely chelating agents of these<br />

mineral micronutrients. <strong>The</strong> complex Fe3Cit3 has been detected in<br />

xylem sap from tomato. Unloading of Ni, Fe <strong>and</strong> Zn from the xylem<br />

takes place via members of the Yellow Stripe-Like family of metal<br />

transporters (YSL). Phloem loading <strong>and</strong> unloading of Fe, Mn, Cu<br />

<strong>and</strong> Zn is also mediated by several members of the YSL family<br />

in rice <strong>and</strong> Arabidopsis. AtOPT3, a member of the oligopeptide<br />

transporter family, is involved in Fe <strong>and</strong> Mn loading into the sieve<br />

tube system. Chemical species of micronutrient minerals in the<br />

phloem sap include complexes of Ni, Cu, Zn <strong>and</strong> Fe with NA. <strong>The</strong><br />

complexes Zn-NA <strong>and</strong> Fe (III)-2 ′ DMA have been recently detected<br />

in phloem sap from rice. Iron Transporter Protein (ITP) <strong>and</strong> Copper<br />

Chaperone (CCH) may have a role in Fe <strong>and</strong> Cu transport within<br />

the phloem, respectively, whereas, Mn <strong>and</strong> Ni have been detected<br />

Vacchina et al. 2003; Ouerdane et al. 2006; Mijovilovich et al.<br />

2009, Trampczynska et al. 2010).<br />

Histidine (His) can function to chelate Zn, Cu <strong>and</strong> Ni in the<br />

xylem sap (Krämer et al. 1996; Salt et al. 1999; Liao et al.<br />

2000; Küpper et al. 2004). An extended X-ray absorption fine<br />

structure (EXAFS) study demonstrated that most of the Zn<br />

in petioles <strong>and</strong> stems of Noccacea caerulescens existed as<br />

a complex with His (Küpper et al. 2004). However, a recent<br />

study performed on the same species proposed His as a Zn<br />

lig<strong>and</strong> within cells, <strong>and</strong> NA as the Zn chelator involved in longdistance<br />

transport (Trampczynska et al. 2010). For Cu, as<br />

commented above, NA plays a key role in xylem transport.<br />

However, xylem transport of Cu in tomato <strong>and</strong> chicory is<br />

efficient even in the absence of NA, provided that His is present,<br />

thus offering support for the existence of both mechanisms<br />

for Cu complexation in xylem sap (Liao et al. 2000). Based<br />

on these findings, Irtelli et al. (2009) proposed that, under Cu<br />

deficiency conditions, NA is responsible for Cu chelation in<br />

xylem sap, whereas His <strong>and</strong> Pro serve as the major chelators<br />

in excess Cu conditions.<br />

<strong>The</strong> involvement of His in Ni chelation in the xylem sap<br />

has been proposed based on studies in Ni-hyperaccumulator<br />

species (Krämer et al. 1996; Kerkeb <strong>and</strong> Krämer 2003; Mari<br />

et al. 2006; Krämer 2010; McNear et al. 2010). In these plants,<br />

there is an enhanced expression of the first enzyme in the<br />

His biosynthetic pathway <strong>and</strong> higher concentrations of His in<br />

xylem sap (Krämer et al. 1996; Ingle et al. 2005). On the other<br />

h<strong>and</strong>, His-overproducing transgenic A. thaliana lines displayed<br />

enhanced Ni tolerance, but did not exhibit increased Ni concentrations<br />

in xylem sap or leaves (Wycisk et al. 2004; Ingle et al.<br />

2005). <strong>The</strong>se studies suggest that, in non-hyperaccumulator<br />

plants, other chelating agents such as NA <strong>and</strong> organic acids,<br />

may also play important roles (Verbruggen et al. 2009; Hassan<br />

<strong>and</strong> Aarts 2011). Accordingly, studies on natural variation<br />

among Arabidopsis accessions indicated that a Ni(II)-malic acid<br />

complex may be involved in translocation of Ni from roots to<br />

shoots (Agrawal et al. 2012).<br />

As mentioned above, organic acids have also been hypothesized<br />

to serve as chelators for Fe, Zn, Ni <strong>and</strong> Mn in xylem sap,<br />

based on in silico calculations using xylem sap composition<br />

(von Wirén et al. 1999; López-Millán et al. 2000; Rellán-Alvárez<br />

et al. 2008). For instance, in silico speciation studies in xylem<br />

sap of the hyperaccumulator Alyssum serpyllifolium found<br />

approximately 18% of Ni bound to organic acids, mainly malate<br />

<strong>and</strong> citrate (Alves et al. 2011), <strong>and</strong> in tomato, Mn was predicted<br />

←−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−<br />

in association with low molecular (LMW) peptides <strong>and</strong> organic<br />

compounds. <strong>The</strong> molybdate anion has been detected in both xylem<br />

<strong>and</strong> phloem sap. Boron is present as borate <strong>and</strong> boric acid in xylem<br />

sap <strong>and</strong> as complexes with sugar alcohols in phloem sap.


Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 353<br />

Table 2. Chemical speciation of cationic microelements present within the xylem transpiration stream <strong>and</strong> the phloem translocation<br />

stream<br />

Xylem sap Phloem sap<br />

Nicotianamine Histidine Organic acids Nicotianamine Peptides/proteins<br />

Fe Rellán-Álvárez et al. Klatte et al. 2009 ITP<br />

2010<br />

Takahashi et al. 2003<br />

a<br />

Krüger et al. 2002<br />

Zn Klatte et al. 2009 Küpper et al. 2004 Salt et al. 1999<br />

Nishiyama et al. 2012<br />

Trampczynska et al.<br />

2010<br />

Takahashi et al. 2003<br />

Cu Pich <strong>and</strong> Scholz 1996<br />

Liao et al. 2000<br />

Pich et al. 1994 CCH<br />

Herbik et al. 1996<br />

Irtelli et al. 2009<br />

Mira et al. 2001<br />

Liao et al. 2000<br />

(metallothioneins)<br />

Guo et al. 2003, 2008<br />

Mn White et al. 1981 LMW peptides<br />

Van Goor <strong>and</strong> Wiersma<br />

1976<br />

Ni Ouerdane et al. 2006 Krämer et al. 1996 Agrawal et al. 2012 Schaumlöffel et al. 2003 LMW compounds<br />

Kerkeb <strong>and</strong> Krämer Alves et al. 2011<br />

Wiersma <strong>and</strong> Van Goor<br />

2003<br />

Krämer 2010<br />

McNear et al. 2010<br />

1979<br />

a ITP, Iron Transport Protein; CCH, Copper Chaperone Protein; LMW, Low Molecular Weight.<br />

as a citrate complex (White et al. 1981). However, these<br />

models did not include other possible chelating agents such<br />

as amino acids or NA. Recently, a tri-Fe(III), tri-citrate complex<br />

(Fe3Cit3) was identified in the xylem sap of tomato plants, using<br />

an integrated mass spectrometry approach (Rellán-Álvárez<br />

et al. 2010). Also, by means of X-ray absorption spectroscopy,<br />

organic acids have been shown to complex Zn in xylem sap of<br />

Noccacea caerulescens (Salt et al. 1999).<br />

With regard to anionic microelements, the soluble molybdate<br />

anion, which is the predominant aqueous species at pH values<br />

above 4.0, has been detected in both xylem <strong>and</strong> phloem<br />

sap, <strong>and</strong> is assumed to be the major chemical species of<br />

Mo delivered by these two long-distance transport systems<br />

(Marschner 1995). <strong>The</strong> fact that the molybdate anion is not<br />

very biologically active may allow for its transport as a free<br />

anion. For boron (B), in addition to boric acid <strong>and</strong> borate, at<br />

least one other yet unidentified B-containing compound has<br />

been described in the xylem sap of squash roots (Iwai et al.<br />

2003). This compound has a lower molecular weight than the<br />

rhamnogalacturan II-B complex, which contributes to cell wall<br />

strengthening (Takano et al. 2008).<br />

Microelement trafficking <strong>and</strong> speciation in the phloem<br />

sap<br />

Metal mobility in the phloem sap depends on the individual<br />

microelement, its chemical species, <strong>and</strong> in some cases on<br />

the nutritional status of the plant. Zn <strong>and</strong> Ni are considered<br />

highly mobile in the phloem translocation stream; for instance,<br />

the loading of these metals into the developing wheat grain<br />

occurs mostly via the phloem, with transfer from xylem to<br />

phloem occurring in the rachis <strong>and</strong> the peduncle (Riesen <strong>and</strong><br />

Feller 2005). In contrast, manganese (Mn) appears to be poorly<br />

mobile in the phloem; it can be translocated out of source<br />

leaves, but the loading of Mn into the developing grain is<br />

poor in most crop species (Riesen <strong>and</strong> Feller 2005; Williams<br />

<strong>and</strong> Pittman 2010). For Cu, its mobility in the phloem sap is<br />

intermediate. For instance, in wheat, translocation from mature<br />

to younger developing leaves does not occur, <strong>and</strong> it has been<br />

proposed that when Cu enters the cell it becomes bound<br />

by chaperones <strong>and</strong>, therefore, is not immediately available<br />

for retranslocation. Later remobilization of Cu appears to be<br />

possible during leaf senescence when proteins are hydrolyzed,<br />

thereby releasing Cu (Puig <strong>and</strong> Penarrubia 2009). Fe is also<br />

considered intermediately mobile in plants, <strong>and</strong> studies have<br />

shown that it is translocated to young barley leaves mainly via<br />

phloem transport (Tsukamoto et al. 2009). Grusak (1994) also<br />

reported phloem transport of Fe from various source tissues to<br />

developing Pisum sativum seeds. And, as occurs with Cu, an<br />

increased remobilization of Fe occurs during leaf senescence<br />

(Waters et al. 2009; Sperotto et al. 2010).<br />

With regard to the chemical species in which these metals<br />

are transported through the phloem (Figure 26), Fe, Cu <strong>and</strong><br />

Zn are considered to move as either NA- or metal-mugineic


354 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

acid complexes (mugineic acids are synthesized from NA),<br />

especially as the neutral-to-basic pH of the phloem sap is<br />

suitable for metal-NA formation (Pich et al. 1994; von Wirén<br />

et al. 1999; Takahashi et al. 2003). Accordingly, Zn(II)-NA <strong>and</strong><br />

Fe(III)-2 ′ -deoxymugineic acid complexes have been detected<br />

in phloem sap from rice (Nishiyama et al. 2012). Furthermore,<br />

mutants defective in NA production display lower Mn, Zn,<br />

Fe <strong>and</strong> Cu concentrations in reproductive organs (Takahashi<br />

et al. 2003; Curie et al. 2009). However, it is still unclear<br />

whether this impediment to metal delivery into sink organs<br />

is due to alterations in phloem loading <strong>and</strong> transport, per se,<br />

or to changes in the intracellular concentrations of metals in<br />

source tissues where NA also plays an important role in metal<br />

chelation.<br />

<strong>The</strong> phloem may also transport other forms of Fe; e.g., an<br />

iron transport protein (ITP) is present in the phloem sap of<br />

Ricinus communis (Krüger et al. 2002), but it has not yet been<br />

found in other species. A copper chaperone protein (CCH) has<br />

also been proposed to play a role in phloem transport of Cu<br />

from senescing to young leaves (Mira et al. 2001). In addition,<br />

metallothioneins (types 1, 2 <strong>and</strong> 3), proteins predominantly<br />

regulated by Cu, also appear to function in Cu accumulation<br />

<strong>and</strong> phloem transport during senescence (Guo et al. 2003,<br />

2008). <strong>The</strong>se proteins are also associated with Cu tolerance<br />

(Murphy <strong>and</strong> Taiz 1997; van Hoof et al. 2001; Jack et al. 2007).<br />

Currently, little information is available concerning the chemical<br />

forms of Ni <strong>and</strong> Mn in the phloem sap. In R. communis, Mn<br />

has been detected in association with low molecular weight<br />

peptides (van Goor <strong>and</strong> Wiersma 1976), whereas Ni can be<br />

complexed with negatively charged organic compounds with<br />

a molecular weight in the range of 1,000–5,000 Da (Wiersma<br />

<strong>and</strong> van Goor 1979).<br />

<strong>The</strong> mobility of molybdenum (Mo) in the phloem varies<br />

depending on its concentration <strong>and</strong> on plant age. Interestingly,<br />

in wheat, Mo has been associated with the existence of “Mobinding<br />

sites” in the phloem that, until saturated, appear to<br />

prevent its long-distance translocation (Yu et al. 2002). B mobility<br />

in the phloem is highly dependent on the plant species. In<br />

plants that transport sugar alcohols, B appears to be complexed<br />

with diols <strong>and</strong> polyols (Brown <strong>and</strong> Hu 1996; Hu et al. 1997;<br />

Takano et al. 2008). Complexes of sorbitol-B-sorbitol, fructose-<br />

B-fructose, sorbitol-B-fructose <strong>and</strong> mannitol-B-mannitol have<br />

been identified in peach <strong>and</strong> celery phloem sap (Hu et al. 1997).<br />

Enhancement of sorbitol production results in an increase of B<br />

translocation from mature leaves to sink tissues as well as<br />

tolerance to B deficiency. In plant species that do not produce<br />

significant amounts of sugar alcohols, B is thought to be phloem<br />

immobile, or only slightly mobile, <strong>and</strong> its distribution in shoots<br />

seems primarily to follow the xylem transpiration stream (Oertli<br />

1993; Bolaños et al. 2004; Lehto et al. 2004; Takano et al.<br />

2008).<br />

<strong>Vascular</strong> loading <strong>and</strong> unloading of cationic<br />

micronutrients<br />

Yellow Stripe-Like (YSL) proteins play important roles in the<br />

short- <strong>and</strong> long-distance transport of microelements <strong>and</strong> their<br />

delivery to sink tissues (Curie et al. 2009). Members of the YSL<br />

family, AtYSL1, AtYSL2, AtYSL3, OsYSL2 <strong>and</strong> OsYSL18, are<br />

expressed in vascular tissues (Table 3) <strong>and</strong> may have a role<br />

in the lateral movement of Fe within the veins <strong>and</strong> in phloem<br />

transport (DiDonato et al. 2004; Koike et al. 2004; Le Jean et al.<br />

2005; Schaaf et al. 2005; Aoyama et al. 2009). <strong>The</strong> rice OsYSL2<br />

can transport Fe(II)-NA <strong>and</strong> Mn(II)-NA to an equal extent (Koike<br />

et al. 2004). OsYSL18 transports Fe(III)-deoxymugineic acid<br />

(Aoyama et al. 2009), whereas there are contradictory reports<br />

concerning the ability of AtYSL2 to transport Fe(II)-NA <strong>and</strong><br />

Cu(II)-NA (DiDonato et al. 2004; Schaaf et al. 2005). AtYSL1<br />

seems to play a role in Fe(II)-NA translocation to seeds (Le<br />

Jean et al. 2005). A study on the Arabidopsis double mutant<br />

ysl1ysl3 reported reduced accumulation of Fe, Cu <strong>and</strong> Zn in<br />

seeds, consistent with involvement of the YSL1 <strong>and</strong> YSL3<br />

transporters in remobilization from leaves (Waters et al. 2006).<br />

<strong>The</strong>re is also evidence for a role of YSLs in the Zn <strong>and</strong><br />

Ni hyperaccumulation of Thlaspi caerulescens, especially for<br />

TcYSL3 <strong>and</strong> TcYSL7, which are highly expressed around<br />

vascular tissues particularly in shoots when compared with<br />

their A. thaliana orthologs (Gendre et al. 2007). TcYSL3 is an<br />

Fe(II)-NA <strong>and</strong> Ni(II)-NA influx transporter that is suggested to<br />

facilitate the movement of these metal-NA complexes from the<br />

xylem into leaf cells.<br />

A number of other transporters involved in vascular loading<br />

<strong>and</strong> unloading of microelements have also been identified<br />

(Figure 26, Table 3). <strong>The</strong> Arabidopsis OPT3 transporter<br />

(OligoPeptide Transporter) appears to be essential for embryo<br />

development (Stacey et al. 2008). This protein transports Mn,<br />

<strong>and</strong> its expression in the vascular tissue suggests a role<br />

in Mn long-distance transport. Although yeast studies have<br />

suggested that it can also transport Cu, OPT3 does not appear<br />

to play a role in Zn or Cu loading, as seeds of opt3-2 plants<br />

actually accumulate increased levels of these two metals<br />

(Stacey et al. 2008). <strong>The</strong> opt3-2 mutant also has reduced<br />

Fe concentrations in its seeds as well as impaired seedling<br />

growth under Fe-deficient conditions, thus suggesting a role in<br />

Fe loading into the seed (<strong>and</strong> perhaps even phloem-mediated<br />

redistribution).<br />

Another Fe efflux transporter, IREG1/FPN1 (Iron Regulated1/Ferroportin1),<br />

is considered to function in Fe loading<br />

into the xylem within the roots (Morrissey et al. 2009). Loss<br />

of FPN1 function results in chlorosis, <strong>and</strong> FPN1-GUS plants<br />

show staining at the plasma membrane of the root vascular<br />

system. However, yeast complementation studies using FPN1<br />

have failed, <strong>and</strong> information on the chemical form of Fe


Table 3. Transporters involved in vascular loading <strong>and</strong> unloading of cationic microelements<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 355<br />

Family Gene Microelement Function Reference<br />

YSL AtYSL1 Fe-NA Seed loading, phloem loading Le Jean et al. 2005<br />

AtYSL2 Fe, Zn Xylem loading/unloading DiDonato et al. 2004<br />

Schaaf et al. 2005<br />

AtYSL3 Fe, Cu, Zn Phloem loading Waters et al. 2006<br />

OsYSL2 Fe-NA Mn-NA Phloem loading/unloading to grains Koike et al. 2004<br />

OsYSL18 Fe(III)-DMA Phloem loading/unloading<br />

to reproductive organs<br />

Aoyama et al. 2009<br />

TcYSL3 Fe-NA, Ni-NA Xylem unloading Gendre et al. 2007<br />

TcTSL7 Ni-NA Xylem unloading Gendre et al. 2007<br />

OPT AtOPT3 Mn Fe Seed loading Stacey et al. 2008<br />

Ferroportin IREG1/FPN1 Fe Xylem loading Morrissey et al. 2009<br />

MATE FRD3 Citrate Xylem loading Durrett et al. 2007<br />

Yokosho et al. 2009<br />

Yokosho et al. 2010<br />

P-type ATPases AtHMA5 Cu Xylem loading André-Colás et al. 2006<br />

Kobayashi et al. 2008<br />

AtHMA2/4 Zn Xylem loading Hussain et al. 2004<br />

Mills et al. 2005<br />

Wong <strong>and</strong> Cobbett 2009<br />

Verret et al. 2004, 2005<br />

AhHMA4 Zn Xylem loading Hanikenne et al. 2008<br />

transported by FPN1 has yet to be established (Morrissey et al.<br />

2009).<br />

<strong>The</strong> Arabidopsis P-type ATPases, AtHMA5 <strong>and</strong> AtHMA2/4,<br />

have been implicated in Cu <strong>and</strong> Zn efflux, respectively, into<br />

the xylem at the root level, for long-distance transport to<br />

the shoots (Hussain et al. 2004; Mills et al. 2005; Andrés-<br />

Colas et al. 2006). Consistent with this model, both hma5<br />

<strong>and</strong> hma2hma4 loss-of-function mutants accumulate increased<br />

levels of the corresponding metal within the root, <strong>and</strong> show<br />

lower levels in their shoots (Hanikenne et al. 2008; Wong <strong>and</strong><br />

Cobbett 2009). HMA5 is predominantly expressed in the root<br />

<strong>and</strong> is specifically induced by excess Cu. Mutants of HMA5<br />

overaccumulate Cu in the root, suggesting a compromised<br />

efflux system. Further evidence in support of the role of HMA5<br />

in xylem transport of Cu from the roots to the shoots comes from<br />

a study of natural variation in Cu tolerance among Arabidopsis<br />

accessions, which identified HMA5 as a major QTL associated<br />

with Cu translocation capacity <strong>and</strong> sensitivity (Kobayashi et al.<br />

2008). HMA2 <strong>and</strong> 4 are present in the plasma membrane of<br />

root <strong>and</strong> shoot vascular tissues (Mills et al. 2003; Hussain<br />

et al. 2004; Verret et al. 2004; Mills et al. 2005; Verret<br />

et al. 2005; Williams <strong>and</strong> Mills 2005; Sinclair et al. 2007;<br />

Blindauer <strong>and</strong> Schmid 2010). In addition, functional analysis<br />

of HMA4 in A. halleri <strong>and</strong> A. thaliana showed that silencing<br />

of AhHMA4, by RNA interference, completely suppressed Zn<br />

hyperaccumulation. <strong>The</strong>se studies provided a clear demonstration<br />

that HMA4 plays a key role in xylem loading <strong>and</strong>,<br />

consequently, in root-to-shoot transport of Zn (Hanikenne et al.<br />

2008).<br />

Organic acids may also have a role in xylem Fe loading.<br />

Citrate has been described as an Fe(III) chelator in the xylem<br />

sap (Rellán-Álvarez et al. 2010) <strong>and</strong> FRD3 (Ferric Reductase<br />

Defective), a transporter of the MATE family, is localized to<br />

the plasma membrane of the pericycle <strong>and</strong> vascular cylinder.<br />

FRD3 proteins facilitate citrate efflux into the xylem of the root<br />

vasculature <strong>and</strong> have been described in Arabidopsis (Durrett<br />

et al. 2007), rice (Yokosho et al. 2009) <strong>and</strong> rye (Yokosho<br />

et al. 2010). Mutant frd3 plants are chlorotic, show reduced<br />

citrate <strong>and</strong> Fe concentrations in the xylem <strong>and</strong> the shoot,<br />

accumulate Fe in the root, <strong>and</strong> exhibit constitutive expression<br />

of the Fe uptake components, thus suggesting that FRD3 is<br />

necessary for efficient Fe transport to the shoot through the<br />

transpiration stream. Also, independent Fe-citrate <strong>and</strong> Fe-NA<br />

xylem loading systems may complement each other, as in the<br />

frd3 mutant, the nicotianamine synthase NAS4 gene is induced,<br />

<strong>and</strong> the double mutant nas4x-2/frd3 shows impaired growth<br />

<strong>and</strong> low Fe levels in the shoot (Schuler et al. 2010). FRD3<br />

is constitutively expressed in the hyperaccumulators A. halleri<br />

<strong>and</strong> N. caerulescens compared to A. thaliana <strong>and</strong> N. arvensis,<br />

<strong>and</strong> may also play a role in Zn transport (Talke et al. 2006;<br />

van de Mortel et al. 2006). However, this overexpression may<br />

be related to an altered Fe homeostasis leading to high Zn<br />

concentrations in the hyperaccumulators (Roschzttardtz et al.<br />

2011).


356 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

<strong>Vascular</strong> loading <strong>and</strong> unloading of anionic<br />

micronutrients<br />

Xylem loading of B is mediated by BOR1 (Takano et al. 2001;<br />

Miwa <strong>and</strong> Fujiwara 2010). BOR1 is an anion efflux system that<br />

is strongly expressed in the root pericycle cells surrounding the<br />

xylem vessels. <strong>The</strong> bor1-1 mutant is defective in xylem loading<br />

of B (Takano et al. 2002). <strong>The</strong> specific chemical form of the<br />

substrate for BOR1 remains unknown, but electrophysiological<br />

analyses in the human homolog, NaBC1, suggest borate anion<br />

as the likely c<strong>and</strong>idate (Park et al. 2004). <strong>The</strong>re are six BOR1<br />

paralogs in the A. thaliana genome which may also have roles<br />

in xylem-phloem loading <strong>and</strong> unloading (Miwa <strong>and</strong> Fujiwara<br />

2010). A second B transporter, NIP6,1, is a channel protein<br />

required for proper distribution of boric acid, particularly to<br />

young developing shoot tissues (Tanaka et al. 2008). This<br />

transporter is predominantly expressed in nodal regions of<br />

shoots, especially the phloem region of vascular tissues, where<br />

it is likely involved in xylem-phloem transfer of boric acid.<br />

<strong>The</strong> mechanisms for Mo loading <strong>and</strong> unloading in the vascular<br />

tissues remain to be elucidated. To date, only one Mo<br />

transporter from Arabidopsis, MOT1, has been identified in<br />

plants. MOT1 is a high-affinity molybdate transporter localized<br />

to the plasma membrane or mitochondrial membranes, <strong>and</strong><br />

plays an important role in efficient Mo uptake from soils <strong>and</strong><br />

accumulation within the plant (Tomatsu et al. 2007; Baxter<br />

et al. 2008). MOT1 belongs to the family of sulphate transporters,<br />

SULTR (Hawkesford 2003), which has 14 members<br />

in A. thaliana. It is tempting to speculate that some of these<br />

transporters may be involved in vascular tissue loading or<br />

unloading.<br />

<strong>System</strong>ic Signaling: Pathogen<br />

Resistance<br />

Like all living organisms, plants have to constantly resist<br />

pathogenic microbes. <strong>The</strong> absence of a circulatory vascular<br />

system <strong>and</strong> their sessile nature can pose particular problems.<br />

<strong>Plant</strong>s have therefore evolved unique defense mechanisms to<br />

ensure survival. <strong>The</strong> multiple modes of plant defense include<br />

both passive <strong>and</strong> active mechanisms that provide defense<br />

against a wide variety of pathogens. Active defense includes<br />

the production of antimicrobial compounds, cell wall reinforcement<br />

via the synthesis of lignin <strong>and</strong> callose, <strong>and</strong> the specific induction<br />

of elaborate defense signaling pathways. <strong>The</strong>se include<br />

species level (non-host) resistance, race-specific resistance<br />

expressed both locally <strong>and</strong> systemically, <strong>and</strong> basal resistance.<br />

Race-specific resistance is induced when strain-specific<br />

avirulent (Avr) proteins from the pathogen associate directly/indirectly<br />

with cognate plant resistance (R) proteins (reviewed<br />

in Jones <strong>and</strong> Dangl 2006; Caplan et al. 2008). Induction<br />

of R-mediated signaling is often accompanied by the onset of<br />

a hypersensitive response (HR), a form of PCD resulting in<br />

necrotic lesions, at the site of pathogen entry (Dangl et al.<br />

1996). HR is one of the first visible manifestations of pathogeninduced<br />

host defenses, <strong>and</strong> is thought to help confine the<br />

pathogen to the dead cells. R-mediated signaling is also often<br />

accompanied by the induction of a robust form of resistance<br />

against secondary pathogens in the systemic parts of the<br />

plants, termed systemic acquired resistance (SAR) (Durrant<br />

<strong>and</strong> Dong 2004; Vlot et al. 2008; Spoel <strong>and</strong> Dong 2012).<br />

Identified as a form of plant immunity nearly 100 years ago,<br />

SAR is a highly desirable form of resistance that protects<br />

against a broad-spectrum of pathogens. SAR involves the generation<br />

of a mobile signal at the site of primary infection, which<br />

moves to <strong>and</strong> arms distal portions of a plant against subsequent<br />

secondary infections (Figure 27). <strong>The</strong> identification of this signal<br />

could greatly facilitate the use of SAR in protecting agriculturally<br />

important plants against a wide range of pathogens. Because of<br />

its unique mechanistic properties <strong>and</strong> its exciting potential applications<br />

in developing sustainable crop protection strategies,<br />

SAR has been one of the most intensely researched areas<br />

of plant biology. <strong>The</strong> last decade has witnessed considerable<br />

progress, <strong>and</strong> a number of signals contributing to SAR have<br />

been isolated <strong>and</strong> characterized. Despite concerted efforts to<br />

harness this mode of plant immunity, the plant defense field<br />

lacks a consensus regarding the identity of the SAR signal,<br />

whether this signal constitutes multiple molecular components,<br />

<strong>and</strong> how these component(s) might coordinate the systemic<br />

induction of broad-spectrum resistance.<br />

Among the signals contributing to SAR are salicylic acid (SA)<br />

<strong>and</strong> several components that feed into the SA pathway, including<br />

the methylated derivative of SA (MeSA; Park et al. 2007),<br />

the diterpenoid dehdryoabietinal (DA; Chaturvedi et al. 2012),<br />

the nine carbon (C9) dicarboxylic acid azelaic acid (AA; Jung<br />

et al. 2009), auxin (Truman et al. 2010), the phosphorylated<br />

sugar glycerol-3-phosphate (G3P; Ch<strong>and</strong>a et al. 2011; M<strong>and</strong>al<br />

et al. 2011), <strong>and</strong> two lipid transfer proteins (LTPs), Defective<br />

in Induced Resistance (DIR1; Maldonado et al. 2002) <strong>and</strong> AA<br />

insensitive (AZI1; Jung et al. 2009). Jasmonic acid (JA) has also<br />

been suggested to participate in SAR (Truman et al. 2007), but<br />

its precise role remains contentious (Attaran et al. 2009). <strong>The</strong><br />

diverse chemical natures of the SAR-inducing molecules have<br />

led to the growing belief that SAR might involve the interplay of<br />

multiple diverse <strong>and</strong> independent signals. In this final section<br />

of the review, we will evaluate the role of SA <strong>and</strong> the recently<br />

identified mobile inducers of SAR.<br />

SA <strong>and</strong> SAR<br />

SA is a central <strong>and</strong> critical component of SAR. <strong>The</strong> biosynthesis<br />

of SA occurs via the shikimic acid pathway, which<br />

bifurcates into two branches after the biosynthesis of chorismic<br />

acid. In one branch, chorismic acid is converted to SA via


Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 357<br />

Figure 27. A simplified model summarizing the known mobile inducers of systemic acquired resistance (SAR).<br />

Pathogen infection induces an increase in the levels of glycerol-3-phosphate (G3P), azelaic acid (AA), dehdryoabietinal (DA), salicylic acid<br />

(SA) <strong>and</strong> methyl SA (MeSA). Some of the SA is converted to MeSA, whereas G3P is converted to an unknown derivative (indicated by<br />

asterisk). Owing to its volatile nature, a significant proportion of the MeSA is thought to escape by emissions. Although both SA <strong>and</strong> MeSA<br />

are phloem mobile, SA likely functions downstream of mobile signal generation. This is based on grafting experiments that indicate SA is not<br />

the SAR signal, yet basal SA is essential for G3P-, DA-, <strong>and</strong> AA-mediated SAR. DA induces SA accumulation in infected <strong>and</strong> distal tissues,<br />

whereas AA primes for SA biosynthesis in response to secondary pathogen stimulus. Neither G3P nor AA induces SA accumulation. G3P-,<br />

AA-, <strong>and</strong> DA-mediated SAR require the endoplasmic reticulum-localized lipid transfer-like protein, DIR1. <strong>System</strong>ic movement of G3P <strong>and</strong><br />

DIR1 is mutually inter-dependent. Upon transport, complex(es) comprising DIR1 <strong>and</strong> the G3P-derivative induce de novo G3P biosynthesis<br />

in the distal tissues.<br />

phenylalanine <strong>and</strong> cinnamic acid intermediates, <strong>and</strong> in the other<br />

branch chorismic acid is converted to SA via isochorismic acid.<br />

Two well characterized enzymes in these branches include<br />

PHENYLALANINE AMMONIA LYASE (PAL), which converts<br />

phenylalanine to cinnamic acid, <strong>and</strong> ISOCHORISMATE SYN-<br />

THASE (ICS), which catalyzes the conversion of chorismic<br />

acid to isochorismic acid (Wildermuth et al. 2001; Strawn et al.<br />

2007).<br />

Transcriptional profiling has shown that expression of hundreds<br />

of genes is altered during the development of SAR<br />

(Schenk et al. 2000; Wang et al. 2006; Truman et al. 2007;<br />

Ch<strong>and</strong>a et al. 2011). This is likely to have wide-ranging effects,<br />

including strengthening of the cell wall <strong>and</strong> production of<br />

reactive oxygen species <strong>and</strong> SA. A hallmark of plants that have<br />

manifested SAR is the induction of pathogenesis-related (PR)<br />

proteins (Carr et al. 1987; Loon et al. 1987; Ward et al. 1991).<br />

<strong>The</strong>se observations <strong>and</strong> the fact that exogenous SA induces<br />

PR expression led to the suggestion that SA was involved in<br />

SAR signaling.<br />

Exogenous application of SA or its synthetic functional<br />

analogues such as BTH (1,2,3-benzothiadiazole-7-carbothioic<br />

acid, S-methyl ester) also induce generalized defense against a<br />

variety of pathogens. Evidence supporting a role for SA in plant<br />

defense came from analysis of transgenic plants expressing<br />

the bacterial gene encoding salicylate hydroxylase, an enzyme<br />

that catalyzes conversion of SA to catechol. <strong>The</strong>se transgenic<br />

plants were unable to accumulate free SA, showed compromised<br />

defense, <strong>and</strong> were unable to induce SAR (Gaffney et al.<br />

1993; Friedrich et al. 1995; Lawton et al. 1995). <strong>The</strong> fact that<br />

pathogen inoculation induces SA accumulation in both local<br />

<strong>and</strong> distal uninoculated tissues led to the hypothesis that SA<br />

might well be the phloem-mobile signal (Vernooij et al. 1994;


358 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Shulaev et al. 1995). However, the timeframe of SAR signal<br />

movement precedes that of SA accumulation in the distal<br />

tissues. Moreover, SA-deficient rootstocks of plants expressing<br />

SA hydroxylase or plants suppressed in PAL expression are<br />

capable of activating SAR in the leaves of WT scions. <strong>The</strong>se<br />

data argued against a role for SA as the phloem-mobile SAR<br />

signal (Vernooij et al. 1994; Pallas et al. 1996). Regardless,<br />

SA is required for the proper induction of SAR, <strong>and</strong> mutations<br />

in either the ICS or PAL pathway are sufficient to compromise<br />

SAR (Vernooij et al. 1994; Pallas et al. 1996; Wildermuth et al.<br />

2001).<br />

Radiolabel feeding experiments suggest that over 50% of<br />

SA in distal leaves is transported from the inoculated leaves<br />

with the remainder being synthesized de novo (Shulaev et al.<br />

1995; Molders et al. 1996). Whether the induced synthesis of<br />

SA in the distal tissues is required for SAR remains unclear.<br />

A marginal difference (∼10 ng/g FW) in SA levels between<br />

SAR-competent WT versus SAR-compromised SA-deficient<br />

scions grafted onto SA-deficient rootstocks suggests that an<br />

increase in SA accumulation may not be a prerequisite for the<br />

normal induction of SAR (Vernooij et al. 1994; Ch<strong>and</strong>a et al.<br />

2011).<br />

In general, most of the endogenous SA is metabolized<br />

to a glucose conjugate, SA 2-O-β-D-glucose (SAG) or SA<br />

glucose ester (SGE), <strong>and</strong> this reaction is catalyzed by SA<br />

glucosyltransferases (Enyedi et al. 1992; Edwards 1994; Lee<br />

<strong>and</strong> Raskin 1998; Lee <strong>and</strong> Raskin 1999; Dean <strong>and</strong> Delaney<br />

2008; Song et al. 2008). Other derivates of SA include its<br />

methylated ester <strong>and</strong> methyl SA. (MeSA), <strong>and</strong> hydroxylated<br />

form gentisic acid (GA), both of which are also present as<br />

glucose conjugates. Exogenous GA induces a specific set of<br />

PR proteins in tomato that are not induced by SA, suggesting<br />

that SA <strong>and</strong> GA differ in their mode of action (Bellés et al.<br />

