Next Article in Journal
Plant Growth Promoting and Biocontrol Activity of Streptomyces spp. as Endophytes
Next Article in Special Issue
PIN7 Auxin Carrier Has a Preferential Role in Terminating Radial Root Expansion in Arabidopsis thaliana
Previous Article in Journal
Time-Resolved Measurement of the ATP-Dependent Motion of the Group II Chaperonin by Diffracted Electron Tracking
Previous Article in Special Issue
Indole-3-Acetic Acid Biosynthesis Pathways in the Plant-Beneficial Bacterium Arthrobacter pascens ZZ21
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Auxin in Cell Wall Expansion

Umeå Plant Science Centre (UPSC), Department of Forest Genetics and Plant Physiology, Swedish University of Agricultural Sciences, 901 83 Umeå, Sweden
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2018, 19(4), 951; https://doi.org/10.3390/ijms19040951
Submission received: 15 March 2018 / Revised: 19 March 2018 / Accepted: 19 March 2018 / Published: 22 March 2018
(This article belongs to the Special Issue Auxin)

Abstract

:
Plant cells are surrounded by cell walls, which are dynamic structures displaying a strictly regulated balance between rigidity and flexibility. Walls are fairly rigid to provide support and protection, but also extensible, to allow cell growth, which is triggered by a high intracellular turgor pressure. Wall properties regulate the differential growth of the cell, resulting in a diversity of cell sizes and shapes. The plant hormone auxin is well known to stimulate cell elongation via increasing wall extensibility. Auxin participates in the regulation of cell wall properties by inducing wall loosening. Here, we review what is known on cell wall property regulation by auxin. We focus particularly on the auxin role during cell expansion linked directly to cell wall modifications. We also analyze downstream targets of transcriptional auxin signaling, which are related to the cell wall and could be linked to acid growth and the action of wall-loosening proteins. All together, this update elucidates the connection between hormonal signaling and cell wall synthesis and deposition.

Graphical Abstract

1. Introduction

Plant cells exhibit a great diversity in size and shape. Meristematic cells are usually isodiametric and then differentiate by developing distinct forms to acquire specific functions. This is easily noticeable in cells such as tip-growing root hairs or multi-lobed pavement cells. In contrast with animal cells, plant cells have the particularity of being tightly connected to each other by their surrounding walls located outside of the plasma membrane. Cell walls are dynamic structures that act as an exoskeleton by participating in the establishment and maintenance of cell shape and by protecting the cell content from biological, chemical and biophysical sources of aggression [1,2]. Large plants such as trees are able to resist external forces due to the strength given by their cell walls [1]. Moreover, cell walls play a significant role in processes such as cell adhesion, intercellular communication and water movement [1,3]. Plant cell walls are classified into two groups; primary and secondary walls. The latter are usually present in specialized, non-growing cells and are beyond the scope of this review [1,2,3,4].
Primary cell walls (around 100–1000 nm thick in young growing cells) are essentially made of glucan-based cellulose microfibrils (CMFs) embedded in a highly hydrated matrix composed of pectins, hemicelluloses, structural proteins and proteoglycans [1,2,3,5]. The cell wall has to be fairly rigid, to provide support and protection, but also extensible, to allow cell expansion, which is driven by a strong intracellular turgor pressure [6,7,8,9,10,11]. A strictly regulated balance between wall rigidity and flexibility is therefore imperative to regulate the differential growth that results in such a diversity of cell sizes and shapes [2,7,12,13]. The plant hormone auxin is identified as a stimulator of cell elongation, as it increases cell wall extensibility [14,15,16]. Specifically, auxin regulates cell wall properties by initiating wall loosening [17,18]. The close link between hormonal action and cell wall synthesis and deposition has been investigated for many years, but much of the details still need to be clarified. Here we summarize what is currently known about the regulation of cell wall properties and the role of auxin in this process.

2. Plant Cell Walls

Plant cell walls are highly heterogeneous and cell wall composites vary among different species and cell types. Walls are very dynamic and their composition changes even within the same cell over time [1,19,20,21,22,23,24]. Nonetheless, the key polysaccharides are usually present and their structure, biosynthesis and interaction are summarized in this chapter.

2.1. Cellulose Microfibrils (CMFs)

Cell wall consists of different polymers including CMFs, which are embedded in components such as non-cellulosic polysaccharides and structural proteins. CMFs are the largest cell wall polysaccharides, composed of (1,4)-β-d-glucan parallel arrays assembled into long cylinders [25,26]. Due to their stiff and load-bearing properties, CMFs are resistant to tensional forces [1,2,3]. CMFs determine the direction of cell expansion. Indeed, their deposition and alignment define cell growth anisotropy [2,27,28], as shown by the characterization of cellulose-deficient Arabidopsis mutants, in which cell elongation is drastically reduced [29]. Cellulose synthesis takes place beneath the cell wall at the plasma membrane via large rosette complexes made of CELLULOSE SYNTHASEs (CESAs), and certainly other components such as KORRIGAN1 (KOR1), the function of which remains elusive [25,26,30,31]. The CMF patterning of the wall is mediated via cortical microtubules (cMT) and CESAs at the plasma membrane, with the orientation of CMFs within the wall following the pattern given by the cMTs [28,32,33,34,35,36,37].

2.2. Hemicelluloses and Pectins

CMFs are embedded in a matrix of hemicelluloses and pectins composed of various carbohydrates that display complex glyosidic linkages. In dicotyledons such as Arabidopsis, pectins and hemicellulose xyloglucans (XyGs) are the most abundant cell wall components [1]. XyGs are found mainly in primary cell walls and are thought to participate in cell wall extension during cell elongation [38,39,40,41]. XyGs are composed of (1,4)-β-d-glucan chains, with side chains consisting of galactose, fucose and xylose residues [42]. XyGs influence wall extensibility and stiffness, as cell walls in an Arabidopsis double xyloglucan xylosyltransferases mutant (xxt1 xxt2) are softer and weaker than walls in the wild type [7,43]. Mannans and heteromannans are hemicelluloses that are abundant in mosses, lycophytes and in the secondary cell walls of gymnosperms [42,44]. Other hemicelluloses such as xylans, heteroxylans and (1,3;1,4)-β-d-glucans are highly represented in monocotyledons (cereals and grasses) and in secondary cell walls [1,5].
Pectins play an important role in the regulation of wall properties, because they control wall porosity and hydration, which causes wall swelling and influences wall thickness. Moreover, pectins adjust wall extensibility by influencing the alignment of CMFs and form the middle lamella, an adhesive compartment between two adjacent cell walls [45,46,47,48]. Pectins are composed of highly heterogeneous polysaccharides, among which four main elements can be distinguished: homogalacturonan (HG), rhamnogalacturonan I (RGI), rhamnogalacturonan II (RGII) and xylogalacturonan (XGA) [45,49,50,51,52]. HG often contains highly methylesterified galacturonic acid residues, while RGI is more complex and is composed of alternating galacturonic acid and rhamnose with galactose, arabinose or arabinogalactans forming side chains [45,47,49,53]. A common feature of RGII is the presence of borate esters between RGII-specific sugar residues [3,45,49,54].
Non-cellulosic cell wall components are synthesized in the Golgi apparatus, packed into vesicles and trafficked along actin filaments (AFs) to the wall [33,55,56,57,58]. The walls of actively growing cells display a porous structure, which allows polysaccharide to move relatively to each other (such as sliding) within the wall [7,59]. Polysaccharide synthesis is carried out by synthases, which catalyze the polymerization of sugar residues, and glycosyltransferases, which connect the monosaccharides and short oligosaccharides to the polymer chains [1,60].

2.3. Structural Proteins

Beside polysaccharides, the cell wall contains various structural (non-enzymatic) proteins, which regulate its formation and growth [2,61]. Among these structural proteins, EXPANSINs (EXPs), EXTENSINs (EXTs), and ARABINOGALACTAN PROTEINs (AGPs) are well-characterized as regulating wall expansion [62,63]. EXPs are defined as wall-loosening proteins, enhancing wall expansion in acidic pH [64], and will later be discussed in the context of auxin-mediated cell wall expansion. Other structural glycoproteins, EXTs, are required for cell wall assembly [65,66,67,68], as the EXT3-defective Arabidopsis mutant root-, shoot-, hypocotyl-defective (rsh) presents defective wall formation [2,69]. AGPs play a role in plant protection against pathogens, and additionally, an increased amount of AGPs can be observed in wounded plants [61,70]. AGPs are known to specifically control pollen tube growth [61], but are also thought to regulate overall plant development [71].

2.4. Interactions within the Cell Wall

Cell wall physical properties are maintained through the interactions among its polysaccharides [3,72]. A new model displaying the interactions between different cell wall polymers has been recently presented, in which “biochemical hotspots” crosslink different polysaccharides [7,73]. These hotspots are present between CMFs and XyGs, but also among different CMFs, connecting them to each other [7,73,74,75]. This interesting model updates the previous theory, which was based on the wall being composed of separated CMFs, which could be cross-linked to either XyGs, in order to reinforce the wall, and/or pectins, in order to soften the wall [5,76].
Crosslinking of CMFs with XyGs increases wall mechanical resistance [77,78,79,80,81,82]. XyGs are important for the separation of CMFs, as the XyG-deficient xxt1 xxt2 mutant is characterized by tightly compact CMFs [7,43]. XyG-CMF interactions are modulated by XYLOGLUCAN ENDOTRANSGLUCOSYLASE/HYDROLASEs (XTHs), which either catalyze the linkage of the XyGs to cellulose (strengthening the wall) or hydrolyze the breaking of the link of XyGs with CMFs (loosening the wall) [83,84,85,86,87,88,89,90]. During cell development, pectins are regularly delivered and inserted into the wall matrix, which suggests that their presence and abundance might regulate wall extensibility. Pectins can either enhance wall expansion by promoting movement of the CMFs or maintain CMFs in non-growing cell wall zones [91,92,93,94,95,96]. Moreover, different pectin domains crosslink to each other via calcium and boron bonds [1,47,49]. These connections are modified by PECTIN METHYLESTERASEs (PMEs), which regulate the crosslinking of pectins to calcium ions. Methyl-esterification (addition of methyl groups) decreases the ability of HGs to form crosslinks with calcium ions, causing softening of the wall. Accordingly, de-methyl-esterification (removal of the methyl groups) increases HG capacity to crosslink to calcium ions, which causes wall stiffening, compaction and enhanced adhesion [97,98]. Intriguingly, auxin has been shown to reduce the stiffness of the cell wall through demethylesterification of pectins in the shoot apex leading to organ outgrowth [99]. On the other hand, RGII chains are connected to each other through borate diester bonds, influencing wall hydration and thickness [47]. Arabinans and arabinogalactans are known to induce cell wall swelling, decreasing its stiffness while increasing its extensibility [100,101]. In summary, the cell wall is composed of a range of different polysaccharides, whose abundance and interactions determine its properties and regulate cell growth.

3. The Role of Auxin in Wall Extension

Water accumulation in the vacuole induces high turgor pressure, which drives plant cell growth. This strong tensile stress presses against the plasma membrane, leading to the stretching of the cell wall polysaccharides. The wall needs to be moderately rigid to oppose this turgor pressure, to avoid breaking. However, the wall also has to adapt its composition by modifying and constantly adding polysaccharides to allow cell extension [7,59,102,103].
Cell wall expansion and overall cell growth is regulated via several factors, including plant hormones. Among them, auxin plays a vital role in controlling plant growth and development via promotion of cell division (proliferation), growth (expansion, elongation) and differentiation [15,16,104,105,106,107,108]. Enlargement of the cell occurs prior to cell division, however, no changes are observed in the vacuole size at this stage. On the other hand, cell expansion includes vacuole extension and is defined as a turgor-driven increase in cell size, which is controlled by the cell wall capacity to extend. Cell expansion is related to an increased ploidy level (endoreduplication), cellular vacuolization and differentiation [106,109]. Almost four decades ago, auxin or indole-3-acetic acid (IAA) was implicated for the first time in cell wall loosening and cell expansion via modifications of cell wall composition. IAA causes pectin polymerization, and increases pectin viscosity and XyG depolymerization [110].
In this second part, we discuss the auxin role during cell expansion and its direct link to the changes occurring in the cell wall [111]. Auxin activates the expression of cell wall-related genes and stimulates the synthesis of proton pumps, which leads to apoplast acidification [106]. Auxin also activates plasma membrane (PM) H+-ATPases through upregulating the phosphorylation of the penultimate of threonine of PM H+-ATPases, leading to apoplast acidification [112]. In an acidic environment, wall-loosening proteins are active and cause wall enlargement. The changes in the wall trigger the cell to activate calcium channels, which pump calcium into the wall and increase the pH, causing growth cessation. Finally, auxin acts on the cytoskeleton (AFs and cMTs) through RHO OF PLANTS (ROP) GUANOSINE-5′-TRIPHOSPHATASES (GTPases) and promotes trafficking of vesicles containing new cell wall material [113,114,115,116].