1999).<br />

Unlike SA, MeSA is biologically inactive <strong>and</strong> only functions<br />

when converted back to SA. MeSA is a well-characterized<br />

volatile organic compound that can function as an airborne<br />

defense signal <strong>and</strong> can mediate plant-plant communication<br />

(Shulaev et al. 1997; Koo et al. 2007). Conversion of SA to<br />

MeSA is mediated by SA methyltransferases (SAMT), also<br />

designated BA (benzoic acid)-/SA-MT because it can utilize<br />

either SA or BA as substrates (Chen et al. 2003; Effmert et al.<br />

2005; Koo et al. 2007). Overexpression of BSMT leads to the<br />

depletion of endogenous SA <strong>and</strong> SAG, as most of the available<br />

SA is converted to MeSA (Koo et al. 2007). This in turn is<br />

associated with increased susceptibility to bacterial <strong>and</strong> fungal<br />

pathogens, suggesting that levels of free SA, but not MeSA,<br />

are critical for plant immunity. Likewise, overexpression of the<br />

Arabidopsis SA glucosyltransferase (AtSGT1) also results<br />

in the depletion of SA <strong>and</strong> an increase in MeSA levels,<br />

which again correlates with increased susceptibility to bacterial<br />

pathogens (Song et al. 2008).<br />

MeSA accumulates in the phloem following induction of SAR,<br />

<strong>and</strong> this requires SAMT activity. Upon translocation to the distal<br />

tissues, MeSA is converted back to SA via MeSA esterase<br />

(Figure 27). Most of the MeSA accumulating in response to<br />

pathogen inoculation was shown to escape by volatile emissions<br />

(Attaran et al. 2009). Furthermore, Arabidopsis BSMT<br />

mutant plants do not accumulate MeSA, but remain SAR<br />

competent. This discrepancy was attributed to the dependency<br />

of MeSA-derived signaling on light (Liu et al. 2011), which is<br />

well-known to play an important role in plant defense (Karpinski<br />

et al. 2003; Roberts <strong>and</strong> Park 2006). Notably, the phloem<br />

translocation time of the SAR signal to distal tissues precedes<br />

the time of MeSA requirement; i.e., 48 h <strong>and</strong> 72 h post primary<br />

infection, respectively (Park et al. 2009; Ch<strong>and</strong>a et al. 2011;<br />

Chaturvedi et al. 2012). This suggests that MeSA is unlikely<br />

to be the primary mobile signal, <strong>and</strong> possibly might act as a<br />

downstream contributor to SAR.<br />

Recent studies also suggest that defective SAR in dir1 plants<br />

is associated with increased expression of BSMT1, which correlates<br />

with increased accumulation of MeSA <strong>and</strong> a reduction<br />

in SA <strong>and</strong> SAG levels (Liu et al. 2011). However, this is in<br />

contrast with two other independent studies that showed normal<br />

SA levels in pathogen inoculated dir1 plants (Maldonado et al.<br />

2002; Chaturvedi et al. 2012). Some possibilities that might<br />

account for these discrepancies are disparate regulation of<br />

BSMT1 expression <strong>and</strong> the associated changes in MeSA <strong>and</strong><br />

SA levels in different ecotypic backgrounds, <strong>and</strong>/or plant growth<br />

conditions, such as light, humidity, temperature, <strong>and</strong> wind.<br />

For example, light intensities could affect SA levels/defense<br />

responses since photoreceptors are well known to regulate<br />

both SA- <strong>and</strong> R-mediated signaling (Genoud et al. 2002; Jeong<br />

et al. 2010).<br />

Components that affect SAR by regulating SA levels<br />

Many proteins known to mediate SA-derived signaling have<br />

been identified as contributors to SAR. <strong>The</strong>se include proteins<br />

involved in SA biosynthesis (including ICS <strong>and</strong> PAL), transport,<br />

<strong>and</strong>/or SA-dependent R-mediated signaling (ENHANCED<br />

DISEASE SUSCEPTIBILITY 1 (EDS1), EDS5, PHYTOALEXIN<br />

DEFICIENT 4 (PAD4), <strong>and</strong> SENESCENCE-ASSOCIATED<br />

gene 101 (SAG101)). <strong>The</strong> Arabidopsis EDS5 (also called<br />

SA INDUCTION-DEFICIENT 1) encodes a plastid-localized<br />

protein that shows homology to the bacterial multidrug <strong>and</strong><br />

toxin extrusion transporter (MATE) proteins. EDS5 is required<br />

for the accumulation of SA after pathogen inoculation (Nawrath<br />

et al. 2002; Ishihara et al. 2008) <strong>and</strong>, consequently, a mutation<br />

in EDS5 causes enhanced susceptibility against oomycete,<br />

bacterial, <strong>and</strong> viral pathogens (Rogers <strong>and</strong> Ausubel 1997;<br />

Nawrath et al. 2002; Ch<strong>and</strong>ra-Shekara et al. 2004). Mutations<br />

in ICS1 <strong>and</strong> EDS5 lead to similar phenotypes (Venugopal<br />

et al. 2009), suggesting that EDS5 might be involved in


the transport of SA <strong>and</strong>/or its precursors across the plastid<br />

membrane.<br />

Currently, EDS5 is thought to act downstream of three other<br />

signaling components, EDS1 <strong>and</strong> PAD4, <strong>and</strong> NON-RACE<br />

SPECIFIC DISEASE RESISTANCE 1 (NDR1), which are<br />

required for basal R protein-mediated signaling <strong>and</strong> the SAR<br />

response (Glazebrook <strong>and</strong> Ausubel 1994; Century et al. 1995;<br />

Glazebrook et al. 1996; Parker et al. 1996; Century et al. 1997;<br />

Aarts et al. 1998; Zhou et al. 1998; Shapiro <strong>and</strong> Zhang 2001;<br />

Liu et al. 2002; Coppinger et al. 2004; Hu et al. 2005; Truman<br />

et al. 2007). Some of the Arabidopsis ecotypes express two<br />

functionally redundant isoforms of EDS1 which interact with<br />

each other as well as with the structurally similar PAD4 <strong>and</strong><br />

SAG101 proteins (Feys et al. 2001; He <strong>and</strong> Gan 2002; Feys<br />

et al. 2005; García et al. 2010; Zhu et al. 2011). EDS1, PAD4,<br />

<strong>and</strong> SAG101 proteins also exist as a ternary complex (Zhu et al.<br />

2011).<br />

EDS1 interacts with several R proteins, suggesting that<br />

EDS1 <strong>and</strong>, by extension, PAD4 <strong>and</strong> SAG101, likely act at<br />

the R protein level (Bhattacharjee et al. 2011; Heidrich et al.<br />

2011; Zhu et al. 2011). Notably, mutations in EDS1, PAD4,<br />

<strong>and</strong> SAG101 lead to overlapping as well as independent<br />

phenotypes, suggesting that these proteins might function as<br />

complex(s) as well as individual proteins. Mutations in EDS1<br />

or PAD4 attenuate the expression of FLAVIN-DEPENDENT<br />

MONOOXYGENASE 1 (FMO1), which is required for SA accumulation<br />

in the distal tissues <strong>and</strong>, thereby, SAR (Mishina<br />

<strong>and</strong> Zeier 2006). <strong>The</strong> pathogen-induced SA levels are also<br />

regulated by AGD2-LIKE DEFENSE 1 (ALD1), which is induced<br />

in distal tissues after avirulent inoculation in a PAD4-dependent<br />

manner (Song et al. 2004a). <strong>The</strong> ALD1 encoded protein shows<br />

aminotransferase activity in vitro, suggesting that an amino<br />

acid-derived signal might participate in the regulation of SA<br />

levels <strong>and</strong>, thereby, SAR (Song et al. 2004b).<br />

SA signaling components that affect SAR<br />

<strong>The</strong> NON-EXPRESSOR OF PR1 (NPR1), an ankyrin repeat<br />

containing protein also called NON-INDUCIBLE IMMUNITY 1<br />

(NIM1; Delaney et al. 1995; Ryals et al. 1997) or SA INSEN-<br />

SITIVE 1) (SAI1; Shah et al. 1997), is considered a central<br />

regulator of SA-derived signaling. A mutation in Arabidopsis<br />

NPR1 abolishes SAR, suggesting that it is a positive regulator<br />

of SAR (Cao et al. 1994; Cao et al. 1997). Besides SA signaling<br />

<strong>and</strong> SAR, NPR1 also functions in induced systemic resistance<br />

(ISR) <strong>and</strong>, possibly, in regulating cross-talk between the SA<br />

<strong>and</strong> jasmonic acid (JA) pathways (van Wees et al. 2000;<br />

Kunkel <strong>and</strong> Brooks 2002; Lavicoli et al. 2003; Spoel <strong>and</strong> Dong<br />

2008). SA <strong>and</strong> NPR1 negatively affect the symbiotic interaction<br />

between Medicago <strong>and</strong> Rhizobium (Peleg-Grossman et al.<br />

2009), suggesting that SA <strong>and</strong> NPR1 are essential components<br />

of multiple signaling pathway(s).<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 359<br />

In the absence of SA, NPR1 exists as an oligomer via intermolecular<br />

disulfide bonding, <strong>and</strong> remains in the cytoplasm (Mou<br />

et al. 2003). Reducing conditions triggered upon activation of<br />

defense responses <strong>and</strong> the accumulation of SA result in the<br />

dissociation of the NPR1 oligomer into monomers, which are<br />

transported into the nucleus (Mou et al. 2003; Tada et al.<br />

2008). Within the nucleus, these NPR1 monomers interact<br />

with members of the TGACG motif binding transcription factors<br />

belonging to the basic leucine zipper (bZIP) protein family<br />

(Zhang et al. 1999; Després et al. 2000; Niggeweg et al.<br />

2000; Zhou et al. 2000; Chern et al. 2001; Fan <strong>and</strong> Dong<br />

2002; Kim <strong>and</strong> Delaney 2002). SA also induces the reduction<br />

of the disulfide bridges in TGA proteins, thereby allowing the<br />

proteins to interact with NPR1 with subsequent activation of<br />

gene expression (Després et al. 2003).<br />

Genetic evidence supporting a role for TGA factors in SAR<br />

was provided by the analysis of the tga2 tga5 tga6 triple mutant,<br />

which was unable to induce PR gene expression in response<br />

to SA <strong>and</strong> was defective in the onset of SAR (Zhang et al.<br />

2003). Recent studies have also shown that, like NPR1, TGA1<br />

also undergoes S-nitrosylation, which promotes the nuclear<br />

translocation of NPR1 <strong>and</strong> increases the DNA binding activity<br />

of TGA1 (Tada et al. 2008; Lindermayr et al. 2010).<br />

<strong>The</strong> monomerization of NPR1 also appears to be important<br />

for the activation of the NPR1 regulated members of the<br />

WRKY transcription factor family (Mou et al. 2003; Wang<br />

et al. 2006). In addition, NPR1 controls the expression of the<br />

protein secretory pathway genes in a TGA2-, TGA5- <strong>and</strong> TGA6independent<br />

manner (Wang et al. 2006). <strong>The</strong> nuclear NPR1<br />

is phosphorylated <strong>and</strong> recycled in a proteasome-dependent<br />

manner (Spoel et al. 2009). This turnover is required for<br />

the establishment of SAR. <strong>The</strong> Arabidopsis genome contains<br />

five paralogs of NPR1 (Liu et al. 2005). Like NPR1,<br />

NPR3 <strong>and</strong> NPR4 also interact with TGA proteins (Zhang<br />

et al. 2006). <strong>The</strong> npr3 npr4 mutant plants accumulate higher<br />

levels of NPR1 <strong>and</strong>, consequently, are unable to induce<br />

SAR.<br />

In a recent study, NPR3 <strong>and</strong> NPR4 were shown to bind SA<br />

<strong>and</strong> to function as adaptors of the Cullin 3 ubiquitin E3 ligase<br />

to mediate NPR1 degradation in an SA-dependent manner (Fu<br />

et al. 2012). NPR3 <strong>and</strong> NPR4 are neither the first nor the only<br />

known SA binding proteins. However, much like most of the<br />

plant hormone receptors, NPR3 <strong>and</strong> NPR4 are the only known<br />

proteins which regulate the proteasome-dependent recycling<br />

of a master regulator of the SA-signaling pathway. For this<br />

reason, these proteins have been suggested to serve as the<br />

long-sought-after SA receptors. In yet another study, NPR1<br />

was also proposed to function as a SA receptor (Wu et al.<br />

2012). Like NPR3/NPR4, NPR1 bound SA <strong>and</strong> the kinetics of<br />

this binding were similar to those of other receptor-hormone<br />

interactions. Thus, SA might well bind to multiple NPRs <strong>and</strong><br />

differentially modulate their function(s).


360 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Other proteins that bind SA include the MeSA esterase<br />

SABP2 (Kumar <strong>and</strong> Klessig 2003). Binding of SA to SABP2<br />

inhibits its esterase activity, resulting in the accumulation of<br />

MeSA. In addition, SA binds catalase (Chen et al. 1993)<br />

<strong>and</strong> carbonic anhydrase (Slaymaker et al. 2002), <strong>and</strong> inhibits<br />

the activities of the heme-iron-containing enzymes catalase,<br />

ascorbate peroxidase, <strong>and</strong> aconitase (Durner <strong>and</strong> Klessig<br />

1995). <strong>The</strong> ability of SA to chelate iron has been suggested<br />

as one of the mechanisms for SA-mediated inhibition of these<br />

enzymes (Rueffer et al. 1995). Tobacco plants silenced for<br />

carbonic anhydrase or aconitase show increased pathogen<br />

susceptibility, suggesting that these proteins are required for<br />

plant defense (Slaymaker et al. 2002; Moeder et al. 2007).<br />

Whereas NPR1 is an essential regulator of SA-derived signaling,<br />

many protein-induced pathways are known to activate<br />

SA-signaling in an NPR1-independent manner (Kachroo et al.<br />

2000; Takahashi et al. 2002; Raridan <strong>and</strong> Delaney 2002; Van<br />

der Biezen et al. 2002). Furthermore, a number of mutants have<br />

been isolated that induce defense SA signaling in an NPR1independent<br />

manner (Kachroo <strong>and</strong> Kachroo 2006). A screen<br />

for npr1 suppressors resulted in the identification of SNI1 (SUP-<br />

PRESSOR OF npr1, INDUCIBLE), a mutation which restores<br />

SAR in npr1 plants by de-repressing NPR1-dependent SA<br />

responsive genes (Li et al. 1999; Mosher et al. 2006). SNI1 has<br />

been suggested to regulate recombination rates through chromatin<br />

remodeling (Durrant et al. 2007). A subsequent screen<br />

for sni1 suppressors identified BRCA2 (BREAST CANCER)<br />

<strong>and</strong> RAD51D, which when mutated abolish the sni1-induced<br />

de-repression of NPR1-dependent gene expression (Durrant<br />

et al. 2007; Wang et al. 2010). SNI1 is therefore thought to<br />

act as a negative regulator that prevents recombination in<br />

the uninduced state. A role for SNI1, BRCA2 <strong>and</strong> RAD51D in<br />

recombination <strong>and</strong> defense suggests a possible link between<br />

these processes. Collectively, such findings support a key role<br />

for chromatin modification in the activation of plant defense <strong>and</strong><br />

SAR (March-Díaz et al. 2008; Walley et al. 2008; Dhawan et al.<br />

2009; Ma et al. 2011).<br />

Mobile inducers of SAR<br />

Recent advances in the SAR field have led to the identification<br />

of four mobile inducers of SAR, including MeSA, AA, DA<br />

<strong>and</strong> G3P. All of these inducers accumulate in the inoculated<br />

leaves after pathogen inoculation <strong>and</strong> translocate systemically<br />

(Figure 27). <strong>The</strong> role of MeSA, a methylated derivative of<br />

SA, was discussed above. <strong>The</strong> dicarboxylic acid AA <strong>and</strong> the<br />

diterpenoid DA induce SAR in an ICS1-, NPR1-, DIR1-, <strong>and</strong><br />

FMO1-dependent manner (Jung et al. 2009; Chaturvedi et al.<br />

2012). <strong>The</strong>ir common requirements for these components suggest<br />

that AA- <strong>and</strong> DA-mediated SAR may represent different<br />

branches of a common signaling pathway. Indeed, exogenous<br />

application of low concentrations of DA <strong>and</strong> AA, that do not<br />

activate SAR, do so when applied together. However, AA <strong>and</strong><br />

DA differ in their mechanism of SAR activation: DA increases<br />

SA levels in local <strong>and</strong> distal tissues, whereas AA primes for<br />

pathogen-induced biosynthesis of SA in the distal tissues. DA<br />

application also induces local accumulation of MeSA. Unlike<br />

DA, AA does not induce SA biosynthesis when applied by itself.<br />

This is intriguing, considering their common requirements for<br />

downstream factors. At present, the biosynthetic pathways for<br />

AA <strong>and</strong> DA <strong>and</strong> the biochemical basis of AA- <strong>and</strong> DA-induced<br />

SAR remain unclear. Furthermore, firm establishment of AA<br />

or DA as mobile SAR inducers awaits the demonstration that<br />

plants unable to synthesize these compounds are defective in<br />

SAR.<br />

G3P is a phosphorylated three-carbon sugar that serves<br />

as an obligatory component of glycolysis <strong>and</strong> glycerolipid<br />

biosynthesis. In the plant, G3P levels are regulated by enzymes<br />

directly/indirectly involved in G3P biosynthesis, as well as those<br />

involved in G3P catabolism. Recent results have demonstrated<br />

a role for G3P in R-mediated defense leading to SAR <strong>and</strong><br />

defense against the hemibiotrophic fungus Colletotrichum higginsianum<br />

(Ch<strong>and</strong>a et al. 2008). Arabidopsis plants containing<br />

the RPS2 gene rapidly accumulate G3P when infected with an<br />

avirulent (Avr) strain of the bacterial pathogen Pseudomonas<br />

syringae (avrRpt2); G3P levels peak within 6 h post-inoculation<br />

(Ch<strong>and</strong>a et al. 2011). Strikingly, accumulation of G3P in the<br />

infected <strong>and</strong> systemic tissues precedes the accumulation of<br />

other metabolites known to be essential for SAR (SA, JA).<br />

Mutants defective in G3P synthesis are compromised in<br />

SAR, <strong>and</strong> this defect can be restored by the exogenous<br />

application of G3P (Ch<strong>and</strong>a et al. 2011). Exogenous G3P<br />

also induces SAR in the absence of primary pathogen, albeit<br />

only in the presence of the LTP-like protein DIR1, which is a<br />

well-known positive regulator of SAR (Maldonado et al. 2002;<br />

Champigny et al. 2011; Ch<strong>and</strong>a et al. 2011; Liu et al. 2011;<br />

Chaturvedi et al. 2012). DIR1 is also required for AA- <strong>and</strong><br />

DA-mediated SAR, suggesting that DIR1 might be a common<br />

node for several SAR signals. Interestingly, G3P <strong>and</strong> DIR1<br />

are interdependent on each other for their translocation to the<br />

distal tissues. However, G3P does not interact directly with<br />

DIR1. Moreover, 14 C-G3P-feeding experiments have shown<br />

that G3P is translocated as a modified derivative during SAR.<br />

<strong>The</strong>se results suggest that DIR1 likely associates with a G3Pderivative<br />

<strong>and</strong>, upon translocation to the distal tissues this<br />

complex, then induces the de novo synthesis of G3P <strong>and</strong><br />

consequently SAR (Figure 27).<br />

This defense-related function of G3P is conserved because<br />

exogenous G3P can also induce SAR in soybean (Ch<strong>and</strong>a<br />

et al. 2011). Exogenous application of G3P on local leaves<br />

induces transcriptional reprogramming in the distal tissues,<br />

which among other changes leads to the induction of the gene<br />

encoding a SABP2-like protein <strong>and</strong> repression of BSMT1. Thus,<br />

it is possible that G3P-mediated signaling functions to prime


the system for SA biosynthesis in the presence of an invading<br />

pathogen. However, exogenous G3P alone is not associated<br />

with increased SA biosynthesis, in either local or distal leaves.<br />

In this regard, it is interesting that similar to G3P, AA does<br />

not induce the expression of the genes normally associated<br />

with SA signaling, or those induced in response to exogenous<br />

SA. Induced SA accumulation diverts carbon, nitrogen <strong>and</strong><br />

energy away from the plant’s primary metabolic pathways,<br />

which negatively impacts growth <strong>and</strong> development (Heil <strong>and</strong><br />

Baldwin 2002; Heidel et al. 2004). Thus, chemicals like AA <strong>and</strong><br />

G3P, which induce SAR without increasing SA levels, could be<br />

tremendously beneficial in improving crop resistance without<br />

affecting plant growth, development <strong>and</strong> ultimately yield.<br />

Fatty acids, lipids, cuticle <strong>and</strong> plant defense<br />

<strong>The</strong> primary role of G3P in plant metabolism is that of an<br />

obligatory precursor for glycerolipid biosynthesis. G3P enters<br />

lipid biosynthesis upon acylation with the fatty acid (FA) oleic<br />

acid (18:1) to form lyso-phosphatidic acid (lyso-PA), via the<br />

activity of the soluble plastidial G3P acyltransferase (GPAT).<br />

Genetically-based reductions in 18:1 levels induce constitutive<br />

defense signaling via the SA pathway (Kachroo et al. 2003,<br />

2004, 2005; Venugopal et al. 2009). Consequently, low 18:1containing<br />

plants exhibit enhanced resistance to bacterial <strong>and</strong><br />

oomycete pathogens (Shah et al. 2001; Kachroo et al. 2001).<br />

Low 18:1 levels also specifically induce the expression of<br />

several R genes, which in turn induces defense signaling.<br />

SA <strong>and</strong> EDS1 regulate this low 18:1-dependent induction of<br />

defense responses in a redundant manner (Ch<strong>and</strong>ra-Shekara<br />

et al. 2007; Venugopal et al. 2009; Xia et al. 2009). Interestingly,<br />

it has also been shown that 18:1 levels regulate the NOA1<br />

(NITRIC OXIDE ASSOCIATED) protein <strong>and</strong> thereby nitric oxide<br />

levels. Thus, the increased NO in low 18:1-containing plants<br />

is responsible for their altered defense related phenotypes<br />

(M<strong>and</strong>al et al. 2012).<br />

A number of cuticle-defective mutants are compromised<br />

in SAR (Xia et al. 2009, 2010). Whereas acp4 plants can<br />

generate the signal required for inducing SAR, they are unable<br />

to respond to it. This loss of ability to “perceive” the SAR signal<br />

appears to be related to the defective cuticle of acp4 plants,<br />

because mechanical abrasion of the cuticle disrupts SAR in<br />

WT plants. This SAR-disruptive effect of cuticle abrasion is<br />

highly specific because it hinders SAR only during the timeframe<br />

of mobile signal generation <strong>and</strong> translocation to distal<br />

tissues; it does not alter local defenses. <strong>The</strong>se observations<br />

suggest that cuticle-derived component(s) likely participate in<br />

processing/perception of the SAR signal(s). <strong>The</strong> requirement<br />

for the plant cuticle in SAR development, the presence of<br />

lipids <strong>and</strong> FAs in petiole exudates (Madey et al. 2002; Behmer<br />

et al. 2011; Guelette et al. 2012), <strong>and</strong> the derivatization of<br />

G3P (a glycerolipid precursor) into an unknown compound<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 361<br />

that translocates with the LTP DIR1, all suggest a role for<br />

lipids/FAs/sugars in SAR.<br />

Clearly, more work is required to dissect the relationships<br />

between these chemically diverse signals. For example, what<br />

factors govern the transport <strong>and</strong> movement of these signals<br />

through the vascular system, <strong>and</strong> their subsequent unloading<br />

into distal tissues? How are these signals processed at<br />

their systemic destinations? What reprogramming of metabolic<br />

events is required to activate defense <strong>and</strong> subsequently depress<br />

the tissues to the resting phase?<br />

Future Perspectives<br />

<strong>The</strong> emergence of the tracheophyte-based vascular system<br />

had major impacts on the evolution of terrestrial biology,<br />

in general, through its role in facilitating the development<br />

of plants with increased stature, photosynthetic output,<br />

<strong>and</strong> ability to colonize a greatly exp<strong>and</strong>ed range of environmental<br />

habitats. Significant insights have been gained<br />

concerning the genetic <strong>and</strong> hormonal networks that cooperate<br />

to orchestrate vascular development in the angiosperms,<br />

<strong>and</strong> progress is currently being made for the<br />

gymnosperms. However, much remains to be learned in<br />

terms of the early molecular events that led to the co-opting<br />

of pre-tracheophyte transcription factors <strong>and</strong> hormone signaling<br />

pathways, in order to establish the developmental<br />

programs that underlay the emergence of the tracheids<br />

as an effective/superior system for water conduction over<br />

the WCCs/hydroids. <strong>The</strong> same situation holds for the<br />

FCCs/leptoids to sieve cell/SE transition. Certainly, future<br />

application of genomic <strong>and</strong> molecular tools should offer<br />

important insights into the relationships between these pre<strong>and</strong><br />

post-tracheophyte/SE programs.<br />

Cost-effective, high-throughput sequencing technologies<br />

are opening the door to studies that integrate plant functional<br />

genomics with physiology <strong>and</strong> ecology. Such studies<br />

will likely provide important insights into novel strategies,<br />

achieved by different plant families, to refine the operational<br />

characteristics of their xylem/phloem transport systems to<br />

meet the challenges imposed by their specific ecological<br />

niches. Much of our current knowledge of vascular development<br />

is built upon studies conducted on “model” systems<br />

such as Arabidopsis. Although the general principles are<br />

likely to apply to most, if not all, advanced tracheophytes,<br />

many surprises are likely to be unearthed as research<br />

exp<strong>and</strong>s to cover plants with increased stature <strong>and</strong> concomitant<br />

challenges in terms of environmental inputs.<br />

Fundamental details are now established in terms of<br />

the mechanics underlying the thermodynamics of bulk flow<br />

though both the xylem <strong>and</strong> phloem. For the xylem, important


362 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

questions remain to be resolved, including the mechanism<br />

by which a single cavitation event can propagate within<br />

the tracheid/vessel system, the processes involved in refilling<br />

of embolized tracheary elements, especially when<br />

the transpiration stream is under tension, <strong>and</strong> the degree<br />

to which pit architecture between species contributes to<br />

ecological fitness. With regard to the phloem, one of the<br />

most fundamental questions that remains to be resolved<br />

relates to the mechanism(s) by which the plant integrates<br />

sink dem<strong>and</strong> with source capacity to optimize growth under<br />

prevailing environmental conditions. <strong>The</strong> phloem manifold<br />

hypothesis <strong>and</strong> the concept of delivery to various sink<br />

tissues being controlled by local PD properties warrants<br />

close attention.<br />

<strong>The</strong> role of the plant vascular system as a long-distance<br />

signaling system for integration of abiotic <strong>and</strong> biotic inputs<br />

is also firmly established. However, much remains to be<br />

learned concerning the nature of the xylem- <strong>and</strong> phloemmobile<br />

signals that function in nutrient homeostasis, environmental<br />

signaling to control stomatal density in emerging<br />

leaves, pathogen-host plant interactions, etc. In addition,<br />

the discovery that angiosperm phloem sap collected from<br />

the enucleate sieve tube system contains a broad spectrum<br />

of proteins <strong>and</strong> RNA species is consistent with the phloem<br />

functioning as a sophisticated communication system. A<br />

number of pioneering studies have demonstrated the role<br />

of protein <strong>and</strong> RNA as long-distance signaling agents.<br />

Future studies are required to both to exp<strong>and</strong> the number<br />

of proteins/RNA investigated as well as to focus on the<br />

molecular mechanisms involved in determining how these<br />

signaling agents are targeted to specific sink tissues.<br />

Commercial applications of knowledge gained on the<br />

development <strong>and</strong> functions of the plant vascular system<br />

are likely to be boundless. Access to methods to control<br />

source-sink relationships would have profound effects over<br />

yield potential <strong>and</strong> biomass production for the biofuels<br />

industry. Modifications to secondary xylem development<br />

will likely allow for engineering of wood that has unique<br />

properties for industrial applications. Engineering of novel<br />

traits for agriculture will likely be achieved by acquiring a<br />

better underst<strong>and</strong>ing of the root-to-shoot <strong>and</strong> shoot-to-root<br />

signaling networks. Thus, the future for research on plant<br />

vascular biology is very bright indeed!<br />

Acknowledgements<br />

Work in the authors’ laboratories was supported in part by the<br />

National Science Foundation (grants IOS-0752997 <strong>and</strong> IOS-<br />

0918433 to WJL; grants IOS#0749731, IOS#051909 to PK), the<br />

Department of Energy, Division of Energy Biosciences (grant<br />

DE-FG02-94ER20134 to WJL), the US Department of Agriculture,<br />

Agricultural Research Service (under Agreement number<br />

58-6250-0-008 to MAG), the Spanish Ministry of Science <strong>and</strong><br />

Innovation (MICINN) (grants AGL2007-61948 <strong>and</strong> AGL2009-<br />

09018 to AFLM), the Ministry of Education, Science, Sports<br />

<strong>and</strong> Culture of Japan (grant 19060009 to HF), from the Japan<br />

Society for the Promotion of Science (JSPS grant 23227001<br />

to HF), <strong>and</strong> from the NC-CARP project (to HF), the National<br />

Key Basic Research Program of China (grant 2012CB114500<br />

to XH), the National Natural Science Foundation of China (grant<br />

31070156 to XH), <strong>and</strong> the NSFC-JSPS cooperation project<br />

(grant 31011140070 to HF <strong>and</strong> XH).<br />

Received 20 Jan. 2013 Accepted 19 Feb. 2013<br />

References<br />

Aarts N, Metz M, Holub E, Staskawicz BJ, Daniels MJ, Parker<br />

JE (1998) Different requirements for EDS1 <strong>and</strong> NDR1 by disease<br />

resistance genes define at least two R gene-mediated signaling<br />

pathways in Arabidopsis. Proc. Natl. Acad. Sci. USA 95, 10306–<br />

10311.<br />

Agrawal B, Lakshmanan V, Kaushik S, Bais HP (2012) Natural<br />

variation among Arabidopsis accessions reveals malic acid<br />

as a key mediator of Nickel (Ni) tolerance. <strong>Plant</strong>a 236, 477–<br />

489.<br />

Aki T, Shigyo M, Nakano R, Yoneyama T, Yanagisawa S (2008)<br />

Nano scale proteomics revealed the presence of regulatory proteins<br />

including three FT-Like proteins in phloem <strong>and</strong> xylem saps from rice.<br />

<strong>Plant</strong> Cell Physiol. 49, 767–790.<br />

Akiyama K, Matsuzaki K, Hayashi H (2005) <strong>Plant</strong> sesquiterpenes<br />

induce hyphal branching in arbuscular mycorrhizal fungi. Nature<br />

435, 824–827.<br />

Alakonya A, Kumar R, Koenig D, Kimura S, Townsley B, Runo<br />

S, Garces HM, Kang J, Yanez A, David-Schwartz R, Machuka<br />

J, Sinha N (2012) Interspecific RNA interference of SHOOT<br />

MERISTEMLESS-Like disrupts Cuscuta pentagona plant parasitism.<br />

<strong>Plant</strong> Cell 4, 3153–3166.<br />

Alder NN, Pockman WT, Sperry JS, Nuismer S (1997) Use of<br />

centrifugal force in the study of xylem cavitation. J. Exp. Bot. 48,<br />

665–674.<br />

Aloni R (1987) Differentiation of vascular tissues. <strong>Plant</strong> Physiol. 38,<br />

179–204.<br />

Aloni R, Schwalm K, Langhans M, Ullrich CI (2003) Gradual shifts in<br />

sites of free-auxin production during leaf-primordium development<br />

<strong>and</strong> their role in vascular differentiation <strong>and</strong> leaf morphogenesis in<br />

Arabidopsis. <strong>Plant</strong>a 216, 841–853.<br />

Alves S, Nabais C, Simões Gonçalves MDL, Correia dos Santos MM<br />

(2011) Nickel speciation in the xylem sap of the hyperaccumulator<br />

Alyssum serpyllifolium ssp. lusitanicum growing on serpentine soils<br />

of northeast Portugal. <strong>Plant</strong> Physiol. 168, 1715–1722.