3.1. Auxin Signaling Stimulates Cell Elongation

Arabidopsis seedling hypocotyls elongate exclusively by cell expansion, making this organ a model system in which to investigate the contribution of auxin signaling to cell elongation [111,117]. Auxin acts through the TRANSPORT INHIBITOR RESISTANT 1/AUXIN SIGNALING F-BOX (TIR1/AFB) nuclear auxin receptor family, the degradation of the transcriptional regulators AUXIN/INDOLE-3-ACETIC ACID (AUX/IAAs) and the AUXIN RESPONSE FACTORs (ARFs), which mediate different transcriptional responses [117,118]. TIR1/AFBs are part of the Skp1/Cullin/F-box (SCF) complex, which promotes degradation of AUX/IAAs, which otherwise repress auxin-mediated transcription [119] through the interactions with ARFs in the absence of auxin. Once the concentration of auxin increases, the hormone mediates the linkage of TIR1/AFBs with AUX/IAAs and the degradation of the latter through proteasomal activity [120,121,122,123]. Different Arabidopsis AUX/IAA mutants such as auxin resistant/indole-3-acetic acid inducible (axr2/iaa7, axr5/iaa1, axr3/iaa17), or short hypocotyl/indole-3-acetic acid inducible (shy2/iaa3) display cell expansion defects [106,124,125], indicating that auxin induces cell expansion through the degradation of AUX/IAAs. ARFs are transcription factors that bind to the promoters of auxin-responsive genes [122,126,127,128]. Among the 22 ARFs in Arabidopsis, ARF7 has been shown to positively regulate the expression of EXP8 [129], thus playing an essential role in extensive cell growth [130].

3.2. Auxin and Cell Wall-Related Genes

Several studies have shown that auxin treatment can specifically change the expression of different genes. In one such study, seedlings treated with exogenous IAA displayed over 790 differentially regulated genes, 55% of which were upregulated [131]. Of these, we have only selected the upregulated genes that were specifically related to cell walls and classified them into groups related to different cell wall components such as cellulose, XyGs, pectins, and structural proteins (EXPs), peroxidases, and components related to secondary cell walls (Table 1). Clearly, auxin treatment results in the upregulation of key genes related to cell wall components, as can be seen in the table. However, it is important to note that the reported genes are not necessarily specifically related to cell elongation (wall expansion) and could be linked to different auxin-driven processes such as cell division, growth or differentiation.
Several auxin-responsive genes have been shown to be upregulated in elongating dark-grown hypocotyls [130]. Interestingly, cell wall-related genes were also found to be upregulated in this elongating organ, among them genes encoding wall-loosening EXPs [64,132], XTHs [86,133], AGPs [134,135] and related to pectin modification [130,136]. The use of etiolated hypocotyls suggested these genes as being specifically related to cell elongation.
The synthetic auxin picloram (4-amino-3,5,6-trichloropicolinic acid) induces hypocotyl elongation [137]. A transcriptional analysis of differentially regulated genes was performed in elongating light-grown hypocotyls upon treatment with the picloram [117], revealing that picloram and IAA signaling act through common downstream transcriptional targets, which are thought to stimulate cell elongation. However, picloram treatment revealed 79% novel differentially regulated genes, which were not differentially regulated in the seedlings treated with IAA, suggesting that they might be specific for elongating cells. Upon picloram treatment, changes in the expression of 1193 auxin-responsive genes (of which 62% were upregulated) preceded the hypocotyl elongation. Moreover, these genes were identified as downstream targets of picloram-stimulated transcriptional auxin signaling [117]. Study of the gene ontology related to auxin-responsive genes showed over-representation of genes related to hormone signaling, cell wall and cell expansion [117,138,139]. We have decided to focus on the genes upregulated by picloram treatment in the hypocotyl and selected those that were specifically related to cell walls (Table 2). Similarly, to IAA treatment of whole seedlings, picloram treatment of hypocotyls induced genes related to cell wall elements such as cellulose, pectins, EXPs, XTHs, and PEROXIDASEs. However, members of these different classes were more widely represented in the hypocotyls, suggesting their potential role in cell elongation and wall extension. Analysis also revealed many upregulated genes related to pectin metabolism as well as novel players related to hemicelluloses, AGPs and other structural proteins. In summary, among differentially regulated genes upon auxin treatment, many are related to post-synthetic cell wall modifications. This indicates that auxin regulates cell expansion via stimulating changes in cell wall properties. However, the auxin concentrations used in these studies is not physiologically relevant and the interpretation of the results should be cautious.

3.3. Auxin Induces Acid Growth

Auxin is known to induce acid growth (Figure 1), which is defined as the loosening of the walls at low pH, leading to an increase in wall extensibility and rapid cell elongation [14,16,140,141,142,143,144,145], through the TIR/AFB signaling machinery [146]. Auxin stimulates the activity of plasma membrane H+-ATPase proton pumps [147,148] (Figure 1(Aa)), which pump out protons (H+) to the wall matrix, leading to apoplast acidification (pH 4.5–6) [15,138,145,149]. This process induces the hyperpolarization of the plasma membrane and is regulated by the auxin-inducible SMALL AUXIN UP-RNA (SAUR) proteins [148]. Activation of potassium channels occurs and potassium ions are pumped into the cytosol (Figure 1(Ab)). The increasing concentration of potassium in the cytosol stimulates water uptake, which generates tensile stress, forcing the cell wall to extend [106,150,151]. Auxin not only stimulates the activity of proton pumps and potassium channels [150,151,152], but also induces the expression of genes encoding these proteins [150,151,152,153,154]. Note that auxin-sensitive proton pumps are mostly located in the epidermis [14,155], which is thought to be limiting for growth and is essential for shaping plant organs [154,155,156,157]. Moreover, different cells display distinct abilities to perceive acid growth; for instance, mature cells are less sensitive to acidic pH and extend less than young cells [158,159].

3.4. EXPANSINs Mediate Acid Growth

Auxin-induced acidic pH is required to activate EXPs (Figure 1(Ac)), which are specific, non-enzymatic wall-loosening proteins. EXPs were identified as inducing the relaxation of the walls in actively expanding hypocotyl cells of Cucumis sativa [64,160,161,162]. EXPs disintegrate polysaccharide networks by cutting and loosening connections between CMFs and non-cellulosic polysaccharides such as XyGs [161,163,164]. As a result, CMFs slide and move apart, promoting wall loosening, hydration and swelling. Interestingly, in plants exposed to gravitropic and light stimuli, EXP-encoding genes (EXP1 and EXP8) are upregulated in elongating cells. This was observed before plant morphological changes appear, suggesting that auxin stimulates EXP expression, leading to the wall property changes [106,129].

3.5. Cellulose and Xyloglucan Modification during Wall Expansion

Auxin acidification induces cell wall modifications mediated by XTH and ENDO-(1,4)-β-d-GLUCANASEs (CELLULASEs), which loosen the connections between different cell wall polysaccharides within the wall matrix (Figure 1(Ac)) [3,165,166,167]. Auxin upregulates the expression of XTH family members (such as XYLOGLUCAN ENDOTRANSGLUCOSYLASE; XET) and CELLULASEs [106,117,168,169,170,171,172,173,174,175,176]. XTH proteins have been found in actively growing cells such as meristematic cells in the shoot apical meristem, leaf primordia and elongating roots, which are known to accumulate auxin [177]. In these cells, XTHs control cell expansion in two distinct ways. First, XTHs mediate the incorporation of nascent XyG chains into already existing XyGs, which suppress cell elongation [39]. Second, auxin-stimulated XTHs induce the modification of the polysaccharide network by cutting XyG backbones and by re-forming glycosidic linkages between different XyG chains within the already existing wall network. XTH-mediated cutting of XyGs provides short XyG fragments, which lead to loosening of the wall and promotion of wall rearrangement for cell elongation [2,39,178]. These short XyG fragments were also shown to interfere with auxin signaling, suggesting a negative feedback loop [179]. The degree of XyG fucosylation seems to also be important for regulation of cell wall expansion. Upon auxin treatment, non-elongating cells display enhanced abundance of fucosylated XyGs [180]. In the absence of exogenous auxin, cells containing fucosylated xyloglucan were shown to be elongating [39,179,181]. Interestingly, the auxin efflux-deficient Arabidopsis mutant pin-formed1 (pin1) displays a progressive decrease of XTH9 gene expression along its inflorescence stem (from the apex to the base) [182], which indicates that auxin might regulate the expression of the XTH9 gene. However, auxin seems to have no influence on XTH action during root hair formation [183].
CELLULASEs hydrolyze glyosidic bonds in CMFs and are involved in cellulose formation/adjustment [173]. Auxin induces CELLULASE activity, leading to cleavage of load-bearing hemicellulose chains, which tether neighboring CMFs, and cleavage of cellulose chains. CELLULASEs modify the interactions between CMFs and XyGs, depolymerize XyG chains, producing short oligosaccharides [3,173], and promote wall loosening and extensibility [175]. In elongating stems of pea, auxin treatment induces the activity of CELLULASEs, which hydrolase the cellulose-XyG network, resulting in the release of wall-bound XyGs and their degradation [184].

3.6. Pectin Methylesterification and Its Consequences in Wall Loosening

Auxin induces low pH, which activates PECTIN METHYLESTERASE (PME) (Figure 1(Ac)) and inhibits PME INHIBITOR (PMEI). PMEs conduct random demethylesterification of initially homogenous HGs. Next, heterogenous HGs are deacetylated via PECTIN ACETYLESTERASEs (PAEs), neutralizing the galacturonyl residues, blocking their interactions with calcium ions. Finally, PECTATE LYASEs (PLs) [185] and POLYGALACTURONASEs (PGs) [186,187] depolymerize pectic chains, leading to loosened walls. As pectin depolymerizing enzymes, PLs and PGs provide short HG products called oligogalacturonides (OGAs) [98], which are small wall signaling molecules that act as potential ligands binding to WALL ASSOCIATED KINASE (WAK) membrane receptors [2,188,189,190]. OGA treatment inhibits pea stem elongation induced by auxin [191] and moreover, exogenously applied OGAs reduce auxin response and adventitious root formation promoted by auxin in Arabidopsis and tobacco, which indicates that OGAs might regulate auxin responses [190,192]. Accordingly, tobacco plants expressing fungal PG (depolymerizing HG and providing OGAs) display reduced sensitivity to auxin [193]. OGAs are also implicated in hydrogen peroxide (H2O2) production [194].
Hydrogen peroxide is a type of reactive oxygen species (ROS), which also includes superoxide anions (reactive oxygen anions) (O2) and hydroxyl radicals (neutral form of the hydroxide ion OH) (•OH), all of which are produced during plant metabolism, development and defense against pathogens. Although ROS cause cell damage and their levels must be strictly controlled by antioxidation, they also play a number of important roles such as in cell signaling and cell wall structure [195]. Auxin-induced PMEs activate the plasma membrane nicotinamide adenine dinucleotide phosphate (NADPH) OXIDASEs (Figure 1(Ad)), [2], which mediate transport of superoxide anions to the cell wall, where they are converted to hydrogen peroxide. PEROXIDASEs are enzymes abundant in the cell walls (Figure 1(Ac)), which use hydrogen peroxide and/or superoxide anions as substrates to catalyze a reaction producing hydroxyl radicals. These different ROS cause polymer breakdown, which leads to wall loosening during auxin-mediated cell extension [196,197,198]. Auxin has been proposed to stimulate the release of superoxide anions and hydroxyl radicals, leading to cell elongation [199]. Moreover, inducing the production of hydroxyl radicals causes an increase in wall extensibility, which indicates their role in inducing cell growth. On the other hand, the induction of superoxide anions causes the inhibition of auxin-induced growth [199].
The activity of wall-loosening proteins (EXPs) and enzymes such as XTHs, CELLULASEs and PMEs results in the sliding and moving apart of the CMFs (Figure 1(Be)) [2,64,160,200]. Loosening within the wall promotes its hydration and swelling. Next, wall porosity increases, creating a physical space for newly synthesized polysaccharides and proteins, which arrive via vesicule trafficking (Figure 1(Bh)). Nascent wall composites are secreted to the wall and integrated within the existing polysaccharide network thanks to modification of the polysaccharide interactions, through enzymatic hydrolysis, ligations and crosslinking. New polysaccharides must be added to compensate for wall stretching and thinning, in order to avoid the breaking of the wall. The cell wall surface area increases and the wall is irreversibly extended, which leads to wall relaxation and growth deceleration [59,102,103,201,202]. After the cell wall extends, information from the wall is transmitted back to the cytosol. Wall extension and relaxation stretch the plasma membrane and trigger calcium channel opening, leading to a calcium influx towards the cytosol (Figure 1(Bf)). The accumulation of cytosolic calcium inhibits the H+-ATPase proton pumps (Figure 1(Bg)) and stimulates a H+ influx towards the cell, causing apoplast alkalization [2,203,204,205].

3.7. Crosslinking of the Wall Polysaccharides

In an alkaline wall environment, PMEs conduct demethylesterification of HGs (Figure 1(Bi)). Next, PAEs deacetylate HGs making them accessible for calcium crosslinking, leading to pectin compaction [98]. PMEs also modify the methyl groups in HGs, which induces the cross-linking of polysaccharides and proteins (EXTs). This interaction causes wall dehydration and compaction, decreasing extensibility and growth [206,207,208,209,210,211,212,213,214]. Cell wall hydration is also regulated by enzymes such as the β-galactosidases (for example MUCILAGE-MODIFIED2 (MUM2) or SALT-OVERLY SENSITIVE5 (SOS5)), which are necessary for proper seed mucilage hydration. Mucilage in the mum2 Arabidopsis mutant contains an increased level of galactoses, which results in hydration defects [2,215,216,217]). ROS are also proposed to cross-link the wall polysaccharides or remove hydrogen atoms from polysaccharides, modifying the cell wall properties (Figure 1(Bj)). Together with PMEs, ROS promote wall dehydration and strengthening, which slows down growth (Figure 1(Ck)) [3,204,218,219,220,221]. However, Cosgrove (2005) [3] discusses the evidence that ROS play only a minor role in cell wall expansion, being responsible for only 1% of the extension. Growing cells produce very low amounts of ROS due to the fact that higher ROS concentrations can cause damage to the cells.

4. Conclusions

As the most external cell compartment, the cell wall is by necessity involved in plant cell growth. This has been demonstrated by analyzing different cell wall deficient mutants that display various growth defects. Indeed, the cell wall is a very dynamic cell composite, which is characterized by complex polysaccharide interactions and various modifications during cell development. Moreover, plant cells grow symplastically and they need to adjust their growth to the neighboring cells. Changes within the wall occur during turgor-driven cell growth, which is mediated via acidification of the wall and loosening of the connections between different cell wall polysaccharides. Acidic growth has been shown to be promoted by auxin, which activates structural proteins and enzymes such as EXPs, XTHs and PMEs, modifying the interactions between different cell wall polymers. Furthermore, progress in molecular biology has allowed us to connect auxin with the activation of the acidifying proteins (proton pumps, [221]) and numerous genes that are related to wall biosynthesis and modification. In summary, auxin plays a major role in regulating cell expansion through the activation of cell wall synthesis and modification-related genes. However, it still remains elusive as to how auxin regulates the modifications in the wall over time. Further development of in muro detection methods, which follow cell wall changes over cell development, will undoubtedly provide more clues about the temporal regulation of cell wall expansion and cell elongation by this master hormone.

Acknowledgements

We gratefully acknowledge Siamsa M. Doyle for critical reading of the manuscript. This work was supported by Vetenskapsrådet and Vinnova (Verket för Innovationssystem), (Mateusz Majda; Stéphanie Robert), Knut och Alice Wallenbergs Stiftelse via “Shapesystem” grant number 2012.0050 (Stéphanie Robert), Swedish Research Council (grant number VR2013-4632; Mateusz Majda).