Aly R, Cholakh H, Joel DM, Leibman D, Steinitz B, Zelcer A, Naglis<br />

A, Yarden O, Gal-On A (2009) Gene silencing of mannose 6phosphate<br />

reductase in the parasitic weed Orobanche aegyptiaca<br />

through the production of homologous dsRNA sequences in the<br />

host plant. <strong>Plant</strong> Biotechnol. J. 7, 487–498.<br />

Amiard V, Mueh K, Demming-Adams B, Ebbert V, Turgeon R (2005)<br />

Anatomical <strong>and</strong> photosynthetic acclimation to the light environment<br />

in species with differing mechanisms of phloem loading. Proc. Natl.<br />

Acad. Sci. USA 102, 12968–12973.<br />

Anderegg WRL, Berry JA, Smith DD, Sperry JS, Anderegg LDL,<br />

Field CB (2012) <strong>The</strong> roles of hydraulic <strong>and</strong> carbon stress in a<br />

widespread climate-induced forest die-off. Proc. Am. Acad. Arts Sci.<br />

109, 233–237.<br />

Andrés-Colás N, Sancenón V, Rodríguez-Navarro S, Mayo S, Thiele<br />

DJ, Ecker JR, Puig S, Peñarrubia L (2006) <strong>The</strong> Arabidopsis<br />

heavy metal P-type ATPase HMA5 interacts with metallochaperones<br />

<strong>and</strong> functions in copper detoxification of roots. <strong>Plant</strong> J. 45,<br />

225–236.<br />

Aoki K, Suzui N, Fujimaki S, Dohmae N, Yonekura-Sakakibara<br />

K, Fujiwara T, Hayashi H, Yamaya T, Sakakibara H (2005)<br />

Destination-selective long-distance movement of phloem proteins.<br />

<strong>Plant</strong> Cell 17, 1801–1814.<br />

Aoyama T, Kobayashi T, Takahashi M, Nagasaka S, Usuda K,<br />

Kakei Y, Ishimaru Y, Nakanishi H, Mori S, Nishizawa NK (2009)<br />

OsYSL18 is a rice iron(III)-deoxymugineic acid transporter specifically<br />

expressed in reproductive organs <strong>and</strong> phloem of lamina joints.<br />

<strong>Plant</strong> Mol. Biol. 70, 681–692.<br />

Atkins CA, Smith PMC, Rodriguez-Medina C (2011) Macromolecules<br />

in phloem exudates – A review. Protoplasma 248, 165–172.<br />

Attaran E, Zeier TE, Griebel T, Zeier J (2009) Methyl salicylate<br />

production <strong>and</strong> jasmonate signaling are not essential for systemic<br />

acquired resistance in Arabidopsis. <strong>Plant</strong> Cell 21, 954–971.<br />

Avci U, Petzold HE, Ismail IO, Beers EP, Haigler CH (2008) Cysteine<br />

proteases XCP1 <strong>and</strong> XCP2 aid micro-autolysis within the intact<br />

central vacuole during xylogenesis in Arabidopsis roots. <strong>Plant</strong> J.<br />

56, 303–315.<br />

Bai SL, Kasai A, Yamada K, Li TZ, Harada T (2011) A mobile<br />

signal transported over a long distance induces systemic transcriptional<br />

gene silencing in a grafted partner. J. Exp. Bot. 62,<br />

4561–4570.<br />

Baima S, Nobili F, Sessa G, Lucchetti S, Ruberti I, Morelli G<br />

(1995) <strong>The</strong> expression of the athb-8 homeobox gene is restricted to<br />

provascular cells in Arabidopsis thaliana. <strong>Development</strong> 121, 4171–<br />

4182.<br />

Bahrun A, Jensen CR, Asch F, Mogensen VO (2002) Droughtinduced<br />

changes in xylem pH, ionic composition, <strong>and</strong> ABA concentration<br />

act as early signals in field-grown maize (Zea mays L.).<br />

J. Exp. Bot. 53, 251–263.<br />

Ball MC, Canny MJ, Huang CX, Egerton JJG, Wolfe J (2006)<br />

Freeze/thaw-induced embolism depends on nadir temperature: <strong>The</strong><br />

heterogeneous hydration hypothesis. <strong>Plant</strong> Cell Environ. 29, 729–<br />

745.<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 363<br />

Banerjee AK, Chatterjee M, Yu YY, Suh SG, Miller WA, Hannapel<br />

DJ (2006) Dynamics of a mobile RNA of potato involved in a longdistance<br />

signaling pathway. <strong>Plant</strong> Cell 18, 3443–3457.<br />

Barnes A, Bale J, Constantinidou C, Ashton P, Jones A, Pritchard<br />

J (2004) Determining protein identity from sieve element sap in<br />

Ricinus communis L. by quadrupole time of flight (Q-TOF) mass<br />

spectrometry. J. Exp. Bot. 55, 1473–1481.<br />

Barnett JR (1982) Plasmodesmata <strong>and</strong> pit development in secondary<br />

xylem elements. <strong>Plant</strong>a. 155, 251–260.<br />

Batailler B, Lemaitre T, Vilaine F, Sanchez C, Renard D, Cayla T,<br />

Beneteau J, Dinant S (2012) Soluble <strong>and</strong> filamentous proteins in<br />

Arabidopsis sieve elements. <strong>Plant</strong> Cell Environ. 35, 1258–1273.<br />

Bauby H, Divol F, Truernit E, Gr<strong>and</strong>jean O, Palauqui JC (2007) Protophloem<br />

differentiation in early Arabidopsis thaliana development.<br />

<strong>Plant</strong> Cell Physiol. 48, 97–109.<br />

Baxter I, Muthukumar B, Hyeong CP, Buchner P, Lahner B, Danku<br />

J, Zhao K, Lee J, Hawkesford MJ, Guerinot ML, Salt DE (2008)<br />

Variation in molybdenum content across broadly distributed populations<br />

of Arabidopsis thaliana is controlled by a mitochondrial<br />

molybdenum transporter (MOT1). PLoS Genetics 4, e1000004.<br />

Behmer ST, Grebenok RJ, Douglas AE (2011) <strong>Plant</strong> sterols <strong>and</strong> host<br />

plant suitability for a phloem-feeding insect. Funct. Ecol. 25, 484–<br />

491.<br />

Belimov AA, Dodd IC, Hontzeas N, <strong>The</strong>obald JC, Safronova<br />

VI, Davies WJ (2009) Rhizosphere bacteria containing<br />

1-aminocyclopropane-1-carboxylate deaminase increase yield<br />

of plants grown in drying soil via both local <strong>and</strong> systemic hormone<br />

signalling. New Phytol. 181, 413–423.<br />

Bellés JM, Garro R, Fayos J, Navarro P, Primo J, Conejero V (1999)<br />

Gentisic acid as a pathogen-inducible signal, additional to salicylic<br />

acid for activation of plant defenses in tomato. Mol. <strong>Plant</strong>-Microbe<br />

Interact. 12, 227–235.<br />

Benschop JJ, Mohammed S, O’Flaherty M, Heck AJR, Slijper<br />

M, Menke FLH (2007) Quantitative phosphoproteomics of<br />

early elicitor signaling in Arabidopsis. Molec. Cell. Proteomics 6,<br />

1198–1214.<br />

Benkova E, Michniewicz M, Sauer M, Teichmann T, Seifertova<br />

D, Jurgens G, Friml J (2003) Local, efflux-dependent auxin<br />

gradients as a common module for plant organ formation. Cell 115,<br />

591–602.<br />

Benning UF, Tamot B, Guelette BS, Hoffmann-Benning S (2012)<br />

New aspects of phloem-mediated long-distance lipid signaling in<br />

plants. Front. <strong>Plant</strong> Sci. 3, 53.<br />

Berleth T, Mattsson J, Hardtke CS (2000) <strong>Vascular</strong> continuity <strong>and</strong><br />

auxin signals. Trends <strong>Plant</strong> Sci. 5, 387–393.<br />

Betsuyaku S, Takahashi F, Kinoshita A, Miwa H, Shinozaki K,<br />

Fukuda H, Sawa S (2011) Mitogen-activated protein kinase regulated<br />

by the CLAVATA receptors contributes to shoot apical<br />

meristem homeostasis. <strong>Plant</strong> Cell Physiol. 52, 14–29.<br />

Beveridge CA, Ross JJ, Murfet IC (1994) Branching mutant rms-2<br />

in Pisum sativum (grafting studies <strong>and</strong> endogenous indole-3-acetic<br />

acid levels). <strong>Plant</strong> Physiol. 104, 953–959.


364 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Bhattacharjee S, Halane MK, Kim SH, Gassmann W (2011)<br />

Pathogen effectors target Arabidopsis EDS1 <strong>and</strong> alter its interaction<br />

with immune regulators. Science 334, 1405–1408.<br />

Bieleski RL (1973) Phosphate pools, phosphate transport, <strong>and</strong> phosphate<br />

availability. Ann. Rev. <strong>Plant</strong> Physiol. 24, 225–252.<br />

Bishopp A, Help H, El-Showk S, Weijers D, Scheres B, Friml<br />

J, Benková E,Mähönen AP, Helariutta Y (2011a) A mutually<br />

inhibitory interaction between auxin <strong>and</strong> cytokinin specifies vascular<br />

pattern in roots. Curr. Biol. 21, 917–926.<br />

Bishopp A, Lehesranta S, Vatén A, Help H, El-Showk S, Scheres B,<br />

Helariutta K, Mähönen AP, Sakakibara H, Helariutta Y (2011b)<br />

Phloem-transported cytokinin regulates polar auxin transport <strong>and</strong><br />

maintains vascular pattern in the root meristem. Curr. Biol. 21, 927–<br />

932.<br />

Blindauer CA, Schmid R (2010) Cytosolic metal h<strong>and</strong>ling in plants:<br />

Determinants for zinc specificity in metal transporters <strong>and</strong> metallothioneins.<br />

Metallomics 2, 510–529.<br />

Bolaños L, Lukaszewski K, Bonilla I, Blevins D (2004) Why boron?<br />

<strong>Plant</strong> Physiol. Biochem. 42, 907–912.<br />

Bollhoner B, Prestele J, Tuominen H (2012) Xylem cell death:<br />

Emerging underst<strong>and</strong>ing of regulation <strong>and</strong> function. J. Exp. Bot.<br />

63, 1081–1094.<br />

Bonke M, Thitamadee S, Mähönen AP, Hauser MT, Helariutta Y<br />

(2003) APL regulates vascular tissue identity in Arabidopsis. Nature<br />

426, 181–186.<br />

Boot KJM, Libbenga KR, Hille SC, Offringa R, van Duijn B (2012)<br />

Polar auxin transport: An early invention. J. Exp. Bot. 63, 4213–<br />

4218.<br />

Bostwick DE, Dannenhoffer JM, Skaggs MI, Lister RM, Larkins BA,<br />

Thompson GA (1992) Pumpkin phloem lectin genes are specifically<br />

expressed in companion cells. <strong>Plant</strong> Cell 4, 1539–1548.<br />

Bowman JL, Floyd SK (2008) Patterning <strong>and</strong> polarity in seed plant<br />

shoots. Annu. Rev. <strong>Plant</strong> Biol. 59, 67–88.<br />

Boyce KC, Brodribb TJ, Feild TS, Zwieniecki MA (2009) Angiosperm<br />

leaf vein evolution was physiologically <strong>and</strong> environmentally transformative.<br />

Proc. R. Soc. Series B 276, 1771–1776.<br />

Boyce KC, Zwieniecki MA (2012) Leaf fossil record suggests limited<br />

influence of atmospheric CO2 on terrestrial productivity prior to<br />

angiosperm evolution. Proc. Natl. Acad. Sci. USA 109, 10403–<br />

10408.<br />

Brady SM, Orl<strong>and</strong>o DA, Lee JY, Wang JY, Koch J, Dinneny<br />

JR, Mace D, Ohler U, Benfey PN (2007) A high-resolution root<br />

spatiotemporal map reveals dominant expression patterns. Science<br />

318, 801–806.<br />

Brady SM, Zhang L, Megraw M, Martinez NJ, Jiang E, Yi CS,<br />

Liu W, Zeng A, Taylor-Teeples M, Kim D, Ahnert S, Ohler U,<br />

Ware D, Walhout AJ, Benfey PN (2011) A stele-enriched gene<br />

regulatory network in the Arabidopsis root. Mol. Syst. Biol. 7,<br />

459.<br />

Brewer PB, Dun EA, Ferguson BJ, Rameau C, Beveridge CA (2009)<br />

Strigolactone acts downstream of auxin to regulate bud outgrowth<br />

in pea <strong>and</strong> Arabidopsis. <strong>Plant</strong> Physiol. 150, 482–493.<br />

Brewer PB, Koltai H, Beveridge CA (2013) Diverse roles of strigolactones<br />

in plant development. Mol. <strong>Plant</strong> 6, 18–28.<br />

Briggs LJ (1950) Limiting negative pressure of water. J. Appl. Phys.<br />

21, 721–722.<br />

Brodersen C, McElrone AJ, Choat B, Matthews MA, Shackel KA<br />

(2010) <strong>The</strong> dynamics of embolism repair in xylem: in vivo visualizations<br />

using high-resolution computed tomography. <strong>Plant</strong> Physiol.<br />

154, 1088–1095.<br />

Brodersen P, Voinnet O (2006) <strong>The</strong> diversity of RNA silencing pathways<br />

in plants. Trends Genet. 22, 268–280.<br />

Brodribb TJ (2009) Xylem hydraulic physiology: <strong>The</strong> functional backbone<br />

of terrestrial plant productivity. <strong>Plant</strong> Sci. 177, 245–251.<br />

Brodribb TJ, Holbrook NM (2005) Water stress deforms tracheids<br />

peripheral peripheral to the leaf vein of a tropical conifer. <strong>Plant</strong><br />

Physiol. 137, 1139–1146.<br />

Brodribb TJ, McAdam SAM (2011) Passive origins of stomatal control<br />

in vascular plants. Science 331, 582–585.<br />

Brown PH, Hu H (1996) Phloem mobility of boron is species dependent:<br />

Evidence for phloem mobility in sorbitol-rich species. Ann. Bot. 77,<br />

497–505.<br />

Bucci SJ, Scholz FG, Goldstein G, Meinzer FC, Sternberg LDSL<br />

(2003) Dynamic changes in hydraulic conductivity in petioles of<br />

two savanna species: Factors <strong>and</strong> mechanisms contributing to the<br />

refilling of embolized vessels. <strong>Plant</strong> Cell Environ. 26, 1633–1645.<br />

Buhtz A, Pieritz J, Springer F, Kehr J (2010) Phloem small RNAs,<br />

nutrient stress responses, <strong>and</strong> systemic mobility. BMC <strong>Plant</strong> Biol.<br />

10, 64.<br />

Buhtz A, Springer F, Chappell L, Baulcombe DC, Kehr J (2008)<br />

Identification <strong>and</strong> characterization of small RNAs from the phloem<br />

of Brassica napus. <strong>Plant</strong> J. 53, 739–749.<br />

Burleigh SH, Harrison MJ (1999) <strong>The</strong> down-regulation of Mt4-like<br />

genes by phosphate fertilization occurs systemically <strong>and</strong> involves<br />

phosphate translocation to the shoots. <strong>Plant</strong> Physiol. 119, 241–248.<br />

Butenko MA, Vie AK, Brembu T, Aalen RB, Bones AM (2009) <strong>Plant</strong><br />

peptides in signalling: Looking for new partners. Trends <strong>Plant</strong> Sci.<br />

14, 255–263.<br />

Butterfield BG (1995) Vessel element differentiation. In: Iqbal M, ed.<br />

<strong>The</strong> Cambial Derivatives. Encyclopedia of <strong>Plant</strong> Anatomy. Gebruder<br />

Borntraeger, Berlin. pp. 93–106.<br />

Cakmak I, Kirkby EA (2008) Role of magnesium in carbon partitioning<br />

<strong>and</strong> alleviating photo-oxidative damage. Physiol. <strong>Plant</strong> 133, 692–<br />

704.<br />

Canny MJ (1975) Mass Transfer. In: MH Zimermann, JA Milburn, eds.<br />

Transport in <strong>Plant</strong>s I. Phloem Transport. Springer-Verlag, Berlin.<br />

pp. 139–153.<br />

Canny MJ (1998) Transporting water in plants. Amer. Sci. 86, 152–159.<br />

Cano-Delgado A, Lee JY, Demura T (2010) Regulatory mechanisms<br />

for specification <strong>and</strong> patterning of plant vascular tissues. Annu. Rev.<br />

Cell Dev. Biol. 26, 605–637.<br />

Cao H, Bowling SA, Gordon S, Dong X (1994) Characterization of an<br />

Arabidopsis mutant that is nonresponsive to inducers of systemic<br />

acquired resistance. <strong>Plant</strong> Cell 6, 1583–1592.


Cao H, Glazebrook J, Clark JD, Volko S, Dong X (1997) <strong>The</strong><br />

Arabidopsis NPR1 gene that controls systemic acquired resistance<br />

encodes a novel protein containing ankyrin repeats. Cell 88,<br />

57–64.<br />

Caplan J, Padmanabhan M, Dinesh-Kumar SP (2008) <strong>Plant</strong> NB-LRR<br />

immune receptors: From recognition to transcriptional reprogramming.<br />

Cell Host Microbe 3, 126–135.<br />

Carlquist S (1984) Vessel grouping in dicotyledon wood: Significance<br />

<strong>and</strong> relationship to imperforate tracheary elements. Aliso 10, 505–<br />

525.<br />

Carlsbecker A, Lee JY, Roberts CJ, Dettmer J, Lehesranta S, Zhou<br />

J, Lindgren O, Moreno-Risueno MA, Vatén A, Thitamadee S,<br />

Campilho A, Sebastian J, Bowman JL, Helariutta Y, Benfey<br />

PN (2010) Cell signalling by microRNA165/6 directs gene dosedependent<br />

root cell fate. Nature 465, 316–321.<br />

Carr JP, Dixon DC, Nikolau BJ, Voelkerding KV, Klessig DF (1987)<br />

Synthesis <strong>and</strong> localization of pathogenesis-related proteins in tobacco.<br />

Mol. Cell Biol. 7, 1580–1583.<br />

Catford JG, Staehelin C, Lerat S, Piché Y, Vierheilig H (2003)<br />

Suppression of arbuscular mycorrhizal colonization <strong>and</strong> nodulation<br />

in split-root systems of alfalfa after pre-inoculation <strong>and</strong> treatment<br />

with Nod factors. J. Exp. Bot. 54, 1481–1487.<br />

Century KS, Holub EB, Staskawicz BJ (1995) NDR1, a locus of<br />

Arabidopsis thaliana that is required for disease resistance to both<br />

a bacterial <strong>and</strong> a fungal pathogen. Proc. Natl. Acad. Sci. USA 92,<br />

6597–6601.<br />

Century KS, Shapiro AD, Repetti PP, Dahlbeck D, Holub E,<br />

Staskawicz BJ (1997) NDR1, a pathogen-induced component<br />

required for Arabidopsis disease resistance. Science 278, 1963–<br />

1965.<br />

Champigny MJ, Shearer H, Mohammad A, Haines K, Neumann M,<br />

Thilmony R, He SY, Fobert P, Dengler N, Cameron RK (2011)<br />

Localization of DIR1 at the tissue, cellular <strong>and</strong> subcellular levels<br />

during systemic acquired resistance in Arabidopsis using DIR1:GUS<br />

<strong>and</strong> DIR1:EGFP reporters. BMC <strong>Plant</strong> Biol. 11, 125.<br />

Ch<strong>and</strong>a B, Venugopal SC, Kulshrestha S, Navarre D, Downie<br />

B, Vaillancourt L, Kachroo A, Kachroo P (2008) Glycerol-3phosphate<br />

levels are associated with basal resistance to the<br />

hemibiotrophic fungus Colletotrichum higginsianum in Arabidopsis.<br />

<strong>Plant</strong> Physiol. 147, 2017–2029.<br />

Ch<strong>and</strong>a B, Ye X, M<strong>and</strong>al M, Yu K, Sekine KT, Gao QM, Selote<br />

D, Hu Y, Stromberg A, Navarre D, Kachroo A, Kachroo P<br />

(2011) Glycerol-3-phosphate is a critical mobile inducer of systemic<br />

immunity in plants. Nat. Genet. 43, 421–427.<br />

Ch<strong>and</strong>ra-Shekara AC, Navarre D, Kachroo A, Kang HG, Klessig<br />

D, Kachroo P (2004) Signaling requirement <strong>and</strong> role of salicylic<br />

acid in HRT- <strong>and</strong> rrt-mediated resistance to turnip crinkle virus in<br />

Arabidopsis. <strong>Plant</strong> J. 40, 647–659.<br />

Ch<strong>and</strong>ra-Shekara AC, Venugopal SC, Kachroo A, Kachroo P (2007)<br />

Plastidal fatty acid levels regulate resistance gene-dependent defense<br />

signaling in Arabidopsis. Proc. Natl. Acad. Sci. USA 104,<br />

7277–7282.<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 365<br />

Chaturvedi R, Krothapalli K, Mak<strong>and</strong>ar R, N<strong>and</strong>i A, Sparks A, Roth<br />

M,WeltiR,ShahJ(2008) Plastid omega3-fatty acid desaturasedependent<br />

accumulation of a systemic acquired resistance inducing<br />

activity in petiole exudates of Arabidopsis thaliana is independent<br />

of jasmonic acid. <strong>Plant</strong> J. 54, 106–117.<br />

Chaturvedi R, Venables B, Petros R, Nalam V, Li M, Wang<br />

X, Takemoto LJ, Shah J (2012) An abietane diterpenoid is a<br />

potent activator of systemic acquired resistance. <strong>Plant</strong> J. 71,<br />

161–172.<br />

Chen F, D’Auria JC, Tholl D, Ross JR, Gershenzon J, Noel JP, Pichersky<br />

E (2003) An Arabidopsis thaliana gene for methylsalicylate<br />

biosynthesis, identified by a biochemical genomics approach, has<br />

a role in defense. <strong>Plant</strong> J. 36, 577–588.<br />

Chen HM, Pang Y, Zeng J, Ding Q, Yin SY, Liu C, Lu MZ, Cui KM, He<br />

XQ (2012a) <strong>The</strong> Ca2+ -dependent DNases are involved in secondary<br />

xylem development in Eucommia ulmoides. J. Integr. <strong>Plant</strong> Biol. 54,<br />

456–470.<br />

Chen LQ, Qu XQ, Hou BH, Sosso D, Osorio S, Fernie AR, Frommer<br />

WB (2012b) Sucrose efflux mediated by SWEET proteins as a key<br />

step for phloem transport. Science 335, 207–211.<br />

Chen Z, Silva H, Klessig DF (1993) Involvement of reactive oxygen<br />

species in the induction of systemic acquired resistance by salicylic<br />

acid in plants. Science 242, 883–886.<br />

Chern MS, Fitzgerald HA, Yadav RC, Canlas PE, Dong X, Ronald<br />

PC (2001) Evidence for a disease-resistance pathway in rice similar<br />

to the NPR1-mediated signaling pathway in Arabidopsis. <strong>Plant</strong> J. 27,<br />

101–113.<br />

Chincinska IA, Liesche J, Krügel U, Michalska J, Geigenberger P,<br />

Grimm B, Kühn C (2008) Sucrose transporter StSUT4 from potato<br />

affects flowering, tuberization, <strong>and</strong> shade avoidance response.<br />

<strong>Plant</strong> Physiol. 146, 515–528.<br />

Chiou TJ, Lin SI (2011) Signaling network in sensing phosphate<br />

availability in plants. Annu. Rev. <strong>Plant</strong> Biol. 62, 185–206.<br />

Choat B, Cobb AR, Jansen S (2008) Structure <strong>and</strong> function of bordered<br />

pits: New discoveries <strong>and</strong> impacts on whole-plant hydraulic<br />

function. New Phytol. 177, 608–626.<br />

Choat B, Drayton WM, Brodersen C, Matthews MA, Schackel KA,<br />

Wada H, McElrone AJ (2010) Measurement of vulnerability to<br />

water stress-induced cavitation in grapevine: A comparison of four<br />

techniques applied to a long-vesseled species. <strong>Plant</strong> Cell Environ.<br />

33, 1502–1512.<br />

Choat B, Gambetta GA, Shakel KA, Matthews MA (2009) <strong>Vascular</strong><br />

function in grape berries across development <strong>and</strong> its relevance to<br />

apparent hydraulic isolation. <strong>Plant</strong> Physiol. 151, 1677–1687.<br />

Choat B, Jansen S, Zwieniecki MA, Smets E, Holbrook NM (2004)<br />

Changes in pit membrane porosity due to deflection <strong>and</strong> stretching:<br />

<strong>The</strong> role of vestured pits. J. Exp. Bot. 55, 1569–1575.<br />

Choat B, Medek DE, Stuart SA, Pasquet-Kok J, Egerton JJG,<br />

Salari H, Sack L, Ball MC (2011) Xylem traits mediate a trade-off<br />

between resistance to freeze-thaw-induced embolism <strong>and</strong> photosynthetic<br />

capacity in overwintering evergreens. New Phytol. 191,<br />

996–1005.


366 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Christman MA, Sperry JS, Adler FR (2009) Testing the rare pit<br />

hypothesis in three species of Acer. New Phytol. 182, 664–674.<br />

Christman MA, Sperry JS (2010) Single vessel flow measurements<br />

indicate scalariform perforation plates confer higher resistance to<br />

flow than previously estimated. <strong>Plant</strong> Cell Environ. 33, 431–443.<br />

Christman MA, Sperry JS, Smith DD (2012) Rare pits, large vessels,<br />

<strong>and</strong> extreme vulnerability to cavitation in a ring-porous tree species.<br />

New Phytol. 193, 713–720.<br />

Christmann A, Weiler EW, Steudle E, Grill E (2007) A hydraulic signal<br />

in root-to-shoot signalling of water shortage. <strong>Plant</strong> J. 52, 167–174.<br />

Clay NK, Nelson T (2005) Arabidopsis thickvein mutation affects vein<br />

thickness <strong>and</strong> organ vascularization, <strong>and</strong> resides in a provascular<br />

cell-specific spermine synthase involved in vein definition <strong>and</strong> in<br />

polar auxin transport. <strong>Plant</strong> Physiol. 138, 767–777.<br />

Cochard H, Cruiziat P, Tyree MT (1992) Use of positive pressures to<br />

establish vulnerability curves: Further support for the air-seeding<br />

hypothesis <strong>and</strong> implications for pressure-volume analysis. <strong>Plant</strong><br />

Physiol. 100, 205–209.<br />

Cochard H, Froux F, Mayr S, Cout<strong>and</strong> C (2004) Xylem wall collapse<br />

in water-stressed pine needles. <strong>Plant</strong> Physiol. 134, 401–408.<br />

Cochard H, Gaelle D, Bodet C, Tharwat I, Poirier M, Ameglio T<br />

(2005) Evaluation of a new centrifuge technique for rapid generation<br />

of xylem vulnerability curves. Physiol. <strong>Plant</strong>. 124, 410–418.<br />

Cochard H, Herbette S, Barigah T, Vilagrosa A (2010a) Does sample<br />

length influence the shape of vulnerability to cavitation curves? A<br />

test with the Cavitron spinning technique. <strong>Plant</strong> Cell Environ. 33,<br />

1543–1552.<br />

Cochard H, Herbette S, Hern<strong>and</strong>ez E, Holtta T, Mencuccini M<br />

(2010b) <strong>The</strong> effects of sap ionic composition on xylem vulnerability<br />

to cavitation. J. Exp. Bot. 61, 275–285.<br />

Cochard H, Holtta T, Herbette S, Delzon S, Mencuccini M (2009)<br />

New insights into the mechanisms of water-stress-induced cavitation<br />

in conifers. <strong>Plant</strong> Physiol. 151, 949–954.<br />

Colombani A, Djerbi S, Bessueille L, Blomqvist K, Ohlsson A,<br />

Berglund T, Teeri TT, Bulone V (2004) In vitro synthesis of<br />

(1 → 3)-β-D-glucan (callose) <strong>and</strong> cellulose by detergent extracts of<br />

membranes from cell suspension cultures of hybrid aspen. Cellulose<br />

11, 313–327.<br />

Conte SS, Walker EL (2011) Transporters contributing to iron trafficking<br />

in plants. Mol. <strong>Plant</strong> 4, 464–476.<br />

Coppinger P, Pepetti PP, Day B, Dahlbeck D, Mehlert A, Staskawicz<br />

BJ (2004) Overexpression of the plasma membrane-localized<br />

NDR1 protein results in enhanced bacterial resistance in Arabidopsis<br />

thaliana. <strong>Plant</strong> J. 40, 225–237.<br />

Corbesier L, Vincent C, Jang SH, Fornara F, Fan QZ, Searle I,<br />

Giakountis A, Farrona S, Gissot L, Turnbull C, Coupl<strong>and</strong> G<br />

(2007) FT protein movement contributes to long-distance signaling<br />

in floral induction of Arabidopsis. Science 316, 1030–1033.<br />

Courtois-Moreau CL, Pesquet E, Sjodin A, Muniz L, Bollhoner B,<br />

Kaneda M, Samuels L, Jansson S, Tuominen H (2009) A unique<br />

program for cell death in xylem fibers of Populus stem. <strong>Plant</strong> J. 58,<br />

260–274.<br />

Cox RM, Malcolm JM (1997) Effects of duration of a simulated winter<br />

thaw on dieback <strong>and</strong> xylem conductivity of Betula papyrifera. Tree<br />

Physiol. 17, 389–396.<br />

Crombie DS, Hipkins MF, Milburn JA (1985) Gas penetration of<br />

pit membranes in the xylem of Rhododendron as the cause of<br />

acoustically detectable sap cavitation. Aust. J. <strong>Plant</strong> Physiol. 12,<br />

445–454.<br />

Cronshaw J (1981) Phloem structure <strong>and</strong> function. Annu Rev. <strong>Plant</strong><br />

Physiol. <strong>Plant</strong> Mol. Biol. 32, 465–484.<br />

Cui H, Hao YM, Stolc V, Deng XW, Sakakibara H, Kojima M (2011)<br />

Genome-wide direct target analysis reveals a role for SHORT-<br />

ROOT in root vascular patterning through cytokinin homeostasis.<br />

<strong>Plant</strong> Physiol. 157, 1221–1231.<br />

Cui H, Levesque MP, Vernoux T, Jung JW, Paquette AJ, Gallagher<br />

KL, Wang JY, Blilou I, Scheres B, Benfey PN (2007)<br />

An evolutionarily conserved mechanism delimiting SHR movement<br />

defines a single layer of endodermis in plants. Science 316,<br />

421–425.<br />

Cui KM, He XQ (2012) <strong>The</strong> Ca2+ -dependent DNases are involved in<br />

secondary xylem development in Eucommia ulmoides. J. Integr.<br />

<strong>Plant</strong> Biol. 54, 456–470.<br />

Curie C, Cassin G, Couch D, Divol F, Higuchi K, Le Jean M, Misson<br />

J, Schikora A, Czernic P, Mari S (2009) Metal movement within<br />

the plant: Contribution of nicotianamine <strong>and</strong> yellow stripe 1-like<br />

transporters. Ann. Bot. 103, 1–11.<br />

Dangl JL, Dietrich RA, Richberg MH (1996) Death don’t have no<br />

mercy: Cell death programs in plant-microbe interactions. <strong>Plant</strong> Cell<br />

8, 1793–1807.<br />

David-Schwartz R, Runo S, Townsley B, Machuka J, Sinha N (2008)<br />

Long-distance transport of mRNA via parenchyma cells <strong>and</strong> phloem<br />

across the host-parasite junction in Cuscuta. New Phytol. 179,<br />

1133–1141.<br />

Davidson A, Keller F, Turgeon R (2011) Phloem loading, plant growth<br />

form, climate. Protoplasma 248, 153–163.<br />

Davis SD, Sperry JS, Hacke UG (1999) <strong>The</strong> relationship between<br />

xylem conduit diameter <strong>and</strong> cavitation caused by freeze-thaw<br />

events. Am. J. Bot. 86, 1367–1372.<br />

Davis WJ, Zhang J (1991) Root signals <strong>and</strong> the regulation of growth<br />

<strong>and</strong> development of plants in drying soil. Annu. Rev. <strong>Plant</strong> Physiol.<br />

Mol. Biol. 42, 55–76.<br />

Dean JV, Delaney SP (2008) Metabolism of salicylic acid in wild-type,<br />

ugt74f1 <strong>and</strong> ugt74f2 glucosyltransferase mutants of Arabidopsis<br />

thaliana. Physiol. <strong>Plant</strong>. 132, 417–425.<br />

Deeken R, Ache P, Kajahan I, Klinkenberg J, Bringmann G, Hedrich<br />

R (2008) Identification of Arabidopsis thaliana phloem RNAs<br />

provides a search criterion for phloem-based transcripts hidden<br />

in complex datasets of microarray experiments. <strong>Plant</strong> J. 55,<br />

746–759.<br />

Deeken R, Geiger D, Fromm J, Koroleva O, Ache P, Langenfeld-<br />

Heyser R, Sauer N, May ST, Hedrich R (2002) Loss of AKT2/3<br />

potassium channel affects sugar loading into the phloem of Arabidopsis.<br />

<strong>Plant</strong>a 216, 334–344.


Delaney TP, Friedrich L, Ryals JA (1995) Arabidopsis signal transduction<br />

mutant defective in chemically <strong>and</strong> biologically induced disease<br />

resistance. Proc. Natl. Acad. Sci. USA 92, 6602–6606.<br />

Demari-Weissler H, Rachamilavech S, Aloni R, German MA, Cohen<br />

S, Zwieniecki MA, Holbrook NM, Granot D (2009) LeFRK2 is<br />

required for phloem <strong>and</strong> xylem differentiation <strong>and</strong> the transport of<br />

both sugar <strong>and</strong> water. <strong>Plant</strong>a 230, 795–805.<br />

Demura T, Tashiro G, Horiguchi G, Kishimoto N, Kubo M, Matsuoka<br />

N, Minami A, Nagata-Hiwatashi M, Nakamura K, Okamura Y,<br />

Sassa N, Suzuki S, Yazaki J, Kikuchi S, Fukuda H (2002) Visualization<br />

by comprehensive microarray analysis of gene expression<br />

programs during transdifferentiation of mesophyll cells into xylem<br />

cells. Proc. Natl. Acad. Sci. USA 99, 15794–15799.<br />

Demura T, Fukuda H (2007) Transcriptional regulation in wood formation.<br />

Trends <strong>Plant</strong> Sci. 12, 64–70.<br />

Dengler N, Kang J (2001) <strong>Vascular</strong> patterning <strong>and</strong> leaf shape. Curr.<br />

Opin. <strong>Plant</strong> Biol. 4, 50–56.<br />

De Smet I, Tetsumura T, De Rybel B, Frey NF, Laplaze L, Casimiro I,<br />

Swarup R, Naudts M, Vanneste S, Audenaert D, Inze D, Bennett<br />

MJ, Beeckman T (2007) Auxin-dependent regulation of lateral root<br />

positioning in the basal meristem of Arabidopsis. <strong>Development</strong> 134,<br />

681–690.<br />

Després C, DeLong C, Glaze S, Liu E, Fobert PR (2000) <strong>The</strong><br />

Arabidopsis NPR1/NIM1 protein enhances the DNA binding activity<br />

of a subgroup of the TGA family of bZIP transcription factors. <strong>Plant</strong><br />

Cell 12, 279–290.<br />

Després C, Chubak C, Rochon A, Clark R, Bethune T, Desveaux<br />

D, Fobert PR (2003) <strong>The</strong> Arabidopsis NPR1 disease resistance<br />

protein is a novel cofactor that confers redox regulation of DNA<br />

binding activity to the basic domain/leucine zipper transcription<br />

factor TGA1. <strong>Plant</strong> Cell 15, 2181–2191.<br />

Dhawan W, Luo H, Foerster AM, AbuQamar S, Du HN, Briggs SC,<br />

Scheid OM, Mengiste T (2009) HISTONE MONOUBIQUITINA-<br />

TION1 interacts with a subunit of the mediator complex <strong>and</strong> regulates<br />

defense against necrotrophic fungal pathogens in Arabidopsis.<br />

<strong>Plant</strong> Cell 21, 1000–1019.<br />

Dhondt S, Coppens F, De Winter F, Swarup K, Merks RMH,<br />

Inze D, Bennett MJ, Beemster GTS (2010) SHORT-ROOT <strong>and</strong><br />

SCARECROW regulate leaf growth in Arabidopsis by stimulating<br />

S-phase progression of the cell cycle. <strong>Plant</strong> Physiol. 154, 1183–<br />

1195.<br />

DiDonato R, Roberts L, S<strong>and</strong>erson T, Eisley R, Walker E (2004)<br />

Arabidopsis yellow stripe-like2 (YSL2): A metal-regulated gene<br />

encoding a plasma membrane transporter of nicotianamine-metal<br />

complexes. <strong>Plant</strong> J. 39, 403–414.<br />

Diévart A, Clark SE (2004) LRR-containing receptors regulating plant<br />

development <strong>and</strong> defense. <strong>Development</strong> 131, 251–261.<br />

Dinant S, Lucas WJ (2012) Sieve elements: Puzzling activities<br />

deciphered by proteomics studies. In: GA Thompson,<br />

AJE van Bel, eds. Phloem: Molecular Cell Biology, <strong>System</strong>ic<br />

Communication, Biotic Interactions. Wiley-Blackwell, Ames. pp.<br />

157–185.<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 367<br />

Doering-Saad C, Newbury HJ, Couldridge CE, Bale JS, Pritchard J<br />

(2006) A phloem-enriched cDNA library from Ricinus: Insights into<br />

phloem function. J. Exp. Bot. 57, 3183–3193.<br />

Dolan L, Janmaat K, Willemsen V, Linstead P, Poethig S, Roberts K,<br />

Scheres B (1993) Cellular organisation of the Arabidopsis thaliana<br />

root. <strong>Development</strong> 119, 71–84.<br />

Domagalska MA, Leyser O (2011) Signal integration in the control of<br />

shoot branching. Nat. Rev. Mol. Cell Biol. 12, 211–221.<br />

Domec JC, Lachenbruch B, Meinzer FC (2006) Bordered pit structure<br />

<strong>and</strong> function determine spatial patterns of air-seeding thresholds in<br />

xylem of douglas-fir (Psuedotsuga menziesii; Pinaceae) trees. Am.<br />

J. Bot. 93, 1588–1600.<br />

Domec JC, Warren JM, Meinzer FC, Lachenbruch B (2009) Safety<br />

factors for xylem failure by implosion <strong>and</strong> air-seeding with roots,<br />

trunks, <strong>and</strong> branches of young <strong>and</strong> old conifer trees. IAWA J. 30,<br />