Author Contributions

Mateusz Majda and Stéphanie Robert wrote the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Burton, R.A.; Gidley, M.J.; Fincher, G.B. Heterogeneity in the chemistry, structure and function of plant cell walls. Nat. Chem. Biol. 2010, 6, 724–732. [Google Scholar] [CrossRef] [PubMed]
  2. Wolf, S.; Hématy, K.; Höfte, H. Growth Control and Cell Wall Signaling in Plants. Annu. Rev. Plant Biol. 2012, 63, 381–407. [Google Scholar] [CrossRef] [PubMed]
  3. Cosgrove, D.J. Growth of the plant cell wall. Nat. Rev. Mol. Cell Biol. 2005, 6, 850–861. [Google Scholar] [CrossRef] [PubMed]
  4. Fry, S.C.; Love, J.; Bryant, J.A. Plant Cell Wall Polymers. In Biofuels and Bioenergy; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2017; pp. 59–87. ISBN 978-1-11-835055-3. [Google Scholar]
  5. Carpita, N.C.; Gibeaut, D.M. Structural models of primary cell walls in flowering plants: Consistency of molecular structure with the physical properties of the walls during growth. Plant J. 1993, 3, 1–30. [Google Scholar] [CrossRef] [PubMed]
  6. Cosgrove, D.J. Wall Extensiblity: Its Nature, Measurement and Relationship to Plant Cell Growth. New Phytol. 1993, 124, 1–23. [Google Scholar] [CrossRef] [PubMed]
  7. Cosgrove, D.J. Diffuse growth of plant cell walls. Plant Physiol. 2018, 176, 16–27. [Google Scholar] [CrossRef] [PubMed]
  8. Cosgrove, D.J. Plant cell wall extensibility: Connecting plant cell growth with cell wall structure, mechanics, and the action of wall-modifying enzymes. J. Exp. Bot. 2016, 67, 463–476. [Google Scholar] [CrossRef] [PubMed]
  9. Guerriero, G.; Hausman, J.F.; Cai, G. No stress! Relax! Mechanisms governing growth and shape in plant cells. Int. J. Mol. Sci. 2014, 15, 5094–5114. [Google Scholar] [CrossRef] [PubMed]
  10. Wei, C.; Lintilhac, P.M. Loss of stability—A new model for stress relaxation in plant cell walls. J. Theor. Biol. 2003, 224, 305–312. [Google Scholar] [CrossRef]
  11. Wei, C.; Lintilhac, P.M. Loss of Stability: A New Look at the Physics of Cell Wall Behavior during Plant Cell Growth. Plant Physiol. 2007, 145, 763–772. [Google Scholar] [CrossRef] [PubMed]
  12. Kroeger, J.H.; Zerzour, R.; Geitmann, A. Regulator or driving force? The role of turgor pressure in oscillatory plant cell growth. PLoS ONE 2011, 6, e18549. [Google Scholar] [CrossRef] [PubMed]
  13. Green, P.B. Mechanism for Plant Cellular Morphogenesis. Science 1962, 138, 1404–1405. [Google Scholar] [CrossRef] [PubMed]
  14. Kutschera, U. The current status of the acid-growth hypothesis. New Phytol. 1994, 126, 549–569. [Google Scholar] [CrossRef]
  15. Rayle, D.L.; Cleland, R.E. Evidence that Auxin-induced Growth of Soybean Hypocotyls Involves Proton Excretion. Plant Physiol. 1980, 66, 433–437. [Google Scholar] [CrossRef] [PubMed]
  16. Rayle, D.L.; Cleland, R.E. The Acid Growth Theory of auxin-induced cell elongation is alive and well. Plant Physiol. 1992, 99, 1271–1274. [Google Scholar] [CrossRef] [PubMed]
  17. Ray, P.M. Cell wall synthesis and cell elongation in oat coleoptile tissue. Am. J. Bot. 1962, 49, 928–939. [Google Scholar] [CrossRef]
  18. Ray, P.M.; Ruesink, A.W. Kinetics experiments on the nature of the growth mechanism in oat coleoptile cellls. Dev. Biol. 1962, 4, 377–397. [Google Scholar] [CrossRef]
  19. Derbyshire, P.; Findlay, K.; McCann, M.C.; Roberts, K. Cell elongation in Arabidopsis hypocotyls involves dynamic changes in cell wall thickness. J. Exp. Bot. 2007, 58, 2079–2089. [Google Scholar] [CrossRef] [PubMed]
  20. Freshour, G.; Clay, R.P.; Fuller, M.S.; Albersheim, P.; Darvill, A.G.; Hahn, M.G. Developmental and Tissue-Specific Structural Alterations of the Cell-Wall Polysaccharides of Arabidopsis thaliana Roots. Plant Physiol. 1996, 110, 1413–1429. [Google Scholar] [CrossRef] [PubMed]
  21. Phyo, P.; Wang, T.; Kiemle, S.N.; O’Neill, H.; Pingali, S.V.; Hong, M.; Cosgrove, D.J. Gradients in wall mechanics and polysaccharides along growing inflorescence stems. Plant Physiol. 2017. [Google Scholar] [CrossRef] [PubMed]
  22. Majda, M.; Grones, P.; Sintorn, I.M.; Vain, T.; Milani, P.; Krupinski, P.; Zagórska-Marek, B.; Viotti, C.; Jönsson, H.; Mellerowicz, E.J.; et al. Mechanochemical Polarization of Contiguous Cell Walls Shapes Plant Pavement Cells. Dev. Cell 2017, 43, 290–304. [Google Scholar] [CrossRef] [PubMed]
  23. Refrégier, G.; Pelletier, S.; Jaillard, D.; Höfte, H. Interaction between wall deposition and cell elongation in dark-grown hypocotyl cells in Arabidopsis. Plant Physiol. 2004, 135, 959–968. [Google Scholar] [CrossRef] [PubMed]
  24. Majda, M. Role of the Cell Wall in Cell Shape Acquisition; Swedish University of Agricultural Sciences: Umeå, Sweden, 2018; pp. 1652–6880. ISBN 978-9-17-760160-9. [Google Scholar]
  25. Doblin, M. Cellulose Biosynthesis in Plants: From Genes to Rosettes. Plant Cell Physiol. 2002, 43, 1407–1420. [Google Scholar] [CrossRef] [PubMed]
  26. Somerville, C. Cellulose Synthesis in Higher Plants. Annu. Rev. Cell Dev. Biol. 2006, 22, 53–78. [Google Scholar] [CrossRef] [PubMed]
  27. Baskin, T.I. Anisotropic Expansion of the Plant Cell Wall. Annu. Rev. Cell Dev. Biol. 2005, 21, 203–222. [Google Scholar] [CrossRef] [PubMed]
  28. Chen, S.; Ehrhardt, D.W.; Somerville, C.R. Mutations of cellulose synthase (CESA1) phosphorylation sites modulate anisotropic cell expansion and bidirectional mobility of cellulose synthase. Proc. Natl. Acad. Sci. USA 2010, 107, 17188–17193. [Google Scholar] [CrossRef] [PubMed]
  29. Fagard, M.; Desnos, T.; Desprez, T.; Goubet, F.; Refregier, G.; Mouille, G.; McCann, M.; Rayon, C.; Vernhettes, S.; Hofte, H. PROCUSTE1 Encodes a Cellulose Synthase Required for Normal Cell Elongation Specifically in Roots and Dark-Grown Hypocotyls of Arabidopsis. Plant Cell 2000, 12, 2409–2424. [Google Scholar] [CrossRef] [PubMed]
  30. Robert, S.; Bichet, A.; Grandjean, O.; Kierzkowski, D.; Satiat-Jeaunemaitre, B.; Pelletier, S.; Hauser, M.-T.; Höfte, H.; Vernhettes, S. An Arabidopsis Endo-1,4-β-d-Glucanase Involved in Cellulose Synthesis Undergoes Regulated Intracellular Cycling. Plant Cell 2005, 17, 3378–3389. [Google Scholar] [CrossRef] [PubMed]
  31. Mansoori, N.; Timmers, J.; Desprez, T.; Kamei, C.L.A.; Dees, D.C.T.; Vincken, J.P.; Visser, R.G.F.; Fte, H.H.; Vernhettes, S.; Trindade, L.M. KORRIGAN1 interacts specifically with integral components of the Cellulose synthase machinery. PLoS ONE 2014, 9, e112387. [Google Scholar] [CrossRef] [PubMed]
  32. Paredez, A.R.; Somerville, C.; Ehrhardt, D. Visualization of Cellulose Synthase with Microtubules. Science 2006, 312, 1491–1495. [Google Scholar] [CrossRef] [PubMed]
  33. Höfte, H.; Voxeur, A. Plant cell walls. Curr. Biol. 2017, 27, R865–R870. [Google Scholar] [CrossRef] [PubMed]
  34. Wasteneys, G.O.; Galway, M.E. Remodeling the cytoskeleton for growth and form: An overview with some new views. Annu. Rev. Plant Biol. 2003, 54, 691–722. [Google Scholar] [CrossRef] [PubMed]
  35. Wasteneys, G.O. Progress in understanding the role of microtubules in plant cells. Curr. Opin. Plant Biol. 2004, 7, 651–660. [Google Scholar] [CrossRef] [PubMed]
  36. Gutierrez, R.; Lindeboom, J.J.; Paredez, A.R.; Emons, A.M.C.; Ehrhardt, D.W. Arabidopsis cortical microtubules position cellulose synthase delivery to the plasma membrane and interact with cellulose synthase trafficking compartments. Nat. Cell Biol. 2009, 11, 797–806. [Google Scholar] [CrossRef] [PubMed]
  37. Crowell, E.F.; Bischoff, V.; Desprez, T.; Rolland, A.; Stierhof, Y.-D.; Schumacher, K.; Gonneau, M.; Hofte, H.; Vernhettes, S. Pausing of Golgi Bodies on Microtubules Regulates Secretion of Cellulose Synthase Complexes in Arabidopsis. Plant Cell 2009, 21, 1141–1154. [Google Scholar] [CrossRef] [PubMed]
  38. Eckardt, N.A. Role of Xyloglucan in Primary Cell Walls. Plant Cell 2008, 20, 1421–1422. [Google Scholar] [CrossRef] [PubMed]
  39. Takeda, T.; Furuta, Y.; Awano, T.; Mizuno, K.; Mitsuishi, Y.; Hayashi, T. Suppression and acceleration of cell elongation by integration of xyloglucans in pea stem segments. Proc. Natl. Acad. Sci. USA 2002, 99, 9055–9060. [Google Scholar] [CrossRef] [PubMed]
  40. Cavalier, D.M.; Lerouxel, O.; Neumetzler, L.; Yamauchi, K.; Reinecke, A.; Freshour, G.; Zabotina, O.A.; Hahn, M.G.; Burgert, I.; Pauly, M.; et al. Disrupting Two Arabidopsis thaliana Xylosyltransferase Genes Results in Plants Deficient in Xyloglucan, a Major Primary Cell Wall Component. Plant Cell 2008, 20, 1519–1537. [Google Scholar] [CrossRef] [PubMed]
  41. Hayashi, T. Xyloglucans in the primary cell wall. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1989, 20, 1519–1537. [Google Scholar] [CrossRef]
  42. Scheller, H.V.; Ulvskov, P. Hemicelluloses. Annu. Rev. Plant Biol. 2010, 61, 263–289. [Google Scholar] [CrossRef] [PubMed]
  43. Xiao, C.; Zhang, T.; Zheng, Y.; Cosgrove, D.J.; Anderson, C.T. Xyloglucan Deficiency Disrupts Microtubule Stability and Cellulose Biosynthesis in Arabidopsis, Altering Cell Growth and Morphogenesis. Plant Physiol. 2016, 170, 234–249. [Google Scholar] [CrossRef] [PubMed]
  44. Moller, I.; Sørensen, I.; Bernal, A.J.; Blaukopf, C.; Lee, K.; Øbro, J.; Pettolino, F.; Roberts, A.; Mikkelsen, J.D.; Knox, J.P.; et al. High-throughput mapping of cell-wall polymers within and between plants using novel microarrays. Plant J. 2007, 50, 1118–1128. [Google Scholar] [CrossRef] [PubMed]
  45. Willats, W.G.T.; Mccartney, L.; Mackie, W.; Knox, J.P. Pectin: Cell biology and prospects for functional analysis. Plant Mol. Biol. 2001, 47, 9–27. [Google Scholar] [CrossRef] [PubMed]
  46. Iwai, H.; Masaoka, N.; Ishii, T.; Satoh, S. A pectin glucuronyltransferase gene is essential for intercellular attachment in the plant meristem. Proc. Natl. Acad. Sci. USA 2002, 99, 16319–16324. [Google Scholar] [CrossRef] [PubMed]
  47. Ridley, B.L.; O’Neill, M.A.; Mohnen, D. Pectins: Structure, biosynthesis, and oligogalacturonide-related signaling. Phytochemistry 2001, 57, 929–967. [Google Scholar] [CrossRef]
  48. Verger, S.; Chabout, S.; Gineau, E.; Mouille, G. Cell adhesion in plants is under the control of putative O-fucosyltransferases. Development 2016, 143, 2536–2540. [Google Scholar] [CrossRef] [PubMed]
  49. Vincken, J.-P. If Homogalacturonan Were a Side Chain of Rhamnogalacturonan I. Implications for Cell Wall Architecture. Plant Physiol. 2003, 132, 1781–1789. [Google Scholar] [CrossRef] [PubMed]
  50. Vincken, J.P.; Schols, H.A.; Oomen, R.J.F.J.; Beldman, G.; Visser, R.G.F.; Voragen, A.G.J. Pectin : The hairy thing : Evidence that homogalacturonan is a side chain of rhamnogalacturonan I. In Advances in Pectin and Pectinase Research; Voragen, A.G.J., Voragen, F., Schols, H., Visser, R., Eds.; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2003; pp. 47–61. [Google Scholar]
  51. Caffall, K.H.; Mohnen, D. The structure, function, and biosynthesis of plant cell wall pectic polysaccharides. Carbohydr. Res. 2009, 344, 1879–1900. [Google Scholar] [CrossRef] [PubMed]
  52. Round, A.N.; Rigby, N.M.; MacDougall, A.J.; Morris, V.J. A new view of pectin structure revealed by acid hydrolysis and atomic force microscopy. Carbohydr. Res. 2010, 345, 487–497. [Google Scholar] [CrossRef] [PubMed]
  53. Cornuault, V.; Posé, S.; Knox, J.P. Disentangling pectic homogalacturonan and rhamnogalacturonan-I polysaccharides: Evidence for sub-populations in fruit parenchyma systems. Food Chem. 2018, 246, 275–285. [Google Scholar] [CrossRef] [PubMed]
  54. Matsunaga, T.; Ishii, T.; Matsumoto, S.; Higuchi, M.; Darvill, A.; Albersheim, P.; O’Neill, M.A. Occurrence of the primary cell wall polysaccharide rhamnogalacturonan II in pteridophytes, lycophytes, and bryophytes. Implications for the evolution of vascular plants. Plant Physiol. 2004, 134, 339–351. [Google Scholar] [CrossRef] [PubMed]
  55. Kim, S.J.; Brandizzi, F. The plant secretory pathway for the trafficking of cell wall polysaccharides and glycoproteins. Glycobiology 2016, 26, 940–949. [Google Scholar] [CrossRef] [PubMed]
  56. Zhu, C.; Ganguly, A.; Baskin, T.I.; McClosky, D.D.; Anderson, C.T.; Foster, C.; Meunier, K.A.; Okamoto, R.; Berg, H.; Dixit, R. The Fragile Fiber1 Kinesin Contributes to Cortical Microtubule-Mediated Trafficking of Cell Wall Components. Plant Physiol. 2015, 167, 780–792. [Google Scholar] [CrossRef] [PubMed]
  57. Rose, J.K.C.; Lee, S.J. Straying off the Highway: Trafficking of Secreted Plant Proteins and Complexity in the Plant Cell Wall Proteome. Plant Physiol. 2010, 153, 433–436. [Google Scholar] [CrossRef] [PubMed]
  58. Toyooka, K.; Goto, Y.; Asatsuma, S.; Koizumi, M.; Mitsui, T.; Matsuoka, K. A Mobile Secretory Vesicle Cluster Involved in Mass Transport from the Golgi to the Plant Cell Exterior. Plant Cell 2009, 21, 1212–1229. [Google Scholar] [CrossRef] [PubMed]
  59. Proseus, T.E.; Boyer, J.S. Turgor pressure moves polysaccharides into growing cell walls of Chara corallina. Ann. Bot. 2005, 95, 967–979. [Google Scholar] [CrossRef] [PubMed]
  60. Scheible, W.R.; Pauly, M. Glycosyltransferases and cell wall biosynthesis: Novel players and insights. Curr. Opin. Plant Biol. 2004, 7, 285–295. [Google Scholar] [CrossRef] [PubMed]
  61. Cassab, G.I. Plant cell wall proteins. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1998, 49, 281–309. [Google Scholar] [CrossRef] [PubMed]
  62. Carpita, N.C. Structure and Biogenesis of the Cell Walls of Grasses. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1996, 47, 445–476. [Google Scholar] [CrossRef] [PubMed]
  63. Re, A.; Plant, P.; Bioi, M.; Cassab, G.I.; Varner, J.E. Cell wall proteins. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1988, 39, 321–353. [Google Scholar]
  64. Cosgrove, D.J. Loosening of plant cell walls by expansins. Nature 2000, 407, 321–326. [Google Scholar] [CrossRef] [PubMed]
  65. Kieliszewski, M.J.; Lamport, D.T.A. Extensin: Repetitive motifs, functional sites, post-translational codes, and phylogeny. Plant J. 1994, 5, 157–172. [Google Scholar] [CrossRef] [PubMed]
  66. Lamport, D.T.A.; Kieliszewski, M.J.; Chen, Y.; Cannon, M.C. Role of the Extensin Superfamily in Primary Cell Wall Architecture. Plant Physiol. 2011, 156, 11–19. [Google Scholar] [CrossRef] [PubMed]
  67. Lamport, D.T. Oxygen Fixation into Hydroxyproline of Oxygen Fixation into Hydroxyproline Plant Cell Wall Protein. J. Biol. Chem. 1963, 238, 1438–1440. [Google Scholar] [PubMed]
  68. Showalter, A.M. Structure and function of plant cell wall proteins. Plant Cell 1993, 5, 9–23. [Google Scholar] [CrossRef] [PubMed]
  69. Cannon, M.C.; Terneus, K.; Hall, Q.; Tan, L.; Wang, Y.; Wegenhart, B.L.; Chen, L.; Lamport, D.T.A.; Chen, Y.; Kieliszewski, M.J. Self-assembly of the plant cell wall requires an extensin scaffold. Proc. Natl. Acad. Sci. USA 2008, 105, 2226–2231. [Google Scholar] [CrossRef] [PubMed]
  70. Kreuger, M.; Van Hoist, G.J. Arabinogalactan proteins and plant differentiation. Plant Mol. Biol. 1996, 30, 1077–1086. [Google Scholar] [CrossRef] [PubMed]
  71. Kreuger, M.; Van Hoist, G. Arabinogalactan proteins are essential in somatic embryogenesis of Daucus carota L. Planta 1993, 189, 243–248. [Google Scholar] [CrossRef]
  72. Veytsman, B.A.; Cosgrove, D.J. A model of cell wall expansion based on thermodynamics of polymer networks. Biophys. J. 1998, 75, 2240–2250. [Google Scholar] [CrossRef]
  73. Zhang, T.; Zheng, Y.; Cosgrove, D.J. Spatial organization of cellulose microfibrils and matrix polysaccharides in primary plant cell walls as imaged by multichannel atomic force microscopy. Plant J. 2016, 85, 179–192. [Google Scholar] [CrossRef] [PubMed]
  74. Cosgrove, D.J. Re-constructing our models of cellulose and primary cell wall assembly. Curr. Opin. Plant Biol. 2014, 22, 122–131. [Google Scholar] [CrossRef] [PubMed]
  75. Wang, T.; Hong, M. Solid-state NMR investigations of cellulose structure and interactions with matrix polysaccharides in plant primary cell walls. J. Exp. Bot. 2016, 67, 503–514. [Google Scholar] [CrossRef] [PubMed]
  76. Keegstra, K.; Kenneth, W.; Bauer, W.D.; Albersheim, P. The Structure of Plant Cell Walls. Plant Physiol. 1973, 188–196. [Google Scholar] [CrossRef]
  77. Whitney, S.E.C.; Wilson, E.; Webster, J.; Bacic, A.; Grant Reid, J.S.; Gidley, M.J. Effects of structural variation in xyloglucan polymers on interactions with bacterial cellulose. Am. J. Bot. 2006, 93, 1402–1414. [Google Scholar] [CrossRef] [PubMed]
  78. Talbott, L.D.; Ray, P.M. Molecular size and separability features of pea cell wall polysaccharides: Implications for models of primary wall structure. Plant Physiol. 1992, 98, 357–368. [Google Scholar] [CrossRef] [PubMed]
  79. Park, Y.B.; Cosgrove, D.J. Xyloglucan and its interactions with other components of the growing cell wall. Plant Cell Physiol. 2015, 56, 180–194. [Google Scholar] [CrossRef] [PubMed]
  80. Hrmova, M.; Farkas, V.; Lahnstein, J.; Fincher, G.B. A barley xyloglucan xyloglucosyl transferase covalently links xyloglucan, cellulosic substrates, and (1,3;1,4)-β-d-glucans. J. Biol. Chem. 2007, 282, 12951–12962. [Google Scholar] [CrossRef] [PubMed]
  81. Hanus, J.; Mazeau, K. The Xyloglucan-Cellulose Assembly at the Atomic Scale. Biopolymers 2006, 82, 59–73. [Google Scholar] [CrossRef] [PubMed]
  82. Zhao, Z.; Crespi, V.H.; Kubicki, J.D.; Cosgrove, D.J.; Zhong, L. Molecular dynamics simulation study of xyloglucan adsorption on cellulose surfaces: Effects of surface hydrophobicity and side-chain variation. Cellulose 2014, 21, 1025–1039. [Google Scholar] [CrossRef]
  83. Vissenberg, K.; Fry, S.C.; Pauly, M.; Höfte, H.; Verbelen, J.P. XTH acts at the microfibril-matrix interface during cell elongation. J. Exp. Bot. 2005, 56, 673–683. [Google Scholar] [CrossRef] [PubMed]
  84. Fry, S.C.; Smith, R.C.; Renwick, K.F.; Martin, D.J.; Hodge, S.K.; Matthews, K.J. Xyloglucan endotransglycosylase, a new wall-loosening enzyme activity from plants. Biochem. J. 1992, 282, 821–828. [Google Scholar] [CrossRef] [PubMed]
  85. Steele, N.M.; Sulová, Z.; Campbell, P.; Braam, J.; Farkas, V.; Fry, S.C. Ten isoenzymes of xyloglucan endotransglycosylase from plant cell walls select and cleave the donor substrate stochastically. Biochem. J. 2001, 355, 671–679. [Google Scholar] [CrossRef] [PubMed]
  86. Rose, J.K.C.; Braam, J.; Fry, S.C.; Nishitani, K. The XTH family of enzymes involved in xyloglucan endotransglucosylation and endohydrolysis: Current perspectives and a new unifying nomenclature. Plant Cell Physiol. 2002, 43, 1421–1435. [Google Scholar] [CrossRef] [PubMed]
  87. Nishitani, K.; Tominaga, R. Endo-xyloglucan transferase, a novel class of glycosyltransferase that catalyzes transfer of a segment of xyloglucan molecule to another xyloglucan molecule. J. Biol. Chem. 1992, 267, 21058–21064. [Google Scholar] [PubMed]
  88. Thompson, J.E.; Fry, S.C. Restructuring of wall-bound xyloglucan by transglycosylation in living plant cells. Plant J. 2001, 26, 23–34. [Google Scholar] [CrossRef] [PubMed]
  89. Strohmeier, M.; Hrmova, M.; Fischer, M.; Harvey, A.J.; Fincher, G.B.; Pleiss, J. Molecular modeling of family GH16 glycoside hydrolases: Potential roles for xyloglucan transglucosylases/hydrolases in cell wall modification in the poaceae. Protein Sci. 2004, 13, 3200–3213. [Google Scholar] [CrossRef] [PubMed]
  90. Thompson, J.E.; Fry, S.C. Trimming and solubilization of xyloglucan after deposition in the walls of cultured rose cells. J. Exp. Bot. 1997, 48, 297–305. [Google Scholar] [CrossRef]
  91. Baluška, F.; Hlavacka, A.; Šamaj, J.; Palme, K.; Robinson, D.G.; Matoh, T.; Mccurdy, D.W.; Menzel, D.; Volkmann, D. F-Actin-Dependent Endocytosis of Cell Wall Pectins in Meristematic Root Cells. Insights from Brefeldin A-Induced Compartments. Plant Physiol. 2002, 130, 422–431. [Google Scholar] [CrossRef] [PubMed]
  92. Wang, T.; Zabotina, O.; Hong, M. Pectin-cellulose interactions in the Arabidopsis primary cell wall from two-dimensional magic-angle-spinning solid-state nuclear magnetic resonance. Biochemistry 2012, 51, 9846–9856. [Google Scholar] [CrossRef] [PubMed]
  93. Wang, T.; Park, Y.B.; Cosgrove, D.J.; Hong, M. Cellulose-Pectin Spatial Contacts Are Inherent to Never-Dried Arabidopsis Primary Cell Walls: Evidence from Solid-State Nuclear Magnetic Resonance. Plant Physiol. 2015, 168, 871–884. [Google Scholar] [CrossRef] [PubMed]
  94. Yoneda, A.; Ito, T.; Higaki, T.; Kutsuna, N.; Saito, T.; Ishimizu, T.; Osada, H.; Hasezawa, S.; Matsui, M.; Demura, T. Cobtorin target analysis reveals that pectin functions in the deposition of cellulose microfibrils in parallel with cortical microtubules. Plant J. 2010, 64, 657–667. [Google Scholar] [CrossRef] [PubMed]
  95. Dick-Pérez, M.; Zhang, Y.; Hayes, J.; Salazar, A.; Zabotina, O.A.; Hong, M. Structure and interactions of plant cell-wall polysaccharides by two- and three-dimensional magic-angle-spinning solid-state NMR. Biochemistry 2011, 50, 989–1000. [Google Scholar] [CrossRef] [PubMed]
  96. Chanliaud, E.; Gidley, M.J. In vitro synthesis and properties of pectin/Acetobacter xylinus cellulose composites. Plant J. 1999, 20, 25–35. [Google Scholar] [CrossRef] [PubMed]
  97. Virk, S.S.; Cleland, R.E. The role of wall calcium in the extension of cell walls of soybean hypocotyls. Planta 1990, 182, 559–564. [Google Scholar] [CrossRef] [PubMed]
  98. Hocq, L.; Pelloux, J.; Lefebvre, V. Connecting Homogalacturonan-Type Pectin Remodeling to Acid Growth. Trends Plant Sci. 2017, 22, 20–29. [Google Scholar] [CrossRef] [PubMed]
  99. Braybrook, S.A.; Peaucelle, A. Mechano-Chemical Aspects of Organ Formation in Arabidopsis thaliana: The Relationship between Auxin and Pectin. PLoS ONE 2013, 8, e57813. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Zykwinska, A.; Thibault, J.F.; Ralet, M.C. Organization of pectic arabinan and galactan side chains in association with cellulose microfibrils in primary cell walls and related models envisaged. J. Exp. Bot. 2007, 58, 1795–1802. [Google Scholar] [CrossRef] [PubMed]
  101. Zykwinska, A.W.; Ralet, M.J.; Garnier, C.D.; Thibault, J.-F.J. Evidence for in vitro binding of pectin side chains to cellulose. Plant Physiol. 2005, 139, 397–407. [Google Scholar] [CrossRef] [PubMed]
  102. Proseus, T.E.; Boyer, J.S. Periplasm turgor pressure controls wall deposition and assembly in growing Chara corallina cells. Ann. Bot. 2006, 98, 93–105. [Google Scholar] [CrossRef] [PubMed]
  103. Schopfer, P. Biomechanics of plant growth. Am. J. Bot. 2006, 93, 1415–1425. [Google Scholar] [CrossRef] [PubMed]
  104. Kutschera, U.; Bergfeld, R.; Schopfer, P. Cooperation of epidermis and inner tissues in auxin-mediated growth of maize coleoptiles. Planta 1987, 170, 168–180. [Google Scholar] [CrossRef] [PubMed]
  105. Yamamoto, R. Stress relaxation property of the cell wall and auxin-induced cell elongation. J. Plant Res. 1996, 109, 75–84. [Google Scholar] [CrossRef]