100–120.<br />

Donner TJ, Sherr I, Scarpella E (2009) Regulation of preprocambial<br />

cell state acquisition by auxin signaling in Arabidopsis leaves.<br />

<strong>Development</strong> 136, 3235–3246.<br />

Du J, Groover A (2010) Transcriptional regulation of secondary growth<br />

<strong>and</strong> wood formation. J. Integr. <strong>Plant</strong> Biol 52, 17–27.<br />

Du S, Yamamoto F (2007) An overview of the biology of reaction wood<br />

formation. J. Integr. <strong>Plant</strong> Biol. 49, 131–143.<br />

Duckett JG, Pressel S, P’ng KMY, Renzaglia, KS (2009) Exploding<br />

a myth: <strong>The</strong> capsule dehiscence mechanism <strong>and</strong> the function of<br />

pseudostomata in Sphagnum. New Phytol. 183, 1053–1063.<br />

Dunoyer P, Brosnan CA, Schott G, Wang Y, Jay F, Alioua A, Himber<br />

C, Voinnet O (2010) An endogenous, systemic RNAi pathway in<br />

plants. EMBO J. 29, 1699–1712.<br />

Durner J, Klessig DF (1995) Inhibition of ascorbate peroxidase by<br />

salicylic acid <strong>and</strong> 2.6-dichloroisonicdinic acid, two inducers of plant<br />

defense responses. Proc. Natl. Acad. Sci. USA 92, 11312–11316.<br />

Durrant WE, Wang S, Dong X (2007) Arabidopsis SNI1 <strong>and</strong> RAD51D<br />

regulate both gene transcription <strong>and</strong> DNA recombination during the<br />

defense response. Proc. Natl. Acad. Sci. USA 104, 4223–4227.<br />

Durrant W, Dong X (2004) <strong>System</strong>ic acquired resistance. Annu. Rev.<br />

Phytopathol. 42, 185–209.<br />

Durrett TP, Gassmann W, Rogers EE (2007) <strong>The</strong> FRD3-Mediated<br />

efflux of citrate into the root vasculature is necessary for efficient<br />

iron translocation. <strong>Plant</strong> Physiol. 144, 197–205.<br />

Edwards D, Davies KL, AXE L (1992) A vascular conducting str<strong>and</strong> in<br />

the early plant Cooksonia. Nature 357, 683–685.<br />

Edwards R (1994) Conjugation <strong>and</strong> metabolism of salicylic acid in<br />

tobacco. J. <strong>Plant</strong> Physiol. 143, 609–614.<br />

Effmert U, Saschenbrecker S, Ross J, Negre F, Fraser CM, Noel JP,<br />

Dudareva N, Piechulla B (2005) Floral benzenoid carboxyl methyltransferases:<br />

From in vitro to in planta function. Phytochemistry 66,<br />

1211–1230.<br />

Ejeta G (2007) Breeding for Striga resistance in sorghum: Exploitation<br />

of an intricate host-parasite biology. Crop Sci. 47, S216–S227.<br />

Emery JF, Floyd SK, Alvarez J, Eshed Y, Hawker NP, Izhaki A,<br />

Baum SF, Bowman JL (2003) Radial patterning of Arabidopsis


368 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

shoots by Class III HD-ZIP <strong>and</strong> KANADI genes. Curr. Biol. 13, 1768–<br />

1774.<br />

Endo S, Demura T, Fukuda H (2001) Inhibition of proteasome activity<br />

by the TED4 protein in extracellular space: A novel mechanism for<br />

protection of living cells from injury caused by dying cells. <strong>Plant</strong> Cell<br />

Physiol. 42, 9–19.<br />

Enyedi AJ, Yalpani N, Silverman P, Raskin I (1992) Localization,<br />

conjugation, <strong>and</strong> function of salicylic acid in tobacco during the<br />

hypersensitive reaction to tobacco mosaic virus. Proc. Natl. Acad.<br />

Sci. USA 89, 2480–2484.<br />

Eom JS, Cho JI, Reinder A, Lee SW, Yoo Y, Tuan PQ, Choi SB,<br />

Bang G, Park YI, Cho MH, Bhoo SH, An G, Hahn TR, Ward JM,<br />

Jeon JS (2011) Impaired function of the tonoplast-localized sucrose<br />

transporter in rice (Oryza sativa), OsSUT2, limits the transport of<br />

vacuolar reserve sucrose <strong>and</strong> affects plant growth. <strong>Plant</strong> Physiol.<br />

157, 109–119.<br />

Esau K (1965a) <strong>Plant</strong> Anatomy. John Wiley & Sons, Inc., New York.<br />

Esau K (1965b) <strong>Vascular</strong> Differentiation in <strong>Plant</strong>s. Holt, Rinehart <strong>and</strong><br />

Winston, New York.<br />

Esau K (1969) <strong>The</strong> Phloem. Encyclopedia of <strong>Plant</strong> Anatomy. Borntaeger,<br />

Berlin.<br />

Esau K, Cheadle VI, Gifford EM (1953) Comparative structure <strong>and</strong><br />

possible trends of specialization of the phloem. Am.J.Bot.40, 9–<br />

19.<br />

Eshed Y, Baum SF, Perea JV, Bowman JL (2001) Establishment of<br />

polarity in lateral organs of plants. Curr. Biol. 11, 1251–1260.<br />

Etchells JP, Provost CM, Turner SR (2012) <strong>Plant</strong> vascular cell<br />

division is maintained by an interaction between PXY <strong>and</strong> ethylene<br />

signalling. PLoS Genet. 8, e1002997.<br />

Etchells JP, Turner SR (2010) <strong>The</strong> pxy-cle41 receptor lig<strong>and</strong> pair defines<br />

a multifunctional pathway that controls the rate <strong>and</strong> orientation<br />

of vascular cell division. <strong>Development</strong> 137, 767–774.<br />

Evert RF (2006) Esau’s <strong>Plant</strong> Anatomy: Meristems, Cells, <strong>and</strong> Tissues<br />

of the <strong>Plant</strong> Body: <strong>The</strong>ir Structure, Function, <strong>and</strong> <strong>Development</strong>. John<br />

Wiley & Sons, Hoboken, NJ.<br />

Evert RF, Warmbrodt RD, Eichhorn SE (1989) Sieve-pore development<br />

in some leptosporangiate ferns. Am. J. Bot. 76, 1404–1413.<br />

Fan RC, Peng CC, Xu YH, Wang XF, Li Y, Shang Y, Du SY, Zhao R,<br />

Zhang XY, Zhang LY, Zhang DP (2009) Apple sucrose transporter<br />

SUT1 <strong>and</strong> sorbitol transporter SOT6 interact with cytochrome b5<br />

to regulate their affinity for substrate sugars. <strong>Plant</strong> Physiol. 150,<br />

1880–1901.<br />

Fan W, Dong X (2002) In vivo interaction between NPR1 <strong>and</strong> transcrip-<br />

tion factor TGA2 leads to salicylic acid-mediated gene activation in<br />

Arabidopsis. <strong>Plant</strong> Cell 14, 1377–1389.<br />

FAO (2008) Forest <strong>and</strong> energy. Food <strong>and</strong> Agriculture Organization of<br />

the United Nations. FAO Forestry Paper 154.<br />

FAO (2009) State of the world’s forests. FAO Report, Food <strong>and</strong><br />

Agriculture Organization of the United Nations. ISBN 978-92-5-<br />

106057-5.<br />

Farrar JF, Jones DL (2000) <strong>The</strong> control of carbon acquisition by roots.<br />

New Phytol. 147, 43–53.<br />

Feys BJ, Moisan LJ, Newman MA, Parker JE (2001) Direct interaction<br />

between the Arabidopsis disease resistance signaling proteins,<br />

EDS1 <strong>and</strong> PAD4. EMBO J. 20, 5400–5411.<br />

Feys BJ, Wiermer M, Bhat RA, Moisan LJ, Medina-Escobar N,<br />

Neu C, Cabral A, Parker JE (2005) Arabidopsis SENESCENCE-<br />

ASSOCIATED GENE101 stablizes <strong>and</strong> signals within an EN-<br />

HANCED DISEASE SUSCEPTIBILITY1 complex in plant innate<br />

immunity. <strong>Plant</strong> Cell 17, 2601–2613.<br />

Fiers M, Ku KL, Liu CM (2007) CLE peptide lig<strong>and</strong>s <strong>and</strong> their roles in<br />

establishing meristems. Curr. Opin. <strong>Plant</strong> Biol. 10, 39–43.<br />

Fisher DB (2000) Long-Distance Transport. In: B Buchanan, W Grissem,<br />

R Jones, eds. Biochemistry & Molecular Biology of <strong>Plant</strong>s.<br />

American Soc. of <strong>Plant</strong> Biologists, Maryl<strong>and</strong>, pp. 730–785.<br />

Fisher DB, Cash-Clarke CE (2000a) Sieve tube unloading, postphloem<br />

transport of fluorescent tracers, proteins injected into sieve<br />

tubes severed aphid stylets. <strong>Plant</strong> Physiol. 123, 125–137.<br />

Fisher DB, Cash-Clarke CE (2000b) Gradients in water potential,<br />

turgor pressure along the translocation pathway during grain filling<br />

in normally watered, water-stressed wheat plants. <strong>Plant</strong> Physiol.<br />

123, 139–147.<br />

Fisher DB, Wang N (1995) Sucrose concentration gradients along the<br />

post-phloem transport pathway in the maternal tissue of developing<br />

wheat grains. <strong>Plant</strong> Physiol. 109, 587–592.<br />

Fisher DB, Wu K, Ku MSB (1992) Turnover of soluble proteins in the<br />

wheat sieve tube. <strong>Plant</strong> Physiol. 100, 1433–1441.<br />

Fisher K, Turner S (2007) Pxy, a receptor-like kinase essential for<br />

maintaining polarity during plant vascular-tissue development. Curr.<br />

Biol. 17, 1061–1066.<br />

Floyd SK, Bowman JL (2004) Gene regulation: Ancient microrna<br />

target sequences in plants. Nature 428, 485–486.<br />

Floyd SK, Zalewski CS, Bowman JL (2006) <strong>Evolution</strong> of class iii<br />

homeodomain-leucine zipper genes in streptophytes. Genetics 173,<br />

373–388.<br />

Franco-Zorrilla JM, Martín AC, Leyva A, Paz-Ares J (2005) Interaction<br />

between phosphate-starvation, sugar, <strong>and</strong> cytokinin signaling<br />

in Arabidopsis <strong>and</strong> the roles of cytokinin receptors CRE1/AHK4 <strong>and</strong><br />

AHK3. <strong>Plant</strong> Physiol. 38, 847–57.<br />

Franco-Zorrilla JM, Martin AC, Solano R, Rubio V, Leyva A, Paz-<br />

Ares J (2002) Mutations at CRE1 impair cytokinin-induced repression<br />

of phosphate starvation responses in Arabidopsis. <strong>Plant</strong> J. 32,<br />

353–360.<br />

Franks P, Brodribb T (2005) Stomatal control <strong>and</strong> water transport in<br />

stems. In: Holbrook NM, Zwieniecki MA, eds. <strong>Vascular</strong> Transport in<br />

<strong>Plant</strong>s. Elsevier, Amsterdam. pp. 69–89.<br />

Fraysse LC, Wells B, McCann MC, Kjellbom P (2005) Specific plasma<br />

membrane aquaporins of the PIP1 subfamily are expressed in sieve<br />

elements <strong>and</strong> guard cells. Biol. Cell 97, 519–534.<br />

Friedrich L, Vernooij E, Gaffney T, Morse A, Ryals J (1995)<br />

Characterization of tobacco plants expressing a bacterial salicylate<br />

hydroxylase gene. <strong>Plant</strong> Mol. Biol. 29, 959–968.<br />

Froelich DR, Mullendore DL, Jensen KH, Ross-Elliot TJ, Anstead<br />

JA, Thompson GA, Pélissier HC, Knoblauch M (2011) Phloem


ultrastructure <strong>and</strong> pressure flow: Sieve-element-occlusion-related<br />

agglomerations do not affect translocation. <strong>Plant</strong> Cell 23, 4428–<br />

4445.<br />

Fu ZQ, Yan S, Saleh A, Wang W, Ruble J, Oka N, Mohan R,<br />

Spoel SH, Tada Y, Zheng Y, Dong X (2012) NPR3 <strong>and</strong> NPR4<br />

are receptors for the immune signal salicylic acid in plants. Nature<br />

16, 228–232.<br />

Fukuda H (2000) Programmed cell death of tracheary elements as a<br />

paradigm in plants. <strong>Plant</strong> Mol. Biol. 44, 245–253.<br />

Fukuda H (2004) Signals that control plant vascular cell differentiation.<br />

Nat. Rev. Mol. Cell Biol. 5, 379–391.<br />

Fukuda H, Hirakawa Y, Sawa S (2007) Peptide signaling in vascular<br />

development. Curr. Opin. <strong>Plant</strong> Biol. 10, 477–482.<br />

Funk V, Kositsup B, Zhao C, Beers EP (2002) <strong>The</strong> Arabidopsis xylem<br />

peptidase XCP1 is a tracheary element vacuolar protein that may<br />

be a papain ortholog. <strong>Plant</strong> Physiol. 128, 84–94.<br />

Gabaldon C, Ros LVG, Pedreno MA, Barcelo AR (2005) Nitric oxide<br />

production by the differentiating xylem of Zinnia elegans. New<br />

Phytol. 165, 121–130.<br />

Gaffney T, Friedrich L, Vernooij B, Negmtto D, Nye G, Uknes S,<br />

Ward E, Kessmann H, Ryals J (1993) Requirement of salicylic<br />

acid for the induction of systemic acquired resistance. Science 261,<br />

754–756.<br />

Gallagher KL, Benfey PN (2009) Both the conserved GRAS domain<br />

<strong>and</strong> nuclear localization are required for SHORT-ROOT movement.<br />

<strong>Plant</strong> J. 57, 785–797.<br />

García AV, Blanvillain-Baufumé S,HuibersRP,WiermerM,LiG,<br />

Gobbato E, Rietz S, Parker JE (2010) Balanced nuclear <strong>and</strong><br />

cytoplasmic activities of EDS1 are required for a complete plant<br />

innate immune response. PLoS Pathogens 6, e1000970<br />

Garcia-Mas J, Benjak A, Sanseverino W, Bourgeois M, Mir G,<br />

González VM, Hénaff E, Câmara F, Cozzuto L, Lowy E, Alioto<br />

T, Capella-Gutiérrez S, Blanca J, Cañizares J, Ziarsolo P,<br />

Gonzalez-Ibeas D, Rodríguez-Moreno L, Droege M, Du L,<br />

Alvarez-Tejado M, Lorente-Galdos B, Melé M, Yang L, Weng<br />

YQ, Navarro A, Marques-Bonet T, Ar<strong>and</strong>a MA, Nuez F, Picó<br />

B, Gabaldón T, Roma G, Guigó R, Casacuberta JM, Arús P,<br />

Puigdomènech P (2012) <strong>The</strong> genome of melon (Cucumis melo L.)<br />

Proc. Natl. Acad. Sci. USA 109, 11872–11877.<br />

Gardiner J, Donner TJ, Scarpella E (2011) Simultaneous activation<br />

of SHR <strong>and</strong> ATHB8 expression defines switch to preprocambial cell<br />

state in Arabidopsis leaf development. Dev. Dynam. 240, 261–270.<br />

Gaupels F, Buhtz A, Knauer T, Deshmurkh S, Waller F, van Bel<br />

AJE, Kogel KH, Kehr J (2008a) Adaption of aphid stylectomy for<br />

analyses of proteins <strong>and</strong> mRNA in barley phloem sap. J. Exp. Bot.<br />

59, 3297–3306.<br />

Gaupels F, Knauer T, van Bel AJE (2008b) A combinatory approach<br />

for analysis of protein sets in barley sieve-tube samples using<br />

EDTA-facilitated exudation <strong>and</strong> aphid stylectomy. J. <strong>Plant</strong> Physiol.<br />

165, 95–103.<br />

Gendre D, Czernic P, Conejero G, Pianelli K, Briat J, Lebrun M, Mari<br />

S (2007) TcYSL3, a member of the YSL gene family from the hyper-<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 369<br />

accumulator Thlaspi caerulescens, encodes a nicotianamine-Ni/Fe<br />

transporter. <strong>Plant</strong> J. 49, 1–15.<br />

Genoud T, Buchala AJ, Chua NH, Metraux JP (2002) Phytochrome<br />

signalling modulates the SA-perceptive pathway in Arabidopsis.<br />

<strong>Plant</strong> J. 31, 87–95.<br />

Ghanem ME, Albacete A, Smigocki AC, Frébort I, Pospísilová<br />

H, Martínez-Andújar C, Acosta M, Sánchez-Bravo J, Lutts S,<br />

Dodd IC, Pérez-Alfocea F (2011) Root-synthesized cytokinins<br />

improve shoot growth <strong>and</strong> fruit yield in salinized tomato (Solanum<br />

lycopersicum L.) plants. J. Exp. Bot. 62, 125–140.<br />

Giaquinta RT (1983) Phloem loading of sucrose. Annu. Rev. <strong>Plant</strong><br />

Physiol. 34, 347–387.<br />

Giavalisco P, Kapitza K, Kolasa A, Buhtz A, Kehr J (2006) Towards<br />

theproteomeofBrassica napus phloem sap. Proteomics 6, 896–<br />

909.<br />

Glazebrook J, Ausubel FM (1994) Isolation of phytoalexin-deficient<br />

mutants of Arabidopsis thaliana <strong>and</strong> characterization of their interactions<br />

with bacterial pathogens. Proc. Natl. Acad. Sci. USA 91,<br />

8955–8959.<br />

Glazebrook J, Rogers EE, Ausubel FM (1996) Isolation of Arabidopsis<br />

mutants with enhanced disease susceptibility by direct screening.<br />

Genetics 143, 973–982.<br />

Golecki B, Schulz A, Carstens-Behrens U, Kollmann R (1998)<br />

Evidence for graft transmission of structural phloem proteins or<br />

their precursors in heterografts of Cucurbitaceae. <strong>Plant</strong>a 206,<br />

630–640.<br />

Golecki B, Schulz A, Thompson GA (1999) Translocation of structural<br />

P proteins in the phloem. <strong>Plant</strong> Cell 11, 127–140.<br />

Gomez-Roldan V, Fermas S, Brewer PB, Puech-Pagès V, Dun<br />

EA, Pillot JP, Letisse F, Matusova R, Danoun S, Portais<br />

JC, Bouwmeester H, Bécard G, Beveridge CA, Rameau C,<br />

Rochange SF (2008) Strigolactone inhibition of shoot branching.<br />

Nature 455, 189–194.<br />

Gould N, Minchin PEH, Thorpe MR (2004a) Direct measurements of<br />

sieve element hydrostatic pressure reveal strong regulation after<br />

pathway blockage. Funct. <strong>Plant</strong> Biol. 31, 987–993.<br />

Gould N, Thorpe MR, Koroleva O, Minchin PEH (2005) Phloem<br />

hydrostatic pressure relates to solute loading rate: A direct test of<br />

the Münch hypothesis. Funct. <strong>Plant</strong> Biol. 32, 1019–1026.<br />

Gould N, Thorpe MR, Minchin PEH, Pritchard J, White PJ (2004b)<br />

Solute is imported to elongating root cells of barley as a pressure<br />

driven-flow of solution. Funct. <strong>Plant</strong> Biol. 31, 391–397.<br />

Grimmer C, Komor E (1999) Assimilate export by leaves of Ricinus<br />

communis L. growing under normal <strong>and</strong> elevated carbon dioxide<br />

concentrations: <strong>The</strong> same rate during the day, a different rate at<br />

night. <strong>Plant</strong>a 209, 275–281.<br />

Grodzinski B, Jiao J, Leonardos ED (1998) Estimating photosynthesis<br />

<strong>and</strong> concurrent export rates in C3 <strong>and</strong> C4 species at ambient<br />

<strong>and</strong> elevated CO2. <strong>Plant</strong> Physiol. 117, 207–215.<br />

Grusak MA (1994) Iron transport to developing ovules of Pisum<br />

sativum. (I. seed import characteristics <strong>and</strong> phloem iron-loading<br />

capacity of source regions). <strong>Plant</strong> Physiol. 104, 649–655.


370 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Guelette BS, Benning UF, Hoffmann-Benning S (2012) Identification<br />

of lipids <strong>and</strong> lipid-binding proteins in phloem exudates from<br />

Arabidopsis thaliana. J. Exp. Bot. 63, 3603–3616.<br />

Guo S, Zhang J, Sun H, Salse J, Lucas WJ, Zhang H, Zheng Y,<br />

Mao L, Ren Y, Wang Z, Min J, Guo X, Murat F, Ham BK, Zhang<br />

Z, Gao S, Huang M, Xu Y, Zhong S, Bombarely A, Mueller LA,<br />

Zhao H, He H, Zhang Y, Zhang Z, Huang S, Tan T, Pang E, Lin<br />

K, Hu Q, Kuang H, Ni P, Wang B, Liu J, Kou Q, Hou W, Zou X,<br />

Jiang J, Gong G, Klee K, Schoof H, Huang Y, Hu X, Dong S,<br />

Liang D, Wang J, Wu K, Xia Y, Zhao X, Zheng Z, Xing M, Liang<br />

X, Huang B, Lv T, Wang J, Yin Y, Yi H, Li R, Wu M, Levi A,<br />

Zhang X, Giovannoni JJ, Wang J, Li Y, Fei Z, Xu Y (2012) <strong>The</strong><br />

draft genome of watermelon (Citrullus lunatus) <strong>and</strong> resequencing of<br />

20 diverse accessions. Nat. Genet. 45, 51–58.<br />

Guo WJ, Bundithya W, Goldsbrough PB (2003) Characterization<br />

of the Arabidopsis metallothionein gene family: Tissue-specific<br />

expression <strong>and</strong> induction during senescence <strong>and</strong> in response to<br />

copper. New Phytol. 159, 369–381.<br />

Guo WJ, Meetam M, Goldsbrough PB (2008) Examining the specific<br />

contributions of individual Arabidopsis metallothioneins to copper<br />

distribution <strong>and</strong> metal tolerance. <strong>Plant</strong> Physiol. 146, 1697–<br />

1706.<br />

Hacke U, Sauter JJ (1996) Xylem dysfunction during winter <strong>and</strong><br />

recovery of hydraulic conductivity in diffuse-porous <strong>and</strong> ring-porous<br />

trees. Oecologia 105, 435–439.<br />

Hacke UG, Sperry JS (2003) Limits to xylem refilling under negative<br />

pressure in Laurus nobilis <strong>and</strong> Acer negundo. <strong>Plant</strong> Cell Environ.<br />

26, 303–311.<br />

Hacke UG, Sperry JS, Pockman WP, Davis SD, McCulloh KA<br />

(2001a) Trends in wood density <strong>and</strong> structure are linked to prevention<br />

of xylem implosion by negative pressure. Oecologia 126,<br />

457–461.<br />

Hacke UG, Sperry JS, Wheeler JK, Castro L (2006) Scaling of<br />

angiosperm xylem structure with safety <strong>and</strong> efficiency. Tree Physiol.<br />

26, 689–701.<br />

Hacke UG, Stiller V, Sperry JS, Pittermann J, McCulloh KA<br />

(2001b) Cavitation fatigue: Embolism <strong>and</strong> refilling cycles can<br />

weaken cavitation resistance of xylem. <strong>Plant</strong> Physiol. 125, 779–<br />

786.<br />

Hacke UG, Jansen S (2009) Embolism resistance of three boreal<br />

conifers varies with pit structure. New Phytol. 182, 675–686.<br />

Ham B-K, Br<strong>and</strong>om J, Xoconostle-Cázares, B, Ringgold V, Lough<br />

TJ, Lucas WJ (2009) Polypyrimidine tract binding protein, Cm-<br />

RBP50, forms the basis of a pumpkin phloem ribonucleoprotein<br />

complex. <strong>Plant</strong> Cell 21, 197–215.<br />

Hamburger D, Rezzonico E, MacDonald-Comber Petétot J,<br />

Somerville C, Poirier Y (2002) Identification <strong>and</strong> characterization<br />

of the Arabidopsis PHO1 gene involved in phosphate loading to the<br />

xylem. <strong>Plant</strong> Cell 14, 889–902.<br />

Hamilton AJ, Baulcombe DC (1999) A species of small antisense<br />

RNA in posttranscriptional gene silencing in plants. Science 286,<br />

950–952.<br />

Han JJ, Lin W, Oda Y, Cui KM, Fukuda H, He XQ (2012) <strong>The</strong><br />

proteasome is responsible for caspase-3-like activity during xylem<br />

development. <strong>Plant</strong> J. 72, 129–141.<br />

Hanikenne M, Talke IN, Haydon MJ, Lanz C, Nolte A, Motte P,<br />

Kroymann J, Weigel D, Krämer U (2008) <strong>Evolution</strong> of metal<br />

hyperaccumulation required cis-regulatory changes <strong>and</strong> triplication<br />

of HMA4. Nature 453, 391–395.<br />

Hannapel DJ (2010) A model system of development regulated by the<br />

long-distance transport of mRNA. J. Integr. <strong>Plant</strong> Biol. 52, 40–52.<br />

Hanzawa Y, Takahashi T, Komeda Y (1997) ACL5: AnArabidopsis<br />

gene required for internodal elongation after flowering. <strong>Plant</strong> J. 12,<br />

863–874.<br />

Hardtke CS, Berleth T (1998) <strong>The</strong> Arabidopsis gene MONOPTEROS<br />

encodes a transcription factor mediating embryo axis formation <strong>and</strong><br />

vascular development. EMBO J. 17, 1405–1411.<br />

Hardtke CS, Ckurshumova W, Vidaurre DP, Singh SA, Stamatiou G,<br />

Tiwari SB, Hagen G, Guilfoyle TJ, Berleth T (2004) Overlapping<br />

<strong>and</strong> non-redundant functions of the Arabidopsis auxin response<br />

factors MONOPTEROS <strong>and</strong> NONPHOTOTROPIC HYPOCOTYL<br />

4. <strong>Development</strong> 131, 1089–1100.<br />

Haritatos E, Medville R, Turgeon R (2000) Minor vein structure <strong>and</strong><br />

sugar transport in Arabidopsis thaliana. <strong>Plant</strong>a 211, 105– 111.<br />

Hassan Z, Aarts MGM (2011) Opportunities <strong>and</strong> feasibilities for<br />

biotechnological improvement of Zn, Cd or Ni tolerance <strong>and</strong> accumulation<br />

in plants. Environ. Exp. Bot. 72, 53–63.<br />

Hawkesford MJ (2003) Transporter gene families in plants: <strong>The</strong><br />

sulphate transporter gene family – redundancy or specialization?<br />

Physiol. <strong>Plant</strong>. 117, 155–163.<br />

Haywood V, Yu TS, Huang NC, Lucas WJ (2005) Phloem longdistance<br />

trafficking of GIBBERELLIC ACID INSENSITIVE RNA<br />

regulates leaf development. <strong>Plant</strong> J. 42, 49–68.<br />

He Y, Gan S (2002) A gene encoding an acyl hydrolase is involved in<br />

leaf senescence in Arabidopsis. <strong>Plant</strong> Cell 14, 805–815.<br />

Heidel AJ, Clarke JD, Antonovics J, Dong X (2004) Fitness costs<br />

of mutations affecting the systemic acquired resistance pathway in<br />

Arabidopsis thaliana. Genetics 168, 2197–2206.<br />

Heidrich K, Wirthmueller L, Tasset C, Pouzet C, Desl<strong>and</strong>es L,<br />

Parker JE (2011) Arabidopsis EDS1 connects pathogen effector<br />

recognition to cell compartment-specific immune responses.<br />

Science 334, 1401–1404.<br />

Heil M, Baldwin IT (2002) Fitness costs of induced resistance: Emerging<br />

experimental support for a slippery concept. Trends <strong>Plant</strong> Sci.<br />

7, 61–66.<br />

Herbette S, Cochard H (2010) Calcium is a major determinant of xylem<br />

vulnerability to cavitation. <strong>Plant</strong> Physiol. 153, 1932–1939.<br />

Herbik A, Giritch A, Horstmann C, Becker R, Balzer HJ, Bäumlein H,<br />

Stephan UW (1996) Iron <strong>and</strong> copper nutrition-dependent changes<br />

in protein expression in a tomato wild type <strong>and</strong> the nicotianaminefree<br />

mutant chloronerva. <strong>Plant</strong> Physiol. 111, 533–540.<br />

Hermans C, Hammond JP, White PJ, Verbuggen N (2006) How do<br />

plants respond to nutrient shortage by biomass allocation? Trends<br />

<strong>Plant</strong> Sci. 11, 610–617.


Hickey L (1973) Classification of the architecture of dicotyledonous<br />

leaves. Amer. J. Bot. 60, 17–33.<br />

Hirakawa Y, Kondo Y, Fukuda H (2010) Regulation of vascular<br />

development by CLE peptide-receptor systems. J. Integr. <strong>Plant</strong> Biol.<br />

52, 8–16.<br />

Hirakawa Y, Kondo Y, Fukuda H (2010) TDIF peptide signaling<br />

regulates vascular stem cell proliferation via the WOX4 homeobox<br />

gene in Arabidopsis. <strong>Plant</strong> Cell 22, 2618–2629.<br />

Hirakawa Y, Kondo Y, Fukuda H (2011) Establishment <strong>and</strong> maintenance<br />

of vascular cell communities through local signaling. Curr.<br />

Opin. <strong>Plant</strong> Biol. 14, 17–23.<br />

Hirakawa Y, Shinohara H, Kondo Y, Inoue A, Nakanomyo I, Ogawa<br />

M, Sawa S, Ohashi-Ito K, Matsubayashi Y, Fukuda H (2008)<br />

Non-cell-autonomous control of vascular stem cell fate by a CLE<br />

peptide/receptor system. Proc. Natl. Acad. Sci. USA 105, 15208–<br />

15213.<br />

Hirose N, Takei K, Kuroha T, Kamada-Nobusada T, Hayashi H,<br />

Sakakibara H (2008) Regulation of cytokinin biosynthesis, compartmentalization<br />

<strong>and</strong> translocation. J. Exp. Bot. 59, 75–83.<br />

Hoad GV (1995) Transport of hormones in the phloem of higher plants.<br />

<strong>Plant</strong> Growth Reg. 16, 173–182.<br />

Hoffman-Thoma G, van Bel AJE, Ehlers K (2001) Ultrastructure of<br />

minor-vein phloem <strong>and</strong> assimilate export in summer <strong>and</strong> winter<br />

leaves of the symplasmically loading evergreens Ajuga reptans L.,<br />

Aucuba japonica Thumb. <strong>and</strong> Hedrix helix L. <strong>Plant</strong>a 212, 231–242.<br />

Holbrook NM, Shashidhar VR, James RA, Munns R (2002) Stomatal<br />

control in tomato with ABA-deficient roots: Response of grafted<br />

plants to soil drying. J. Exp. Bot. 53, 1503–1514.<br />

Holtta T, Mencuccini M, Nikinmaa E (2011) A carbon cost-gain model<br />

explains the observed patterns of xylem safety <strong>and</strong> efficiency. <strong>Plant</strong><br />

Cell Environ. 34, 1819–1834.<br />

Holtta T, Juurola E, Lindfors L, Porcar-Castell A (2012) Cavitation<br />

induced by a surfactant leads to transient release of water stress<br />

<strong>and</strong> subsequent “run away” embolism in scots pine (Pinus sylvestris)<br />

seedlings. J. Exp. Bot. 63, 1057–1067.<br />

Hu G, deHart AKA, Li Y, Ustach C, H<strong>and</strong>ley V, Navarre R, Hwang<br />

CF, Aegerter BJ, Williamson VM, Baker B (2005) EDS1 in tomato<br />

is required for resistance mediated by TIR-class R genes <strong>and</strong> the<br />

receptor-like R gene Ve. <strong>Plant</strong> J. 42, 376–391.<br />

Hu H, Penn SG, Lebrilla CB, Brown PH (1997) Isolation <strong>and</strong> characterization<br />

of soluble boron complexes in higher plants: <strong>The</strong> mechanism<br />

of phloem mobility of boron. <strong>Plant</strong> Physiol. 113, 649–655.<br />

Hu L, Sun H, Li R, Zhang L, Wang S, Sui X, Zhang Z (2011) Phloem<br />

unloading follows an extensive apoplasmic pathway in cucumber<br />

(Cucumis sativa L.) fruit from anthesis to marketable maturing stage.<br />

<strong>Plant</strong> Cell Environ. 40, 743–748.<br />

Huang SW, Li RQ, Zhang ZH, Li L, Gu XF, Fan W, Lucas WJ,<br />

Wang XW, Xie BY, Ni PX, Ren YY, Zhu HM, Li J, Lin K, Jin<br />

WW, Fei ZJ, Li GC, Staub J, Kilian A, van der Vossen EAG,<br />

Wu Y, Guo J, He J, Jia ZQ, Ren Y, Tian G, Lu Y, Ruan J, Qian<br />

WB, Wang MW, Huang QF, Li B, Xuan ZL, Cao JJ, Asan, Wu<br />

ZG, Zhang JB, Cai QL, Bai YQ, Zhao BW, Han YH, Li Y, Li<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 371<br />

XF, Wang SH, Shi QX, Liu SQ, Cho WK, Kim JY, Xu Y, Heller-<br />

Uszynska K, Miao H, Cheng ZC, Zhang SP, Wu J, Yang YH,<br />

KangHX,LiM,LiangHQ,RenXL,ShiZB,WenM,JianM,<br />

Yang HL, Zhang GJ, Yang ZT, Chen R, Liu SF, Li JW, Ma LJ,<br />

Liu H, Zhou Y, Zhao J, Fang XD, Li GQ, Fang L, Li YR, Liu DY,<br />

Zheng HK, Zhang Y, Qin N, Li Z, Yang GH, Yang S, Bolund L,<br />

Kristiansen K, Zheng HC, Li SC, Zhang XQ, Yang HM, Wang J,<br />

Sun RF, Zhang BX, Jiang SZ, Wang J, Du YC, Li SG (2009) <strong>The</strong><br />

genome of the cucumber, Cucumis sativus L. Nat. Genet. 41, 1275–<br />

1281.<br />

Huang NC, Yu TS (2009) <strong>The</strong> sequences of Arabidopsis GA-<br />

INSENSITIVE RNA constitute the motifs that are necessary <strong>and</strong><br />

sufficient for RNA long-distance trafficking. <strong>Plant</strong> J. 59, 921–929.<br />

Huang NC, Jane WN, Chen J, Yu TS (2012) Arabidopsis thaliana<br />

CENTRORADIALIS homologue (ATC) acts systemically to inhibit<br />

floral initiation in Arabidopsis. <strong>Plant</strong> J. 72, 175–184.<br />

Hubbard RM, Stiller V, Ryan MG, Sperry JS (2001) Stomatal conductance<br />

<strong>and</strong> photosynthesis vary linearly with plant hydraulic<br />

conductance in ponderosa pine. <strong>Plant</strong> Cell Environ. 24, 113–<br />

121.<br />

Husb<strong>and</strong>s AY, Chitwood DH, Plavskin Y, Timmermans MC (2009)<br />

Signals <strong>and</strong> prepatterns: New insights into organ polarity in plants.<br />

Genes Dev. 23, 1986–1997.<br />

Hussain D, Haydon MJ, Wang Y, Wong E, Sherson SM, Young<br />

J, Camakaris J, Harper JF, Cobbett CS (2004) P-type ATPase<br />

heavy metal transporters with roles in essential zinc homeostasis in<br />

Arabidopsis. <strong>Plant</strong> Cell 16, 1327–1339.<br />

Ilegems M, Douet V, Meylan-Bettex M, Uyttewaal M, Br<strong>and</strong> L,<br />

Bowman JL, Stieger PA (2010) Interplay of auxin, kanadi <strong>and</strong><br />

class iii hd-zip transcription factors in vascular tissue formation.<br />

<strong>Development</strong> 137, 975–984.<br />

Imai A, Hanzawa Y, Komura M, Yamamoto KT, Komeda Y, Takahashi<br />

T (2006) <strong>The</strong> dwarf phenotype of the Arabidopsis acl5 mutant<br />

is suppressed by a mutation in an upstream ORF of a bHLH gene.<br />

<strong>Development</strong> 133, 3575–3585.<br />

Imai A, Komura M, Kawano E, Kuwashiro Y Takahashi T (2008)<br />

A semi-dominant mutation in the ribosomal protein L10 gene<br />

suppresses the dwarf phenotype of the acl5 mutant in Arabidopsis<br />

thaliana. <strong>Plant</strong> J. 56, 881–890.<br />

Ingle RA, Mugford ST, Rees JD, Campbell MM, Smith JAC (2005)<br />

Constitutively high expression of the histidine biosynthetic pathway<br />

contributes to nickel tolerance in hyperaccumulator plants. <strong>Plant</strong><br />

Cell 17, 2089–2106.<br />

Ingram P, Dettmer J, Helariutta Y, Malamy JE (2011) Arabidopsis<br />

lateral root development 3 is essential for early phloem development<br />

<strong>and</strong> function, <strong>and</strong> hence for normal root system development. <strong>Plant</strong><br />

J. 68, 455–467.<br />

Irtelli B, Petrucci WA, Navari-Izzo F (2009) Nicotianamine <strong>and</strong><br />

histidine/proline are, respectively, the most important copper<br />

chelators in xylem sap of Brassica carinata under conditions<br />

of copper deficiency <strong>and</strong> excess. J. Exp. Bot. 60, 269–<br />

277.