  106. Perrot-Rechenmann, C. Cellular responses to auxin: Division versus expansion. Cold Spring Harb. Perspect. Biol. 2010, 2, 1–15. [Google Scholar] [CrossRef] [PubMed]
  107. Masuda, Y. Auxin-induced cell elongation and cell wall changes. Bot. Mag. 1990, 103, 345–370. [Google Scholar] [CrossRef]
  108. Velasquez, S.M.; Barbez, E.; Kleine-Vehn, J.; Estevez, J. Auxin and cellular elongation. Plant Physiol. 2016, 170. [Google Scholar] [CrossRef] [PubMed]
  109. Del Pozo, J.C. The Balance between Cell Division and Endoreplication Depends on E2FC-DPB, Transcription Factors Regulated by the Ubiquitin-SCFSKP2A Pathway in Arabidopsis. Plant Cell 2006, 18, 2224–2235. [Google Scholar] [CrossRef] [PubMed]
  110. Nishitani, K.; Masuda, Y. Auxin-induced changes in the cell wall structure: Changes in the sugar compositions, intrinsic viscosity and molecular weight distributions of matrix polysaccharides. Physiol. Plant. 1981, 52, 482–494. [Google Scholar] [CrossRef]
  111. Gendreau, E.; Traas, J.; Desnos, T.; Grandjean, O.; Caboche, M.; Höfte, H. Cellular basis of hypocotyl growth in Arabidopsis thaliana. Plant Physiol. 1997, 114, 295–305. [Google Scholar] [CrossRef] [PubMed]
  112. Ren, H.; Gray, W.M. SAUR Proteins as Effectors of Hormonal and Environmental Signals in Plant Growth. Mol. Plant 2015, 8, 1153–1164. [Google Scholar] [CrossRef] [PubMed]
  113. Blancaflor, E.B.; Hasenstein, K.H. Time course and auxin sensitivity of cortical microtubule reorientation in maize roots. Protoplasma 1995, 185, 72–82. [Google Scholar] [CrossRef] [PubMed]
  114. Takesue, K.; Shibaoka, H. Auxin-induced longitudinal-to-transverse reorientation of cortical microtubules in nonelongating epidermal cells of azuki bean epicotyls. Protoplasma 1999, 206, 27–30. [Google Scholar] [CrossRef]
  115. Fu, Y.; Li, H.; Yang, Z. The ROP2 GTPase Controls the Formation of Cortical Fine F-Actin and the Early Phase of Directional Cell Expansion during Arabidopsis Organogenesis. Plant Cell 2002, 14, 777–794. [Google Scholar] [CrossRef] [PubMed]
  116. Gu, Y.; Wang, Z.; Yang, Z. ROP/RAC GTPase: An old new master regulator for plant signaling. Curr. Opin. Plant Biol. 2004, 7, 527–536. [Google Scholar] [CrossRef] [PubMed]
  117. Chapman, E.J.; Greenham, K.; Castillejo, C.; Sartor, R.; Bialy, A.; Sun, T.; Estelle, M. Hypocotyl transcriptome reveals auxin regulation of growth-promoting genes through GA-dependent and -independent pathways. PLoS ONE 2012, 7. [Google Scholar] [CrossRef] [PubMed]
  118. Dharmasiri, N.; Dharmasiri, S.; Weijers, D.; Lechner, E.; Yamada, M.; Hobbie, L.; Ehrismann, J.S.; Jürgens, G.; Estelle, M. Plant development is regulated by a family of auxin receptor F box proteins. Dev. Cell 2005, 9, 109–119. [Google Scholar] [CrossRef] [PubMed]
  119. Berleth, T.; Krogan, N.T.; Scarpella, E. Auxin signals—Turning genes on and turning cells around. Curr. Opin. Plant Biol. 2004, 7, 553–563. [Google Scholar] [CrossRef] [PubMed]
  120. Reed, J.W. Roles and activities of Aux/IAA proteins in Arabidopsis. Trends Plant Sci. 2001, 6, 420–425. [Google Scholar] [CrossRef]
  121. Ramos, J.A.; Zenser, N.; Leyser, O.; Callis, J. Rapid degradation of auxin/indoleacetic acid proteins requires conserved amino acids of domain II and is proteasome dependent. Plant Cell 2001, 13, 2349–2360. [Google Scholar] [CrossRef] [PubMed]
  122. Overvoorde, P.J.P.; Okushima, Y.; Alonso, J.J.M.; Chan, A.; Chang, C.; Ecker, J.R.; Hughes, B.; Liu, A.; Onodera, C.; Quach, H.; et al. Functional genomic analysis of the AUXIN/INDOLE-3-ACETIC ACID gene family members in Arabidopsis thaliana. Plant Cell 2005, 17, 3282–3300. [Google Scholar] [CrossRef] [PubMed]
  123. Sato, A.; Yamamoto, K.T. Overexpression of the non-canonical Aux/IAA genes causes auxin-related aberrant phenotypes in Arabidopsis. Physiol. Plant. 2008, 133, 397–405. [Google Scholar] [CrossRef] [PubMed]
  124. Liscum, E.; Reed, J.W. Genetics of Aux/IAA and ARF action in plant growth and development. Plant Mol. Biol. 2002, 49, 387–400. [Google Scholar] [CrossRef] [PubMed]
  125. Mockaitis, K.; Estelle, M. Auxin Receptors and Plant Development: A New Signaling Paradigm. Annu. Rev. Cell Dev. Biol. 2008, 24, 55–80. [Google Scholar] [CrossRef] [PubMed]
  126. Ulmasov, T.; Hagen, G.; Guilfoyle, T.J. Activation and repression of transcription by auxin-response factors. Proc. Natl. Acad. Sci. USA 1999, 96, 5844–5849. [Google Scholar] [CrossRef] [PubMed]
  127. Tiwari, S.B.; Hagen, G.; Guilfoyle, T. The roles of auxin response factor domains in auxin-responsive transcription. Plant Cell 2003, 15, 533–543. [Google Scholar] [CrossRef] [PubMed]
  128. Hagen, G.; Guilfoyle, T. Auxin-responsive gene expression: Genes, promoters and regulatory factors. Plant Mol. Biol. 2002, 49, 373–385. [Google Scholar] [CrossRef] [PubMed]
  129. Esmon, C.A.; Tinsley, A.G.; Ljung, K.; Sandberg, G.; Hearne, L.B.; Liscum, E. A gradient of auxin and auxin-dependent transcription precedes tropic growth responses. Proc. Natl. Acad. Sci. USA 2006, 103, 236–241. [Google Scholar] [CrossRef] [PubMed]
  130. Pelletier, S.; Van Orden, J.; Wolf, S.; Vissenberg, K.; Delacourt, J.; Ndong, Y.A.; Pelloux, J.; Bischoff, V.; Urbain, A.; Mouille, G.; et al. A role for pectin de-methylesterification in a developmentally regulated growth acceleration in dark-grown Arabidopsis hypocotyls. New Phytol. 2010, 188, 726–739. [Google Scholar] [CrossRef] [PubMed]
  131. Nemhauser, J.L.; Hong, F.; Chory, J. Different Plant Hormones Regulate Similar Processes through Largely Nonoverlapping Transcriptional Responses. Cell 2006, 126, 467–475. [Google Scholar] [CrossRef] [PubMed]
  132. Cosgrove, D.J. Enzymes and Other Agents That Enhance Cell Wall Extensibility. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1999, 50, 391–417. [Google Scholar] [CrossRef] [PubMed]
  133. Eklof, J.M.; Brumer, H. The XTH Gene Family: An Update on Enzyme Structure, Function, and Phylogeny in Xyloglucan Remodeling. Plant Physiol. 2010, 153, 456–466. [Google Scholar] [CrossRef] [PubMed]
  134. Motose, H.; Sugiyama, M.; Fukuda, H. A proteoglycan mediates inductive interaction during plant vascular development. Nature 2004, 429, 873–878. [Google Scholar] [CrossRef] [PubMed]
  135. Seifert, G.J.; Roberts, K. The Biology of Arabinogalactan Proteins. Annu. Rev. Plant Biol. 2007, 58, 137–161. [Google Scholar] [CrossRef] [PubMed]
  136. Peaucelle, A.; Braybrook, S.A.; Le Guillou, L.; Bron, E.; Kuhlemeier, C.; Höfte, H. Pectin-induced changes in cell wall mechanics underlie organ initiation in Arabidopsis. Curr. Biol. 2011, 21, 1720–1726. [Google Scholar] [CrossRef] [PubMed]
  137. Savaldi-Goldstein, S.; Baiga, T.J.; Pojer, F.; Dabi, T.; Butterfield, C.; Parry, G.; Santner, A.; Dharmasiri, N.; Tao, Y.; Estelle, M.; et al. New auxin analogs with growth-promoting effects in intact plants reveal a chemical strategy to improve hormone delivery. Proc. Natl. Acad. Sci. USA 2008, 105, 15190–15195. [Google Scholar] [CrossRef] [PubMed]
  138. Hong, F.; Breitling, R.; McEntee, C.W.; Wittner, B.S.; Nemhauser, J.L.; Chory, J. A bioconductor package for detecting differentially expressed genes in meta-analysis. Bioinformatics 2006, 22, 2825–2827. [Google Scholar] [CrossRef] [PubMed]
  139. Breitling, R.; Armengaud, P.; Amtmann, A.; Herzyk, P. Rank products: A simple, yet powerful, new method to detect differentially regulated genes in replicated microarray experiments. FEBS Lett. 2004, 573, 83–92. [Google Scholar] [CrossRef] [PubMed]
  140. Hager, A.; Menzel, H.; Krauss, A. Versuche und Hypothese zur Primärwirkung des Auxins beim Streckungswachstum. Planta 1971, 100, 47–75. [Google Scholar] [CrossRef] [PubMed]
  141. Rayle, D.L.; Cleland, R. Enhancement of wall loosening and elongation by Acid solutions. Plant Physiol. 1970, 46, 250–253. [Google Scholar] [CrossRef] [PubMed]
  142. Lüthen, H.; Bigdon, M.; Böttger, M. Reexamination of the Acid growth theory of auxin action. Plant Physiol. 1990, 93, 931–939. [Google Scholar] [CrossRef] [PubMed]
  143. Cleland, R.E.; Buckley, G.; Nowbar, S.; Lew, N.M.; Stinemetz, C.; Evans, M.L.; Rayle, D.L. The pH profile for acid-induced elongation of coleoptile and epicotyl sections is consistent with the acid-growth theory. Planta 1991, 186, 70–74. [Google Scholar] [CrossRef] [PubMed]
  144. Karcz, W.; Stolarek, J.; Pietruszka, M.; Malkowski, E. The dose-response curves for IAA induced elongation growth and acidification of the incubation medium of Zea mays coleoptile segments. Physiol. Plant. 1990, 80, 257–261. [Google Scholar] [CrossRef]
  145. Arsuffi, G.; Braybrook, S.A. Acid growth: An ongoing trip. J. Exp. Bot. 2017, 2, 137–146. [Google Scholar] [CrossRef] [PubMed]
  146. Fendrych, M.; Leung, J.; Friml, J. Tir1/AFB-Aux/IAA auxin perception mediates rapid cell wall acidification and growth of Arabidopsis hypocotyls. Elife 2016, 5, 1–19. [Google Scholar] [CrossRef] [PubMed]
  147. Takahashi, K.; Hayashi, K.; Kinoshita, T. Auxin Activates the Plasma Membrane H+-ATPase by Phosphorylation during Hypocotyl Elongation in Arabidopsis. Plant Physiol. 2012, 159, 632–641. [Google Scholar] [CrossRef] [PubMed]
  148. Spartz, A.K.; Ren, H.; Park, M.Y.; Grandt, K.N.; Lee, S.H.; Murphy, A.S.; Sussman, M.R.; Overvoorde, P.J.; Gray, W.M. SAUR Inhibition of PP2C-D Phosphatases Activates Plasma Membrane H+-ATPases to Promote Cell Expansion in Arabidopsis. Plant Cell 2014, 26, 2129–2142. [Google Scholar] [CrossRef] [PubMed]
  149. Hager, A.; Debus, G.; Edel, H.G.; Stransky, H.; Serrano, R. Auxin induces exocytosis and the rapid synthesis of a high-turnover pool of plasma-membrane H+-ATPase. Planta 1991, 185, 527–537. [Google Scholar] [CrossRef] [PubMed]
  150. Thiel, G.; Weise, R. Auxin augments conductance of K+ inward rectifier in maize coleoptile protoplasts. Planta 1999, 208, 38–45. [Google Scholar] [CrossRef]
  151. Philippar, K.; Ivashikina, N.; Ache, P.; Christian, M.; Lüthen, H.; Palme, K.; Hedrich, R. Auxin activates KAT1 and KAT2, two K+-channel genes expressed in seedlings of Arabidopsis thaliana. Plant J. 2004, 37, 815–827. [Google Scholar] [CrossRef] [PubMed]
  152. Frías, I.; Caldeira, M.T.; Pérez-Castiñeira, J.R.; Navarro-Aviñó, J.P.; Culiañez-Maciá, F.; Kuppinger, O.; Stransky, H.; Pagés, M.; Hager, A.; Serrano, R. A major isoform of the maize plasma membrane H+-ATPase: Characterization and induction by auxin in coleoptiles. Plant Cell 1996, 8, 1533–1544. [Google Scholar] [PubMed]
  153. Kutschera, U.; Niklas, K.J. The epidermal-growth-control theory of stem elongation: An old and a new perspective. J. Plant Physiol. 2007, 164, 1395–1409. [Google Scholar] [CrossRef] [PubMed]
  154. Marcotrigiano, M. A role for leaf epidermis in the control of leaf size and the rate and extent of mesophyll cell division. Am. J. Bot. 2010, 97, 224–233. [Google Scholar] [CrossRef] [PubMed]
  155. Philippar, K.; Fuchs, I.; Luthen, H.; Hoth, S.; Bauer, C.S.; Haga, K.; Thiel, G.; Ljung, K.; Sandberg, G.; Bottger, M.; et al. Auxin-induced K+ channel expression represents an essential step in coleoptile growth and gravitropism. Proc. Natl. Acad. Sci. USA 1999, 96, 12186–12191. [Google Scholar] [CrossRef] [PubMed]
  156. Savaldi-Goldstein, S.; Chory, J. Growth coordination and the shoot epidermis. Curr. Opin. Plant Biol. 2008, 11, 42–48. [Google Scholar] [CrossRef] [PubMed]
  157. Van Volkenburgh, E.; Schmidt, M.G.; Cleland, R.E. Loss of capacity for acid-induced wall loosening as the principal cause of the cessation of cell enlargement in light-grown bean leaves. Planta 1985, 163, 500–505. [Google Scholar] [CrossRef] [PubMed]
  158. Winch, S.; Pritchard, J. Acid-induced wall loosening is confined to the accelerating region of the root growing zone. J. Exp. Bot. 1999, 50, 1481–1487. [Google Scholar] [CrossRef]
  159. McQueen-Mason, S.; Durachko, D.M.; Cosgrove, D.J. Two Endogenous Proteins That Induce Cell Wall Extension in Plants. Plant Cell 1992, 4, 1425. [Google Scholar] [CrossRef] [PubMed]
  160. McQueen-Mason, S.; Cosgrove, D.J. Disruption of hydrogen bonding between plant cell wall polymers by proteins that induce wall extension. Proc. Natl. Acad. Sci. USA 1994, 91, 6574–6578. [Google Scholar] [CrossRef] [PubMed]
  161. McQueen-Mason, S.J.; Cosgrove, D.J. Expansin Mode of Action on Cell Walls. Plant Physiol. 1995, 107, 87–100. [Google Scholar] [CrossRef] [PubMed]
  162. Link, B.M.; Cosgrove, D.J. Acid-growth response and α-expansins in suspension cultures of bright yellow 2 tobacco. Plant Physiol. 1998, 118, 907–916. [Google Scholar] [CrossRef] [PubMed]
  163. Yennawar, N.H.; Li, L.-C.; Dudzinski, D.M.; Tabuchi, A.; Cosgrove, D.J. Crystal structure and activities of EXPB1 (Zea m 1), a β-expansin and group-1 pollen allergen from maize. Proc. Natl. Acad. Sci. USA 2006, 103, 14664–14671. [Google Scholar] [CrossRef] [PubMed]
  164. Wang, L.; Ruan, Y.-L. Regulation of cell division and expansion by sugar and auxin signaling. Front. Plant Sci. 2013, 4, 1–9. [Google Scholar] [CrossRef] [PubMed]
  165. Purugganan, M.M.; Braam, J.; Fry, S.C. The Arabidopsis TCH4 xyloglucan endotransglycosylase. Substrate specificity, pH optimum, and cold tolerance. Plant Physiol. 1997, 115, 181–190. [Google Scholar] [CrossRef] [PubMed]
  166. Shipp, M.; Nadella, R.; Gao, H.; Farkas, V.; Sigrist, H.; Faik, A. Glyco-array technology for efficient monitoring of plant cell wall glycosyltransferase activities. Glycoconj. J. 2008, 25, 49–58. [Google Scholar] [CrossRef] [PubMed]
  167. Kotake, T.; Nakagawa, N.; Takeda, K.; Sakurai, N. Auxin-Induced Elongation Growth and Expressions of Cell Wall-Bound Exo- and Endo-β-Glucanases in Barley Coleoptiles. Plant Cell Physiol. 2000, 41, 1272–1278. [Google Scholar] [CrossRef] [PubMed]
  168. Yokoyama, R.; Nishitani, K. A comprehensive expression analysis of all members of a gene family encoding cell-wall enzymes allowed us to predict cis-regulatory regions involved in cell-wall construction in specific organs of Arabidopsis. Plant Cell Physiol. 2001, 42, 1025–1033. [Google Scholar] [CrossRef] [PubMed]
  169. Swarup, K.; Benková, E.; Swarup, R.; Casimiro, I.; Péret, B.; Yang, Y.; Parry, G.; Nielsen, E.; De Smet, I.; Vanneste, S.; et al. The auxin influx carrier LAX3 promotes lateral root emergence. Nat. Cell Biol. 2008, 10, 946–954. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Xu, W. Arabidopsis TCH4, Regulated by Hormones and the Environment, Encodes a Xyloglucan Endotransglycosylase. Plant Cell 1995, 7, 1555–1567. [Google Scholar] [CrossRef] [PubMed]
  171. Xu, W.; Campbell, P.; Vargheese, A.K.; Braam, J. The Arabidopsis XET-related gene family: Environmental and hormonal regulation of expression. Plant J. 1996, 9, 879–889. [Google Scholar] [CrossRef] [PubMed]
  172. Ohmiya, Y.; Samejima, M.; Shiroishi, M.; Amano, Y.; Kanda, T.; Sakai, F.; Hayashi, T. Evidence that endo-1,4-β-glucanases act on cellulose in suspension- cultured poplar cells. Plant J. 2000, 24, 147–158. [Google Scholar] [CrossRef] [PubMed]
  173. Verma, D.P.S.; Maclachlan, G.A.; Byrne, H.; Ewings, D. Regulation In Vitro Translation of Messenger Ribonucleic Acid for Cellulase from Auxin-Treated Pea Epicotyls. J. Biol. Chem. 1975, 250, 1019–1026. [Google Scholar] [PubMed]
  174. Fry, S.C. Cellulases, hemicelluloses and auxin-stimulated growth: A possible relationship. Physiol. Plant. 1989, 75, 532–536. [Google Scholar] [CrossRef]
  175. Catala, C.; Rose, J.K.; Bennett, A.B. Auxin regulation and spatial localization of an endo-1,4-β-d-glucanase and a xyloglucan endotransglycosylase in expanding tomato hypocotyls. Plant J. 1997, 12, 417–426. [Google Scholar] [CrossRef] [PubMed]
  176. Antosiewicz, D.M.; Purugganan, M.M.; Polisensky, D.H.; Braam, J. Cellular localization of Arabidopsis xyloglucan endotransglycosylase-related proteins during development and after wind stimulation. Plant Physiol. 1997, 115, 1319–1328. [Google Scholar] [CrossRef] [PubMed]
  177. Van Sandt, V.S.T.; Suslov, D.; Verbelen, J.P.; Vissenberg, K. Xyloglucan endotransglucosylase activity loosens a plant cell wall. Ann. Bot. 2007, 100, 1467–1473. [Google Scholar] [CrossRef] [PubMed]
  178. McDougall, G.J.; Fry, S.C. Inhibition of auxin-stimulated growth of pea stem segments by a specific nonasaccharide of xyloglucan. Planta 1988, 175, 412–416. [Google Scholar] [CrossRef] [PubMed]
  179. Paque, S.; Mouille, G.; Grandont, L.; Alabadi, D.; Gaertner, C.; Goyallon, A.; Muller, P.; Primard-Brisset, C.; Sormani, R.; Blazquez, M.A.; et al. AUXIN BINDING PROTEIN1 Links Cell Wall Remodeling, Auxin Signaling, and Cell Expansion in Arabidopsis. Plant Cell 2014, 26, 280–295. [Google Scholar] [CrossRef] [PubMed]
  180. York, W.S.; Darvill, A.G.; Albersheim, P. Inhibition of 2,4-dichlorophenoxyacetic Acid-stimulated elongation of pea stem segments by a xyloglucan oligosaccharide. Plant Physiol. 1984, 75, 295–297. [Google Scholar] [CrossRef] [PubMed]
  181. Hyodo, H.; Yamakawa, S.; Takeda, Y.; Tsuduki, M.; Yokota, A.; Nishitani, K.; Kohchi, T. Active gene expression of a xyloglucan endotransglucosylase/hydrolase gene, XTH9, in inflorescence apices is related to cell elongation in Arabidopsis thaliana. Plant Mol. Biol. 2003, 52, 473–482. [Google Scholar] [CrossRef] [PubMed]
  182. Vissenberg, K.; Fry, S.C.; Verbelen, J.P. Root hair initiation is coupled to a highly localized increase of xyloglucan endotransglycosylase action in Arabidopsis roots. Plant Physiol. 2001, 127, 1125–1135. [Google Scholar] [CrossRef] [PubMed]
  183. Fry, S.C. The Structure and Functions of Xyloglucan. J. Exp. Bot. 1989, 40, 1–11. [Google Scholar] [CrossRef]
  184. Marin-Rodriguez, M.C.; Orchard, J.; Seymour, G.B. Pectate lyases, cell wall degradation and fruit softening. J. Exp. Bot. 2002, 53, 2115–2119. [Google Scholar] [CrossRef] [PubMed]
  185. Wakabayashi, K.; Chun, J.P.; Huber, D.J. Extensive solubilization and depolymerization of cell wall polysaccharides during avocado (Persea americana) ripening involves concerted action of polygalacturonase and pectinmethylesterase. Physiol. Plant. 2000, 108, 345–352. [Google Scholar] [CrossRef]
  186. Wakabayashi, K.; Hoson, T.; Huber, D.J. Methyl de-esterification as a major factor regulating the extent of pectin depolymerization during fruit ripening: A comparison of the action of avocado (Persea americana) and tomato (Lycopersicon esculentum) polygalacturonases. J. Plant Physiol. 2003, 160, 667–673. [Google Scholar] [CrossRef] [PubMed]
  187. Osorio, S.; Castillejo, C.; Quesada, M.A.; Medina-Escobar, N.; Brownsey, G.J.; Suau, R.; Heredia, A.; Botella, M.A.; Valpuesta, V. Partial demethylation of oligogalacturonides by pectin methyl esterase 1 is required for eliciting defence responses in wild strawberry (Fragaria vesca). Plant J. 2008, 54, 43–55. [Google Scholar] [CrossRef] [PubMed]
  188. Ferrari, S. Oligogalacturonides: Plant damage-associated molecular patterns and regulators of growth and development. Front. Plant Sci. 2013, 4, 1–9. [Google Scholar] [CrossRef] [PubMed]
  189. Savatin, D.V.; Ferrari, S.; Sicilia, F.; De Lorenzo, G. Oligogalacturonide-Auxin Antagonism Does Not Require Posttranscriptional Gene Silencing or Stabilization of Auxin Response Repressors in Arabidopsis. Plant Physiol. 2011, 157, 1163–1174. [Google Scholar] [CrossRef] [PubMed]
  190. Branca, C.; De Lorenzo, G.; Cervone, F. Competitive inhibition of the auxin-induced elongation by α-d-oligogalacturonides in pea stem segments. Physiol. Plant. 1988, 72, 499–504. [Google Scholar] [CrossRef]
  191. Bellincampi, D.; Cardarelli, M.; Zaghi, D.; Serino, G.; Salvi, G.; Gatz, C.; Cervone, F.; Altamura, M.M.; Costantino, P.; Lorenzoai, G.D. Oligogalacturonides Prevent Rhizogenesis in rolB-Transformed Tobacco Explants by lnhibiting Auxin-lnduced Expression of the rolB Gene. Plant Cell 1996, 8, 477–487. [Google Scholar] [CrossRef] [PubMed]
  192. Ferrari, S.; Galletti, R.; Pontiggia, D.; Manfredini, C.; Lionetti, V.; Bellincampi, D.; Cervone, F.; De Lorenzo, G. Transgenic expression of a fungal endo-polygalacturonase increases plant resistance to pathogens and reduces auxin sensitivity. Plant Physiol. 2008, 146, 669–681. [Google Scholar] [CrossRef] [PubMed]
  193. Galletti, R.; Denoux, C.; Gambetta, S.; Dewdney, J.; Ausubel, F.M.; De Lorenzo, G.; Ferrari, S. The AtrbohD-Mediated Oxidative Burst Elicited by Oligogalacturonides in Arabidopsis Is Dispensable for the Activation of Defense Responses Effective against Botrytis cinerea. Plant Physiol. 2008, 148, 1695–1706. [Google Scholar] [CrossRef] [PubMed]
  194. Noctor, G.; Reichheld, J.P.; Foyer, C.H. ROS-related redox regulation and signaling in plants. Semin. Cell Dev. Biol. 2017. [Google Scholar] [CrossRef] [PubMed]
  195. Tenhaken, R. Cell wall remodeling under abiotic stress. Front. Plant Sci. 2015, 5, 1–9. [Google Scholar] [CrossRef] [PubMed]
  196. Francoz, E.; Ranocha, P.; Nguyen-Kim, H.; Jamet, E.; Burlat, V.; Dunand, C. Roles of cell wall peroxidases in plant development. Phytochemistry 2015, 112, 15–21. [Google Scholar] [CrossRef] [PubMed]
  197. Bailey-Serres, J.; Mittler, R. The roles of reactive oxygen species in plant cells. Plant Physiol. 2006, 141, 900191. [Google Scholar] [CrossRef] [PubMed]
  198. Schopfer, P.; Liszkay, A.; Bechtold, M.; Frahry, G.; Wagner, A. Evidence that hydroxyl radicals mediate auxin-induced extension growth. Planta 2002, 214, 821–828. [Google Scholar] [CrossRef] [PubMed]
  199. Marga, F.; Grandbois, M.; Cosgrove, D.J.; Baskin, T.I. Cell wall extension results in the coordinate separation of parallel microfibrils: Evidence from scanning electron microscopy and atomic force microscopy. Plant J. 2005, 43, 181–190. [Google Scholar] [CrossRef] [PubMed]
  200. Thompson, D.S. Space and time in the plant cell wall: Relationships between cell type, cell wall rheology and cell function. Ann. Bot. 2008, 101, 203–211. [Google Scholar] [CrossRef] [PubMed]
  201. Lockhart, J.A. An analysis of irreversible plant cell elongation. J. Theor. Biol. 1965, 8, 264–275. [Google Scholar] [CrossRef]
  202. Nakagawa, Y.; Katagiri, T.; Shinozaki, K.; Qi, Z.; Tatsumi, H.; Furuichi, T.; Kishigami, A.; Sokabe, M.; Kojima, I.; Sato, S.; et al. Arabidopsis plasma membrane protein crucial for Ca2+ influx and touch sensing in roots. Proc. Natl. Acad. Sci. USA 2007, 104, 3639–3644. [Google Scholar] [CrossRef] [PubMed]
  203. Monshausen, G.B.; Bibikova, T.N.; Weisenseel, M.H.; Gilroy, S. Ca2+ Regulates Reactive Oxygen Species Production and pH during Mechanosensing in Arabidopsis Roots. Plant Cell 2009, 21, 2341–2356. [Google Scholar] [CrossRef] [PubMed]
  204. Knight, M.R.; Campbell, A.K.; Smith, S.M.; Trewavas, A.J. Transgenic plant aequorin reports the effects of touch and cold-shock and elicitors on cytoplasmic calcium. Nature 1991, 352, 524–526. [Google Scholar] [CrossRef] [PubMed]