372 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Ishihara T, Sekine KT, Hase S, Kanayama Y, Seo S, Ohashi Y,<br />

Kusano T, Shibata D, Shah J, Takahashi H (2008) Overexpression<br />

of the Arabidopsis thaliana EDS5 gene enhances resistance to<br />

viruses. <strong>Plant</strong> Biol. 10, 451–461.<br />

Ito J, Fukuda H (2002) ZEN1 is a key enzyme in the degradation of<br />

nuclear DNA during programmed cell death of tracheary elements.<br />

<strong>Plant</strong> Cell 14, 3201–3211.<br />

Ito Y, Nakanomyo I, Motose H, Iwamoto K, Sawa S, Dohmae N,<br />

Fukuda H (2006) Dodeca-cle peptides as suppressors of plant stem<br />

cell differentiation. Science 313, 842–845.<br />

Ivanov R, Brumbarova T, Bauer P (2012) Fitting into the harsh reality:<br />

Regulation of iron-deficiency responses in dicotyledonous plants.<br />

Mol. <strong>Plant</strong> 5, 27–42.<br />

Iwai H, Usui M, Hoshino H, Kamada H, Matsunaga T, Kakegawa K,<br />

IshiiT,SatohS(2003) Analysis of sugars in squash xylem sap.<br />

<strong>Plant</strong> Cell Physiol. 44, 582–587.<br />

Izhaki A, Bowman JL (2007) KANADI <strong>and</strong> Class III HD-ZIP gene<br />

families regulate embryo patterning <strong>and</strong> modulate auxin flow during<br />

embryogenesis in Arabidopsis. <strong>Plant</strong> Cell 19, 495–508.<br />

Jack E, Hakvoort HWJ, Reumer A, Verkleij JAC, Schat H,<br />

Ernst WHO (2007) Real-time PCR analysis of metallothionein-<br />

2b expression in metallicolous <strong>and</strong> non-metallicolous populations<br />

of Silene vulgaris (Moench) Garcke. Environ. Exp. Bot. 59,<br />

84–91.<br />

Jacobson AL, Pratt RB (2012) No evidence for an open vessel effect in<br />

centrifuge-based vulnerability curves of a long-vesseled liana (Vitis<br />

vinifera). New Phytol. 194, 982–990.<br />

Jacobson AL, Pratt RB, Ewers FW, Davis SD (2007) Cavitation<br />

resistance among 26 chaparral species of southern California. Ecol.<br />

Mon. 77, 99–115.<br />

Jansen S, Baas P, Smets E (2001) Vestured pits: <strong>The</strong>ir occurrence<br />

<strong>and</strong> systematic importance in eudicots. Taxon 50, 135–167.<br />

Jansen S, Choat B, Pletsers A (2009) Morphological variation in<br />

intervessel pit membranes <strong>and</strong> implications to xylem function in<br />

angiosperms. Am. J. Bot. 96, 409–419.<br />

Jansen S, Lamy JB, Burlett R, Cochard H, Gasson P, Delzon S<br />

(2012) Plasmodesmatal pores in the torus of bordered pit membranes<br />

affect cavitation resistance of conifer xylem. <strong>Plant</strong> Cell<br />

Environ. 35, 1109–1120.<br />

Jarbeau JA, Ewers FW, Davis SD (1995) <strong>The</strong> mechanism of waterstress-induced<br />

embolism in two species of chaparral shrubs. <strong>Plant</strong><br />

Cell Environ. 18, 189–196.<br />

Jensen KH, Lee J, Bohr T, Bruus H, Holbrook NM, Zwieniecki<br />

MA (2011) Optimality of the Münch mechanism for translocation<br />

of sugars in plants. J. Royal Soc. Interface 8, 1155–1165.<br />

Jensen KH, Liesche J, Bohr T, Schulz A (2012) Universality of phloem<br />

transport in seed plants. <strong>Plant</strong> Cell Environ. 35, 1065–1076.<br />

Jeong R-D, Ch<strong>and</strong>ra-Shekara AC, Barman SR, Navarre DA, Klessig<br />

D, Kachroo A, Kachroo P (2010) CRYPTOCHROME 2 <strong>and</strong> PHO-<br />

TOTROPIN 2 regulate resistance protein mediated viral defense by<br />

negatively regulating a E3 ubiquitin ligase. Proc. Natl. Acad. Sci.<br />

USA 107, 13538–13543.<br />

Ji J, Strable J, Shimizu R, Koenig D, Sinha N, Scanlon MJ (2010)<br />

WOX4 promotes procambial development. <strong>Plant</strong> Physiol. 152,<br />

1346–1356.<br />

Jia W, Davies WJ (2008) Modification of leaf apoplastic pH in relation<br />

to stomatal sensitivity to root-sourced abscisic acid signals. <strong>Plant</strong><br />

Physiol. 143, 68–77.<br />

Jiang F, Hartung W (2008) Long-distance signalling of abscisic acid<br />

(ABA): <strong>The</strong> factors regulating the intensity of the ABA signal. J. Exp.<br />

Bot. 59, 37–43.<br />

Johri M (2008) Hormonal regulation in green plant lineage families.<br />

Physiol. Mol. Biol. <strong>Plant</strong>s 14, 23–38.<br />

Jones JD, Dangl JL (2006) <strong>The</strong> plant immune system. Nature 444,<br />

323–329.<br />

Jorgensen RA, Atkinson RG, Forster RLS, Lucas WJ (1998)<br />

An RNA-based information superhighway in plants. Science 279,<br />

1486–1487.<br />

Jun JH, Fiume E, Fletcher JC (2008) <strong>The</strong> CLE family of plant<br />

polypeptide signaling molecules. Cell Mol. Life Sci. 65, 743–755.<br />

Jung HW, Tschaplinkski TJ, Wang L, Glazebrook J, Greenberg JT<br />

(2009) Priming in systemic plant immunity. Science 324, 89–91.<br />

Kachroo A, Kachroo P (2006) Salicylic acid-, jasmonic acid- <strong>and</strong><br />

ethylene-mediated regulation of plant defense signaling. In: Setlow<br />

J, ed. Genetic Engineering, Principles <strong>and</strong> Methods. 28, 55–83.<br />

Kachroo A, Lapchyk L, Fukushigae H, Hildebr<strong>and</strong> D, Klessig D,<br />

Kachroo P (2003) Plastidial fatty acid signaling modulates salicylic<br />

acid- <strong>and</strong> jasmonic acid-mediated defense pathways in the Arabidopsis<br />

ssi2 mutant. <strong>Plant</strong> Cell 12, 2952–2965.<br />

Kachroo A, Venugopal SC, Lapchyk L, Falcone D, Hildebr<strong>and</strong><br />

D, Kachroo P (2004) Oleic acid levels regulated by glycerolipid<br />

metabolism modulate defense gene expression in Arabidopsis.<br />

Proc. Natl. Acad. Sci. USA 101, 5152–5257.<br />

Kachroo P, Kachroo A, Lapchyk L, Hildebr<strong>and</strong> D, Klessig D (2003)<br />

Restoration of defective cross talk in ssi2 mutant: Role of salicylic<br />

acid, jasmonic acid, <strong>and</strong> fatty acids in SSI2-mediated signaling. Mol.<br />

<strong>Plant</strong>-Microbe Interact. 11, 1022–1029.<br />

Kachroo P, Shanklin J, Shah J, Whittle E, Klessig D (2001)<br />

A fatty acid desaturase modulates the activation of defense<br />

signaling pathways in plants. Proc. Natl. Acad. Sci. USA 98,<br />

9448–9453.<br />

Kachroo P, Srivathsa CV, Navarre DA, Lapchyk L, Kachroo A (2005)<br />

Role of salicylic acid <strong>and</strong> fatty acid desaturation pathways in ssi2mediated<br />

signaling. <strong>Plant</strong> Physiol. 139, 1717–1735.<br />

Kachroo P, Yoshioka K, Shah J, Dooner H, Klessig DF (2000)<br />

Resistance to turnip crinkle virus in Arabidopsis requires two host<br />

genes <strong>and</strong> is salicylic acid dependent but NPR1, ethylene <strong>and</strong><br />

jasmonate independent. <strong>Plant</strong> Cell 12, 677–690.<br />

Kakehi J, Kuwashiro Y, Motose H, Igarashi K, Takahashi T (2010)<br />

Norspermine substitutes for thermospermine in the control of stem<br />

elongation in Arabidopsis thaliana. FEBS Lett. 584, 3042–3046.<br />

Kakehi JI, Kuwashiro Y, Niitsu Y, Takahashi T (2008) <strong>The</strong>rmospermine<br />

is required for stem elongation in Arabidopsis thaliana. <strong>Plant</strong><br />

Cell Physiol. 49, 1342–1349.


Kallarackal J, Milburn JA (1984) Specific mass transfer, sinkcontrolled<br />

phloem translocation in castor bean. Aust. J. <strong>Plant</strong><br />

Physiol. 10, 561–568.<br />

Kang J, Dengler N (2002) Cell cycling frequency <strong>and</strong> expression<br />

of the homeobox gene ATHB-8 during leaf vein development in<br />

Arabidopsis. <strong>Plant</strong>a 216, 212–219.<br />

Kang J, Dengler N (2004) Vein pattern development in adult leaves of<br />

Arabidopsis thaliana. Int. J. <strong>Plant</strong> Sci. 165, 231–242.<br />

Kang J, Mizukami Y, Wang H, Fowke L, Dengler NG (2007) Modification<br />

of cell proliferation patterns alters leaf vein architecture in<br />

Arabidopsis thaliana. <strong>Plant</strong>a 226, 1207–1218.<br />

Karpinski S, Gabrys H, Mateo A, Karpinska B, Mullineaux P (2003)<br />

Light perception in plant disease defense signaling. Curr. Opin.<br />

<strong>Plant</strong> Biol. 6, 390–396.<br />

Kaufman I, Schulze-Till T, Schneider H, Zimmermann U, Jakob P,<br />

Wegner L (2009) Function repair of embolized vessels in maize<br />

roots after temporal drought stress, as demonstrated by magnetic<br />

resonance imaging. New Phytol. 184, 245–256.<br />

Kempers R, van Bel AJE (1997) Symplasmic connections between<br />

sieve element <strong>and</strong> companion cell in the stem of Vicia fava L.<br />

have a molecular exclusion limit of at least 10 kD. <strong>Plant</strong>a 201,<br />

195–201.<br />

Kenrick P, Crane PR (1997) <strong>The</strong> origin <strong>and</strong> early evolution of plants<br />

on l<strong>and</strong>. Nature 389, 33–39.<br />

Kerkeb L, Krämer U (2003) <strong>The</strong> role of free histidine in xylem loading<br />

of nickel in Alyssum lesbiacum <strong>and</strong> Brassica juncea. <strong>Plant</strong> Physiol.<br />

131, 716–724.<br />

Kerstetter RA, Bollman K, Taylor RA, Bomblies K, Poethig RS<br />

(2001) KANADI regulates organ polarity in Arabidopsis. Nature 411,<br />

706–709.<br />

Kiba T, Yamada H, Sato S, Kato T, Tabata S, Yamashino T,<br />

Mizuno T (2003) <strong>The</strong> type-A response regulator, ARR15, acts as a<br />

negative regulator in the cytokinin-mediated signal transduction in<br />

Arabidopsis thaliana. <strong>Plant</strong> Cell Physiol. 44, 868–874.<br />

Kim HS, Delaney TP (2002) Over-expression of TGA5, which encodes<br />

a bZIP transcription factor that interacts with NIM1/NPR1,<br />

confers SAR-independent resistance in Arabidopsis thaliana to<br />

Peronospora parasitica. <strong>Plant</strong> J. 32, 151–163.<br />

Kim M, Canio W, Kessler S, Sinha N (2001) <strong>Development</strong>al changes<br />

due to long-distance movement of a homeobox fusion transcript in<br />

tomato. Science 293, 287–289.<br />

Klatte M, Schuler M, Wirtz M, Fink-Straube C, Hell R, Bauer P<br />

(2009) <strong>The</strong> analysis of Arabidopsis nicotianamine synthase mutants<br />

reveals functions for nicotianamine in seed iron loading <strong>and</strong> iron<br />

deficiency responses. <strong>Plant</strong> Physiol. 150, 257–271.<br />

Kobayashi Y, Kuroda K, Kimura K, Southron-Francis JL, Furuzawa<br />

A, Iuchi S, Kobayashi M, Taylor GJ, Koyama H (2008) Amino acid<br />

polymorphisms in strictly conserved domains of a P-type ATPase<br />

HMA5 are involved in the mechanism of copper tolerance variation<br />

in Arabidopsis. <strong>Plant</strong> Physiol. 148, 969–980.<br />

Kohlen W, Charnikhova T, Liu Q, Bours R, Domagalska MA,<br />

Beguerie S, Verstappen F, Leyser O, Bouwmeester H, Ruyter-<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 373<br />

Spira C (2011) Strigolactones are transported through the xylem<br />

<strong>and</strong> play a key role in shoot architectural response to phosphate<br />

deficiency in nonarbuscular mycorrhizal host Arabidopsis. <strong>Plant</strong><br />

Physiol. 155, 974–987.<br />

Kohonen M, Hell<strong>and</strong> A (2009) On the function of wall sculpturing in<br />

xylem conduits. J. Bionics Eng. 6, 324–329.<br />

Koike S, Inoue H, Mizuno D, Takahashi M, Nakanishi H, Mori<br />

S, Nishizawa N (2004) OsYSL2 is a rice metal-nicotianamine<br />

transporter that is regulated by iron <strong>and</strong> expressed in the phloem.<br />

<strong>Plant</strong> J. 39, 415–424.<br />

Kondo Y, Hirakawa Y, Fukuda H (2011) CLE peptides can negatively<br />

regulate protoxylem vessel formation via cytokinin signaling. <strong>Plant</strong><br />

Cell Physiol. 52, 37–48.<br />

Koo YJ, Kim MA, Kim EH, Song JT, Jung C, Moon JK, Kim JH,<br />

Seo HS, Song SI, Kim JK, Lee JS, Cheong JJ, Choi YD (2007)<br />

Overexpression of salicylic acid carboxyl methyltransferase reduces<br />

salicylic acid-mediated pathogen resistance in Arabidopsis thaliana.<br />

<strong>Plant</strong> Mol. Biol. 64, 1–15.<br />

Koroleva OA, Tomos AD, Farrar J, Pollock CJ (2002) Changes in<br />

osmotic <strong>and</strong> turgor pressure in response to sugar accumulation in<br />

barley source leaves. <strong>Plant</strong>a 215, 210–219.<br />

Kramer EM, Lew<strong>and</strong>owski M, Beri S, Bernard J, Borkowski M,<br />

Borkowski MH, Burchfield LA, Mathisen B, Normanly J (2008)<br />

Auxin gradients are associated with polarity changes in trees.<br />

Science 320, 1610.<br />

Krämer U (2010) Metal hyperaccumulation in plants. Annu. Rev. <strong>Plant</strong><br />

Biol. 61, 517–534.<br />

Krämer U, Cotter-Howells JD, Charnock JM, Baker AJM, Smith JAC<br />

(1996) Free histidine as a metal chelator in plants that accumulate<br />

nickel. Nature 379, 635–638.<br />

Krecek P, Skupa P, Libus J, Naramoto S, Tejos R, Friml J,<br />

Zazimalova E (2009) <strong>The</strong> pin-formed (pin) protein family of auxin<br />

transporters. Genome Biol. 10, 249.<br />

Krügel U, Veenhoff LM, Langbein J, Wiederhold E, Liesche J,<br />

Friedrich T, Grimm B, Martinoia E, Poolman B, Kühn C (2008)<br />

Transport <strong>and</strong> sorting of Solanum tuberosum sucrose transporter<br />

SUT1 is affected by posttranslational modification. <strong>Plant</strong> Cell 20,<br />

2497–2513.<br />

Krüger C, Berkowitz O, Stephan U, Hell R (2002) A metal-binding<br />

member of the late embryogenesis abundant protein family transports<br />

iron in the phloem of Ricinus communis L. J. Biol. Chem. 277,<br />

25062–25069.<br />

Kubo M, Udagawa M, Nishikubo N, Horiguchi G, Yamaguchi M, Ito<br />

J, Mimura T, Fukuda H, Demura T (2005) Transcription switches<br />

for protoxylem <strong>and</strong> metaxylem vessel formation. Genes Dev. 19,<br />

1855–1860.<br />

Kudo T, Kiba T, Sakakibara H (2010) Metabolism <strong>and</strong> long-distance<br />

translocation of cytokinins. J. Integr. <strong>Plant</strong> Biol. 52, 53–60.<br />

Kumar D, Klessig DF (2003) High-affinity salicylic acid-binding protein<br />

2 is required for plant innate immunity <strong>and</strong> has salicylic acid<br />

stimulated lipase activity. Proc. Natl. Acad. Sci. USA 100, 16101–<br />

16106.


374 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Kunkel BN, Brooks DM (2002) Cross talk between signaling pathways.<br />

Curr. Opin. <strong>Plant</strong> Biol. 5, 325–331.<br />

Küpper H, Mijovilovich A, Meyer-Klaucke W, Kroneck PMH (2004)<br />

Tissue- <strong>and</strong> age-dependent differences in the complexation of<br />

cadmium <strong>and</strong> zinc in the cadmium/zinc hyperaccumulator Thlaspi<br />

caerulescens (Ganges Ecotype) revealed by X-ray absorption spectroscopy.<br />

<strong>Plant</strong> Physiol. 134, 748–757.<br />

Kuriyama H, Fukuda H (2002) <strong>Development</strong>al programmed cell death<br />

in plants. Curr. Opin. <strong>Plant</strong> Biol. 5, 568–573.<br />

Kwon SI, Cho HJ, Jung JH, Yoshimoto K, Park OK (2010) <strong>The</strong><br />

RabGTPase RabG3b functions in autophagy <strong>and</strong> contributes to<br />

tracheary element differentiation in Arabidopsis. <strong>Plant</strong> J. 64, 151–<br />

164.<br />

Lachaud S, Maurousset L (1996) Occurrence of plasmodesmata<br />

between differentiating vessels <strong>and</strong> other xylem cells in Sorbus<br />

torminalis L. Crantz <strong>and</strong> their fate during xylem maturation. Protoplasma<br />

191, 220–226.<br />

Lai F, Thacker J, Li Y, Doerner P (2007) Cell division activity<br />

determines the magnitude of phosphate starvation responses in<br />

Arabidopsis. <strong>Plant</strong> J. 50, 545–556.<br />

Lalonde S, Tegeder M, Throne-Holst M, Frommer WB, Patrick JW<br />

(2003) Phloem loading, unloading of amino acids, sugars. <strong>Plant</strong> Cell<br />

Environ. 26, 37–56.<br />

Langan S, Ewers FW, Davis SD (1997) Xylem dysfunction caused by<br />

water stress <strong>and</strong> freezing in two species of co-occurring chaparral<br />

shrubs. <strong>Plant</strong> Cell Environ. 20, 425–437.<br />

Lau S, Shao N, Bock R, Jürgens G, De Smet I (2009) Auxin<br />

signaling in algal lineages: Fact or myth? Trends <strong>Plant</strong> Sci. 14, 182–<br />

188.<br />

Lavicoli A, Boutet E, Buchala A, Metraux JP (2003) Induced resistance<br />

in Arabidopsis thaliana in response to root inoculation with<br />

Pseudomonas fluorescens CHA0. Mol. <strong>Plant</strong>-Microbe Interact. 16,<br />

851–858.<br />

Lawton K, Weymann K, Friedrich L, Vernooij B, Uknes S, Ryals<br />

J (1995) <strong>System</strong>ic acquired resistance in Arabidopsis requires<br />

salicylic acid but not ethylene. Mol. <strong>Plant</strong>-Microbe Interact. 8, 863–<br />

870.<br />

Le Jean ML, Schikora A, Mari S, Briat JF, Curie C (2005) A loss-offunction<br />

mutation in AtYSL1 reveals its role in iron <strong>and</strong> nicotianamine<br />

seed loading. <strong>Plant</strong> J. 44, 769–782.<br />

Lee DK, Sieburth LE (2012) <strong>The</strong> bps signal: Embryonic arrest from<br />

an auxin-independent mechanism in bypass triple mutants. <strong>Plant</strong><br />

Signal. Behav. 7, 698–700.<br />

Lee DK, Van Norman JM, Murphy C, Adhikari E, Reed JW, Sieburth<br />

LE (2012) In the absence of BYPASS1-related gene function,<br />

the bps signal disrupts embryogenesis by an auxin-independent<br />

mechanism. <strong>Development</strong> 139, 805–815.<br />

Lee HI, Raskin I (1998) Glucosylation of salicylic acid in Nicotiana<br />

tabacum cv. Xanthi-nc. Phytopathol. 88, 692–697.<br />

Lee HI, Raskin I (1999) Purification, cloning, <strong>and</strong> expression of a<br />

pathogen inducible UDP-glucose:salicylic acid glucosuyltransferase<br />

from tobacco. J. Biol. Chem. 274, 36637–36642.<br />

Lee JY, Colinas J, Wang JY, Mace D, Ohler U, Benfey PN (2006)<br />

Transcriptional <strong>and</strong> posttranscriptional regulation of transcription<br />

factor expression in Arabidopsis roots. Proc. Natl. Acad. Sci. USA<br />

103, 6055–6060.<br />

Lee J, Holbrook NM, Zwieniecki MA (2012) Ion-induced changes<br />

in the structure of bordered pit membranes. Front. <strong>Plant</strong> Sci. 3,<br />

1–4.<br />

Lee JY, Yoo BC, Rojas M, Gomez Ospina N, Staehelin LA, Lucas<br />

WJ (2003) Selective trafficking of non-cell-autonomous proteins<br />

mediated by NtNCAPP1. Science 299, 392–396.<br />

Leggewie G, Kolbe A, Lemoine R, Roessner U, Lytovchenko<br />

A, Zuther E, Kehr J, Frommer WB, Riemeier JW, Willmitzer<br />

L, Fernie AR (2003) Overexpression of the sucrose transporter<br />

SoSUT1 in potato results in alterations in leaf carbon partitioning<br />

<strong>and</strong> in tuber metabolism but has little impact on tuber morphology.<br />

<strong>Plant</strong>a 217, 158–167.<br />

Lehesranta SJ, Lichtenberger R, Helariutta Y (2010) Cell-to-cell<br />

communication in vascular morphogenesis. Curr. Opin. <strong>Plant</strong> Biol.<br />

13, 59–65.<br />

Lehmann K, Hause B, Altmann D, Kock M (2001) Tomato ribonuclease<br />

LX with the functional endoplasmic reticulum retention motif<br />

HDEF is expressed during programmed cell death processes,<br />

including xylem differentiation, germination, <strong>and</strong> senescence. <strong>Plant</strong><br />

Physiol. 127, 436–449.<br />

Lehto T, Lavola A, Julkunen-Tiitto R, Aphalo PJ (2004) Boron<br />

retranslocation in scots pine <strong>and</strong> norway spruce. Tree Physiol. 24,<br />

1011–1017.<br />

Lens F, Sperry JS, Christman MA, Choat B, Rabaey D, Jansen<br />

S (2010) Testing hypotheses that link wood anatomy to cavitation<br />

resistance <strong>and</strong> hydraulic conductivity in the genus Acer. New Phytol.<br />

190, 709–723.<br />

Leonardos ED, Micallef BJ, Micallef MC, Grodzinski B (2006) Diel<br />

patterns of leaf export <strong>and</strong> of main shoot growth for Flaveria linearis<br />

with altered leaf sucrose-starch partitioning. J. Exp. Bot. 57, 801–<br />

814.<br />

Levesque MP, Vernoux T, Busch W, Cui H, Wang JY, Blilou I,<br />

Hassan H, Nakajima K, Matsumoto N, Lohmann JU, Scheres B,<br />

Benfey PN (2006) Whole-genome analysis of the SHORT-ROOT<br />

developmental pathway in Arabidopsis. PLoS Biol. 4, e143.<br />

Li CY, Zhang K, Zeng XW, Jackson S, Zhou Y, Hong YG (2009) A<br />

cis element within Flowering Locus T mRNA determines its mobility<br />

<strong>and</strong> facilitates trafficking of heterologous viral RNA. J. Virol. 83,<br />

3540–3548.<br />

Li P, Ham BK, Lucas WJ (2011) CmRBP50 phosphorylation is<br />

essential for assembly of a stable phloem-mobile high-affinity ribonucleoprotein<br />

complex. J. Biol. Chem. 286, 23142–23149.<br />

Li X, Zhang Y, Clarke JD, Li Y, Dong X (1999) Identification<br />

<strong>and</strong> cloning of a negative regulator of systemic acquired<br />

resistance, SNI1, through a screen for supressors of npr1-1. Cell<br />

3, 329–339.<br />

Li Y, Beisson F, Koo AJ, Molina I, Pollard M, Ohlrogge J (2007)<br />

Identification of acyltransferases required for cutin biosynthesis <strong>and</strong>


production of cutin with suberin-like monomers. Proc. Natl. Acad.<br />

Sci. USA 104, 18339–18344.<br />

Liao MT, Hedley MJ, Woolley DJ, Brooks RR, Nichols MA (2000)<br />

Copper uptake <strong>and</strong> translocation in chicory (Cichorium intybus L.<br />

cv Grassl<strong>and</strong>s Puna) <strong>and</strong> tomato (Lycopersicon esculentum Mill. cv<br />

Rondy) plants grown in NFT system. II. <strong>The</strong> role of nicotianamine<br />

<strong>and</strong> histidine in xylem sap copper transport. <strong>Plant</strong> Soil 223, 243–<br />

252.<br />

Liesche J, Schulz A (2012) In vivo quantification of cell coupling in<br />

plants with different phloem loading strategies. <strong>Plant</strong> Physiol. 159,<br />

355–365.<br />

Lifschitz E, Eviatar T, Rozman A, Shalit A, Goldshmidt A, Amsellem<br />

Z, Alvarez JP, Eshed Y (2006) <strong>The</strong> tomato FT ortholog triggers<br />

systemic signals that regulate growth <strong>and</strong> flowering <strong>and</strong> substitute<br />

for diverse environmental stimuli. Proc. Natl. Acad. Sci. USA 103,<br />

6398–6403.<br />

Ligrone R, Duckett JG, Renzaglia KS (2000) Conducting tissues <strong>and</strong><br />

phyletic relationships of bryophytes. Phil. Trans. R. Soc. Lond. B.<br />

Biol. Sci. 355, 795–813.<br />

Ligrone R, Duckett JG, Renzaglia KS (2012) Major transitions in the<br />

evolution of early l<strong>and</strong> plants: A bryological perspective. Ann. Bot.<br />

109, 851–871.<br />

Lin MK, Belanger H, Lee YJ, Varkonyi-Gasic E, Taoka KI, Miura<br />

E, Xoconostle-Cázares B, Gendler K, Jorgensen RA, Phinney<br />

B, Lough TJ, Lucas WJ (2007) FT protein may act as the longdistance<br />

florigenic signal in the cucurbits. <strong>Plant</strong> Cell 19, 1488–1506.<br />

Lin MK, Lee YJ, Lough TJ, Phinney BS, Lucas WJ (2009) Analysis<br />

of the pumpkin phloem proteome provides insights into angiosperm<br />

sieve tube function. Mol. Cell. Proteomics 8, 343–356.<br />

Lin SI, Chiang SF, Lin WY, Chen JW, Tseng CY, Wu PC, Chiou TJ<br />

(2008) Regulatory network of microRNA399 <strong>and</strong> PHO2 by systemic<br />

signaling. <strong>Plant</strong> Physiol. 147, 732–746.<br />

Lindermayr C, Sell S, Müller B, Leister D, Durner J (2010) Redox<br />

regulation of the NPR1-TGA1 system of Arabidopsis thaliana by<br />

nitric oxide. <strong>Plant</strong> Cell 22, 2894–2907.<br />

Linkohr BI, Williamson LC, Fitter AH, Leyser HMO (2002) Nitrate<br />

<strong>and</strong> phosphate availability <strong>and</strong> distribution have different effects on<br />

root system architecture of Arabidopsis. <strong>Plant</strong> J. 29, 751–760.<br />

Liu G, Holub EB, Alonso JM, Ecker JR, Fobert PR (2005) An Arabidopsis<br />

NPR1-like gene, NPR4, is required for disease resistance.<br />

<strong>Plant</strong> J. 41, 304–318.<br />

Liu PP, Dahl CC, Klessig DF (2011) <strong>The</strong> extent to which methyl<br />

salicylate is required for systemic acquired resistance is dependent<br />

on exposure to light after infection. <strong>Plant</strong> Physiol. 157, 2216–2226.<br />

LiuTY,ChangCY,ChiouTJ(2009) <strong>The</strong> long-distance signaling of<br />

mineral macronutrients. Curr. Opin. <strong>Plant</strong> Biol. 12, 312–319.<br />

Liu Y, Schiff M, Marathe R, Dinesh-Kumar SP (2002) Tobacco Rar1,<br />

EDS1 <strong>and</strong> NPR1/NIM1 like genes are required for N-mediated<br />

resistance to tobacco mosaic virus. <strong>Plant</strong> J. 30, 415–429.<br />

Loepfe L, Martinez-Vilalta J, Pinol J, Mencuccini M (2007) <strong>The</strong><br />

relevance of xylem network structure for plant hydraulic efficiency<br />

<strong>and</strong> safety. J. <strong>The</strong>or. Biol. 247, 788–803.<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 375<br />

Lomax TL, Hicks GR (1992) Specific auxin-binding proteins in the<br />

plasma membrane: Receptors or transporters? Biochem. Soc.<br />

Trans. 20, 64–69.<br />

Loon LCV, Gerritsen YAM, Ritter CE (1987) Identification, purification,<br />

<strong>and</strong> characterization of pathogenesis-related proteins from virusinfected<br />

Samsun NN tobacco leaves. <strong>Plant</strong> Mol. Biol. 9, 593–609.<br />

López-Bucio J, Hernández-Abreu E, Sánchez-Calderón L, Nieto-<br />

Jacobo MF, Simpson J, Herrera-Estrella L (2002) Phosphate<br />

availability alters architecture <strong>and</strong> causes changes in hormone<br />

sensitivity in the Arabidopsis root system. <strong>Plant</strong> Physiol. 129, 244–<br />

256.<br />

López-Millán AF, Morales F, Abadía A, Abadía J (2000) Effects of<br />

iron deficiency on the composition of the leaf apoplastic fluid <strong>and</strong><br />

xylem sap in sugar beet. Implications for iron <strong>and</strong> carbon transport.<br />

<strong>Plant</strong> Physiol. 124, 873–884.<br />

Lough TJ, Lucas WJ (2006) Integrative plant biology: Role of phloem<br />

long-distance macromolecular trafficking. Annu. Rev. <strong>Plant</strong> Biol. 57,<br />

203–232.<br />

Lu KJ, Huang NC, Liu YS, Lu CA, Yu TS (2012) Long-distance<br />

movement of Arabidopsis FLOWERING LOCUS T RNA participates<br />

in systemic floral regulation. RNA Biology 9, 653–662.<br />

Lucas WJ (2006) <strong>Plant</strong> viral movement proteins: Agents for cell-to-cell<br />

trafficking of viral genomes. Virology 344, 169–184.<br />

Lucas WJ, Ding B, van der Schoot C (1993) Plasmodesmata <strong>and</strong> the<br />

supracellular nature of plants. New Phytol. 125, 435–476.<br />

Lucas WJ, Yoo BC, Kragler F (2001) RNA as a long-distance<br />

information macromolecule in plants. Nat. Rev. Molec. Cell Biol.<br />

2, 849–857.<br />

Lundmark M, Cavaco AM, Trevanior S, Hurry V (2006) Carbon<br />

partitioning <strong>and</strong> export in transgenic Arabidopsis thaliana with<br />

altered capacity for sucrose synthesis grown at low temperature:<br />

A role for metabolite transporters. <strong>Plant</strong> Cell Environ. 29, 1703–<br />

1714.<br />

MaKW,FloresC,MaW(2011) Chromatin configuration as a battlefield<br />

in plant-bacteria interactions. <strong>Plant</strong> Physiol. 157, 535–543.<br />

Ma Y, Miura E, Ham BK, Cheng HW, Lee YJ, Lucas WJ (2010)<br />

Pumpkin eIF5A isoforms interact with components of the translational<br />

machinery in the cucurbit sieve tube system. <strong>Plant</strong> J. 64,<br />

536–550.<br />

Madey E, Nowack LM, Thompson JE (2002) Isolation <strong>and</strong> characterization<br />

of lipid in phloem sap of canola. <strong>Plant</strong>a 214, 625–634.<br />

Mähönen AP, Bishopp A, Higuchi M, Nieminen KM, Kinoshita K,<br />

Törmäkangas K, Ikeda Y, Oka A, Kakimoto T, Helariutta Y (2006)<br />

Cytokinin signaling <strong>and</strong> its inhibitor AHP6 regulate cell fate during<br />

vascular development. Science 311, 94–98.<br />

Mähönen AP, Bonke M, Kauppinen L, Riikonen M, Benfey PN,<br />

Helariutta Y (2000) A novel two-component hybrid molecule regulates<br />

vascular morphogenesis of the Arabidopsis root. Genes Dev.<br />

14, 2938–2943.<br />

Mahajan A, Bhogale S, Kang IH, Hannapel DJ, Banerjee AK (2012)<br />

<strong>The</strong> mRNA of a knotted1-like transcription factor of potato is phloem<br />

mobile. <strong>Plant</strong> Mol. Biol. 79, 595–608.


376 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Maherali H, Pockman WT, Jackson RB (2003) Adaptive variation in<br />

the vulnerability of woody plants to xylem cavitation. Ecology 85,<br />

2184–2199.<br />

Maldonado AM, Doerner P, Dixon RA, Lamb CJ, Cameron RKA<br />

(2002) Putative lipid transfer protein involved in systemic resistance<br />

signaling in Arabidopsis. Nature 419, 399–403.<br />

M<strong>and</strong>al M, Ch<strong>and</strong>a B, Xia Y, Yu K, Sekine K, Gao QM, Selote D,<br />

Kachroo A, Kachroo P (2011) Glycerol-3-phosphate <strong>and</strong> systemic<br />

immunity. <strong>Plant</strong> Signal. Behav. 6, 1871–1874.<br />

M<strong>and</strong>al MK, Ch<strong>and</strong>ra-Shekara AC, Jeong RD, Yu K, Zhu S, Ch<strong>and</strong>a<br />

B, Navarre D, Kachroo A, Kachroo P (2012) Oleic acid-dependent<br />

modulation of NITRIC OXIDE ASSOCIATED 1 protein levels<br />

regulates nitric oxide-mediated defense signaling in Arabidopsis.<br />

<strong>Plant</strong> Cell 24, 1654–1674.<br />

March-Díaz R, García-Domínguez M, Lozano-Juste J, León J,<br />

Florencio FJ, Reyes JC (2008) Histone H2A.Z <strong>and</strong> homologues of<br />

components of the SWR1 complex are required to control immunity<br />

in Arabidopsis. <strong>Plant</strong> J. 53, 475–487.<br />

Mari S, Gendre D, Pianelli K, Ouerdane L, Lobinski R, Briat JF,<br />

Lebrun M, Czernic P (2006) Root-to-shoot long-distance circulation<br />

of nicotianamine <strong>and</strong> nicotianamine-nickel chelates in the metal<br />

hyperaccumulator Thlaspi caerulescens. J. Exp. Bot. 57, 4111–<br />

4122.<br />

Markesteijn L, Poorter L, Paz H, Sack L, Bongers F (2011) Ecological<br />

differentiation in xylem cavitation is associated with stem <strong>and</strong> leaf<br />

structural traits. <strong>Plant</strong> Cell Environ. 34, 137–148.<br />

Marschner H (1995) Mineral Nutrition of Higher <strong>Plant</strong>s. Academic<br />

Press, London.<br />

Martin A, Adam H, Diaz-Mendoza M, Zurczak M, Gonzalez-Schain<br />

ND, Suarez-Lopez P (2009) Graft-transmissible induction of potato<br />

tuberization by the microRNA miR172. <strong>Development</strong> 136, 2873–<br />

2881.<br />

Martin AC, del Pozo J, Iglesias J, Rubio V, Solano R, de La Pena<br />

A, Leyva A, Paz-Ares J (2000) Influence of cytokinins on the expression<br />

of phosphate starvation responsive genes in Arabidopsis.<br />

<strong>Plant</strong> J. 24, 559–567.<br />

Mathieu J, Warthmann N, Kuttner F, Schmid M (2007) Export of FT<br />

protein from phloem companion cells is sufficient for floral induction<br />

in Arabidopsis. Curr. Biol. 17, 1055–1060.<br />

Matsubayashi Y, Sakagami Y (2006) Peptide hormones in plants.<br />

Annu Rev. <strong>Plant</strong> Biol. 57, 649–674.<br />

Matsumoto-Kitano M, Kusumoto T, Tarkowski P, Kinoshita-<br />

Tsujimura K, Václavíková K, Miyawaki K, Kakimoto T (2008)<br />

Cytokinins are central regulators of cambial activity. Proc. Natl.<br />

Acad. Sci. USA 105, 20027–20031.<br />

Matte Risopatron JP, Sun Y, Jones BJ (2010) <strong>The</strong> vascular cambium:<br />

Molecular control of cellular structure. Protoplasma 247, 145–161.<br />

Mattsson J, Ckurshumova W, Berleth T (2003) Auxin signaling in<br />

Arabidopsis leaf vascular development. <strong>Plant</strong> Physiol. 131, 1327–<br />

1339.<br />

Mayr S, Sperry JS (2010) Freeze-thaw induced embolism in Pinus<br />

contorta: Centrifuge experiments validate the “thaw-expansion”<br />

hypothesis but conflict with ultrasonic data. New Phytol. 18, 1016–<br />

1024.<br />

Mayr S, Zublasing V (2010) Ultrasonic emissions from conifer xylem<br />

exposed to repeated freezing. <strong>Plant</strong> Physiol. 167, 34–40.<br />

Mayzlish-Gati E, De-Cuyper C, Goormachtig S, Beeckman T, Vuylsteke<br />

M, Brewer PB, Beveridge CA, Yermiyahu U, Kaplan<br />

Y, Enzer Y, Wininger S, Resnick N, Cohen M, Kapulnik Y,<br />

Koltai H (2012) Strigolactones are involved in root response to<br />

low phosphate conditions in Arabidopsis. <strong>Plant</strong> Physiol. 160, 1329–<br />

1341.<br />

McAdam SAM, Brodribb TJ (2012) Stomatal innovation <strong>and</strong> the rise<br />

of seed plants. Ecol. Lett. 15, 1–8.<br />

McCarthy RL, Zhong R, Ye ZH (2009) MYB83 is a direct target of<br />

SND1 <strong>and</strong> acts redundantly with MYB46 in the regulation of secondary<br />

cell wall biosynthesis in Arabidopsis. <strong>Plant</strong> Cell Physiol. 50,<br />

1950–1964.<br />

McConnell JR, Barton MK (1998) Leaf polarity <strong>and</strong> meristem formation<br />

in Arabidopsis. <strong>Development</strong> 125, 2935–2942.<br />

McConnell JR, Emery J, Eshed Y, Bao N, Bowman J, Barton<br />

MK (2001) Role of PHABULOSA <strong>and</strong> PHAVOLUTA in determining<br />

radial patterning in shoots. Nature 411, 709–713.<br />

McNear DH Jr, Chaney RL, Sparks DL (2010) <strong>The</strong> hyperaccumulator<br />

Alyssum murale uses complexation with nitrogen <strong>and</strong> oxygen<br />

donor lig<strong>and</strong>s for Ni transport <strong>and</strong> storage. Phytochemistry 71,<br />

188–200.<br />

Meinzer FC, Goldstein G, Jackson P, Holbrook NM, Gutierrez<br />

MV, Cavelier J (1995) Environmental <strong>and</strong> physiological regulation<br />

of transpiration in tropical forest gap species: <strong>The</strong> influence<br />

of boundary layer <strong>and</strong> hydraulic properties. Oecologia 101,<br />

514–522.<br />

Melnyk CW, Molnar A, Baulcombe DC (2011) Intercellular <strong>and</strong><br />

systemic movement of RNA silencing signals. EMBO J. 30, 3553–<br />

3563.<br />

Mencuccini M, Hölttä T(2010) <strong>The</strong> significance of phloem transport<br />

for the speed with which canopy photosynthesis <strong>and</strong> belowground<br />

respiration are linked. New Phytol. 185, 189–203.<br />

Mercury L, Tardy Y (2001) Negative pressure of stretched liquid water.<br />

Geochemistry of soil capillaries. Geochem. Cosmochem. Acta 65,<br />

3391–3408.<br />

Mijovilovich A, Leitenmaier B, Meyer-Klaucke W, Kroneck PMH,<br />

Gotz B, Küpper H (2009) Complexation <strong>and</strong> toxicity of copper in<br />

higher plants. II. Different mechanisms for copper versus cadmium<br />

detoxification in the copper-sensitive cadmium/zinc hyperaccumulator<br />

Thlaspi caerulescens (Ganges ecotype). <strong>Plant</strong> Physiol. 151,<br />

715–731.<br />

Milburn JA, Kallarackal J (1989) Physiological aspects of phloem<br />

translocation. In: Baker DA, Milburn JA, eds. Transport of Photoassimilates.<br />

Lonhman Scientific & Technical, Essex. pp. 262–305.<br />

Milioni D, Sado PE, Stacey NJ, Roberts K, McCann MC (2002) Early<br />

gene expression associated with the commitment <strong>and</strong> differentiation<br />

of a plant tracheary element is revealed by cDNA-amplified fragment<br />

length polymorphism analysis. <strong>Plant</strong> Cell 14, 2813–2824.


Mills RF, Krijgert GC, Baccarini PJ, Hall JL, Williams LE (2003)<br />

Functional expression of AtHMA4, a P1B-type ATPase of the<br />

Zn/Co/Cd/Pb subclass. <strong>Plant</strong> J. 35, 164–176.<br />

Mills RF, Francini A, Ferreira Da Rocha PSC, Baccarini PJ, Aylett<br />

M, Krijger GC, Williams LE (2005) <strong>The</strong> plant P1B-type ATPase<br />

AtHMA4 transports Zn <strong>and</strong> Cd <strong>and</strong> plays a role in detoxification<br />

of transition metals supplied at elevated levels. FEBS Lett. 579,<br />

783–791.<br />

Minchin PEH, Thorpe MR (1983) A rate of cooling response in phloem<br />

translocation. J. Exp. Bot. 34, 529–536.<br />

Mira H, Martínez-García F, Peñarrubia L (2001) Evidence for the<br />

plant-specific intercellular transport of the Arabidopsis copper chaperone<br />

CCH. <strong>Plant</strong> J. 25, 521–528.<br />

Mishina TE, Zeier J (2006) <strong>The</strong> Arabidopsis flavin-dependent<br />

monooxygenase FMO1 is an essential component of biologically<br />

induced systemic acquired resistance. <strong>Plant</strong> Physiol. 141, 1666–<br />

1675.<br />

Mishler BD, Churchill SP (1984) A cladistic approach to the phylogeny<br />

of the “bryophytes.” Brittonia 36, 406–424.<br />

Misra NB, Varshini YP (1961) Viscosity-temperature relation to solutions.<br />

J. Chem. Eng. Data 6, 194–196.<br />

Mitsuda N, Seki M, Shinozaki K, Ohme-Takagi M (2005) <strong>The</strong> NAC<br />

transcription factors NST1 <strong>and</strong> NST2 of Arabidopsis regulate secondary<br />

wall thickenings <strong>and</strong> are required for anther dehiscence.<br />

<strong>Plant</strong> Cell 17, 2993–3006.<br />

Mitsuda N, Iwas A, Yamamoto H, Yoshida M, Seki M, Shinozaki<br />

K, Ohme-Takagi M (2007) NAC transcription factors, NST1 <strong>and</strong><br />

NST3, are key regulators of the formation of secondary walls in<br />

woody tissues of Arabidopsis. <strong>Plant</strong> Cell 19, 270–280.<br />

Mittelheuser CJ, Van Steveninck RFM (1969) Stomatal closure <strong>and</strong><br />

inhibition of transpiration induced by (RS)-abscisic acid. Nature 221,<br />

281–282.<br />

Miwa H, Betsuyaku S, Iwamoto K, Kinoshita A, Fukuda H, Sawa<br />

S (2008) <strong>The</strong> receptor-like kinase SOL2 mediates CLE signaling in<br />

Arabidopsis. <strong>Plant</strong> Cell Physiol. 49, 1752–1757.<br />

Miwa K, Fujiwara T (2010) Boron transport in plants: Co-ordinated<br />

regulation of transporters. Ann. Bot. 105, 1103–1108.<br />

Miyashima S, Nakajima K (2011) <strong>The</strong> root endodermis: A hub of<br />

developmental signals <strong>and</strong> nutrient flow. <strong>Plant</strong> Signal. Behav. 6,<br />

1954–1958.<br />

Miyashima S, Koi S, Hashimoto T, Nakajima K (2011) Non-cellautonomous<br />

microRNA165 acts in a dose-dependent manner to<br />

regulate multiple differentiation status in the Arabidopsis root. <strong>Development</strong><br />

138, 2303–2313.<br />

Miyashima S, Sebastian J, Lee JY, Helariutta Y (2012) Stem cell<br />

function during plant vascular development. EMBO J. 32, 178–193.<br />

Miyazawa H, Oka-Kira E, Sato N, Takahashi H, Wu GJ, Sato S,<br />

Hayashi M, Betsuyaku S, Nakazono M, Tabata S, Harada K,<br />

Sawa S, Fukuda H, Kawaguchi M (2010) <strong>The</strong> receptor-like kinase<br />

KLAVIER mediates systemic regulation of nodulation <strong>and</strong> nonsymbiotic<br />

shoot development in Lotus japonicus. <strong>Development</strong> 137,<br />

4317–4325.<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 377<br />

Moeder W, Pozo OD, Navarre DA, Martin GB, Klessig DF (2007)<br />

Aconitase plays a role in regulating resistance to oxidative stress<br />

<strong>and</strong> cell death in Arabidopsis <strong>and</strong> Nicotiana benthamiana. <strong>Plant</strong> Mol.<br />

Biol. 63, 273–287.<br />

Mok DW, Mok MC (2001) Cytokinin metabolism <strong>and</strong> action. Annu. Rev.<br />

<strong>Plant</strong> Physiol. <strong>Plant</strong> Mol. Biol. 52, 89–118.<br />

Molders W, Buchala A, Metraux J-P (1996) Transport of salicylic acid<br />

in tobacco necrosis virus-infected cucumber plants. <strong>Plant</strong> Physiol.<br />

112, 787–792.<br />

Molnar A, Melnyk CW, Bassett A, Hardcastle TJ, Dunn R,<br />

Baulcombe DC (2010) Small silencing RNAs in plants are mobile<br />

<strong>and</strong> direct epigenetic modification in recipient cells. Science 328,<br />

872–875.<br />

Moreno-Risueno MA, Van Norman JM, Moreno A, Zhang JY, Ahnert<br />

SE, Benfey PN (2010) Oscillating gene expression determines<br />

competence for periodic Arabidopsis root branching. Science 329,<br />

1306–1311.<br />

Morrissey J, Baxter IR, Lee J, Li L, Lahner B, Grotz N, Kaplan J,<br />

Salt DE, Guerinot ML (2009) <strong>The</strong> ferroportin metal efflux proteins<br />

function in iron <strong>and</strong> cobalt homeostasis in Arabidopsis. <strong>Plant</strong> Cell<br />

21, 3326–3338.<br />

Mosher RA, Durrant WE, Wang D, Song J, Dong X (2006) A comprehensive<br />

structure-function analysis of Arabidopsis SNI1 defines<br />

essential regions <strong>and</strong> transcriptional supppressor activity. <strong>Plant</strong> Cell<br />

18, 1750–1765.<br />

Mosher RA, Schwach F, Studhollme D, Baulcombe DC (2008)<br />

PolIVb influences RNA-directed DNA-methylation independently of<br />

its role in siRNA biogenesis. Proc. Natl. Acad. Sci USA 105, 3145–<br />

3150.<br />

Motose H, Sugiyama M, Fukuda H (2004) A proteoglycan mediates<br />

inductive interaction during plant vascular development. Nature 429,<br />

873–878.<br />

Mott KA, Peak D (2011) Alternative perspective on the control of<br />

transpiration by radiation. Proc. Natl. Acad. Sci. USA 108, 19820–<br />

19823.<br />

Mou Z, Fan W, Dong X (2003) Inducers of plant systemic acquired<br />

resistance regulate NPR1 function through redox changes. Cell 113,<br />

935–944.<br />

Mullendore DL, Windt CW, Van As H, Knoblauch M (2010) Sieve<br />

tube geometry in relation to phloem flow. <strong>Plant</strong> Cell 22, 579–<br />

593.<br />

Müller R, Bleckmann A, Simon R (2008) <strong>The</strong> receptor kinase<br />

CORYNE of Arabidopsis transmits the stem cell-limiting signal<br />

CLAVATA3 independently of CLAVATA1. <strong>Plant</strong> Cell 20, 934–946.<br />

Münch E (1930) Material Flow in <strong>Plant</strong>s. Translated 2003 by Milburn<br />

JA <strong>and</strong> Kreeb KH, University of Bremen, Germany. Gustav Fischer<br />

Verlag, Jena Germany.<br />

Muñiz L, Minguet EG, Singh SK, Pesquet E, Vera-Sirera F, Moreau-<br />

Courtois CL, Carbonell J, Blázquez MA, Tuominen H (2008)<br />

ACAULIS5 controls Arabidopsis xylem specification through the<br />

prevention of premature cell death. <strong>Development</strong> 135, 2573–<br />

2582.


378 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Murphy A, Taiz L (1997) Correlation between potassium efflux <strong>and</strong><br />

copper sensitivity in 10 Arabidopsis ecotypes. New Phytol. 136,<br />

211–222.<br />

NagahashiG,DoudsDD(2000) Partial separation of root exudate<br />

components <strong>and</strong> their effects upon the growth of germinated spores<br />

of AM fungi. Mycol Res. 104, 1453–1464.<br />

Nagawa S, Sawa S, Sato S, Kato T, Tabata S, Fukuda H (2006)<br />

Gene trapping in Arabidopsis reveals genes involved in vascular<br />

development. <strong>Plant</strong> Cell Physiol. 47, 1394–1405.<br />

Naik SN, Goud VV, Rout PK, Dalai AK (2010) Production of first <strong>and</strong><br />

second generation biofuels: A comprehensive review. Renew. Sust.<br />

Energ. Rev. 14, 578–597.<br />

Napoli C (1996) Highly branched phenotype of the petunia dad1-1<br />

mutant is reversed by grafting. <strong>Plant</strong> Physiol. 111, 27–37.<br />

Nardini A, LoGullo MA, Salleo S (2011a) Refilling embolized xylem<br />

conduits: Is it a matter of phloem unloading? <strong>Plant</strong> Sci. 180, 604–<br />

611.<br />

Nardini A, Salleo S, Jansen S (2011b) More than just a vulnerable<br />

pipeline: Xylem physiology in the light of ion-mediated regulation of<br />

plant water transport. J. Exp. Bot. 62, 4701–4718.<br />

Nawrath C, Heck S, Parinthawong N, Métraux JP (2002) EDS5,<br />

an essential component of salicylic acid-dependent signaling for<br />

disease resistance in Arabidopsis, is a member of the MATE<br />

transporter family. <strong>Plant</strong> Cell 14, 275–286.<br />

Neale DB, Kremer A (2011) Forest tree genomics: Growing resources<br />

<strong>and</strong> applications. Nat. Rev. Genet. 12, 111–122.<br />

Nelson T, Dengler N (1997) Leaf vascular pattern formation. <strong>Plant</strong> Cell<br />

9, 1121–1135.<br />

Niggeweg R, Thurow C, Weigel R, Pfitzner U, Gatz C (2000)<br />

Tobacco TGA factors differ with respect to interaction with NPR1,<br />

activation potential <strong>and</strong> DNA-binding properties. <strong>Plant</strong> Mol. Biol. 42,<br />

775–788.<br />

Nilsson J, Karlberg A, Antti H, Lopez-Vernaza M, Mellerowicz<br />

E, Perrot-Rechenmann C, S<strong>and</strong>berg G, Bhalerao RP (2008)<br />

Dissecting the molecular basis of the regulation of wood formation<br />

by auxin in hybrid aspen. <strong>Plant</strong> Cell 20, 843–855.<br />

Nishiyama R, Kato M, Nagata S, Yanagisawa S, Yoneyama T (2012)<br />

Identification of Zn-nicotianamine <strong>and</strong> Fe-2 ′ -deoxymugineic acid in<br />

the phloem sap from rice plants (Oryza sativa L.). <strong>Plant</strong> Cell Physiol.<br />

53, 381–390.<br />

Notaguchi M, Abe M, Kimura T, Daimon Y, Kobayashi T, Yamaguchi<br />

A, Tomita Y, Dohi K, Mori M, Araki T (2008) Long-distance, grafttransmissible<br />

action of Arabidopsis FLOWERING LOCUS T protein<br />

to promote flowering. <strong>Plant</strong> Cell Physiol. 49, 1645–1658.<br />

Notaguchi M, Wolf S, Lucas WJ (2012) Phloem-mobile Aux/IAA<br />

transcripts target to the root tip <strong>and</strong> modify root architecture. J.<br />

Integr. <strong>Plant</strong> Biol. 54, 760–772.<br />

Oda Y, Fukuda H (2012a) How xylem cells design wood cell walls:<br />

Secondary cell wall patterning by microtubule-associated proteins.<br />

Curr. Opin. <strong>Plant</strong> Biol. 15, 38–44.<br />

Oda Y, Fukuda H (2012b) Initiation of cell wall pattern by a Rho- <strong>and</strong><br />

microtubule-driven symmetry breaking. Science 337, 11333–11336.<br />

Oda Y, Hasezawa S (2006) Cytoskeletal organization during xylem cell<br />

differentiation. J. <strong>Plant</strong> Res. 119, 167–177.<br />

Oda Y, Iida Y, Kondo Y, Fukuda H (2010) Wood cell-wall structure<br />

requires local 2D-microtubule disassembly by a novel plasma<br />

membrane-anchored protein. Curr. Biol. 20, 1197–1202.<br />

Oda Y, Mimura T, Hasezawa S (2005) Regulation of secondary cell<br />

wall development by cortical microtubules during tracheary element<br />

differentiation in Arabidopsis cell suspensions. <strong>Plant</strong> Physiol. 137,<br />

1027–1036.<br />

Oertli JJ (1993) <strong>The</strong> mobility of boron in plants. <strong>Plant</strong> Soil 155, 301–<br />

304.<br />

Ohashi-Ito K, Fukuda H (2003) HD-zip III homeobox genes that include<br />

a novel member, ZeHB-13 (Zinnia)/ATHB-15 (Arabidopsis), are<br />

involved in procambium <strong>and</strong> xylem cell differentiation. <strong>Plant</strong> Cell<br />

Physiol. 44, 1350–1358.<br />

Ohashi-Ito K, Oda Y, Fukuda H (2010) Arabidopsis VASCULAR-<br />

RELATED NAC-DOMAIN6 directly regulates the genes that govern<br />

programmed cell death <strong>and</strong> secondary wall formation during xylem<br />

differentiation. <strong>Plant</strong> Cell 22, 3461–3473.<br />

Oka-Kira E, Tateno K, Miura K-I, Haga T, Hayashi M, Harada<br />

K, Sato S, Tabata S, Shikazono N, Tanaka A, Watanabe Y,<br />

Fukuhara I, Nagata T, Kawaguchi M (2005) klavier (klv), A novel<br />

hypernodulation mutant of Lotus japonicus affected in vascular<br />

tissue organization <strong>and</strong> floral induction. <strong>Plant</strong> J. 44, 505–515.<br />

Okamoto S, Ohnishi E, Sato S, Takahashi H, Nakazono M, Tabata<br />

S, Kawaguchi M (2009) Nod factor/nitrate-induced CLE genes that<br />

drive HAR1-mediated systemic regulation of nodulation. <strong>Plant</strong> Cell<br />

Physiol. 50, 67–77.<br />

Omid A, Keilin T, Glass A, Leshkowitz D, Wolf S (2007) Characterization<br />

of phloem-sap transcription profile in melon plants. J. Exp.<br />

Bot. 58, 3645–3656.<br />

Oren R, Sperry JS, Katul GG, Pataki DE, Ewers BE, Phillips N,<br />

Schafer KVR (1999) Survey <strong>and</strong> synthesis of intra- <strong>and</strong> interspecific<br />

variation in stomatal sensitivity to vapour pressure deficit. <strong>Plant</strong> Cell<br />

Environ. 22, 1515–1526.<br />

Orl<strong>and</strong>o DA, Brady SM, Fink TM, Benfey PN, Ahnert SE (2010)<br />

Detecting separate time scales in genetic expression data. BMC<br />

Genomics 11, 381.<br />

Oross JW, Lucas WJ (1985) Sugar-beet petiole structure – vascular<br />

anastomoses <strong>and</strong> phloem ultrastructure. Can. J. Bot. 63, 2295–<br />

2304.<br />

Osipova MA, Mortier V, Demchenko KN, Tsyganov VE,<br />

Tikhonovich IA, Lutova LA, Dolgikh EA, Goormachtig S (2012)<br />

Wuschel-related homeobox5 gene expression <strong>and</strong> interaction of<br />

CLE peptides with components of the systemic control add two<br />

pieces to the puzzle of autoregulation of nodulation. <strong>Plant</strong> Physiol.<br />

158, 1329–1341.<br />

Ouerdane L, Mari S, Czernic P, Lebrun M, Łobiński R (2006)<br />

Speciation of non-covalent nickel species in plant tissue extracts<br />

by electrospray Q-TOFMS/MS after their isolation by 2D size<br />

exclusion-hydrophilic interaction LC (SEC-HILIC) monitored by ICP-<br />

MS. J. Anal. At. Spec. 21, 676–683.


Palauqui JC, Elmayan T, deBorne FD, Crete P, Charles C,<br />

Vaucheret H (1996) Frequencies, timing, <strong>and</strong> spatial patterns of cosuppression<br />

of nitrate reductase <strong>and</strong> nitrite reductase in transgenic<br />

tobacco plants. <strong>Plant</strong> Physiol. 112, 1447–1456.<br />

Palauqui JC, Elmayan T, Pollien JM, Vaucheret H (1997) <strong>System</strong>ic<br />

acquired silencing: Transgene-specific post-transcriptional silencing<br />

is transmitted by grafting from silenced stocks to non-silenced<br />

scions. EMBO J. 16, 4738–4745.<br />

Pallas JA, Paiva NL, Lamb C, Dixon RA (1996) Tobacco plants epigenetically<br />

suppressed in phenylalanine ammonia-lyase expression<br />

do not develop systemic acquired resistance in response to infection<br />

by tobacco mosaic virus. <strong>Plant</strong> J. 10, 281–293.<br />

Pant BD, Buhtz A, Kehr J, Scheible WR (2008) MicroRNA399 is a<br />

long-distance signal for the regulation of plant phosphate homeostasis.<br />

<strong>Plant</strong> J. 53, 731–738.<br />

Parizot B, Laplaze L, Ricaud L, Boucheron-Dubuisson E, Bayle<br />

V, Bonke M, De Smet I, Poethig SR, Helariutta Y, Haseloff J,<br />

Chriqui D, Beeckman T, Nussaume L (2008) Diarch symmetry of<br />

the vascular bundle in Arabidopsis root encompasses the pericycle<br />

<strong>and</strong> is reflected in distich lateral root initiation. <strong>Plant</strong> Physiol. 146,<br />

140–148.<br />

Park M, Li Q, Shcheynikov N, Zeng W, Muallem S (2004) NaBC1 is<br />

a ubiquitous electrogenic Na + -coupled borate transporter essential<br />

for cellular boron homeostasis <strong>and</strong> cell growth <strong>and</strong> proliferation. Mol.<br />

Cell 16, 331–341.<br />

Park SW, Kaimoyo E, Kumar D, Mosher S, Klessig D (2007) Methyl<br />

salicylate is a critical mobile signal for plant systemic acquired<br />

resistance. Science 318, 113–116.<br />

Park SW, Liu PP, Forouhar F, Vlot AC, Tong L, Tietjen K, Klessig D<br />

(2009) Use of a synthetic salicylic acid analog to investigate the roles<br />

of methyl salicylate <strong>and</strong> its esterases in plant disease resistance. J.<br />

Biol. Chem. 284, 7307–7317.<br />

Parker J, Holub E, Frost L, Falk A, Gunn N, Daniels M (1996)<br />

Characterization of eds1, a mutation in Arabidopsis suppressing<br />

resistance to Peronospora parasitica specified by several different<br />

RPP genes. <strong>Plant</strong> Cell 8, 2033–2046.<br />

Passioura JB, Ashford AE (1974) Rapid translocation in the phloem<br />

of wheat roots. Aust. J. <strong>Plant</strong> Physiol. 1, 521–527.<br />

Peguero-Pina JS, Alquezar-Alquezar JM, Mayr S, Cochard H, Gil-<br />

Pelegrin D (2011) Embolism induced by winter drought in Pinus<br />

sylvestris L.: A possible reason for dieback near its southern<br />

distribution limit? Ann. For. Sci. 38, 565–574.<br />

Peleg-Grossman S, Golani Y, Kaye Y, Melamed-Book N, Levine<br />

A (2009) NPR1 protein regulates pathogenic <strong>and</strong> symbiotic interactions<br />

between Rhizobium <strong>and</strong> legumes <strong>and</strong> non-legumes. PLoS<br />

ONE 4, e8399.<br />

Pérez-Alfocea F, Ghanem ME, Gómez-Cadenas A, Dodd IC (2011)<br />

Omics of root-to-shoot signaling under salt stress <strong>and</strong> water deficit.<br />

OMIC 15, 893–901.<br />

Pesacreta TC, Groom LH, Rials TG (2005) Atomic force microscopy<br />

of the intervessel pit membrane in the stem of Sapium sebiferum<br />

(Euphorbiaceae). IAWA J. 26, 397–426.<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 379<br />

Pesquet E, Ranocha P, Legay S, Digonnet C, Barbier O, Pichon<br />

M, Goffner D (2005) Novel markers of xylogenesis in zinnia are<br />

differentially regulated by auxin <strong>and</strong> cytokinin. <strong>Plant</strong> Physiol. 139,<br />

1821–1839.<br />

Pesquet E, Korolev AV, Calder G, Lloyd CW (2010) <strong>The</strong> microtubuleassociated<br />

protein AtMAP70–5 regulates secondary wall patterning<br />

in Arabidopsis wood cells. Curr. Biol. 20, 744–749.<br />

Petráˇsek J, Mravec J, Bouchard R, Blakeslee JJ, Abas M, Seifertova<br />

D, Wisniewska J, Tadele Z, Kubes M, Covanova M,<br />

Dhonukshe P, Skupa P, Benkova E, Perry L, Krecek P, Lee<br />

OR, Fink GR, Geisler M, Murphy AS, Luschnig C, Zazimalova<br />

E, Friml J (2006) PIN proteins perform a rate-limiting function in<br />

cellular auxin efflux. Science 312, 914–918.<br />

Petráˇsek J, Friml J (2009) Auxin transport routes in plant development.<br />

<strong>Development</strong> 136, 2675–2688.<br />

Petty JA (1972) <strong>The</strong> aspiration of bordered pits in conifer wood. Proc.<br />

R. Soc. Ser. B. Biol. Sci. 181, 395–406.<br />

Petty JA, Preston RD (1969) <strong>The</strong> dimensions <strong>and</strong> number of pit<br />

membrane pores in conifer wood. Proc. R. Soc. Ser. B. Biol. Sci.<br />

172, 137–151.<br />

Peuke AD, Windt C, van As H (2006) Effects of cold girdling on flows in<br />

the transport phloem in Ricinus communis: Is mass flow inhibited?<br />

<strong>Plant</strong> Cell Environ. 29, 15–25.<br />

Pich A, Scholz G (1996) Translocation of copper <strong>and</strong> other micronutrients<br />

in tomato plants (Lycopersicon esculentum Mill.):<br />

Nicotianamine-stimulated copper transport in the xylem. J. Exp. Bot.<br />

47, 41–47.<br />

Pich A, Scholz G, Stephan U (1994) Iron-dependent changes of<br />

heavy-metals, nicotianamine, <strong>and</strong> citrate in different plant organs<br />

<strong>and</strong> in the xylem exudate of 2 tomato genotypes. Nicotianamine<br />

as possible copper translocator. <strong>Plant</strong> Soil 165, 189–<br />

196.<br />

Pickard WF (1981) <strong>The</strong> ascent of sap in plants. Progr. Biophys. Mol.<br />

Biol. 37, 181–229.<br />

Pieruschka R, Huber G, Berry JA (2010) Control of transpiration by<br />

radiation. Proc. Natl. Acad. Sci. USA 107, 13372–13377.<br />

Piquemal J, LaPierre C, Myton K, O’Connel A, Schuch W, Grima-<br />

Pettenati J, Boudet AM (1998) Down-regulation of cinnamoyl-CoA<br />

reductase induces significant changes of lignin profiles in transgenic<br />

tobacco plants. <strong>Plant</strong> J. 13, 71–83.<br />

Pires ND, Dolan L (2012) Morphological evolution in l<strong>and</strong> plants: New<br />

designs with old genes. Phil Trans. R. Soc. 367, 508–518.<br />

Pittermann J (2010) <strong>The</strong> evolution of water transport in plants: An<br />

integrated approach. Goebiology 8, 112–139.<br />

Pittermann J, Sperry JS (2003) Tracheid diameter is the key trait<br />

determining extent of freezing-induced cavitation in conifers. Tree<br />

Physiol. 23, 907–914.<br />

Pittermann J, Sperry JS (2006) Analysis of freeze-thaw embolism in<br />

conifers: <strong>The</strong> interaction between cavitation pressure <strong>and</strong> tracheid<br />

size. <strong>Plant</strong> Physiol. 140, 374–382.<br />

Pittermann J, Sperry JS, Hacke UG, Wheeler JK, Sikkema EH (2005)<br />

Torus-margo pits help conifers compete with angiosperms. Science


380 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

310, 1924.<br />

Pittermann J, Sperry JS, Hacke UG, Wheeler JK, Sikkema EH<br />

(2006a) Inter-tracheid pitting <strong>and</strong> the hydraulic efficiency of conifer<br />

wood: <strong>The</strong> role of tracheid allometry <strong>and</strong> cavitation protection. Am.<br />

J. Bot. 93, 1105–1113.<br />

Pittermann J, Sperry JS, Wheeler JK, Hacke UG, Sikkema EH<br />

(2006b) Mechanical reinforcement of tracheids compromises the<br />

hydraulic efficiency of conifer xylem. <strong>Plant</strong> Cell Environ. 29, 1618–<br />

1628.<br />

Plaut JA, Yepez EA, Hill J, Pangle R, Sperry JS, Pockman WT,<br />

McDowell NG (2012) Hydraulic limits preceeding mortality in a<br />

pinon-juniper woodl<strong>and</strong> under experimental drought. <strong>Plant</strong> Cell<br />

Environ. 35, 1601–1617.<br />

Pockman WT, Sperry JS (1997) Freezing-induced xylem cavitation<br />

<strong>and</strong> the northern limit of Larrea tridentata. Oecologia 109, 19–27.<br />

Poirier Y, Thoma S, Somerville C, Schiefelbein J (1991) Mutant of<br />

Arabidopsis deficient in xylem loading of phosphate. <strong>Plant</strong> Physiol.<br />

97, 1087–1093.<br />

Pommerrenig B, Papini-Terzi FS, Sauer N (2007) Differential regulation<br />

of sorbitol <strong>and</strong> sucrose loading into the phloem of <strong>Plant</strong>ago<br />

major in response to salt stress. <strong>Plant</strong> Physiol. 144, 1029–1038.<br />

Pressel S, Ligrone R, Duckett JG (2006) Effects of de- <strong>and</strong> rehydration<br />

on food-conducting cells in the moss Polytrichum formosum: A<br />

cytological study. Ann. Bot. 98, 67–76.<br />

Prigge MJ, Clark SE (2006) <strong>Evolution</strong> of the Class III HDd-ZIP gene<br />

family in l<strong>and</strong> plants. Evol. Dev. 8, 350–361.<br />

Prigge MJ, Otsuga D, Alonso JM, Ecker JR, Drews GN, Clark SE<br />

(2005) Class III homeodomain-leucine zipper gene family members<br />

have overlapping, antagonistic, <strong>and</strong> distinct roles in Arabidopsis<br />

development. <strong>Plant</strong> Cell 17, 61–76.<br />

Pritchard J (1996) Aphid stylectomy reveals an osmotic step between<br />

sieve tube, cortical cells in barley roots. J. Exp. Bot. 47, 1519–<br />

1524.<br />

Pritchard J, Tomos AD, Farrar JF, Minchin PEH, Gould N, Paul MJ,<br />

MacRae EA, Ferrier RA, Gray DW, Thorpe MR (2004) Turgor,<br />

solute import, growth in maize roots treated with galactose. Funct.<br />

<strong>Plant</strong> Biol. 31, 1095–1103.<br />

Puig S, Penarrubia L (2009) Placing metal micronutrients in context:<br />

Transport <strong>and</strong> distribution in plants. Curr. Op. <strong>Plant</strong> Biol. 12, 299–<br />

306.<br />

Pyo H, Demura T, Fukuda H (2007) TERE; a novel cis-element responsible<br />

for a coordinated expression of genes related to programmed<br />

cell death <strong>and</strong> secondary wall formation during differentiation of<br />

tracheary elements. <strong>Plant</strong> J. 51, 955–965.<br />

Rahayu YS, Walch-Liu P, Neumann G, Römheld V, von Wirén N,<br />

Bangerth F (2005) Root-derived cytokinins as long-distance signals<br />

for NO −<br />

3 -induced stimulation of leaf growth. J. Exp. Bot. 56, 1143–<br />

1152.<br />

Ransom-Hodgkins WD, Vaughn MW, Bush DR (2003) Protein phosphorylation<br />

plays a key role in sucrose-mediated transcriptional<br />

regulation of a phloem-specific proton-sucrose symporter. <strong>Plant</strong>a<br />

217, 483–489.<br />

Raridan GJ, Delaney TP (2002) Role of salicylic acid <strong>and</strong> NIM1/NPR1<br />

in race-specific resistance in Arabidopsis. Genetics 161, 803–811.<br />

Raven JA (1991) Long-term functioning of enucleate sieve elements<br />

- possible mechanisms of damage avoidance <strong>and</strong> damage repair.<br />

<strong>Plant</strong> Cell Environ. 14, 139–146.<br />

Raven JA (2003) Long-distance transport in non-vascular plants. <strong>Plant</strong><br />

Cell Environ. 26, 73–85.<br />

Reid DE, Ferguson BJ, Gresshoff PM (2011) Inoculation- <strong>and</strong> nitrateinduced<br />

CLE peptides of soybean control NARK-dependent nodule<br />

formation. Mol. <strong>Plant</strong>-Microbe Interact. 24, 606–618.<br />

Reinhardt D, M<strong>and</strong>el T, Kuhlemeier C (2000) Auxin regulates the<br />

initiation <strong>and</strong> radial position of plant lateral organs. <strong>Plant</strong> Cell 12,<br />

507–518.<br />

Reinhardt D, Pesce ER, Stieger P, M<strong>and</strong>el T, Baltensperger K,<br />

Bennett M, Traas J, Friml J, Kuhlemeier C (2003) Regulation<br />

of phyllotaxis by polar auxin transport. Nature 426, 255–260.<br />

Rellán-Álvarez R, Abadía J, Álvarez-Fernández A (2008) Formation<br />

of metal-nicotianamine complexes as affected by pH, lig<strong>and</strong> exchange<br />

with citrate <strong>and</strong> metal exchange. A study by electrospray<br />

ionization time-of-flight mass spectrometry. Rapid Commun. Mass<br />

Sp. 22, 1553–1562.<br />

Rellán-Álvarez R, Giner-Martínez-Sierra J, Orduna J, Orera I,<br />

Rodríguez-Castrillón JA, García-Alonso JI, Abadía J, Álvarez-<br />

Fernández A (2010) Identification of a tri-iron(III), tri-citrate complex<br />

in the xylem sap of iron-deficient tomato resupplied with iron: New<br />

insights into plant iron long-distance transport. <strong>Plant</strong> Cell Physiol.<br />

51, 91–102.<br />

Rennie EA, Turgeon R (2009) A comprehensive picture of phloem<br />

loading strategies. Proc. Natl. Acad. Sci. USA 106, 14162–14167.<br />

Riesen O, Feller U (2005) Redistribution of nickel, cobalt, manganese,<br />

zinc, <strong>and</strong> cadmium via the phloem in young <strong>and</strong> maturing wheat.<br />

J. <strong>Plant</strong> Nutr. 28, 421–430.<br />

Roberts MR, Paul ND (2006) Seduced by the dark side: Integrating<br />

molecular <strong>and</strong> ecological perspectives on the influence of light on<br />

plant defense against pests <strong>and</strong> pathogens. New Phytol. 170, 677–<br />

699.<br />

Robischon M, Du J, Miura E, Groover A (2011) <strong>The</strong> Populus<br />

Class III HD ZIP, popREVOLUTA, influences cambium initiation<br />

<strong>and</strong> patterning of woody stems. <strong>Plant</strong> Physiol. 155, 1214–<br />

1225.<br />

Rodriguez-Medina C, Atkins CA, Mann AJ, Jordan ME, Smith PMC<br />

(2011) Macromolecular composition of phloem exudate from white<br />

lupin (Lupinus albus L.). BMC <strong>Plant</strong> Biol. 11, 36.<br />

Rogers EE, Ausubel FM (1997) Arabidopsis enhanced disease susceptibility<br />

mutants exhibit enhanced susceptibility to several bacterial<br />

pathogens <strong>and</strong> alterations in PR-1 gene expression. <strong>Plant</strong> Cell<br />

9, 305–316.<br />

Roney JK, Khatibi PA, Westwood JH (2007) Cross-species translocation<br />

of mRNA from host plants into the parasitic plant dodder.<br />

<strong>Plant</strong> Physiol. 143, 1037–1043.<br />

Rosche E, Blackmore D, Tegeder M, Richardson T, Schroeder<br />

H, Higgins TJV, Frommer WB, Offler CE, Patrick JW (2002)


Seed-specific expression of a potato sucrose transporter increases<br />

sucrose uptake, growth rates of developing pea cotyledons. <strong>Plant</strong><br />

J. 30, 165–175.<br />

Roschzttardtz H, Séguéla-Arnaud M, Briat J, Vert G, Curie C (2011)<br />

<strong>The</strong> FRD3 citrate effluxer promotes iron nutrition between symplastically<br />

disconnected tissues throughout Arabidopsis development.<br />

<strong>Plant</strong> Cell 23, 2725–2737.<br />

Rothwell GW, Lev-Yadun S (2005) Evidence of polar auxin flow in 375<br />

million-year-old fossil wood. Am. J. Bot. 92, 903–906.<br />

Rouached H, Stefanovic A, Secco D, Bulak Arpat A, Gout E,<br />

Bligny R, Poirier Y (2011) Uncoupling phosphate deficiency from<br />

its major effects on growth <strong>and</strong> transcriptome via PHO1 expression<br />

in Arabidopsis. <strong>Plant</strong> J. 65, 557–570.<br />

Rueffer M, Steipe B, Zenk MH (1995) Evidence against specific<br />

binding of salicylic acid to plant catalase. FEBS Lett. 377, 175–180.<br />

Ruffel S, Krouk G, Ristova D, Shasha D, Birnbaum KD, Coruzzi<br />

GM (2011) Nitrogen economics of root foraging: Transitive closure<br />

of the nitrate-cytokinin relay <strong>and</strong> distinct systemic signaling for N<br />

supply vs. dem<strong>and</strong>. Proc. Natl. Acad. Sci. USA 108, 18524–18529.<br />

Ruiz-Medrano R, Xoconostle-Cázares B, Lucas WJ (1999) Phloem<br />

long-distance transport of CmNACP mRNA: Implications for supracellular<br />

regulation in plants. <strong>Development</strong> 126, 4405–4419.<br />

Ruszala EM, Beerling DJ, Franks PJ, Chater C, Casson SA, Gray<br />

JE, Hetherington AM (2011) L<strong>and</strong> plants acquired active stomatal<br />

control early in their evolutionary history. Curr. Biol. 21, 1030–1035.<br />

Ruyter-Spira C, Al-Babili S, van der Krol S, Bouwmeester H (2012)<br />

<strong>The</strong> biology of strigolactones. Trends <strong>Plant</strong> Sci. 12, 1360–1385.<br />

Ryals J, Weymann K, Lawton K, Friedrich L, Ellis D, Steiner H-<br />

Y, Johnson J, Delaney TP, Jesse T, Vos P, Uknes S (1997)<br />

<strong>The</strong> Arabidopsis NIM1 protein shows homology to the mammalian<br />

transcription factor inhibitor IκB. <strong>Plant</strong> Cell 9, 425–439.<br />

Sachs T (1981) <strong>The</strong> control of patterned differentiation of vascular<br />

tissues. Adv. Bot. Res. 9, 152–262.<br />

Sachs T (1991) Pattern Formation in <strong>Plant</strong>s. Cambridge University<br />

Press, Cambridge UK.<br />

Salim E, Ullsten O (1999) Our forests our future. Report of the World<br />

Commission on Forests <strong>and</strong> Sustainable <strong>Development</strong>, Cambridge<br />

University Press, Cambridge UK.<br />

Salleo S, Lo Gullo MA, De Paoli D, Zippo M (1996) Xylem recovery<br />

from cavitation-induced embolism in young plants of Laurus nobilis:<br />

A possible mechanism. New Phytol. 132, 47–56.<br />

Salleo S, Trifilo P, Esposito S, Nardini A, LoGullo MA (2009) Starchto-sugar<br />

conversion in wood parenchyma of field-growing Laurus<br />

nobilis plants: A component of the signal pathway for embolism<br />

repair? Funct. <strong>Plant</strong> Biol. 36, 815–825.<br />

Salleo S, Trifilo P, LoGullo MA (2006) Phloem as a possible major<br />

determinant of rapid cavitation reversal in Laurus nobilis (laurel).<br />

Funct. <strong>Plant</strong> Biol. 33, 1063–1074.<br />

Salt DE, Prince RC, Baker AJM, Raskin I, Pickering IJ (1999)<br />

Zinc lig<strong>and</strong>s in the metal hyperaccumulator Thlaspi caerulescens<br />

as determined using x-ray absorption spectroscopy. Environ. Sci.<br />

Technol. 33, 713–717.<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 381<br />

Sannigrahi P, Ragauskas AJ, Tuskan GA (2010) Poplar as a feedstock<br />

for biofuels: A review of compositional characteristics. Biofuels<br />

Bioprod. Bioref. 4, 209–226.<br />

Sasaki T, Chino M, Hayashi H, Fujiwara T (1998) Detection of several<br />

mRNA species in rice phloem sap. <strong>Plant</strong> Cell Physiol. 39, 895–897.<br />

Sawa S, Ito T, Shimura Y, Okada K (1999) Filamentous flower controls<br />

the formation <strong>and</strong> development of Arabidopsis inflorescences <strong>and</strong><br />

floral meristems. <strong>Plant</strong> Cell 11, 69–86.<br />

Sawchuck T, Head P, Donner TJ, Scarpella E (2007) Time-lapse<br />

imaging of Arabidopsis leaf development shows dynamic patterns<br />

of procambium formation. New Phytol. 176, 560–571.<br />

Sawchuk MG, Donner TJ, Scarpella E (2008) Auxin transportdependent,<br />

stage-specific dynamics of leaf vein formation. <strong>Plant</strong><br />

Signal. Behav. 3, 286–289.<br />

Scarpella E, Barkoulas M, Tsiantis M (2010) Control of leaf <strong>and</strong><br />

vein development by auxin. Cold Spring Harb. Perspect. Biol. 2,<br />

a001511.<br />

Scarpella E, Francis P, Berleth T (2004) Stage-specific markers<br />

define early steps of procambium development in Arabidopsis<br />

leaves <strong>and</strong> correlate termination of vein formation with mesophyll<br />

differentiation. <strong>Development</strong> 131, 3445–3455.<br />

Scarpella E, Helariutta Y (2010) <strong>Vascular</strong> pattern formation in plants.<br />

Curr. Op. Dev. Biol. 91, 221–265.<br />

Scarpella E, Marcos D, Friml JÃ, Berleth T (2006) Control of leaf<br />

vascular patterning by polar auxin transport. Genes Dev. 20, 1015–<br />

1027.<br />

Scarpella E, Meijer AH (2004) Pattern formation in the vascular system<br />

of monocot <strong>and</strong> dicot plant species. New Phytol. 164, 209–242.<br />

Schaaf G, Schikora A, Haberle J, Vert G, Ludewig U, Briat J, Curie<br />

C, von Wirén N (2005) A putative function for the Arabidopsis<br />

Fe-phytosiderophore transporter homolog AtYSL2 in Fe <strong>and</strong> Zn<br />

homeostasis. <strong>Plant</strong> Cell Physiol. 46, 762–774.<br />

Schachtman DP, Goodger JQD (2008) Chemical root to shoot signaling<br />

under drought. Trends <strong>Plant</strong> Sci. 6, 281–287.<br />

Schaumlöffel D, Ouerdane L, Bouyssiere B, Łobiński R (2003)<br />

Speciation analysis of nickel in the latex of a hyperaccumulating<br />

tree Sebertia acuminata by HPLC <strong>and</strong> CZE with ICP MS <strong>and</strong><br />

electrospray MS-MS detection. J. Anal. At. Spect. 18, 120–127.<br />

Schenk PM, Kazan K, Wilson I, Anderson JP, Richmond T,<br />

Somerville SC, Manners JM (2000) Coordinated plant defense<br />

responses in Arabidopsis revealed by microarray analysis. Proc.<br />

Natl. Acad. Sci. USA 97, 11655–11660.<br />

Scheres B (2010) <strong>Development</strong>al biology: Roots respond to an inner<br />

calling. Nature 465, 299–300.<br />

Scheres B, Di Laurenzio L, Willemsen V, Hauser MT, Janmaat<br />

K, Weisbeek P, Benfey PN (1995) Mutations affecting the radial<br />

organisation of the Arabidopsis root display specific defects throughout<br />

the embryonic axis. <strong>Development</strong> 121, 53–62.<br />

Schlereth A, Moller B, Liu W, Kientz M, Flipse J, Rademacher EH,<br />

Schmid M, Jurgens G, Weijers D (2010) MONOPTEROS controls<br />

embryonic root initiation by regulating a mobile transcription factor.<br />

Nature 464, 913–916.


382 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Schlesinger WH (1997) Biogeochemistry. Academic Press, San<br />

Diego.<br />

Schrader J, Baba K, May ST, Palme K, Bennett M, Bhalerao RP,<br />

S<strong>and</strong>berg G (2003) Polar auxin transport in the wood-forming<br />

tissues of hybrid aspen is under simultaneous control of developmental<br />

<strong>and</strong> environmental signals. Proc. Natl. Acad. Sci. USA 100,<br />

10096–10101.<br />

Schreiber SG, Hacke UG, Hamann A, Thomas BR (2011) Genetic<br />

variation of hydraulic <strong>and</strong> wood anatomical traits in hybrid poplar<br />

<strong>and</strong> trembling aspen. New Phytol. 190, 150–160.<br />

Schuetz M, Berleth T, Mattsson J (2008) Multiple MONOPTEROSdependent<br />

pathways are involved in leaf initiation. <strong>Plant</strong> Physiol.<br />

148, 870–880.<br />

Schuler M, Lehmann M, Fink-Straube C, Bauer P (2010) <strong>The</strong><br />

interaction of NAS genes <strong>and</strong> FRD3 in the long-distance transport of<br />

iron in Arabidopsis thaliana. 15th International Symposium on Iron<br />

Nutrition <strong>and</strong> Interactions in <strong>Plant</strong>s. Budapest, Hungary.<br />

Schulz A (1992) Living sieve cells of conifers as visualized by confocal,<br />

laser-scanning fluorescence microscopy. Protoplasma 166, 153–<br />

164.<br />

Searle IR, Men AE, Laniya TS, Buzas DM, Iturbe-Ormaetxe I, Carroll<br />

BJ, Gresshoff PM (2003) Long-distance signaling in nodulation<br />

directed by a CLAVATA1-like receptor kinase. Science 299, 109–<br />

112.<br />

Secchi F, Zwieniecki MA (2010) Patterns of PIP gene expression in<br />

Populus trichocarpa during recovery from xylem embolism suggest<br />

a major role for the PIP1 aquaporin subfamily as moderators of the<br />

refilling process. <strong>Plant</strong> Cell Environ. 33, 1285–1297.<br />

Sevanto S, Holbrook NM, Ball MC (2012) Freeze/thaw-induced embolism:<br />

Probability of critical bubble formation depends on speed of<br />

ice formation. <strong>Plant</strong> Sci. 3, 1–12.<br />

Shah J, Tsui F, Klessig DF (1997) Characterization of a salicylic<br />

acid-insensitive mutant (sai1) ofArabidopsis thaliana identified in<br />

a selective screen utilizing the SA-inducible expression of the tms2<br />

gene. Mol. <strong>Plant</strong>-Microbe Interact. 10, 69–78.<br />

Shah J, Kachroo P, N<strong>and</strong>i A, Klessig D (2001) A recessive mutation<br />

in the Arabidopsis ssi2 gene confers SA- <strong>and</strong> NPR1-independent<br />

expression of PR genes <strong>and</strong> resistance against bacterial <strong>and</strong><br />

oomycete pathogens. <strong>Plant</strong> J. 25, 563–574.<br />

Shalit A, Rozman A, Goldshmidt A, Alvarez JP, Bowman JL, Eshed<br />

Y, Lifschitz E (2009) <strong>The</strong> flowering hormone florigen functions as<br />

a general systemic regulator of growth <strong>and</strong> termination. Proc. Natl.<br />

Acad. Sci. USA 106, 8392–8397.<br />

Shapiro AD, Zhang C (2001) <strong>The</strong> role of NDR1 in avirulence genedirected<br />

signaling <strong>and</strong> control of programmed cell death in Arabidopsis.<br />

<strong>Plant</strong> Physiol. 127, 1089–1101.<br />

Sharp RG, Davies WJ (2009) Variability among species in the apoplastic<br />

pH signalling response to drying soils. J. Exp. Bot. 60, 4363–<br />

4370.<br />

Shulaev V, Leon J, Raskin I (1995) Is salicylic-acid a translocated<br />

signal of systemic acquired-resistance in tobacco. <strong>Plant</strong> Cell 7,<br />

1691–1701.<br />

Shulaev V, Silverman P, Raskin I (1997) Airborne signaling by methyl<br />

salicylate in plant pathogen resistance. Nature 385, 718–721.<br />

Sieburth LE (1999) Auxin is required for leaf vein pattern in Arabidopsis.<br />

<strong>Plant</strong> Physiol. 121, 1179–1190.<br />

Sieburth LE, Lee DK (2010) BYPASS1: How a tiny mutant tells a big<br />

story about root-to-shoot signaling. J. Integr. <strong>Plant</strong> Biol. 52, 77–85.<br />

Sinclair SA, Sherson SM, Jarvis R, Camakaris J, Cobbett CS (2007)<br />

<strong>The</strong> use of the zinc-fluorophore, Zinpyr-1, in the study of zinc<br />

homeostasis in Arabidopsis roots. New Phytol. 174, 39–45.<br />

Slaymaker DH, Navarre DA, Clark D, Pozo OD, Martin GB, Klessig<br />

DF (2002) <strong>The</strong> tobacco salicylic acid-binding protein 3 (SABP3)<br />

is the chloroplast carbonic anhydrase, which exhibits antioxidant<br />

activity <strong>and</strong> plays a role in the hypersensitive defense response.<br />

Proc. Natl. Acad. Sci. USA 99, 11640–11645.<br />

Smith AM, Stitt M (2007) Coordination of carbon supply <strong>and</strong> plant<br />

growth. <strong>Plant</strong> Cell Environ. 30, 1126–1149.<br />

Smith JAC, Milburn JA (1980a) Osmoregulation, the control of<br />

phloem-sap composition in Ricinus communis. <strong>Plant</strong>a 148, 28–34.<br />

Smith JAC, Milburn JA. (1980b) Phloem turgor, the regulation of<br />

sucrose loading in Ricinus communis. <strong>Plant</strong>a 148, 42–48.<br />

Snow R (1935) Activation of cambial growth by pure hormones. New<br />

Phytol. 34, 347–360.<br />

Song JT, Koo YJ, Seo HK, Kim MC, CHoi YD, Kim JH (2008)<br />

Overexpression of AtSGT1, an Arabidopsis salicylic acid glucosyltransferase,<br />

leads to increased susceptibility to Pseudomonas<br />

syringae. Phytochemistry 69, 1128–1134.<br />

Song JT, Lu H, Greenberg JT (2004b) Divergent roles in Arabidopsis<br />

thaliana development <strong>and</strong> defense of two homologous genes, aberrant<br />

growth <strong>and</strong> death 2 <strong>and</strong> AGD2-LIKE DEFENSE RESPONSE<br />

PROTEIN 1, encoding novel aminotransferases. <strong>Plant</strong> Cell 16, 353–<br />

366.<br />

Song JT, Lu H, McDowell JM, Greenberg JT (2004a) A key role for<br />

ALD1 in activation of local <strong>and</strong> systemic defenses in Arabidopsis.<br />

<strong>Plant</strong> J. 40, 200–212.<br />

Soyano T, Thitamadee S, Machida Y, Chua NH (2008) ASYMMETRIC<br />

LEAVES2-LIKE19/LATERAL ORGAN BOUNDARIES DOMAIN30<br />

<strong>and</strong> ASL20/LBD18 regulate tracheary element differentiation in<br />

Arabidopsis. <strong>Plant</strong> Cell 20, 3359–3373.<br />

Sperotto RA, Boff T, Duarte GL, Santos LS, Grusak MA, Fett<br />

JP (2010) Identification of putative target genes to manipulate Fe<br />

<strong>and</strong> Zn concentrations in rice grains. J. <strong>Plant</strong> Physiol. 167, 1500–<br />

1506.<br />

Sperotto RA, Ricachenevsky FK, Waldow VA, Fett JP (2012) Iron<br />

biofortification in rice: It’s a long way to the top. <strong>Plant</strong> Sci. 190,<br />

24–39.<br />

Sperry JS (1993) Winter xylem embolism <strong>and</strong> spring recovery in Betula<br />

cordifolia, Fagus gr<strong>and</strong>ifolia, Abies balsamea, <strong>and</strong> Picea rubens.<br />

In: Borghetti M, Grace J, Raschi A, eds. Water Transport in <strong>Plant</strong>s<br />

under Climatic Stress. Cambridge University Press, Cambridge. pp.<br />

86–98.<br />

Sperry JS (2000) Hydraulic constraints on plant gas exchange. Ag.<br />

For. Meteorol. 2831, 1–11.


Sperry JS (2003) <strong>Evolution</strong> of water transport <strong>and</strong> xylem structure. Int.<br />

J. <strong>Plant</strong> Sci. 164, S115–S127.<br />

Sperry JS (2011) Hydraulics of vascular water transport In: P. W, ed.<br />

Mechanical integration of plant cells <strong>and</strong> plants. Springer-Verlag<br />

Berlin. pp. 303–327.<br />

Sperry JS, Christman MA, Smith DD (2012) Vulnerability curves by<br />

centrifugation: Is there an open vessel artifact, <strong>and</strong> are “r” shaped<br />

curves necessarily invalid? <strong>Plant</strong> Cell Environ. 35, 601–610.<br />

Sperry JS, Hacke UG, Oren R, Comstock JP (2002) Water deficits<br />

<strong>and</strong> hydraulic limits to leaf water supply. <strong>Plant</strong> Cell Environ. 25,<br />

251–263.<br />

Sperry JS, Holbrook NM, Zimmermann MH, Tyree MT (1987) Spring<br />

filling of xylem vessels in wild grapevine. <strong>Plant</strong> Physiol. 83, 414–417.<br />

Sperry JS, Tyree MT (1988) Mechanism of water stress-induced xylem<br />

embolism. <strong>Plant</strong> Physiol. 88, 581–587.<br />

Sperry JS, Tyree MT (1990) Water-stress-induced xylem embolism in<br />

three species of conifers. <strong>Plant</strong> Cell Environ. 13, 427–436.<br />

Spicer R, Groover A (2010) <strong>The</strong> evolution of development of the<br />

vascular cambium <strong>and</strong> secondary growth. New Phytol. 186, 577–<br />

592.<br />

Spoel SH, Dong X (2008) Making sense of hormone cross talk during<br />

plant immune response. Cell Host Microbe 3, 348–351.<br />

Spoel SH, Dong X (2012) How do plants achieve immunity? Defence<br />

without specialized immune cell. Nat. Rev. Immunol. 12,<br />

89–100.<br />

Spoel SH, Mou Z, Tada Y, Spivey NW, Genshik P, Dong X<br />

(2009) Proteasome-mediated turnover of the transcription coactivator<br />

NPR1 plays dual roles in regulating plant immunity. Cell 137,<br />

860–872.<br />

Stacey MG, Patel A, McClain WE, Mathieu M, Remley M, Rogers<br />

EE, Gassmann W, Blevins DG, Stacey G (2008) <strong>The</strong> Arabidopsis<br />

AtOPT3 protein functions in metal homeostasis <strong>and</strong> movement of<br />

iron to developing seeds. <strong>Plant</strong> Physiol. 146, 589–601.<br />

Stadler R, Wright KM, Lauterbach C, Amon G, Gahrtz M, Feuerstein<br />

A, Oparka KJ, Sauer N (2005) Expression of GFP-fusions in<br />

Arabidopsis companion cells reveals non-specific protein trafficking<br />

into sieve elements, identifies a novel post-phloem domain in roots.<br />

<strong>Plant</strong> J. 41, 319–331.<br />

Staehelin C, Xie ZP, Illana A, Vierheilig H (2011) Long-distance<br />

transport of signals during symbiosis: Are nodule formation <strong>and</strong><br />

mycorrhization autoregulated in a similar way? <strong>Plant</strong> Signal. Behav.<br />

6, 372–377.<br />

Stefanovic A, Ribot C, Rouached H, Wang Y, Chong J, Belbahri<br />

L, Delessert S, Poirier Y (2007) Members of the PHO1 gene<br />

family show limited functional redundancy in phosphate transfer<br />

to the shoot, <strong>and</strong> are regulated by phosphate deficiency via distinct<br />

pathways. <strong>Plant</strong> J. 50, 982–994.<br />

Stiller V, Lafitte HR, Sperry JS (2005) Embolized conduits of rice<br />

(Oryza sativa L.) refill despite negative xylem pressure. Am. J. Bot.<br />

92, 1970–1974.<br />

Stiller V, Sperry JS (2002) Cavitation fatigue <strong>and</strong> its reversal in intact<br />

sunflower plants. J. Exp. Bot. 53, 1155–1161.<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 383<br />

Strawn MA, Marr SK, Inoue K, Inada N, Zubieta C, Wildermuth MC<br />

(2007) Arabidopsis isochorismate synthase functional in pathogeninduced<br />

salicylate biosynthesis exhibits properties consistent with<br />

a role in diverse stress responses. J. Biol. Chem. 282, 5919–<br />

5933.<br />

Street N, Jansson S, Hvidsten T (2011) A systems biology model<br />

of the regulatory network in populus leaves reveals interacting<br />

regulators <strong>and</strong> conserved regulation. BMC <strong>Plant</strong> Biol. 11, 13.<br />

Suer S, Agusti J, Sanchez P, Schwarz M, Greb T (2011) WOX4<br />

imparts auxin responsiveness to cambium cells in Arabidopsis.<br />

<strong>Plant</strong> Cell 23, 3247–3259.<br />

Sussex IM (1954) Experiments on the cause of dorsiventrality in leaves.<br />

Nature 167, 651–652.<br />

Svistoonoff S, Creff A, Reymond M, Sigoillot-Claude C, Ricaud<br />

L, Blanchet A, Nussaume L, Desnos T (2007) Root tip contact<br />

with low-phosphate media reprograms plant root architecture. Nat.<br />

Genet. 39, 792–796.<br />

Sweetlove LJ, Kossmann J, Riesmeier JW, Riesmeier JW,<br />

Trethewey RN, Hill SA (1998) <strong>The</strong> control of source to sink carbon<br />

flux during tuber development in potato. <strong>Plant</strong> J. 15, 697–706.<br />

Tada Y, Spoel SH, Pajerowska-Muhktar K, Mou Z, Song J, Wang<br />

C, Zuo J, Dong, X (2008) <strong>Plant</strong> immunity requires conformational<br />

changes of NPR1 via S-nitrosylation <strong>and</strong> thoredoxins. Science 321,<br />

952–956.<br />

Takahashi H, Miller J, Nozaki Y, Takeda M, Shah J, Hase S,<br />

Ikegami M, Ehara Y, Dinesh-Kumar SP (2002) RCY1, anArabidopsis<br />

thaliana RPP8/HRT family resistance gene, conferring<br />

resistance to cucumber mosaic virus requires salicylic acid, ethylene<br />

<strong>and</strong> a novel signal transduction mechanism. <strong>Plant</strong> J. 32, 655–<br />

667.<br />

Takahashi M, Terada Y, Nakai I, Nakanishi H, Yoshimura E, Mori<br />

S, Nishizawa N (2003) Role of nicotianamine in the intracellular<br />

delivery of metals <strong>and</strong> plant reproductive development. <strong>Plant</strong> Cell<br />

15, 1263–1280.<br />

Takano J, Miwa K, Fujiwara T (2008) Boron transport mechanisms:<br />

Collaboration of channels <strong>and</strong> transporters. Trends <strong>Plant</strong> Sci. 13,<br />

451–457.<br />

Takano J, Noguchi K, Yasumori M, Kobayashi M, Gajdos Z, Miwa<br />

K, Hayashi H, Yoneyama T, Fujiwara T (2002) Arabidopsis boron<br />

transporter for xylem loading. Nature 420, 337–340.<br />

Takano J, Yamagami M, Noguchi K, Hayashi H, Fujiwara T (2001)<br />

Preferential translocation of boron to young leaves in Arabidopsis<br />

thaliana regulated by the BOR1 gene. Soil Sci. <strong>Plant</strong> Nutr. 47, 345–<br />

357.<br />

Takei K, Takahashi T, Sugiyama T, Yamaya T, Sakakibara H (2002)<br />

Multiple routes communicating nitrogen availability from roots to<br />

shoots: A signal transduction pathway mediated by cytokinin.<br />

J. Exp. Bot. 53, 971–977.<br />

Talke IN, Hanikenne M, Krämer U (2006) Zinc-dependent global<br />

transcriptional control, transcriptional deregulation, <strong>and</strong> higher gene<br />

copy number for genes in metal homeostasis of the hyperaccumulator<br />

Arabidopsis halleri. <strong>Plant</strong> Physiol. 142, 148–167.


384 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Tamaki S, Matsuo S, Wong HL, Yokoi S, Shimamoto K (2007)<br />

Hd3a protein is a mobile flowering signal in rice. Science 316,<br />

1033–1036.<br />

Tan Q, Zhang L, Grant J, Cooper P, Tegeder M (2010) Altered<br />

phloem transport of S-methylmethionine affects plant metabolism,<br />

seed number in pea plants. <strong>Plant</strong> Physiol. 154, 1886–1896.<br />

Tanaka M, Wallace IS, Takano J, Roberts DM, Fujiwara T (2008)<br />

NIP6;1 is a boric acid channel for preferential transport of boron to<br />

growing shoot tissues in Arabidopsis. <strong>Plant</strong> Cell 20, 2860–2875.<br />

Taoka KI, Ham BK, Xoconostle-Cázares B, Rojas MR, Lucas WJ<br />

(2007) Reciprocal phosphorylation <strong>and</strong> glycosylation recognition<br />

motifs control NCAPPl interaction with pumpkin phloem proteins<br />

<strong>and</strong> their cell-to-cell movement. <strong>Plant</strong> Cell 19, 1866–1884.<br />

Teo G, Suzuki Y, Uratsu SL, Lamoinen B, Ormonde N, Hu WK,<br />

DeJong TM, D<strong>and</strong>ekar AM (2006) Silencing leaf sorbitol synthesis<br />

alters long-distance partitioning <strong>and</strong> apple fruit quality. Proc. Natl.<br />

Acad. Sci. USA 103, 18842–18847.<br />

Thibaud MC, Arrighi JF, Bayle V, Chiarenza S, Creff A, Bustos<br />

R, Paz-Ares J, Poirier Y, Nussaume L (2010) Dissection of local<br />

<strong>and</strong> systemic transcriptional responses to phosphate starvation in<br />

Arabidopsis. <strong>Plant</strong> J. 64, 775–789.<br />

Thomas RJ (1972) Bordered pit aspiration in angiosperms. Wood Fiber<br />

3, 236–237.<br />

Thompson AJ, Mulholl<strong>and</strong> BJ, Jackson AC, McKee JM, Hilton HW,<br />

Symonds RC, Sonneveld T, Burbidge A, Stevenson P, Taylor IB<br />

(2007) Regulation <strong>and</strong> manipulation of ABA biosynthesis in roots.<br />

<strong>Plant</strong> Cell Environ. 1, 67–78.<br />

Thompson MV (2006) Phloem: <strong>The</strong> long, the short of it. Trends <strong>Plant</strong><br />

Sci. 11, 26–32.<br />

Thompson MV, Wolniak SM (2008) A plasma membrane-anchored<br />

fluorescent protein fusion illuminates sieve element plasma membranes<br />

in Arabidopsis, tobacco. <strong>Plant</strong> Physiol. 146, 1599–1610.<br />

To JP, Deruère J, Maxwell BB, Morris VF, Hutchison CE, Ferreira<br />

FJ, Schaller GE, Kieber JJ (2007) Cytokinin regulates type-A<br />

Arabidopsis Response Regulator activity <strong>and</strong> protein stability via<br />

two-component phosphorelay. <strong>Plant</strong> Cell 19, 3901–3914.<br />

To JP, Haberer G, Ferreira FJ, Deruère J, Mason MG, Schaller<br />

GE, Alonso JM, Ecker JR, Kieber JJ (2004) Type-A Arabidopsis<br />

response regulators are partially redundant negative regulators of<br />

cytokinin signaling. <strong>Plant</strong> Cell 16, 658–671.<br />

Tomatsu H, Takano J, Takahashi H, Watanabe-Takahashi A, Shibagaki<br />

N, Fujiwara T (2007) An Arabidopsis thaliana high-affinity<br />

molybdate transporter required for efficient uptake of molybdate<br />

from soil. Proc. Natl. Acad. Sci USA 104, 18807–18812.<br />

Trampczynska A, Küpper H, Meyer-Klaucke W, Schmidt H,<br />

Clemens S (2010) Nicotianamine forms complexes with Zn(ii) in<br />

vivo. Metallomics 2, 57–66.<br />

Truernit E, Bauby H, Dubreucq B, Gr<strong>and</strong>jean O, Runions J,<br />

Barthélémy J, Palauqui JC (2008) High-resolution whole-mount<br />

imaging of three-dimensional tissue organization <strong>and</strong> gene expression<br />

enables the study of phloem development <strong>and</strong> structure in<br />

Arabidopsis. <strong>Plant</strong> Cell 20, 1494–1503.<br />

Truernit E, Bauby H, Belcram K, Barthélémy J, Palauqui JC (2012)<br />

OCTOPUS, a polarly localised membrane-associated protein, regulates<br />

phloem differentiation entry in Arabidopsis thaliana. <strong>Development</strong><br />

139, 1306–1315.<br />

Truman W, Bennett MH, Turnbull CGN, Grant MR (2010) Arabidopsis<br />

auxin mutants are compromised in systemic acquired resistance<br />

<strong>and</strong> exhibit aberrant accumulation of various indolic compounds.<br />

<strong>Plant</strong> Physiol. 152, 1562–1573.<br />

Truman W, Bennett MH, Kubigsteltig I, Turnbull C, Grant M (2007)<br />

Arabidopsis systemic immunity uses conserved defense signaling<br />

pathways <strong>and</strong> is mediated by jasmonates. Proc. Natl. Acad. Sci.<br />

USA 104, 1075–1080.<br />

Tsukamoto T, Nakanishi H, Uchida H, Watanabe S, Matsuhashi<br />

S, Mori S, Nishizawa NK (2009) 52Fe translocation in barley as<br />

monitored by a positron-emitting tracer imaging system (PETIS):<br />

Evidence for the direct translocation of Fe from roots to young leaves<br />

via phloem. <strong>Plant</strong> Cell Physiol. 50, 48–57.<br />

Tudela D, Primo-Millo E (1992) 1-Aminocyclopropane-1-carboxylic<br />

acid transported from roots to shoots promotes leaf abscission in<br />

cleopatra m<strong>and</strong>arin (Citrus reshni Hort. ex Tan.) seedlings rehydrated<br />

after water stress. <strong>Plant</strong> Physiol. 100, 131–137.<br />

Tuominen H, Puech L, Fink S, Sundberg B (1997) A radial concentration<br />

gradient of indole-3-acetic acid is related to secondary xylem<br />

development in hybrid aspen. <strong>Plant</strong> Physiol. 115, 577–585.<br />

Turgeon R (2010a) <strong>The</strong> role of phloem loading reconsidered. <strong>Plant</strong><br />

Physiol. 152, 1817–1823.<br />

Turgeon R (2010b) <strong>The</strong> puzzle of phloem pressure. <strong>Plant</strong> Physiol. 154,<br />

578–581.<br />

Turgeon R, Wolfe S (2009) Phloem transport: Cellular pathways,<br />

molecular trafficking. Annu. Rev. <strong>Plant</strong> Biol. 60, 207–221.<br />

Turnbull CGN, Booker JP, Leyser HMO (2002) Micrografting techniques<br />

for testing long-distance signalling in Arabidopsis. <strong>Plant</strong> J.<br />

32, 255–262.<br />

Turner S, Gallois P, Brown D (2007) Tracheary element differentiation.<br />

Annu. Rev. <strong>Plant</strong> Biol. 58, 407–433.<br />

Tuskan GA, Difazio S, Jansson S, Bohlmann J, Grigoriev I, Hellsten<br />

U, Putnam N, Ralph S, Rombauts S, Salamov A, Schein J, Sterck<br />

L, Aerts A, Bhalerao RR, Bhalerao RP, Blaudez D, Boerjan W,<br />

Brun A, Brunner A, Busov V, Campbell M, Carlson J, Chalot<br />

M, Chapman J, Chen GL, Cooper D, Coutinho PM, Couturier<br />

J, Covert S, Cronk Q, Cunningham R, Davis J, Degroeve<br />

S, Déjardin A, Depamphilis C, Detter J, Dirks B, Dubchak<br />

I, Duplessis S, Ehlting J, Ellis B, Gendler K, Goodstein D,<br />

Gribskov M, Grimwood J, Groover A, Gunter L, Hamberger B,<br />

Heinze B, Helariutta Y, Henrissat B, Holligan D, Holt R, Huang<br />

W, Islam-Faridi N, Jones S, Jones-Rhoades M, Jorgensen R,<br />

Joshi C, Kangasjärvi J, Karlsson J, Kelleher C, Kirkpatrick<br />

R, Kirst M, Kohler A, Kalluri U, Larimer F, Leebens-Mack J,<br />

Leplé JC, Locascio P, Lou Y, Lucas S, Martin F, Montanini B,<br />

Napoli C, Nelson DR, Nelson C, Nieminen K, Nilsson O, Pereda<br />

V, Peter G, Philippe R, Pilate G, Poliakov A, Razumovskaya<br />

J, Richardson P, Rinaldi C, Ritl<strong>and</strong> K, Rouzé P, Ryaboy D,


Schmutz J, Schrader J, Segerman B, Shin H, Siddiqui A, Sterky<br />

F, Terry A, Tsai CJ, Uberbacher E, Unneberg P, Vahala J, Wall K,<br />

Wessler S, Yang G, Yin T, Douglas C, Marra M, S<strong>and</strong>berg G, Van<br />

de Peer Y, Rokhsar D (2006) <strong>The</strong> genome of black cottonwood,<br />

Populus trichocarpa (Torr. & Gray). Science 313, 1596–1604.<br />

Twumasi P, Iakimova ET, Qian T, van Ieperen W, Schel JH, Emons<br />

AM, van Kooten O, Woltering EJ (2010) Caspase inhibitors<br />

affect the kinetics <strong>and</strong> dimensions of tracheary elements in xylogenic<br />

Zinnia (Zinnia elegans) cell cultures. BMC <strong>Plant</strong> Biol. 10,<br />

162.<br />

Tyree MT, Sperry JS (1988) Do woody plants operate near the point<br />

of catastrophic xylem dysfunction caused by dynamic water stress?<br />

Answers from a model. <strong>Plant</strong> Physiol. 88, 574–580.<br />

Tyree MT, Sperry JS (1989) Vulnerability of xylem to cavitation <strong>and</strong><br />

embolism. Annu. Rev. <strong>Plant</strong> Physiol. <strong>Plant</strong> Mol. Biol. 40, 19–38.<br />

Tyree MT, Zimmermann MH (2002) Xylem Structure <strong>and</strong> the Ascent<br />

of Sap. Springer, Berlin.<br />

Uggla C, Moritz T, S<strong>and</strong>berg G, Sundberg B (1996) Auxin as a<br />

positional signal in pattern formation in plants. Proc. Natl. Acad.<br />

Sci. USA 93, 9282–9286.<br />

Umehara M, Hanada A, Yoshida S, Akiyama K, Arite T, Takeda-<br />

Kamiya N, Magome H, Kamiya Y, Shirasu K, Yoneyama K,<br />

Kyozuka J, Yamaguchi S (2008) Inhibition of shoot branching by<br />

new terpenoid plant hormones. Nature 455, 195–200.<br />

Umehara M, Hanada A, Magome H, Takeda-Kamiya N, Yamaguchi<br />

S (2010) Contribution of strigolactones to the inhibition of tiller bud<br />

outgrowth under phosphate deficiency in rice. <strong>Plant</strong> Cell Physiol.<br />

51, 1118–1126.<br />

Vacchina V, Mari S, Czernic P, Marquès L, Pianelli K, Schaumlöffel<br />

D, Lebrun M, Łobiński R (2003) Speciation of nickel in a hyperaccumulating<br />

plant by high-performance liquid chromatographyinductively<br />

coupled plasma mass spectrometry <strong>and</strong> electrospray<br />

MS/MS assisted by cloning using yeast complementation. Anal.<br />

Chem. 75, 2740–2745.<br />

van Bel AJE (2003) <strong>The</strong> phloem, a miracle of ingenuity. <strong>Plant</strong> Cell<br />

Environ. 26, 125–149.<br />

van Bel AJE, Hafke JB (2005) Physicochemical determinants of<br />

phloem transport. In: Holbrook NM, Zwieniecki M, eds. <strong>Vascular</strong><br />

Transport in <strong>Plant</strong>s. Elsevier, Amsterdam. pp. 19–44.<br />

van de Mortel JE, Villanueva LA, Schat H, Kwekkeboom J, Coughlan<br />

S, Moerl<strong>and</strong> PD, van <strong>The</strong>maat EVL, Koornneef M, Aarts<br />

MGM (2006) Large expression differences in genes for iron <strong>and</strong> zinc<br />

homeostasis, stress response, <strong>and</strong> lignin biosynthesis distinguish<br />

roots of Arabidopsis thaliana <strong>and</strong> the related metal hyperaccumulator<br />

Thlaspi caerulescens. <strong>Plant</strong> Physiol. 142, 1127–1147.<br />

van der Biezen EA, Freddie CT, Kahn K, Parker JE, Jones<br />

JDG (2002) Arabidopsis RPP4 is a member of the RPP5 multigene<br />

family of TIR-NB-LRR genes <strong>and</strong> confers downy mildew<br />

resistance through multiple signalling components. <strong>Plant</strong> J. 29, 439–<br />

451.<br />

Van Doorn W, Beers E, Dangl J, Franklin-Tong V, Gallois P, Hara-<br />

Nishimura I, Jones A, Kawai-Yamada M, Lam E, Mundy J<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 385<br />

(2011a) Morphological classification of plant cell deaths. Cell Death<br />

Differ. 18, 1241–1246.<br />

Van Doorn WG, Hiemstra T, Fanourakis D (2011b) Hydrogel regulation<br />

of xylem flow: An alternative hypothesis. <strong>Plant</strong> Physiol. 157,<br />

1642–1649.<br />

Van Goor BJ, Wiersma D (1976) Chemical forms of manganese <strong>and</strong><br />

zinc in phloem exudate. Physiol. <strong>Plant</strong>. 36, 213–216.<br />

Van Hoof NALM, Hassinen VH, Hakvoort HWJ, Ballintijn KF, Schat<br />

H, Verkleij JAC, Ernst WHO, Karenlampi SO, Tervahauta AI<br />

(2001) Enhanced copper tolerance in Silene vulgaris (Moench)<br />

Garcke populations from copper mines is associated with increased<br />

transcript levels of a 2b-type metallothionein gene. <strong>Plant</strong> Physiol.<br />

126, 1519–1526.<br />

Van Iepeeren W (2007) Ion-mediated changes of xylem hydraulic<br />

conductivity in planta: Fact or fiction? Trends <strong>Plant</strong> Sci. 12, 137–<br />

142.<br />

Van Iepeeren W, Van Gelder A (2006) Ion-mediated flow changes<br />

suppressed by minimal calcium presence in Chrysanthemum <strong>and</strong><br />

Prunus laurocerasus. J. Exp. Bot. 57, 2743–2750.<br />

Van Norman JM, Frederick RL, Sieburth LE (2004) BYPASS1 negatively<br />

regulates a root-derived signal that controls plant architecture.<br />

Curr. Biol. 14, 1739–1746.<br />

Van Norman JM, Sieburth LE (2007) Dissecting the biosynthetic<br />

pathway for the bypass1 root-derived signal. <strong>Plant</strong> J. 49, 619–628.<br />

Van Wees SCM, De Swart EAM, Van Pelt JA, Van Loon LC,<br />

Pieterse CMJ (2000) Enhancement of induced disease resistance<br />

by simultaneous activation of salicylate- <strong>and</strong> jasmonate-dependent<br />

defense pathways in Arabidopsis thaliana. Proc. Natl. Acad. Sci.<br />

USA 97, 8711–8716.<br />

Vanneste S, Coppens F, Lee E, Donner TJ, Xie Z, Van Isterdael G,<br />

Dhondt S, De Winter F, De Rybel B, Vuylsteke M, De Veylder L,<br />

Friml J, Inze D, Grotewold E, Scarpella E, Sack F, Beemster<br />

GT, Beeckman T (2011) <strong>Development</strong>al regulation of CYCA2s<br />

contributes to tissue-specific proliferation in Arabidopsis. EMBO J.<br />

30, 3430–3441.<br />

Vatén A, Dettmer J, Wu S, Stierhof YD, Miyashima S, Yadav SR,<br />

Roberts CJ, Campilho A, Bulone V, Lichtenberger R, Lehesranta<br />

S, Mähönen AP, Kim JY, Jokitalo E, Sauer N, Scheres B,<br />

Nakajima K, Carlsbecker A, Gallagher KL, Helariutta Y (2011)<br />

Callose biosynthesis regulates symplastic trafficking during root<br />

development. Dev. Cell 21, 1144–1155.<br />

Venugopal SC, Jeong RD, Zhu S, Ch<strong>and</strong>ra-Shekara AC, Navarre D,<br />

Kachroo A, Kachroo P (2009) ENHANCED DISEASE SUSCEP-<br />

TIBILITY 1 <strong>and</strong> salicylic acid act redundantly to regulate resistance<br />

gene expression <strong>and</strong> low oleate-induced defense signaling. PLoS<br />

Genetics 5, e1000545.<br />

Vera-Sirera F, Minguet EG, Singh SK, Ljung K, Tuominen HCL,<br />

Blázquez MA, Carbonell J (2010) Role of polyamines in plant<br />

vascular development. <strong>Plant</strong> Physiol. Biochem. 48, 534–539.<br />

Verbruggen N, Hermans C, Schat H (2009) Molecular mechanisms<br />

of metal hyperaccumulation in plants. New Phytol. 181, 759–<br />

776.


386 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Vercammen D, van de Cotte B, De Jaeger G, Eeckhout D, Casteels<br />

P, V<strong>and</strong>epoele K, V<strong>and</strong>enberghe I, Van Beeumen J, InzéD,Van<br />

Breusegem F (2004) Type II metacaspases Atmc4 <strong>and</strong> Atmc9 of<br />

Arabidopsis thaliana cleave substrates after arginine <strong>and</strong> lysine. J.<br />

Biol. Chem. 279, 45329–45336.<br />

Verma DP, Hong Z (2001) <strong>Plant</strong> callose synthase complexes. <strong>Plant</strong><br />

Mol. Biol. 47, 693–701.<br />

Vernooij B, Friedrich L, Morse A, Reist R, Kolditz-Jawhar R, Ward<br />

E, Uknes S, Kessmann H, Ryals J (1994) Salicylic acid is not<br />

the translocated signal responsible for inducing systemic acquired<br />

resistance but is required in signal transduction. <strong>Plant</strong> Cell 6, 959–<br />

965.<br />

Verret F, Gravot A, Auroy P, Leonhardt N, David P, Nussaume<br />

L, Vavasseur A, Richaud P (2004) Overexpression of AtHMA4<br />

enhances root-to-shoot translocation of zinc <strong>and</strong> cadmium <strong>and</strong> plant<br />

metal tolerance. FEBS Lett. 576, 306–312.<br />

Verret F, Gravot A, Auroy P, Preveral S, Forestier C, Vavasseur A,<br />

Richaud P (2005) Heavy metal transport by AtHMA4 involves the<br />

N-terminal degenerated metal binding domain <strong>and</strong> the C-terminal<br />

His11 stretch. FEBS Lett. 579, 1515–1522.<br />

Vierheilig H, Lerat S, PichéY(2003) <strong>System</strong>ic inhibition of arbuscular<br />

mycorrhiza development by root exudates of cucumber plants<br />

colonized by Glomus mosseae. Mycorrhiza 13, 167–170.<br />

Vlot AC, Klessig DF, Park S-W (2008) <strong>System</strong>ic acquired resistance:<br />

<strong>The</strong> elusive signal(s). Curr. Op. <strong>Plant</strong> Biol. 11, 436–442.<br />

Voesenek LACJ, Jackson MB, Toebes AHW, Huibers W, Vriezen<br />

WH, Colmer TD (2003) De-submergence-induced ethylene production<br />

in Rumex palustris: Regulation <strong>and</strong> ecophysiological significance.<br />

<strong>Plant</strong> J. 33, 341–352.<br />

Voinnet O (2009) Origin, biogenesis, <strong>and</strong> activity of plant microRNAs.<br />

Cell 136, 669–687.<br />

Voitsekhovskaja OV, Koroleva OA, Batashev DR, Knop C, Tomos<br />

AD, Gamalei YV, Heldt HW, Lohaus G (2006) Phloem loading in<br />

two Scrophulariaceae species. What can drive symplastic flow via<br />

plasmodesmata? <strong>Plant</strong> Physiol. 140, 383–395.<br />

von Wirén N, Klair S, Bansal S, Briat J, Khodr H, Shioiri T, Leigh<br />

R, Hider R (1999) Nicotianamine chelates both Fe-III <strong>and</strong> Fe-II.<br />

Implications for metal transport in plants. <strong>Plant</strong> Physiol. 119, 1107–<br />

1114.<br />

Walley JW, Rowe HC, Xiao Y, Chehab EW, Kliebenstein DJ, Wagner<br />

D, Dehesh K (2008) <strong>The</strong> chromatin remodeler SPLAYED regulates<br />

specific stress signaling pathways. PLoS Pathol. 4, e1000237.<br />

Walter A, Silk WK, Schurr (2009) Environmental effects on spatial <strong>and</strong><br />

temporal patterns of leaf <strong>and</strong> root growth. Annu. Rev. <strong>Plant</strong> Biol. 60,<br />

279–304.<br />

Walz C, Juenger M, Schad M, Kehr J (2002) Evidence for the presence<br />

<strong>and</strong> activity of a complete antioxidant defence system in mature<br />

sieve tubes. <strong>Plant</strong> J. 31, 189–197.<br />

Wang D, Amornsiripanitch N, Dong X (2006) A genomic approach to<br />

identify regulatory nodes in the transcriptional network of systemic<br />

acquired resistance in plants. PLoS Pathog. 2, e123.<br />

Wang S, Durrant WE, Song J, Spivey NW, Dong X (2010) Arabidopsis<br />

BRCA2 <strong>and</strong> RAD51 proteins are specifically involved in defense<br />

gene transcription during plant immune responses. Proc. Natl. Acad.<br />

Sci. USA 107, 22716–22721.<br />

Ward ER, Uknes SJ, Williams SC, Dincher SS, Wiederhold DL,<br />

Alex<strong>and</strong>er DC, Ahl-Goy P, Metraux JP, Ryals JA (1991) Coordinate<br />

gene activity in response to agents that induce systemic<br />

acquired resistance. <strong>Plant</strong> Cell 3, 1085–1094.<br />

Wardlaw IF (1974) Temperature control of translocation. In: Bieleski<br />

RL, Fergupon R, Cresswell MM, eds. Mechanisms <strong>and</strong> Regulation<br />

of <strong>Plant</strong> Growth. Bull. Royal Soc., New Zeal<strong>and</strong>. pp. 533–387.<br />

Wardlaw IF (1990) <strong>The</strong> control of carbon partitioning in plants. New<br />

Phytol. 116, 341–381.<br />

Wardlaw IF, Bagnall D (1981) Phloem transport <strong>and</strong> the regulation<br />

of growth of Sorghum bicolor (Moench) at low temperature. <strong>Plant</strong><br />

Physiol. 68, 411–414.<br />

Wardlaw IF, Moncur L (1976) Source, sink <strong>and</strong> control of translocation<br />

in wheat. <strong>Plant</strong>a 128, 93–100.<br />

Warmbrodt RD (1987) Solute concentrations in the phloem <strong>and</strong> apex<br />

of the root of Zea mays. Am. J. Bot. 74, 394–402.<br />

Waters BM, Chu H, DiDonato R, Roberts L, Eisley R, Lahner B, Salt<br />

D, Walker E (2006) Mutations in Arabidopsis yellow stripe-like1 <strong>and</strong><br />

yellow stripe-like3 reveal their roles in metal ion homeostasis <strong>and</strong><br />

loading of metal ions in seeds. <strong>Plant</strong> Physiol. 141, 1446–1458.<br />

Waters BM, Sankaran RP (2011) Moving micronutrients from the soil to<br />

the seeds: Genes <strong>and</strong> physiological processes from a biofortification<br />

perspective. <strong>Plant</strong> Sci. 180, 562–574.<br />

Waters BM, Uauy C, Dubcovsky J, Grusak MA (2009) Wheat<br />

(Triticum aestivum) NAM proteins regulate the translocation of iron,<br />

zinc, <strong>and</strong> nitrogen compounds from vegetative tissues to grain.<br />

J. Exp. Bot. 60, 4263–4274.<br />

Weichert N, Saalbach I, Weichert H, Kohl S, Erban A, Kopka J,<br />

Hause B, Varshney A, Sreenivasulu N, Strickert M, Kumlehn J,<br />

Weschke W, Weber H (2010) Increasing sucrose uptake capacity<br />

of wheat grains stimulates storage protein synthesis. <strong>Plant</strong> Physiol.<br />

152, 698–710.<br />

Weigel D, Jürgens G (2002) Stem cells that make stems. Nature 415,<br />

751–754.<br />

Weir IE, Maddumage R, Allan AC, Ferguson IB (2005) Flow cytometric<br />

analysis of tracheary element differentiation in Zinnia elegans<br />

cells. Cytometry A. 68, 81–91.<br />

Wenzel CL, Schuetz M, Yu Q, Mattsson J (2007) Dynamics of<br />

MONOPTEROS <strong>and</strong> PIN-FORMED1 expression during leaf vein<br />

pattern formation in Arabidopsis thaliana. <strong>Plant</strong> J. 49, 387–398.<br />

Werner T, Schmülling T (2009) Cytokinin action in plant development.<br />

Curr. Opin. <strong>Plant</strong> Biol. 12, 527–538.<br />

West AG, Hultine KR, Sperry JS, Bush SE, Ehleringer JR (2008)<br />

Transpiration <strong>and</strong> hydraulic strategies in a pinyon-juniper woodl<strong>and</strong>.<br />

Ecol. Appl. 18, 911–927.<br />

Wheeler JK, Sperry JS, Hacke UG, Hoang N (2005) Inter-vessel<br />

pitting <strong>and</strong> cavitation in woody Rosaceae <strong>and</strong> other vesselled plants:<br />

A basis for a safety vs. efficiency trade-off in xylem transport. <strong>Plant</strong><br />

Cell Environ. 28, 800–812.


White MC, Decker AM, Chaney RL (1981) Metal complexation in<br />

xylem fluid: I. Chemical composition of tomato <strong>and</strong> soybean stem<br />

exudate. <strong>Plant</strong> Physiol. 67, 292–300.<br />

Wiersma D, Van Goor BJ (1979) Chemical forms of nickel <strong>and</strong> cobalt<br />

in phloem of Riccinus communis. Physiol. <strong>Plant</strong>. 45, 440–451.<br />

Wildermuth MC, Dewdney J, Wu G, Ausubel FM (2001) Isochorismate<br />

synthase is required to synthesize salicylic acid for plant<br />

defense. Nature 414, 562–571.<br />

Williams LE, Mills RF (2005) P1B-ATPases – An ancient family of<br />

transition metal pumps with diverse functions in plants. Trends <strong>Plant</strong><br />

Sci. 10, 491–502.<br />

Williams LE, Pittman JK (2010) Dissecting pathways involved in<br />

manganese homeostasis <strong>and</strong> stress in higher plant cells. <strong>Plant</strong> Cell<br />

Monographs 17, 95–117.<br />

Windt CW, Vergeldt FJ, de Jager PA, van As H (2006) MRI of longdistance<br />

water transport: A comparison of the phloem <strong>and</strong> xylem<br />

flow characteristics <strong>and</strong> dynamics in poplar, castor bean, tomato<br />

<strong>and</strong> tobacco. <strong>Plant</strong> Cell Environ. 29, 1715–1729.<br />

Woffenden BJ, Freeman TB, Beers EP (1998) Proteasome inhibitors<br />

prevent tracheary element differentiation in zinnia mesophyll cell<br />

cultures. <strong>Plant</strong> Physiol. 118, 419–430.<br />

Wong CKE, Cobbett CS (2009) HMA P-type ATPases are the major<br />

mechanism for root-to-shoot Cd translocation in Arabidopsis<br />

thaliana. New Phytol. 181, 71–78.<br />

Wu Y, Zhang D, Chu JY, Boyle P, Wang Y, Brindle ID, De Luca<br />

V, Després C (2012) <strong>The</strong> Arabidopsis NPR1 protein is a receptor<br />

for the plant defense hormone salicylic acid. Cell Rep. 28, 639–<br />

647.<br />

Wycisk K, Kim EJ, Schroeder JI, Krämer U (2004) Enhancing the<br />

first enzymatic step in the histidine biosynthesis pathway increases<br />

the free histidine pool <strong>and</strong> nickel tolerance in Arabidopsis thaliana.<br />

FEBS Lett. 578, 128–134.<br />

Xia Y, Gao QM, Yu K, Lapchyk L, Navarre D, Hildebr<strong>and</strong> D, Kachroo<br />

A, Kachroo P (2009) An intact cuticle in distal tissues is essential<br />

for the induction of systemic acquired resistance in plants. Cell Host<br />

Microbe 5, 151–165.<br />

Xia Y, Yu K, Navarre D, Seebold K, Kachroo A, Kachroo, P (2010)<br />

<strong>The</strong> glabra1 mutation affects cuticle formation <strong>and</strong> plant responses<br />

to microbes. <strong>Plant</strong> Physiol. 154, 833–846.<br />

Xie B, Wang X, Zhu M, Zhang Z, Hong Z (2011) CalS7 encodes a<br />

callose synthase responsible for callose deposition in the phloem.<br />

<strong>Plant</strong> J. 65, 1–14.<br />

Xoconostle-Cázares B, Xiang Y, Ruiz-Medrano R, Wang HL,<br />

Monzer J, Yoo BC, McFarl<strong>and</strong> KC, Franceschi VR, Lucas WJ<br />

(1999) <strong>Plant</strong> paralog to viral movement protein that potentiates<br />

transport of mRNA into the phloem. Science 283, 94–98.<br />

Yaginuma H, Hirakawa Y, Kondo Y, Ohashi-Ito K, Fukuda H (2011)<br />

A novel function of TDIF-related peptides: Promotion of axillary bud<br />

formation. <strong>Plant</strong> Cell Physiol. 52, 1354–1364.<br />

Yamaguchi M, Kubo M, Fukuda H, Demura T (2008) <strong>Vascular</strong>-related<br />

NAC-DOMAIN7 is involved in the differentiation of all types of xylem<br />

vessels in Arabidopsis roots <strong>and</strong> shoots. <strong>Plant</strong> J. 55, 652–664.<br />

Insights into <strong>Plant</strong> <strong>Vascular</strong> Biology 387<br />

Yamaguchi M, Mitsuda N, Ohtani M, Ohme-Takagi M, Demura T<br />

(2011) VASCULAR-RELATED NAC-DOMAIN7 directly regulates<br />

the expression of a broad range of genes for xylem vessel formation.<br />

<strong>Plant</strong> J. 66, 579–590.<br />

Yamaguchi M, Ohtani M, Mitsuda N, Kubo M, Ohme-Takagi M,<br />

Fukuda H, Demura T (2010) VND-INTERACTING2, a NAC domain<br />

transcription factor, negatively regulates xylem vessel formation in<br />

Arabidopsis. <strong>Plant</strong> Cell 22, 1249–1263.<br />

Yamamoto R, Demura T, Fukuda H (1997) Brassinosteroids induce<br />

entry into the final stage of tracheary element differentiation in<br />

cultured zinnia cells. <strong>Plant</strong> Cell Physiol. 38, 980–983.<br />

Yang SJ, Zhang YJ, Sun M, Goldstein G, Cao KF (2012) Recovery<br />

of diurnal depression of leaf hydraulic conductance in a subtropical<br />

woody bamboo species: Embolism refilling by nocturnal root<br />

pressure. Tree Physiol. 32, 414–422.<br />

Yoder JI, Gunathilake P, Wu B, Tomilova N, Tomilov AA (2009)<br />

Engineering host resistance against parasitic weeds with RNA<br />

interference. Pest Manag. Sci. 65, 460–466.<br />

Yokosho K, Yamaji N, Ma J (2010) Isolation <strong>and</strong> characterisation of<br />

two MATE genes in rye. Funct. <strong>Plant</strong> Biol. 37, 296–303.<br />

Yokosho K, Yamaji N, Ueno D, Mitani N, Ma J-F (2009) OSFRDL1<br />

is a citrate transporter required for efficient translocation of iron in<br />

rice. <strong>Plant</strong> Physiol. 149, 297–305.<br />

Yoo BC, Kragler F, Varkonyi-Gasic E, Haywood V, Archer-Evans<br />

S, Lee YM, Lough TJ, Lucas WJ (2004) A systemic small RNA<br />

signaling system in plants. <strong>Plant</strong> Cell 16, 1979–2000.<br />

Yoshida S, Iwamoto K, Demura T, Fukuda H (2009) Comprehensive<br />

analysis of the regulatory roles of auxin in early transdifferentiation<br />

into xylem cells. <strong>Plant</strong> Mol. Biol. 70, 457–469.<br />

Yoshimoto K, Noutoshi Y, Hayashi K, Shirasu K, Takahashi T,<br />

Motose H (2012) A chemical biology approach reveals an opposite<br />

action between thermospermine <strong>and</strong> auxin in xylem development in<br />

Arabidopsis thaliana. <strong>Plant</strong> Cell Physiol. 53, 635–645.<br />

Yu M, Hu CX, Wang YH (2002) Molybdenum efficiency in winter wheat<br />

cultivars as related to molybdenum uptake <strong>and</strong> distribution. <strong>Plant</strong><br />

Soil 245, 287–293.<br />

Zeevaart JAD (1976) Physiology of flower formation. Annu. Rev. <strong>Plant</strong><br />

Physiol. 27, 321–348.<br />

Zeevaart JAD (2006) Florigen coming of age after 70 years. <strong>Plant</strong> Cell<br />

18, 1783–1789.<br />

Zhang LY, Peng YB, Pelleschi-Travier S, Fan Y, Lu YF, Lu YM, Gao<br />

XP, Shen YY, Delrot S, Zhang DP (2004) Evidence for apoplasmic<br />

phloem unloading in developing apple fruit. <strong>Plant</strong> Physiol. 135, 1–<br />

13.<br />

Zhang WH, Zhou Y, Dibley KE, Tyerman SD, Furbank RT, Patrick<br />

JW (2007) Nutrient loading of developing seeds. Funct. <strong>Plant</strong> Biol.<br />

34, 314–331.<br />

Zhang XY, Wang XL, Wang XF, Xia GH, Pan QH, Fan RC, Wu<br />

FQ, Yu XC, Zhang DP (2006) A shift of phloem unloading from<br />

symplasmic to apoplasmic pathway is involved in developmental<br />

onset of ripening in grape berry. <strong>Plant</strong> Physiol. 142, 220–<br />

232.


388 Journal of Integrative <strong>Plant</strong> Biology Vol. 55 No. 4 2013<br />

Zhang Y, Cheng YT, Qu N, Zhao Q, Bi D, Li X (2006) Negative regulation<br />

of defense response in Arabidopsis by two NPR1 paralogs.<br />

<strong>Plant</strong> J. 48, 647–656.<br />

ZhangY,FanW,KinkemaM,LiX,DongX(1999) Interaction of<br />

NPR1 with basic leucine zipper protein transcription factors that<br />

bind sequences required for salicylic acid induction of the PR-1<br />

gene. Proc. Natl. Acad. Sci. USA 96, 6523–6528.<br />

Zhang Y, Tessaro MJ, Lassner M, Li X (2003) Knockout analysis of<br />

Arabidopsis transcription factors TGA2, TGA5, <strong>and</strong> TGA6 reveals<br />

their redundant <strong>and</strong> essential roles in systemic acquired resistance.<br />

<strong>Plant</strong> Cell 15, 2647–2653.<br />

Zhao C, Johnson BJ, Kositsup B, Beers EP (2000) Exploiting<br />

secondary growth in Arabidopsis. Construction of xylem <strong>and</strong> bark<br />

cDNA libraries <strong>and</strong> cloning of three xylem endopeptidases. <strong>Plant</strong><br />

Physiol. 123, 1185–1196.<br />

Zheng Q, Durben DJ, Wolf GH, Angell CA (1991) Liquids at large<br />

negative pressures: Water at the homogeneous nucleation limit.<br />

Science 254, 829–832.<br />

Zhong R, Demura T, Ye ZH (2006) SND1, a NAC domain transcription<br />

factor, is a key regulator of secondary wall synthesis in fibers of<br />

Arabidopsis. <strong>Plant</strong> Cell 18, 3158–3170.<br />

Zhong R, Lee C, Ye ZH (2010) Global analysis of direct targets of<br />

secondary wall NAC master switches in Arabidopsis. Mol. <strong>Plant</strong> 3,<br />

1087–1103.<br />

Zhong R, Lee C, Zhou J, McCarthy RL, Ye ZH (2008) A battery of<br />

transcription factors involved in the regulation of secondary cell wall<br />

biosynthesis in Arabidopsis. <strong>Plant</strong> Cell 20, 2763–2782.<br />

Zhong R, Richardson EA, Ye ZH (2007) <strong>The</strong> MYB46 transcription<br />

factor is a direct target of SND1 <strong>and</strong> regulates secondary wall<br />

biosynthesis in Arabidopsis. <strong>Plant</strong> Cell 19, 2776–2792.<br />

Zhong R, Ye ZH (2007) Regulation of cell wall biosynthesis. Curr. Opin.<br />

<strong>Plant</strong> Biol. 10, 564–572.<br />

Zhong R, Ye ZH (1999) Ifl1, a gene regulating interfascicular fiber<br />

differentiation in Arabidopsis, encodes a homeodomain-leucine<br />

zipper protein. <strong>Plant</strong> Cell 11, 2139–2152.<br />

Zhong R, Ye ZH (2001) Alteration of auxin polar transport in the<br />

Arabidopsis ifl1 mutants. <strong>Plant</strong> Physiol. 126, 549–563.<br />

Zhou J, Lee C, Zhong R, Ye ZH (2009) MYB58 <strong>and</strong> MYB63 are<br />

transcriptional activators of the lignin biosynthetic pathway during<br />

secondary cell wall formation in Arabidopsis. <strong>Plant</strong> Cell 21, 248–<br />

266.<br />

Zhou JM, Trifa Y, Silva H, Pontier D, Lam E, Shah J, Klessig DF<br />

(2000) NPR1 differentially interacts with members of the TGA/OBF<br />

family of transcription factors that bind an element of the PR-1 gene<br />

required for induction by salicylic acid. Mol. <strong>Plant</strong>-Microbe Interact.<br />

13, 191–202.<br />

Zhou N, Tootle TL, Tsui F, Klessig DF, Glazebrook J (1998) PAD4<br />

functions upstream from salicylic acid to control defense responses<br />

in Arabidopsis. <strong>Plant</strong> Cell 10, 1021–1030.<br />

Zhou Y, Setz N, Niemietz C, Qu H, Offler CE, Tyerman SD, Patrick<br />

JW (2007) Aquaporins <strong>and</strong> unloading of phloem-imported water<br />

in coats of developing bean seeds. <strong>Plant</strong> Cell Environ. 30, 1566–<br />

1577.<br />

Zhu S, Jeong RD, Venugopal SC, Lapchyk L, Navarre D, Kachroo<br />

A, Kachroo P (2011) SAG101 forms a ternary complex with EDS1<br />

<strong>and</strong> is required for resistance signaling against turnip crinkle virus.<br />

PLoS Pathog. 7, e1002318.<br />

Zimmermann U, Schneider H, Wegner L, Haase A (2004)<br />

Water ascent in tall trees: Does evolution of l<strong>and</strong> plants<br />

rely on a highly metastable state? New Phytol. 162, 575–<br />

615.<br />

Zwieniecki MA, Holbrook NM (2000) Bordered pit structure <strong>and</strong> vessel<br />

wall surface properties. Implications for embolism repair. <strong>Plant</strong><br />

Physiol. 123, 1015–1020.<br />

Zwieniecki MA, Holbrook NM (2009) Confronting Maxwell’s demon:<br />

Biophysics of xylem embolism repair. Trends <strong>Plant</strong> Sci. 14, 530–<br />

534.<br />

Zwieniecki MA, Melcher PJ, Feild TS, Holbrook NM (2004) A potential<br />

role for xylem-phloem interactions in the hydraulic architecture<br />

of trees: Effects of phloem girdling on xylem hydraulic conductance.<br />

Tree Physiol. 24, 911–917.<br />

Zwieniecki MA, Melcher PJ, Holbrook NM (2001a) Hydraulic properties<br />

of individual xylem vessels of Fraxinus americana. J. Exp. Bot.<br />

52, 257–264.<br />

Zwieniecki MA, Melcher PJ, Holbrook NM (2001b) Hydrogel control<br />

of xylem hydraulic resistance in plants. Science 291, 1059–<br />

1062.<br />

(Co-Editor: Li-Jia Qu)

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!