  205. Valentin, R.; Cerclier, C.; Geneix, N.; Aguié-Béghin, V.; Gaillard, C.; Ralet, M.C.; Cathala, B. Elaboration of extensin-pectin thin film model of primary plant cell wall. Langmuir 2010, 26, 9891–9898. [Google Scholar] [CrossRef] [PubMed]
  206. Wolf, S.; Mouille, G.; Pelloux, J. Homogalacturonan methyl-esterification and plant development. Mol. Plant 2009, 2, 851–860. [Google Scholar] [CrossRef] [PubMed]
  207. Willats, W.G.T.; Orfila, C.; Limberg, G.; Buchholt, H.C.; Van Alebeek, G.J.W.M.; Voragen, A.G.J.; Marcus, S.E.; Christensen, T.M.I.E.; Mikkelsen, J.D.; Murray, B.S.; et al. Modulation of the degree and pattern of methyl-esterification of pectic homogalacturonan in plant cell walls: Implications for pectin methyl esterase action, matrix properties, and cell adhesion. J. Biol. Chem. 2001, 276, 19404–19413. [Google Scholar] [CrossRef] [PubMed]
  208. Cameron, R.G.; Luzio, G.A.; Goodner, K.; Williams, M.A.K. Demethylation of a model homogalacturonan with a salt-independent pectin methylesterase from citrus: I. Effect of pH on demethylated block size, block number and enzyme mode of action. Carbohydr. Polym. 2008, 71, 287–299. [Google Scholar] [CrossRef]
  209. Catoire, L.; Pierron, M.; Morvan, C.; Du Penhoat, C.H.; Goldberg, R. Investigation of the action patterns of pectinmethylesterase isoforms through kinetic analyses and NMR spectroscopy: Implications in cell wall expansion. J. Biol. Chem. 1998, 273, 33150–33156. [Google Scholar] [CrossRef] [PubMed]
  210. Denès, J.M.; Baron, A.; Renard, C.M.G.C.; Péan, C.; Drilleau, J.F. Different action patterns for apple pectin methylesterase at pH 7.0 and 4.5. Carbohydr. Res. 2000, 327, 385–393. [Google Scholar] [CrossRef]
  211. Jolie, R.P.; Duvetter, T.; Van Loey, A.M.; Hendrickx, M.E. Pectin methylesterase and its proteinaceous inhibitor: A review. Carbohydr. Res. 2010, 345, 2583–2595. [Google Scholar] [CrossRef] [PubMed]
  212. Kim, Y.; Teng, Q.; Wicker, L. Action pattern of Valencia orange PME de-esterification of high methoxyl pectin and characterization of modified pectins. Carbohydr. Res. 2005, 340, 2620–2629. [Google Scholar] [CrossRef] [PubMed]
  213. Wolf, S.; Rausch, T.; Greiner, S. The N-terminal pro region mediates retention of unprocessed type-I PME in the Golgi apparatus. Plant J. 2009, 58, 361–375. [Google Scholar] [CrossRef] [PubMed]
  214. Western, T.L.; Burn, J.; Tan, W.L.; Skinner, D.J.; Martin-McCaffrey, L.; Moffatt, B.A.; Haughn, G.W. Isolation and Characterization of Mutants Defective in Seed Coat Mucilage Secretory Cell Development in Arabidopsis. Plant Physiol. 2001, 127, 998–1011. [Google Scholar] [CrossRef] [PubMed]
  215. Griffiths, J.S.; Tsai, A.Y.-L.; Xue, H.; Voiniciuc, C.; Ola, K.; Seifert, G.J.; Mansfield, S.D.; Haughn, G.W. SALT-OVERLY SENSITIVE5 Mediates Arabidopsis Seed Coat Mucilage Adherence and Organization through Pectins. Plant Physiol. 2014, 165, 991–1004. [Google Scholar] [CrossRef] [PubMed]
  216. Griffiths, J.S.; Crepeau, M.-J.; Ralet, M.-C.; Seifert, G.J.; North, H.M. Dissecting Seed Mucilage Adherence Mediated by FEI2 and SOS5. Front. Plant Sci. 2016, 7, 1–13. [Google Scholar] [CrossRef] [PubMed]
  217. Raggi, S.; Ferrarini, A.; Delledonne, M.; Dunand, C.; Ranocha, P.; De Lorenzo, G.; Cervone, F.; Ferrari, S. The Arabidopsis thaliana Class III Peroxidase AtPRX71 Negatively Regulates Growth under Physiological Conditions and in Response to Cell Wall Damage. Plant Physiol. 2015, 169, 2513–2525. [Google Scholar] [CrossRef] [PubMed]
  218. Pereira, C.S.; Ribeiro, J.M.L.; Vatulescu, A.D.; Findlay, K.; MacDougall, A.J.; Jackson, P.A.P. Extensin network formation in Vitis vinifera callus cells is an essential and causal event in rapid and H2O2-induced reduction in primary cell wall hydration. BMC Plant Biol. 2011, 11, 106. [Google Scholar] [CrossRef] [PubMed]
  219. Swanson, S.; Gilroy, S. ROS in plant development. Physiol. Plant. 2010, 138, 384–392. [Google Scholar] [CrossRef] [PubMed]
  220. Fry, S.C. Oxidative scission of plant cell wall polysaccharides by ascorbate-induced hydroxyl radicals. Biochem. J. 1998, 332, 507–515. [Google Scholar] [CrossRef] [PubMed]
  221. Li, L.; Krens, S.F.G.; Fendrych, M.; Friml, J. Real-time Analysis of Auxin Response, Cell Wall pH and Elongation in Arabidopsis thaliana Hypocotyls. Bio-Protocol 2018, 7, 1–13. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The role of auxin in cell wall expansion. Isodiametric plant cell preparing for elongation (A), undergoing elongation (B) and fully elongated (C). The cell contains intracellular structures such as nucleus (n) and vacuole(s) (v) in the cytosol (c) and is surrounded by plasma membrane (PM). Outside of the PM the cell wall (CW) is present (AC). The PM consists of a phospholipid bilayer (in blue), while the cell wall consists of different polysaccharides such as cellulose microfibrils (CMFs in yellow), pectins (green double line), XyGs (red line) and other polysaccharides (not shown). Auxin activates plasma membrane H+-ATPase proton pumps, which pump protons (H+) into the wall matrix, leading to wall acidification (a). Acidification of the apoplast activates potassium channels, which transport potassium ions (K+) to the cytosol, stimulating water (H2O) uptake and maintaining tensile stress (yellow arrows in A and B) (b). Acidic pH activates wall-loosening proteins and enzymes, which loosen the connections between different cell wall polysaccharides (c). PMEs activate plasma membrane nicotinamide adenine dinucleotide phosphate (NADPH), transporting superoxide anions to the cell wall where they are converted to hydrogen peroxide (d). Wall-loosening proteins and enzymes cause CMF sliding and moving apart, which increases wall porosity (e). Cell wall extension leads to the activation of calcium channels and calcium efflux into the cytosol (f). Accumulation of cytosolic calcium inhibits H+-ATPase proton pumps and protoplast alkalization (g). Newly synthesized polysaccharides are inserted into the wall, where they arrive via vesicular trafficking (h). Wall alkalization activates PMEs, which in turn activate wall-degrading enzymes (i) and NADPH (j) causing crosslinking of the wall polysaccharides and growth cessation (k).
Figure 1. The role of auxin in cell wall expansion. Isodiametric plant cell preparing for elongation (A), undergoing elongation (B) and fully elongated (C). The cell contains intracellular structures such as nucleus (n) and vacuole(s) (v) in the cytosol (c) and is surrounded by plasma membrane (PM). Outside of the PM the cell wall (CW) is present (AC). The PM consists of a phospholipid bilayer (in blue), while the cell wall consists of different polysaccharides such as cellulose microfibrils (CMFs in yellow), pectins (green double line), XyGs (red line) and other polysaccharides (not shown). Auxin activates plasma membrane H+-ATPase proton pumps, which pump protons (H+) into the wall matrix, leading to wall acidification (a). Acidification of the apoplast activates potassium channels, which transport potassium ions (K+) to the cytosol, stimulating water (H2O) uptake and maintaining tensile stress (yellow arrows in A and B) (b). Acidic pH activates wall-loosening proteins and enzymes, which loosen the connections between different cell wall polysaccharides (c). PMEs activate plasma membrane nicotinamide adenine dinucleotide phosphate (NADPH), transporting superoxide anions to the cell wall where they are converted to hydrogen peroxide (d). Wall-loosening proteins and enzymes cause CMF sliding and moving apart, which increases wall porosity (e). Cell wall extension leads to the activation of calcium channels and calcium efflux into the cytosol (f). Accumulation of cytosolic calcium inhibits H+-ATPase proton pumps and protoplast alkalization (g). Newly synthesized polysaccharides are inserted into the wall, where they arrive via vesicular trafficking (h). Wall alkalization activates PMEs, which in turn activate wall-degrading enzymes (i) and NADPH (j) causing crosslinking of the wall polysaccharides and growth cessation (k).
Ijms 19 00951 g001
Table 1. Selected cell wall-related genes upregulated by IAA treatment in Arabidopsis seedlings (genes from [131]).
Table 1. Selected cell wall-related genes upregulated by IAA treatment in Arabidopsis seedlings (genes from [131]).
Cellulose Related
CELLULOSE SYNTHASE-LIKECSLC4; CSLC5
EXPANSIN related
EXPANSINEXPA4; EXPA11
EXPANSIN-LIKEEXLA3
XTH related
XYLOGLUCAN ENDOTRANSGLUCOSYLASE/HYDROLASEXTH18; XTH19; XTH23; XTH33
TOUCHTCH2; TCH3; TCH4
XYLOSYLTRANSFERASEXT1
ACT DOMAIN REPEAT 7ACR7
Pectin related
PECTIN METHYLESTERASEPME1; PME34
PLANT INVERTASE/PECTIN METHYLESTERASE INHIBITOR SUPERFAMILY
PECTIN ACETYLESTERASEPAE11
GALACTURONOSYLTRANSFERASE-LIKE 10GATL10
GALACTURONOSYLTRANSFERASE-LIKEGATL3
GALACTAN SYNTHASEGALS3
POLYGALACTURONASE INHIBITING PROTEIN 1PGIP1
PEROXIDASE related
PEROXIDASE SUPERFAMILY PROTEINS
Secondary cell wall related
OVATE FAMILY PROTEIN 1OFP1
REDUCED IN LATERAL GROWTH 1RUL1
Other/biosynthesis related
EXORDIUM LIKE 2EXL2
Table 2. Selected cell wall-related genes upregulated by picloram treatment in elongating Arabidopsis hypocotyls (genes from [117]).
Table 2. Selected cell wall-related genes upregulated by picloram treatment in elongating Arabidopsis hypocotyls (genes from [117]).
Cellulose Related
CELLULOSE SYNTHASE-LIKECSLC04; CSLC12; CSLD2; CSLD3
CELLULOSE SYNTHASE-INTERACTIVE PROTEIN 1CSI1
EXPANSIN related
EXPANSINEXPA1; XPA7; EXPA10; EXPA12; EXPA18
EXPANSIN-LIKEEXLA1; EXLA2; EXLA3
XTH related
XYLOGLUCAN ENDOTRANSGLUCOSYLASE/HYDROLASEXTH8; XTH17; XTH18; XTH19 XTH23; XTH33
TOUCH 3TCH3
XYLOSYLTRANSFERASE 1XT1
Pectin related
PECTIN METHYLESTERASEPME2; PME41
PLANT INVERTASE/PECTIN METHYLESTERASE INHIBITOR SUPERFAMILY
PECTIN METHYLESTERASE INIHIBITOR 7PMEI7
PECTIN ACETYLESTERASE 11PAE9; PAE11
PECTIN LYASE-LIKE SUPERFAMILY PROTEIN
POLYGALACTURONASE INHIBITING PROTEIN 1PGIP1; PGIP2
ARABINOXYLAN PECTIN ARABINOGALACTAN PROTEIN 1APAP1
FRA8 HOMOLOGF8H
GALACTAN SYNTHASEGALS2; GALS3
BETA-GALACTOSIDASEBGAL10; BGAL12
GALACTURONOSYLTRANSFERASE-LIKEGATL7; GATL10
MALE GAMETOPHYTE DEFECTIVE 4MGP4
Hemicelluloserelated
ENDO-BETA-MANNASE 7MAN7
ALPHA-XYLOSIDASE 1XYL1
GLYCOSYLTRANSFERASE 18GT18
MURUS 3MUR3
AGPrelated
ARABINOGALACTAN PROTEINAGP2; AGP9
PROLINE/LEUCIN-RICH PROTEIN related
PROLINE-RICH PROTEIN 1PRP1; PRP2
PROLINE-RICH PROTEIN-LIKE 1PRPL1
LEUCINE-RICH REPEAT/EXTENSIN 2LRX2
LEUCINE-RICH REPEAT PROTEINLRR1
Peroxidase related
PEROXIDASE SUPERFAMILY PROTEIN
PEROXIDASE 7PER7; PRX25; PRX33
Secondary cell wall related
PARVUSPARVUS
TRANSPARENT TESTA 8TT8
GLABRA 2GL2
Signal Perception
FORMIN HOMOLOGYFH1; FH5
WALL ASSIOCIATED KINASEWAK
THESEUS 1THE1

Share and Cite

MDPI and ACS Style

Majda, M.; Robert, S. The Role of Auxin in Cell Wall Expansion. Int. J. Mol. Sci. 2018, 19, 951. https://doi.org/10.3390/ijms19040951

AMA Style

Majda M, Robert S. The Role of Auxin in Cell Wall Expansion. International Journal of Molecular Sciences. 2018; 19(4):951. https://doi.org/10.3390/ijms19040951

Chicago/Turabian Style

Majda, Mateusz, and Stéphanie Robert. 2018. "The Role of Auxin in Cell Wall Expansion" International Journal of Molecular Sciences 19, no. 4: 951. https://doi.org/10.3390/ijms19040951

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop