You are on page 1of 14

Journal of Experimental Botany Advance Access published November 25, 2015

Journal of Experimental Botany


doi:10.1093/jxb/erv511

REVIEW PAPER

Plant cell wall extensibility: connecting plant cell growth


with cell wall structure, mechanics, and the action of wall-
modifying enzymes
Daniel J. Cosgrove*
Department of Biology, 208 Mueller Lab, Pennsylvania State University, University Park, PA 16802, USA

Downloaded from http://jxb.oxfordjournals.org/ at Universidad Autonoma de Chiapas on July 11, 2016


*  To whom correspondence should be addressed. E-mail: dcosgrove@psu.edu

Received 18 September 2015; Revised 6 November 2015; Accepted 9 November 2015

Editor: Nadav Sorek, University of California Berkeley

Abstract
The advent of user-friendly instruments for measuring force/deflection curves of plant surfaces at high spatial resolu-
tion has resulted in a recent outpouring of reports of the ‘Young’s modulus’ of plant cell walls. The stimulus for these
mechanical measurements comes from biomechanical models of morphogenesis of meristems and other tissues, as
well as single cells, in which cell wall stress feeds back to regulate microtubule organization, auxin transport, cellu-
lose deposition, and future growth directionality. In this article I review the differences between elastic modulus and
wall extensibility in the context of cell growth. Some of the inherent complexities, assumptions, and potential pitfalls
in the interpretation of indentation force/deflection curves are discussed. Reported values of elastic moduli from
surface indentation measurements appear to be 10- to >1000-fold smaller than realistic tensile elastic moduli in the
plane of plant cell walls. Potential reasons for this disparity are discussed, but further work is needed to make sense
of the huge range in reported values. The significance of wall stress relaxation for growth is reviewed and connected
to recent advances and remaining enigmas in our concepts of how cellulose, hemicellulose, and pectins are assem-
bled to make an extensible cell wall. A comparison of the loosening action of α-expansin and Cel12A endoglucanase
is used to illustrate two different ways in which cell walls may be made more extensible and the divergent effects on
wall mechanics.

Key words:  Atomic force microscopy, Cel12A endoglucanase, cell growth, cell wall, expansin, extensibility, indentation modulus,
Young’s modulus of elasticity.

Introduction
Interest in cell wall mechanics has received a major boost Kierzkowski et  al., 2012; Robinson et  al., 2013; Bassel
from recent elegant models of plant morphogenesis in which et  al., 2014; Sampathkumar et  al., 2014b; Boudon et  al.,
mechanical stresses in growing cell walls guide a dynamic 2015; Yanagisawa et  al., 2015). These ideas, stimulated by
supracellular system that involves microtubule reorganiza- advances in fluorescent tagging of cell components and con-
tion, oriented cellulose deposition, and auxin transport. The focal imaging of living shoot apical meristems (Meyerowitz
models simulate a highly choreographed process regulating et  al., 2005; Bainbridge et  al., 2008; Hamant et  al., 2008),
the location and directionality of cell growth and cell wall have revived the concepts of biophysical control of meristem
reinforcement in the shoot apical meristem or other tis- dynamics pioneered by Green (1999) and extended by sub-
sues or single cells such as the trichome (Laufs et al., 2009; sequent advances in cell biology and genetics (Landrein and

© The Author 2015. Published by Oxford University Press on behalf of the Society for Experimental Biology. All rights reserved.
For permissions, please email: journals.permissions@oup.com
Page 2 of 14 | Cosgrove

Hamant, 2013; Braidwood et  al., 2014). Additional techni- non-growing cells, for example when cell turgor pressure is
cal advances include methods based on micro-indentation modulated or when an external force, such as from an AFM
and atomic force microscopy (AFM) to assess wall stiffness tip, suddenly impinges on a cell. As reviewed two decades
at cellular and subcellular resolution in living meristems as ago (Cosgrove, 1993), wall extensibility depends on the rate
well as in other tissues and single cells (Milani et  al., 2013; of wall-loosening processes acting on the cell wall to induce
Bonilla et  al., 2015; Braybrook, 2015; Weber et  al., 2015). stress relaxation, which creates the conditions for sustained
Mechanical measures of cell walls are most informative when cell water uptake and irreversible physical enlargement of the
interpreted in terms of the structure of cell walls, which are growing cell. A  central message of this update is that wall
comprised of relatively stiff cellulose microfibrils embedded extensibility and wall elasticity are not the same, and in some
in a hydrated matrix of pectins, hemicellulose, and structural instances are not correlated with each other. A second mes-
protein. Concepts of how these components are linked to one sage is that methods to measure non-elastic properties such as
another have changed considerably since early views of the extensibility of cell walls at cellular resolution in complex tis-
primary cell wall as a fiber-reinforced polymer composite, sues need to be developed to probe deeper into the physical,
and concepts of cell wall structure are currently in a state of cellular, and genetic control of plant morphogenesis.
flux (Cosgrove, 2014b).
Thus it is perhaps timely to review the notion of cell

Downloaded from http://jxb.oxfordjournals.org/ at Universidad Autonoma de Chiapas on July 11, 2016


wall extensibility and how it relates to cell growth, cell wall What is wall extensibility and how is it
structure, the action of wall-loosening processes, and purely
measured?
mechanical measures of wall viscoelasticity such as wall stiff-
ness. I use ‘viscoelasticity’ here rather loosely to encompass Early concepts of wall extensibility and its relationship to
the range of material properties of cell walls that include cell growth were summarized by Heyn (1940) and updated by
elastic, viscous, plastic, and retarded elastic deformations, Cleland (1971). These ideas, in which cell growth was thought
but not to include enzyme-dependent, or chemorheological, to result from yielding of a plastic wall to the forces gener-
flows. Although integration of these inter-related concepts ated by cell turgor pressure (P), pre-dated the first detailed
into a comprehensive theory of plant cell growth is still some molecular model of a growing cell wall (Keegstra et al., 1973)
way off, recent movement in this direction is refining our and did not offer a specific molecular concept of wall extensi-
appreciation of plant growth as a subtle and dynamic pro- bility other than the general notion that the physical viscosity
cess. Complementing these developments are genetic studies or plasticity of the wall matrix controlled wall extensibility.
of ‘cell wall integrity sensing’ (Wolf et  al., 2014; Hamann, Although ample evidence has accrued in the meantime that
2015; Hofte, 2015), which have identified signal transduction this purely viscoelastic concept of cell wall growth falls short
elements in this defense response, but leave unanswered ques- of reality (Taiz, 1984; Cosgrove, 1993), it seems to be the
tions such as what ‘wall integrity’ means at the structural level assumption of many recent studies that use micro-mechani-
and what is actually being sensed. This topic is also linked cal mapping of surface stiffness to infer morphogenetic con-
to a resurgence of interest in pectins and their methylation trol mechanisms for meristems and single cells.
state in wall structure and wall sensing (Peaucelle et al., 2008, A major (but often misunderstood) conceptual advance in
2011; Palin and Geitmann, 2012; Wolf and Greiner, 2012; the field was taken by Lockhart (1965) who worked out a con-
Braybrook and Peaucelle, 2013). sistent biophysical theory connecting cell water uptake and
In this review, I  use the term ‘wall extensibility’ to mean yielding of the cell wall, eventually leading to recognition of
the ability of the cell wall to increase in surface area irrevers- wall stress relaxation as the seminal biophysical event under-
ibly during growth. Unlike growing bacterial cells, where an lying cell growth (Ray et al., 1972; Cosgrove, 1985; Hamant
increase in wall surface is directly coupled to addition of new and Traas, 2010). Wall stress relaxation results from rear-
wall components, plant cells grow in surface area by a spread- rangements (yielding) of the load-bearing polymers in the cell
ing movement of cellulose microfibrils and associated matrix wall, consequently reducing cell turgor pressure and creating
components in the plane of the wall without necessary addi- the impetus for simultaneous water uptake into the cell, volu-
tion of new wall polymers (Cosgrove, 2014a). It involves a metric expansion of the cell wall chamber, and restoration of
coupled process of wall stress relaxation and water uptake, wall stress (Fig. 1A). The molecular nature of stress relaxa-
described below. In the long term, addition of new polymers tion is key to understanding how plants regulate their growth
is needed to maintain wall integrity, but wall synthesis and and will be discussed below, but a complete understanding of
surface expansion are separable processes, at least in cells with the mechanism still eludes us, largely because the stress distri-
diffuse growth patterns. In tip-growing cells such as the pol- bution and load-bearing components of the growing cell wall
len tube (Rojas et al., 2011; Sanati Nezhad et al., 2014), wall are not well understood.
synthesis and surface expansion are more tightly coupled and Wall stress relaxation in vivo has been measured by use of
thus a modified conceptual framework may be needed, which the pressure probe, psychrometer, or pressure chamber to
is beyond the scope of this review, but many of the concepts monitor the decay in water potential or turgor pressure in
discussed here also apply in modified form to tip growth. growing tissues isolated from an external supply of water, a
Wall extensibility, as used here, is characteristic of grow- condition that blocks the uptake of water needed to restore
ing cell walls and differs profoundly from elastic (=revers- turgor and wall stress following wall relaxation (Cosgrove,
ible) extensibility, which is a property of both growing and 1987a, b). Under these conditions, turgor pressure ideally
Wall extensibility and elastic modulus   |  Page 3 of 14

A (1) (2) (3)

Reduced
Turgor Water influx
wall stress
and
pressure
turgor

cell expands

wall stress wall loosening

stress relaxaon

Downloaded from http://jxb.oxfordjournals.org/ at Universidad Autonoma de Chiapas on July 11, 2016


restoring turgor and wall stress

B Time course for stress relaxaon C Growth rate v. turgor


Turgor pressure

Growth rate
Rate constant = φ εv

0
Y

Time Turgor pressure


Fig. 1.  Biophysics of plant cell growth. (A) Diagram illustrating the concept of wall stress relaxation and its connection with turgor pressure and induced
water flows. (1) In a well-hydrated non-growing cell, the cell reaches osmotic equilibrium, with wall stresses counter-balancing the outward force of turgor
pressure against the wall. (2) Growing cell walls are loosened, resulting in a reduction in cell wall stress and turgor pressure. This means that elastic
elements in the cell wall shrink as a result of the slippage or bond breakage caused by wall loosening. (3) In response to the reduced cell turgor and water
potential, water flows into the cell, elastically expanding the wall and restoring turgor and wall stress. This process is illustrated as discrete steps, but
relaxation, water influx, volume enlargement, and turgor restoration occur simultaneously. (B) Time course for stress relaxation measured, for example,
with the pressure probe in a growing tissue that is isolated from an external water supply. Turgor pressure decays gradually to a yield threshold Y with a
rate constant given by the cell wall extensibility coefficient ϕ times the cell volumetric elastic modulus εv. Based on Cosgrove (1987a). (C) Graph of growth
rate as a function of turgor, idealized in the equation growth rate=ϕ (P–Y), where ϕ is wall extensibility, P is turgor pressure. and Y is yield threshold. From
such a plot, estimates of Y and ϕ may be made.

decays to a steady value, the yield threshold, with a rate inevitable physical consequence of the facts that (i) wall stress
constant that depends on the wall extensibility coefficient is essential for stress relaxation (and ensuing water uptake)
(Fig.  1B). Hence the general notion of wall extensibility is and (ii) the stress relaxation rate is a function of the wall stress
defined in this formulation with two parameters, the yield itself. When growth is measured as % h−1, extensibility (ϕ) has
threshold and the extensibility coefficient. These parameters units of % h−1 MPa−1, while the yield threshold (Y) is usu-
are also revealed in plots of growth rate as a function of turgor ally measured in units of cell turgor pressure (MPa). Living
pressure (Fig. 1C). Actual plots of growth rate versus turgor cells may complicate this relatively simple ideal behavior by
are not so linear and display a more gradual transition at the modulating wall-loosening processes dynamically in response
yield threshold, as does creep of isolated cell walls subjected to reduced turgor, reduced growth rate, or other conditions
to a series of tensile stresses (Takahashi et  al., 2006). Thus (Green et al., 1977; Nakahori et al., 1991).
the idealized equation representing the turgor dependence of Unfortunately, these methods for measuring wall relaxation
the rate of cell wall enlargement [GR=ϕ(P–Y); see Fig.  1C do not currently lend themselves to micromechanical map-
legend] is only a rough approximation of the actual behavior ping of meristems. Kierzkowski et al. (2012) and Nakayama
of growing tissues. Turgor dependence of plant growth is an et al. (2012) took a step in this direction by measuring osmotic
Page 4 of 14 | Cosgrove

inflation and deflation of cells on the surface of the shoot api- How does wall stiffness relate to wall
cal meristem, but did not map extensibility or yield thresh- extensibility?
olds. Their approach might be extended by use of a series
of osmotic solutions to calculate growth rate versus turgor With the advent of user-friendly atomic force microscopes
for each surface cell (as in Fig. 1C) and plotting extensibil- and other micro-indentation devices coupled to microscopes,
ity coefficient and yield threshold values across the meristem reports of cell wall stiffness or modulus for shoot apical mer-
surface. istems, leaf epidermal layers, single cells, and other plant
Contrary to early ideas that wall growth and extensibility tissues have become common (Milani et  al., 2013; Routier-
arise from passive polymer motions of a viscoelastic wall, Kierzkowska and Smith, 2013). The principle of the inden-
various observations indicate that there is more to it and that tation method is deceptively simple: a cantilever or other
it depends on continual wall loosening. mechanical device is maneuvered to the surface of a cell or
tissue, and a force–deflection curve is measured as the tip of
(i)  W hen isolated cell walls are heat treated so as to inacti- the device makes contact with the surface, pushes against it,
vate wall-bound enzymes and are clamped at constant and then retracts. Probes vary from relatively sharp, pyrami-
force (i.e. in wall creep assays), they undergo only a tran- dal shapes of AFM tips as small as ~4 nm diameter (Xi
sient viscoelastic extension that decays with a half-time et al., 2015) to spherical beads or cylinders with contacts of
of a few minutes or less, in contrast to the walls of living 1–10  μm diameter (Braybrook, 2015). The deformation is

Downloaded from http://jxb.oxfordjournals.org/ at Universidad Autonoma de Chiapas on July 11, 2016


cells that extend steadily for many hours or days (Taiz, typically fast, in the millisecond range for some AFM-based
1984). The viscoelastic extension is a combination of measurements to seconds for other types of indentation pro-
elastic and retarded elastic deformations that are distinct tocols, so wall loosening during the measurement is unlikely
from wall loosening related to growth (Cosgrove, 1993; to affect the force–deflection response. There is a trade-off in
Nolte and Schopfer, 1997; Schopfer, 2006). Likewise, the speed: too fast and hydraulic flows of the bathing fluid and
stress relaxation behavior of inactivated cell walls differs within the wall may influence the deflection rate; too slow and
from that of living cells; compare Cosgrove (1987b) with water outflows from the compressed cells can influence the
Yamamoto et al. (1970). process, if the wall is pressed into the cell as in ball tonometry
(ii)  When isolated walls that are not inactivated are clamped (Cappella and Dietler, 1999; Lintilhac et al., 2000; Forouzesh
at constant force at acidic pH, they extend steadily et al., 2013; Beauzamy et al., 2015).
for many hours under the continued action of endog- While the appearance of such force–deflection curves may
enous expansins bound to the cell wall (Cleland et  al., resemble the curves generated by uniaxial stress–strain assays
1987; Cosgrove, 1989; McQueen-Mason et  al., 1992). of hypocotyl or coleoptile wall specimens (i.e. in an ‘Instron’
Extensions of 50–100% occur before the walls thin to the tester) (Cleland, 1967), such resemblance belies the many com-
point of breakage. Such breakage does not occur in walls plications inherent in micro-scale indentation of living plant
of living cells, presumably because the wall is continually surfaces. One major difference is the direction of the applied
reinforced by newly incorporated polymers. force: a tensile force in the plane of the cell wall, as used in
(iii) When respiration of living cells is inhibited by cold or a uniaxial tensile tester, is resisted by cellulose microfibrils in
metabolic inhibitors, growth ceases almost immediately the direction of the force, whereas this is not the case for an
but the viscoelastic properties of the isolated cell wall do indentation force normal to the plane of the wall for indenta-
not change [i.e. as measured by stress/strain assays (Taiz, tion assays. Some of the assumptions and complicating issues
1984)]. In a similar vein, when growth is rapidly inhibited of the indentation measurements have been discussed in the
by light (Cosgrove, 1988b) or rapidly stimulated by auxin literature (McKee et al., 2011; Milani et al., 2013; Braybrook,
(Cleland, 1984), wall viscoelastic properties change only 2015). Tip size, shape, and penetration depth can influence
slowly after the change in growth. whether the probe penetrates into the cell wall, deforms the
Thus, while growing cell walls clearly have viscoelastic prop- cell wall locally, or changes the whole cell shape (and adjacent
erties (Taiz, 1984; Schopfer, 2006) and viscoelasticity some- cells) more globally. Usually it is unclear how the deflection
times tracks changes in wall extensibility (with some delay), is partitioned into these three different aspects of wall dis-
they are distinctive properties. Ray (1987) proposed exten- placement. Extracting information about cell wall material
sibility to be the result of wall loosening in series with sub- properties (usually quantified as a modulus or its reciprocal, a
sequent viscoelastic extension, each governed by separate compliance) from indentation measurements always involves
rate constants k1 and k2, respectively (Fig. 2). The observa- fitting the force–deflection data to some kind of model and
tions cited above suggest that k2 is larger than k1. In such a making assumptions about wall structure, cell shape, and the
case, the wall deformation following sudden changes in wall influence of turgor pressure, which stretches the cell wall and
stress will be (transiently) dominated by purely viscoelastic applies a pre-stress to the load-bearing wall polymers.
extensions (k2) whereas under steady-state conditions wall Turgor pressure is an important factor to be considered
enlargement is controlled by loosening (k1). This formula- because it increases the effective stiffness of the cell, much
tion may oversimplify the behavior of growing cell walls, as air pressure in a bicycle tire stiffens the tire (stiffness
but it provides a handy conceptual framework for thinking measures the resistance to deformation). There are at least
of loosening and viscoelasticity as distinctive aspects of the three potential components to this effect: (i) non-linearity in
growing cell wall. the stress–strain curve (‘strain hardening’), which generally
Wall extensibility and elastic modulus   |  Page 5 of 14

viscoelasc
wall loosening extension
sff wall extensible wall extended, sff wall
k k
1 2
(stress relaxaon) (water uptake)

Fig. 2.  Conceptual model of Ray (1987) linking cell wall loosening and wall viscoelasticity with rate constants k1 and k2 for biochemical and viscoelastic
steps. This concept was formulated for wall extension but can be extended for parallel hydraulic steps (shown in the bottom line), with stress relaxation
originating from wall loosening and subsequent water uptake driving the passive viscoelastic extension.

increases with wall strain; (ii) resistance to indentation due 1988a; Tomos and Leigh, 1999). With an εv of 10 MPa, a
to pre-stresses in the plane of the wall (‘tightening of the cell would lose 1% of its volume for a 0.1 MPa reduction in
drum head’); and (iii) resistance stemming from the outward turgor pressure. In contrast, the bulk modulus of water (the
force of turgor pressure, when the wall is perceptibly pressed main component of plant cells) is ~2000 MPa, so an increase
into the cell (the latter is the basis for tonometry assessments in atmospheric pressure by 0.1 MPa would compress the cell
of turgor). Non-linear stress–strain behavior has long been negligibly (0.005%).

Downloaded from http://jxb.oxfordjournals.org/ at Universidad Autonoma de Chiapas on July 11, 2016


known in the field of plant wall relations, where the cell volu- A more commonly cited modulus is the Young’s modulus
metric elastic modulus generally increases as turgor pressure of elasticity (E), an engineering unit that quantifies the stress/
increases (see, for eample, Buchner et  al., 1981; Tomos and strain ratio in a uniaxially extended strip of ‘Hookean’ mate-
Leigh, 1999). Separating turgor effects from inherent wall rial, defined as material displaying linear elasticity (i.e. strain
properties can be done, but may be complicated, involving is proportional to stress and is fully and immediately revers-
osmotic treatments, finite element modeling of the cell, and ible upon reduction of stress). Growing plant cell walls—with
many assumptions for tractable mechanical analysis (Hayot their hierarchical, multilamellate, and anisotropic arrange-
et  al., 2012; Forouzesh et  al., 2013; Bonilla et  al., 2015; ment of stiff cellulose microfibrils bonded to each other and
Digiuni et al., 2015; Weber et al., 2015). For surface indenta- to hydrated matrix polymers—exhibit elastic deformation
tion measurements, the tensioning effect of turgor is likely plus complicated, non-linear, and time-dependent strains
to have a large effect on the force–deflection curves, even at that fall outside the Hookean ideal. The non-linear character
small deflections. It is for this reason that some AFM-based of plant cell walls during mechanical deformation was high-
studies have plasmolyzed the shoot apical meristem before lighted in a recent study (Digiuni et al., 2015), but it has been
measurement of wall mechanics (e.g. Peaucelle et al., 2011). recognized for many years.
While this procedure removes complications due to turgor, it Engineers generally avoid non-linear behavior by working
also alters the pattern of wall stresses across the meristem and at small strains (say, <0.1%) where non-linearities become
the mechanical state of the cell walls. inconsequential. This is in contrast to the larger strains often
It is also worth noting at this point that there are many imposed during micro-indentation studies of cell walls and
different kinds of moduli in the world of plant cell walls, and whole-tissue stress–strain assays. Non-linearity becomes a
they measure different properties. Accessible introductions to troublesome issue when modulus values are calculated from
this topic include Wainwright et al. (1976), Niklas (1992), and different parts of a stress–strain curve. The anisotropic con-
Burgert (2006). Generally speaking, a modulus is a measure struction of growing cell walls necessitates recognition of at
of the stress/strain ratio where stress is conventionally defined least three distinct elastic moduli (Fig. 3): EL for changes in
as force per unit area and strain as relative change in length, length, EW for changes in width, and ET for changes in thick-
width, thickness, volume, etc. There are many different ways ness of a cell wall strip [this is an incomplete description of
to apply a stress and to measure the concomitant strain, and the full mechanics of the wall, which is beyond the scope of
consequently there are many different kinds of moduli; they this review; see Niklas (1992) and Weber et al. (2015) for more
may all be expressed in the same units, but they measure dif- detailed treatment]. Cellulose microfibrils, which are depos-
ferent things and ought not to be mixed up. In the case of the ited in the plane of the cell wall, greatly influence EL and EW,
cell volumetric elastic modulus mentioned above, the modu- whereas pectins probably exert dominant control of ET (Ha
lus (εv) is defined as dP/dV where dP is a change in cell turgor et al., 1997). The complex structure of plants cell walls means
pressure (MPa) and dV is the relative change in cell volume that these moduli are first-order approximations of the actual
(dimensionless). This modulus measures the elastic stretch- elastic behavior of cell walls. Additionally, the propensity of
ability of the cell wall chamber and is not the bulk modu- growing cell walls to deform irreversibly in a time-dependent
lus (=1/compressibility) used in mechanics, with which it has manner under the tensile stresses generated by cell turgor
sometimes been confused (Cosgrove, 1988a). When cells lose or external forces complicates their mechanical properties
water, cell volume and turgor pressure decrease and εv quanti- even further. It is, of course, exactly this non-linear, time-
fies this relationship. In cells with very stiff walls (high modu- dependent, irreversible behavior of growing cell walls that is
lus), turgor pressure falls steeply as cells lose water. Values of greatest relevance for growth and morphogenesis. At low
of εv for non-lignified plant cells typically range from 1 MPa strains (<0.1% deformation) and rapid measurements, elastic
to 50  MPa (Zimmermann and Steudle, 1978; Cosgrove, behavior probably dominates, but at larger strains (>0.1%)
Page 6 of 14 | Cosgrove

EL (load-bearing) portion of the cell wall. Because the outer part


ET of the cell wall is the oldest region, it has undergone very large
extensions in the course of its lifetime, possibly fragmenting
EW its structure and uncoupling it from turgor-generated tensile
stresses in the plane of the wall. This ought to be the case
for the outer epidermal wall in the shoot apical meristem,
but probably not for internal cell walls within the meristem
Fig. 3.  Schematic diagram of the anisotropic construction of plant cell that are relatively young (i.e. formed during a relatively recent
walls, made of multiple lamellae (only two are shown), each with aligned cell division). This interpretation would be consistent with
cellulose (longitudinal in the top lamella, transverse the bottom lamella), the concept summarized by Taiz (1984) that the inner half
and embedded in a hydrated matrix of pectins and hemicellulose. The
three major elastic moduli (EL, EW, ET) govern the stress–strain responses
of the cell wall bears turgor-generated wall stresses and that
of walls to forces applied in the three orthogonal axes of the wall. Note that the outer half may not be important for wall enlargement.
cellulose microfibrils contribute to EL and EW, but have less influence on ET. If true, it casts doubt on the interpretation of AFM meas-
Wall indentation measurements may be most sensitive to ET. urements that mechanically probe only the outermost part of
the epidermal wall. Differences in the young and old regions
and longer time scales the non-elastic behaviors may become of the wall could lead to differential contraction and shear

Downloaded from http://jxb.oxfordjournals.org/ at Universidad Autonoma de Chiapas on July 11, 2016


significant. For an example of the issues involved, see Hansen stresses across the wall when turgor pressure is reduced, but
et al. (2011) who have reviewed the background for the time apparently this possibility has not been studied in a quantita-
dependence of viscoelastic measurements of cell walls and tive manner.
described a procedure of dynamic testing for living, but not Turgor dependence of plant cell stiffness is reflected in the
growing, potato explants. everyday experience that carrots become limp upon dehy-
Notwithstanding the foregoing complications, Young’s dration and has been documented many times: 50 years ago
elastic moduli of cell walls based on indentation assays are as measurements of whole-tissue modulus (Virgin, 1955;
commonly reported. From the force–deflection curves gener- Nilsson et al., 1958), a generation later as measurements of
ated by micro-scale indentations of plant cell surfaces, one single-cell εv (Huesken et al., 1978; Buchner et al., 1981), and
may estimate an elastic modulus by fitting the data to an more recently in micro-indentation studies (Wang et al., 2004;
appropriate model. The type of model depends on the size of Hayot et al., 2012; Forouzesh et al., 2013; Weber et al., 2015).
the tip, the depth of probe penetration into the wall, and the In the later set of studies, large-scale wall deformations were
kind of wall deformation that is envisaged. employed and the deflection data were fit to mechanical mod-
For nanometer-scale measurements made with a typi- els of whole-cell deformation. A detailed example illustrating
cal atomic force microscope using a sharp probe (~4–40 nm the modeling that is required and the assumptions involved
diameter), a common assumption is that wall deformation in such models is the analysis of tobacco cell suspension
is local, that large-scale bending of the cell wall is minimal, cells indented with a cylindrical probe of ~3 μm diameter by
and that the influence of turgor pressure on cell wall stiff- Weber et al. (2015). Two elastic moduli for these cylindrical
ness is negligible. The force–deflection curve is then fitted to cells were estimated: EL (153 MPa) for changes in cell length
a mechanical model, usually a variant of the Hertz–Sneddon and EW (571 MPa) for changes in cell circumference (the third
model (Sneddon, 1965) or the DMT model (Derjaguin et al., modulus, for cell wall thickness, was not estimated). Using
1975), to estimate an elastic modulus of the deformed mate- these elastic moduli and assuming the cells are ideal cylin-
rial (Routier-Kierzkowska and Smith, 2013; Braybrook, ders with a turgor pressure of 0.6 MPa, I  estimated εv for
2015). Complications in the analysis and interpretation of these cells to be ~15 MPa, which is within the range of values
the complex curves obtained with living cells are outlined by reported in the literature. Change in length made the largest
Bonilla et al. (2015) who describe an empirical approach to contribution to the change in volume for the εv estimate. This
distinguish different modes of wall deformation that occur calculation serves as a check that the wall moduli reported
during whole-cell indentation with small tips. in this study are consistent with the expected pressure–vol-
In a pioneering report of micro-mechanical mapping ume relations of such cells and also gives a point of compari-
of the shoot apical meristem, Milani et  al. (2011) observed son for values of EL, EW, and εv. Wang et al. (2004) provide
that reduction of cell turgor pressure by addition of 0.25 M another point of comparison for spherical tomato suspen-
NaCl had no effect on mechanical measurements, at least sion cells compressed between two surfaces. They report a
within the error of the measurement. This seems a perplex- Young’s modulus of 2.3 GPa and an estimated εv of 7 MPa.
ing result because tensioning of the wall by turgor pressure They considered the wall to be isotropic and 126 nm thick
should make the wall appear stiffer. The authors estimated and they also took into account the outflow of water from the
the wall deformation of their measurements to be ~300 nm cell when it is compressed in this way (during the 1 s compres-
in diameter. This is ~20-fold larger than the spaces between sion). A theoretical sensitivity analysis performed by Weber
cellulose microfibrils, which are believed to bear much of et al. (2015) showed that their calculations were particularly
the turgor-generated wall stress; thus I  would expect a pro- sensitive to estimates of cell radius, indentation depth, and
nounced turgor dependence of the indentation curves. The cell wall thickness, assumed to be 1250 nm, which is ~10-fold
lack of turgor dependence in their measurements suggests to thicker than estimates from other studies of suspension cells
me that the AFM tip did not probe the mechanically relevant (Samuels et al., 1995; Sabba et al., 1999; Wang et al., 2004).
Wall extensibility and elastic modulus   |  Page 7 of 14

Re-parameterization of geometrical quantities reduced some pectin-rich middle lamella can make a substantial contribu-
of these dependencies. One has to wonder whether the ~10- tion to the mechanical behavior of the multicellular epidermal
fold difference in Young’s moduli in these two studies is a surface (Zamil et al., 2014, 2015; Kim et al., 2015). Hence, the
reflection of different assumptions of the thickness of the mechanical behavior of a subcellular patch of epidermal cell
cell walls rather than differences in the material properties of wall may not fully account for the mechanics of a multicel-
the walls. lular sheet of epidermal cell wall. A related point was made
Compared with these estimates of elastic moduli in the by Chan et al. (2011) and Crowell et al. (2011) in studies of
plane of the plant cell wall, elastic moduli based on indenta- cellulose reinforcement of epidermal cell walls in Arabidopsis
tion using a small probe (10 nm to 1  μm) yield values that hypocotyls. They inferred from microscopic observations (but
are much smaller, by factors of 10- to >1000-fold (Milani not mechanical measurements) that the outer epidermal wall
et al., 2011; Peaucelle et al., 2011; Radotic et al., 2012; Ma is not the major determinant of the directionality of hypoco-
et  al., 2013; Zdunek and Kurenda, 2013; Sampathkumar tyl growth, which instead was linked to the anisotropic orien-
et al., 2014a; Wang et al., 2014; Braybrook, 2015). The rea- tation of cellulose in the side walls and inside walls. Previous
son for the wide divergence in indentation values does not work supporting this concept was summarized by Baskin
seem to have been explored, but it appears too large to be (2005). Hence, the contribution of internal cell walls to mor-
based on real differences in cell wall structure. The systematic phogenesis of other growing organs and meristems is worth

Downloaded from http://jxb.oxfordjournals.org/ at Universidad Autonoma de Chiapas on July 11, 2016


difference between in-plane tensile moduli and values from further scrutiny (Baskin and Jensen, 2013), perhaps using
indentation assays arises in part from the nature of the nano- the combination of mechanical measurements and modeling
indentation method, which applies a force that is normal to employed for shoot apical meristems.
the plane of the cell wall. In other words, they are measur-
ing different properties of an anisotropic wall (note that cell
walls, which are traditionally called isotropic because of a Extensibility, cell wall structure, and wall
random direction of cellulose in the plane of the wall, are still loosening
highly anisotropic when the third dimension, wall thickness,
is considered). The calculation of a modulus from force– A grand challenge in this field is to relate cell wall structure
deflection curves by micro-indentation typically assumes that to the mechanics of cell walls and the action of cell wall-loos-
the indented material (i) is isotropic; (ii) shows linear elastic- ening agents that induce wall stress relaxation and provoke
ity; and (iii) is >10 times thicker than the depth of indenta- water uptake needed for cell growth. For more than two dec-
tion. The construction of epidermal walls deviates from these ades the dominant concept of the growing cell wall has been
assumptions, and mechanical contributions due to cell turgor of a scaffold of well-separated cellulose microfibrils mechani-
pressure add to the difficulty. As noted by Milani et al. (2013), cally tethered together by xyloglucans and embedded in a
‘it is unclear which properties of the cell wall are measured, pliant pectin gel (Hayashi, 1989; Pauly et al., 1999; Scheller
because the force is roughly perpendicular to cellulose micro- and Ulvskov, 2010). Evidence for and against the tethered-
fibrils’. In cross-sections of dry woody cell walls, where the network concept was summarized by Scheller and Ulvskov
nano-indentation method was first applied to plant cell walls, (2010), but recent years have witnessed mounting dissatis-
experimental complications and problems of interpretation faction with the concept (Cosgrove, 2014b). It falls short in
were previously identified (Gindl and Schöberl, 2004). The view of the mild phenotype of xyloglucan-deficient mutants
mechanical situation for indentations of growing tissues pre- (Cavalier et  al., 2008; Park and Cosgrove, 2012a), nuclear
sents a different set of problems, but I  hazard a guess that magetic resonance (NMR) analysis of matrix–cellulose
the method at its best measures a type of wall compressive interactions in the wall (Dick-Perez et al., 2011; Wang et al.,
modulus, dominated by pectin properties, but with some 2015), and enzymatic probes of cell wall mechanics (Park and
(poorly defined) influence of the microfibrils at right angles Cosgrove, 2012b). In addition, recent AFM imaging of never-
to the principle force vector. The term Young’s modulus or dried primary cell walls shows that cellulose microfibrils are
even ‘apparent’ Young’s modulus for these measurements not evenly separated, as commonly depicted, but frequently
seems misleading. The term ‘indentation modulus’ (Eder make close contact with each other, forming junctions or
et al., 2013) may be a clearer way of describing these values bundles that may be important for wall mechanics (Zhang
and avoiding confusion with Young’s moduli in the plane of et al., 2014). Cellulose bundles are also seen in cellulose mats
the wall. made by Acetobacter xylinus where they may contribute to the
From the foregoing discussion, it is evident that the mod- mechanics of these cellulosic hydrogels (Whitney et al., 1999;
ulus calculated from micro-indentation measurements dif- Martinez-Sanz et  al., 2015). Thompson (2005) challenged
fers from the in-plane tensile moduli of the cell wall. This the tethered-network concept by a calculation that hydrogen
disparity may not affect some conclusions based on relative bonding between xyloglucans and cellulose was too weak
changes in stiffness across a surface, but it could have sig- to resist turgor-generated wall stresses (however, I  believe
nificant ramifications for dynamical models of plant mor- the calculation made questionable assumptions); instead, he
phogenesis that depend on use of correct values for in-plane offered a purely polymer-based view of wall extensibility as
stresses and tensile moduli of the wall. Chickarmane et  al. controlled by hydration, spacing, and entanglement of matrix
(2010) and Sampathkumar et  al. (2014a) review some of polymers. Although cell wall hydration contributes greatly to
these models. Additionally, recent studies indicate that the primary wall mechanics (Edelmann, 1995; Tang et al., 1999;
Page 8 of 14 | Cosgrove

Evered et al., 2007; Kennedy et al., 2007; White et al., 2014; The biomechanical hotspot model was tested by atomis-
Kim et  al., 2015), the concept seems unlikely to provide a tic simulation (Zhao et  al., 2014), which indicated that the
mechanism by which a cell regulates its growth. strength of cellulose–xyloglucan–cellulose junctions may
A recent alternative to the tethered-network concept pro- contribute substantially to the mechanical strength of the cell
poses that wall mechanics and wall extensibility are con- wall. This was a nanometer-scale test, but the more interest-
trolled at limited sites where cellulose microfibrils come into ing mechanical properties of the hotspot model may emerge
close contact (Park and Cosgrove, 2012b). In agreement with from simulation of larger scale behavior of the wall, much
experimental results, most of the xyloglucan is bound to cel- as the mechanical properties of a fishnet depend not just on
lulose but does not serve a load-bearing role under static the strength of the individual junctions between two ropes
conditions. This concept accounts for lack of wall loosening but on the spacing between junctions as well as the thickness
by xyloglucan-specific endoglucanase and cellulose-specific and flexibility of the ropes. A nominal attempt in this direc-
endoglucanase, alone or in combination, in contrast to the tion was a finite element model (FEM) that simulated a single
effectiveness of bifunctional endoglucanases that cut both of lamella of nearly parallel cellulose microfibrils connected by
these wall components (Park and Cosgrove, 2012b). It also limited xyloglucan linkers (Nili et al., 2015). The linkers were
accounts for the lack of appreciable loosening by xyloglu- implemented in the model as short tethers between micro-
can endotransglucosylase (Saladie et  al., 2006). The revised fibrils rather than as the monolayer of xyloglucan adhesive

Downloaded from http://jxb.oxfordjournals.org/ at Universidad Autonoma de Chiapas on July 11, 2016


model proposes limited, inaccessible junctions (dubbed bio- suggested by the enzyme results of Park and Cosgrove (2012b)
mechanical hotspots) where cellulose and xyloglucan chains and modeled by Zhao et  al. (2014). The study concluded
are tightly adherent to each other (Fig. 4). Such sites may also that relatively few xyloglucan linkers per cellulose microfi-
be the target of expansin (Wang et al., 2013). Suggestions of bril were needed to maintain mechanical stability, but that
direct cellulose–cellulose junctions in primary walls may also additional load-bearing elements were needed to maintain
be found in earlier reports (Boyd and Foster, 1975; Scheller realistic elastic moduli. This elastic weakness might be elimi-
and Ulvskov, 2010). nated by including multiple lamellae with cellulose oriented

Fig. 4.  An artistic rendition of the arrangement of cellulose, xyloglucan and pectins based upon recent results (Cosgrove, 2014b). Cellulose microfibrils
(~3 nm diameter) are shown in blue and their arrangement is traced from an AFM image of the most recently deposited surface of an onion epidermal
cell wall (Zhang et al., 2014). Xyloglucans (green) are shown as random coil structures interspersed with pectins and with limited contacts with cellulose
(based on NMR results). Pectins (yellow) are shown as two forms, a rigid form in contact with cellulose microfibrils and a more mobile form filling the
spaces between microfibrils. Potential cellulose–cellulose junctions are highlighted in red. Image adapted from Current Opinion in Plant Biology, 22C,
Cosgrove DJ. Re-constructing our models of cellulose and primary cell wall assembly, 122–131, Copyright (2014), with permission from Elsevier.
Wall extensibility and elastic modulus   |  Page 9 of 14

in different directions in the lamellae. The authors speculated vitro binding results cited above. Pectin contributed to the
that pectins might contribute to mechanical strength of the compression resistance of the cellulose–pectin composite by
wall, but pectins were not included in their simulation and limiting hydraulic flows out of the composites, whereas xylo-
likewise were missing in another FEM of a multilamellate cell glucan interactions with the cellulosic pellicles were more sta-
wall (Kha et al., 2010). One problem is that the mechanism by ble and altered the pellicle compressibility in different ways
which pectins may interact with cellulose is still unclear, an (Lopez-Sanchez et  al., 2015). Mobility-resolved 13C-NMR
issue that merits deeper study, as discussed next. indicated that the pectin molecules were highly mobile, not
tightly bound to cellulose.
Pectins These results with artificial composites are consistent with
the common view that pectins are held in the wall only weakly,
The recent re-evaluation of the role of xyloglucans in cell but are in contrast to 13C-NMR studies of native pectin in
wall models coincides with renewed attention to the role of plant primary cell walls, which show abundant pectin–cellu-
pectins (Palin and Geitmann, 2012; Peaucelle et  al., 2012; lose cross-peaks in 2D cross-polarization measurements, indi-
Wang et  al., 2012; Wolf and Greiner, 2012; Braybrook and cating extensive, stable pectin–cellulose interactions (Wang
Peaucelle, 2013). There are diverse views on the structure et al., 2012, 2015). As judged by solid-state NMR, cellulose
and the mechanical role of pectins in growing cell walls. The contacts with pectins were more abundant than contacts with

Downloaded from http://jxb.oxfordjournals.org/ at Universidad Autonoma de Chiapas on July 11, 2016


dominant and traditional view imagines that they form a xyloglucan. Moreover, although pectins are the most mobile
soft, gel-like, viscous matrix in which the cellulose–hemicel- polymers in the primary cell wall, nearly half of the pectin
lulose network is embedded, but otherwise not binding cel- was in a rigid form, with a 13C T1 relaxation time similar to
lulose (Carpita and Gibeaut, 1993; Cosgrove, 2005). In this that of cellulose. These results, supported by previous obser-
view, pectins may contribute to cell wall stiffness during vations (e.g. Ryden and Selvendran, 1990; Foster et al., 1996),
rapid deformations, but their viscosity would cause pectin- were interpreted to mean that a substantial pectin compo-
borne stresses to decay rapidly, displacing forces onto more nent in native cell walls is stably bound to cellulose surfaces
rigid components of the cell wall. If the gel were to become or entrapped within cellulose bundles. However, the molecu-
sufficiently rigid, for example by calcium-cross-linking of lar nature of the cellulose–pectin interaction still is not clear
unesterified regions of homogalacturonan, it could in theory and needs further investigation. Moreover, these results do
lock the wall into an inextensible state; however, it is unclear not establish whether or not pectins have a significant static
whether this actually occurs. Pectin de-esterification is likely load-bearing role in cell wall mechanics.
to affect many physico-chemical properties of the wall. In biomechanical assays (creep assays), wild-type
A different view, first proposed in the pioneering molecular Arabidopsis cell walls showed a relatively modest loosen-
model of the cell wall by the Albersheim group (Keegstra ing effect when treated with pectin-degrading enzymes, with
et al., 1973; Albersheim, 1975), considers pectins to be part a somewhat larger effect in a xyloglucan-deficient mutant
of a massive macromolecular matrix that contains covalently (Park and Cosgrove, 2012a). The results indicated that pec-
linked domains of pectin, hemicellulose, and protein and that tins do not bear static loads in wild-type cell walls, but may
extensively binds to cellulose surfaces through xyloglucan take on a larger mechanical role in the absence of xyloglucan.
domains. Although this model was largely abandoned in the One limitation of this enzymatic approach is that if pectins
1980s in favor of the tethered network concept (Fry, 1989; are entrapped among microfibril bundles, they might not be
Hayashi, 1989; McCann et al., 1990), linkages between pec- accessible to enzymatic digestion, potentially missing a cryp-
tin and xyloglucan or pectin and proteins have been reported tic load-bearing component. Moreover, the essential role of
recently (e.g. Popper and Fry, 2008; Tan et al., 2013). A recent pectins in cell–cell adhesion (i.e. the middle lamella) limits
review (Park and Cosgrove, 2015) concludes that the abun- the extent of pectin weakening treatments feasible for these
dance of pectin–xyloglucan linkages is low in most plant tis- creep experiments where the sample is held under constant
sues, with the notable exception of cell suspension cultures. tensile load (multicellular wall samples fall apart when pec-
Low abundance, however, does not necessarily mean they are tins are significantly degraded). This technical limitation may
insignificant for wall mechanics. This point remains an open be circumvented by micro-indentation assays with isolated
question. cell walls. In such assays with onion epidermal walls, Xi et al.
Unlike xyloglucans, which bind tightly and irreversibly to (2015) used a sharp pyramidal probe and a Hertz model to
cellulose (Hayashi, 1989), pectins bind to cellulose relatively calculate an elastic modulus from the force–deflection curves.
weakly or not at all, at least during binding assays in vitro Removal of calcium with a chelator reduced wall stiffness,
(Zykwinska et  al., 2005, 2008). Lin et  al. (2015b) reported decreasing the calculated indentation modulus from 23 MPa
weak, reversible pectin binding to bacterial cellulose pellicles, to 16 MPa, presumably as a result of reduced ionic cross-link-
also called hydrogels because they are made of a loose, ran- ing of homogalacturonan within the wall (the middle lamella
dom entanglement of cellulose fibrils containing ~99% water. was not probed). This result confirms that pectins contribute
The pectin interactions with cellulose were weak and included to the compression resistance of cell walls to fast deforma-
components of mere inclusion and entrapment within the tions. This still leaves open the question of their contribution
large free space of the entangled cellulosic mat. Neutral pec- to static tensile loads and to slow creep in the plane of grow-
tins (arabinan, galactan) showed stronger interactions than ing cell walls. For comparison, the tensile (Young’s) modulus
homogalacturonan (Lin et al., 2015a), consistent with the in of comparable wall material was estimated as 400 MPa (Kim
Page 10 of 14 | Cosgrove

et al., 2015)—a clear example of the difference between the do not know of plant enzymes with similar wall-loosening
indentation modulus and tensile modulus of the wall. action. Characterizations of two family-9 endoglucanases
In an elegant study of fast and reversible polymer motions from plants indicate that they have a wide substrate specificity
in onion epidermal walls stretched in the plane of the wall, (Ohmiya et al., 1995; Yoshida and Komae, 2006), suggesting
Wilson et al. (2000) combined dynamic mechanical analysis a wall-loosening action similar to Cel12A, but such action
with polarized infrared (IR) spectroscopy. They concluded has not actually been documented. In support of this notion,
that homogalacturonan motions were uncoupled in time ectopic expression of one of these enzymes in Arabidopsis
from cellulose and xyloglucan motions. To explain this tem- promoted plant growth (Park et al., 2003), and transcript pro-
poral shift, they suggested that homogalacturonan motions filing identified several family-9 genes that are expressed in
were dominated by strains in the middle lamella which were growing tissues of grasses (Buchanan et al., 2012).
subsequently transmitted to cellulose in the cell wall proper. Cel12A loosens plant cell walls by its hydrolytic activity,
A subcellular version of this experiment that eliminated the whereas no enzymatic activity has been found for α-expansin.
pectin signal from the middle lamella would be informative Both proteins induce creep of isolated, heat-inactivated cell
for the question of pectins as load-bearing elements of the walls from cucumber hypocotyls and other plant materials,
wall proper, particularly if time constants for decays in wall but with very different kinetics: α-expansin induces creep
stress could be assessed for the different structural compo- within seconds after application whereas Cel12A induced

Downloaded from http://jxb.oxfordjournals.org/ at Universidad Autonoma de Chiapas on July 11, 2016


nents of the primary cell wall. From dynamic mechanical creep only after a substantial lag time that depended on
measurements of (living) potato explants in combination with enzyme concentration: a minimum of 6 min at saturating
a mathematic model, Ulvskov et al. (2005) inferred that pec- enzyme concentration, >60 min for low enzyme concentra-
tins can transmit dynamic, but not static, stresses to cellulose. tions (Yuan et al., 2001). Cel12A reduced the elastic and plas-
Turning to purely computational approaches, Dyson et al. tic moduli of cucumber walls in uniaxial tensile tests, whereas
(2012) developed a noteworthy mathematical model of cell α-expansin did not change these moduli. Note that the plas-
wall yielding based on continuum mechanics. Their model tic modulus in these assays measured the irreversible compo-
considers hemicellulose and pectin to be acting as co-load- nent of uniaxial extension, which Nolte and Schopfer (1997)
bearing elements, modeled with a mathematical function giv- have re-interpreted to be the result of delayed viscoelasticity
ing strongly non-linear stress/strain rate behavior resembling rather than true plasticity. Under this interpretation, Cel12A
a yield threshold. The yield threshold was simulated by cross- increases the viscoelastic hysteresis in the stress–strain curves,
links that stretch up to a limit, then break suddenly, with a whereas α-expansin does not. By either interpretation, the
role for pectins in the post-breakage yielding. The model results suggest that α-expansin can cause stress relaxation
mimics a number of the behaviors of growing cell walls; by and creep by a mechanism (sliding at cellulose junctions?)
its nature it cannot offer insights into the structural or bio- that does not perceptibly alter cell wall structure, whereas
chemical bases for wall stress relaxation, but it does present the cutting action of Cel12A substantially reduces the num-
an attractive approach for further efforts to relate structural ber or strength of cellulose–cellulose contacts, resulting in a
models of cell walls to their mechanical behaviors beyond mechanically weaker cell wall. Attempts to simulate cell wall
simple elastic moduli. creep by cutting cross-links (e.g. Dyson et al., 2012) may not
properly mimic expansin’s loosening activity, so a refined
Wall loosening model may be needed.
The loosening action of α-expansin bears some resem-
As a final point, it is instructive to compare the wall-loos- blance to the ‘stick–slip’ or ‘molecular Velcro’ model of
ening actions of α-expansin and Cel12A, a family-12 endo- plastic deformation of wet wood (Keckes et  al., 2003;
glucanase that cuts both cellulose and xyloglucan (Table 1). Cosgrove and Jarvis, 2012; Adler and Buehler, 2013) in
These are the only two classes of proteins known to induce which hydrogen-bonded contact regions between cellu-
stress relaxation and creep of plant cell walls, although lose layers slip and then stick again, restoring mechanical
it should be noted that Cel12A is a fungal enzyme and we strength until local stresses exceed the yield threshold and
another local slip event occurs. In woody tissues this is a
purely physical process, whereas in growing cells the selec-
Table 1.  Comparison of the characteristics of cell wall loosening tive slippage, relaxation, and creep of primary walls are
by α-expansins and Cel12A, based on their activity with cucumber mediated by expansins (and potentially other agents). The
hypocotyl walls, based on published results (Yuan et al., 2001; fact that α-expansin can loosen cell walls without reducing
Park and Cosgrove, 2012b) the plastic or elastic moduli is a likely explanation for the
fact that wall extensibility and wall viscoelasticity are often
uncorrelated, as discussed above. Another example of such
Property α-Expansin Cel12A
a discrepancy is a micro-indentation study of plasticity
Hydrolytic action No Yes along the growing root of Arabidopsis thaliana (Fernandes
Increased plasticity No Yes et  al., 2012). Plasticity was measured as hysteresis in the
Increase elasticity No Yes approach and retraction phases of the force–deflection
Induces creep Yes Yes
curves. This is a micro-scale version of the hysteresis seen in
Lag time for induction of creep Seconds 6 min to >60 min
uniaxial tests of cell walls, which is the basis for estimating
Wall extensibility and elastic modulus   |  Page 11 of 14

elastic and plastic moduli (Cleland, 1967, 1971). In the Bonilla MR, Stokes JR, Gidley MJ, Yakubov GE. 2015. Interpreting
atomic force microscopy nanoindentation of hierarchical biological
study by Fernandes et al. (2012), no correlation was found materials using multi-regime analysis. Soft Matter 11, 1281–1292.
between the ‘index of plasticity’ on the living root surface Boudon F, Chopard J, Ali O, Gilles B, Hamant O, Boudaoud A,
and the growth rate along the root axis, despite the fact Traas J, Godin C. 2015. A computational framework for 3D mechanical
that cell growth and wall extensibility vary greatly along the modeling of plant morphogenesis with cellular resolution. PLoS
Computational Biology 11, e1003950.
axis (Pritchard, 1994). It might be useful to try this micro
method with wall samples known to vary in wall plastic- Boyd JD, Foster RC. 1975. Microfibirls in primary and secondary wall
growth develop trellis configurations. Canadian Journal of Botany/Revue
ity (or delayed viscoelastic hysteresis), for example after Canadienne de Botanique 53, 2687–2701.
Cel12A treatment, to validate that the method can indeed Braidwood L, Breuer C, Sugimoto K. 2014. My body is a cage:
detect useful plasticity values for cell walls. mechanisms and modulation of plant cell growth. New Phytologist 201,
388–402.
Braybrook SA. 2015. Measuring the elasticity of plant cells with atomic
force microscopy. Methods in Cell Biology 125, 237–254.
Concluding remarks Braybrook SA, Peaucelle A. 2013. Mechano-chemical aspects of organ
There is a growing recognition that stresses and strains in the formation in Arabidopsis thaliana: the relationship between auxin and
pectin. PLos One 8, e57813.
growing cell wall feed back to cellular systems to regulate
Buchanan M, Burton RA, Dhugga KS, Rafalski AJ, Tingey SV,
the cytoskeleton, vesicular trafficking of auxin transporters,

Downloaded from http://jxb.oxfordjournals.org/ at Universidad Autonoma de Chiapas on July 11, 2016


Shirley NJ, Fincher GB. 2012. Endo-(1,4)-beta-glucanase gene families
and the cellular machinery involved in cellulose deposition. in the grasses: temporal and spatial co-transcription of orthologous genes.
This provides the impetus for insightful dynamical models BMC Plant Biology 12, 235.
connecting wall mechanics with microtubule organization Buchner KH, Zimmermann U, Bentrup FW. 1981. Turgor pressure and
water transport properties of suspension-cultured cells of Chenopodium
and studies to identify the molecular connections of these rubrum L. Planta 151, 95–102.
elements of the model. Careful thought needs to be given Burgert I. 2006. Exploring the micromechanical design of plant cell walls.
to the meaning of wall moduli based on surface indentation American Journal of Botany 93, 1391–1401.
assays. They differ from the in-plane tensile elastic modu- Cappella B, Dietler G. 1999. Force–distance curves by atomic force
lus, which is the parameter of greatest value for assessing microscopy. Surface Science Reports 34, 1–104.
the stress/strain properties of the cell wall. A deeper under- Carpita NC, Gibeaut DM. 1993. Structural models of primary cell walls
in flowering plants: consistency of molecular structure with the physical
standing of wall deformations—both reversible and irre- properties of the walls during growth. The Plant Journal 3, 1–30.
versible—will require more extensive experimental testing in Cavalier DM, Lerouxel O, Neumetzler L, et al. 2008. Disrupting two
combination with quantitative models of how the structural Arabidopsis thaliana xylosyltransferase genes results in plants deficient
components of the cell wall are linked to one another and in xyloglucan, a major primary cell wall component. The Plant Cell 20,
1519–1537.
what kinds of polymer motions occur during rapid, revers-
Chan J, Eder M, Crowell EF, Hampson J, Calder G, Lloyd C.
ible deformations versus the slow irreversible enlargement 2011. Microtubules and CESA tracks at the inner epidermal wall align
of the growing wall. independently of those on the outer wall of light-grown Arabidopsis
hypocotyls. Journal of Cell Science 124, 1088–1094.
Chickarmane V, Roeder AHK, Tarr PT, Cunha A, Tobin C, Meyerowitz
Acknowledgements EM. 2010. Computational morphodynamics: a modeling framework to
understand plant growth. Annual Review of Plant Biology 61, 65–87.
Preparation of this review was supported by the US Department of Cleland R. 1967. Extensibility of isolated cell walls: measurement and
Energy, Office of Science, Basic Energy Sciences as part of The Center for changes during cell elongation. Planta 74, 197–209.
LignoCellulose Structure and Formation, an Energy Frontier Research
Center under Award # DE-SC0001090. Cleland R. 1971. Cell wall extension. Annual Review of Plant Physiology
22, 197–222.
Cleland RE. 1984. The Instron technique as a measure of immediate-past
wall extensibility. Planta 160, 514–520.
References Cleland RE, Cosgrove DJ, Tepfer M. 1987. Long-term acid-induced
Adler DC, Buehler MJ. 2013. Mesoscale mechanics of wood cell walls wall extension in an in vitro system. Planta 170, 379–385.
under axial strain. Soft Matter 9, 7138–7144. Cosgrove DJ. 1985. Cell wall yield properties of growing tissues.
Albersheim P. 1975. The walls of growing plant cells. Scientific American Evaluation by in vivo stress relaxation. Plant Physiology 78, 347–356.
232, 80–95. Cosgrove DJ. 1987a. Wall relaxation and the driving forces for cell
Bainbridge K, Guyomarc’h S, Bayer E, Swarup R, Bennett M, expansive growth. Plant Physiology 84, 561–564.
Mandel T, Kuhlemeier C. 2008. Auxin influx carriers stabilize phyllotactic Cosgrove DJ. 1987b. Wall relaxation in growing stems: comparison of
patterning. Genes and Development 22, 810–823. four species and assessment of measurement techniques. Planta 171,
Baskin TI. 2005. Anisotropic expansion of the plant cell wall. Annual 266–278.
Review of Cell and Developmental Biology 21, 203–222. Cosgrove DJ. 1988a. In defense of the cell volumetric elastic modulus.
Baskin TI, Jensen OE. 2013. On the role of stress anisotropy in the Plant, Cell and Environment 11, 67–69.
growth of stems. Journal of Experimental Botany 64, 4697–4707. Cosgrove DJ. 1988b. Mechanism of rapid suppression of cell expansion
Bassel GW, Stamm P, Mosca G, Barbier de Reuille P, Gibbs DJ, in cucumber hypocotyls after blue-light irradiation. Planta 176, 109–116.
Winter R, Janka A, Holdsworth MJ, Smith RS. 2014. Mechanical Cosgrove DJ. 1989. Characterization of long-term extension of isolated
constraints imposed by 3D cellular geometry and arrangement modulate cell walls from growing cucumber hypocotyls. Planta 177, 121–130.
growth patterns in the Arabidopsis embryo. Proceedings of the National
Acadmey of Sciences, USA 111, 8685–8690. Cosgrove DJ. 1993. Wall extensibility: its nature, measurement and
relationship to plant cell growth. New Phytologist 124, 1–23.
Beauzamy L, Derr J, Boudaoud A. 2015. Quantifying hydrostatic
pressure in plant cells by using indentation with an atomic force Cosgrove DJ. 2005. Growth of the plant cell wall. Nature Reviews:
microscope. Biophysical Journal 108, 2448–2456. Molecular Cell Biology 6, 850–861.
Page 12 of 14 | Cosgrove
Cosgrove DJ. 2014a. Plant cell growth and elongation. eLS DOI: Hayot CM, Forouzesh E, Goel A, Avramova Z, Turner JA. 2012.
10.1002/9780470015902.a0001688. Viscoelastic properties of cell walls of single living plant cells determined by
Cosgrove DJ. 2014b. Re-constructing our models of cellulose and dynamic nanoindentation. Journal of Experimental Botany 63, 2525–2540.
primary cell wall assembly. Current Opinion in Plant Biology 22C, 122–131. Heyn ANJ. 1940. The physiology of cell elongation. Botanical Review 6,
Cosgrove DJ, Jarvis MC. 2012. Comparative structure and 515–574.
biomechanics of plant primary and secondary cell walls. Frontiers in Plant Hofte H. 2015. The yin and yang of cell wall integrity control:
Science 3, 204. brassinosteroid and FERONIA signaling. Plant and Cell Physiology 56,
Crowell EF, Timpano H, Desprez T, Franssen-Verheijen T, Emons 224–231.
AM, Hofte H, Vernhettes S. 2011. Differential regulation of cellulose Huesken D, Steudle E, Zimmermann U. 1978. Pressure probe
orientation at the inner and outer face of epidermal cells in the Arabidopsis technique for measuring water relations of cells in higher plants. Plant
hypocotyl. The Plant Cell 23, 2592–2605. Physiology 61, 158–163.
Derjaguin BV, Muller VM, Toporov YP. 1975. Effect of contact Keckes J, Burgert I, Fruhmann K, Muller M, Kolln K, Hamilton M,
deformations on adhesion of particles. Journal of Colloid and Interface Burghammer M, Roth SV, Stanzl-Tschegg S, Fratzl P. 2003. Cell-
Science 53, 314–326. wall recovery after irreversible deformation of wood. Nature Materials 2,
Dick-Perez M, Zhang Y, Hayes J, Salazar A, Zabotina OA, Hong 810–814.
M. 2011. Structure and interactions of plant cell-wall polysaccharides Keegstra K, Talmadge KW, Bauer WD, Albersheim P. 1973. The
by two- and three-dimensional magic-angle-spinning solid-state NMR. structure of plant cell walls. III. A model of the walls of suspension-cultured
Biochemistry 50, 989–1000. sycamore cells based on the interconnections of the macromolecular
Digiuni S, Berne-Dedieu A, Martinez-Torres C, Szecsi J, components. Plant Physiology 51, 188–196.
Bendahmane M, Arneodo A, Argoul F. 2015. Single cell wall nonlinear Kennedy CJ, Sturcova A, Jarvis MC, Wess TJ. 2007. Hydration

Downloaded from http://jxb.oxfordjournals.org/ at Universidad Autonoma de Chiapas on July 11, 2016


mechanics revealed by a multiscale analysis of AFM force–indentation effects on spacing of primary-wall cellulose microfibrils: a small angle X-ray
curves. Biophysical Journal 108, 2235–2248. scattering study. Cellulose 14, 401–408.
Dyson RJ, Band LR, Jensen OE. 2012. A model of crosslink kinetics in Kha H, Tuble SC, Kalyanasundaram S, Williamson RE. 2010.
the expanding plant cell wall: yield stress and enzyme action. Journal of WallGen, software to construct layered cellulose–hemicellulose networks
Theoretical Biology 307C, 125–136. and predict their small deformation mechanics. Plant Physiology 152,
Edelmann HG. 1995. Water potential modulates extensibility of rye 774–786.
coleoptile cell walls. Botanica Acta 108, 374–380. Kierzkowski D, Nakayama N, Routier-Kierzkowska AL, Weber A,
Eder M, Arnould O, Dunlop JWC, Hornatowska J, Salmen L. 2013. Bayer E, Schorderet M, Reinhardt D, Kuhlemeier C, Smith RS. 2012.
Experimental micromechanical characterisation of wood cell walls. Wood Elastic domains regulate growth and organogenesis in the plant shoot
Science and Technology 47, 163–182. apical meristem. Science 335, 1096–1099.
Evered C, Majevadia B, Thompson DS. 2007. Cell wall water content Kim K, Yi H, Zamil MS, Haque MA, Puri VM. 2015. Multiscale stress–
has a direct effect on extensibility in growing hypocotyls of sunflower strain characterization of onion outer epidermal tissue in wet and dry
(Helianthus annuus L.). Journal of Experimental Botany 58, 3361–3371. states. American Journal of Botany 102, 12–20.
Fernandes AN, Chen X, Scotchford CA, Walker J, Wells DM, Landrein B, Hamant O. 2013. How mechanical stress controls
Roberts CJ, Everitt NM. 2012. Mechanical properties of epidermal cells microtubule behavior and morphogenesis in plants: history, experiments
of whole living roots of Arabidopsis thaliana: an atomic force microscopy and revisited theories. The Plant Journal 75, 324–338.
study. Physical Review, E 85, 021916. Laufs P, Peaucelle A, Hofte H. 2009. Shaping the meristem by
Forouzesh E, Goel A, Mackenzie SA, Turner JA. 2013. In vivo mechanical forces. F1000 Biology Reports 1, 45.
extraction of Arabidopsis cell turgor pressure using nanoindentation in Lin D, Lopez-Sanchez P, Gidley MJ. 2015a. Binding of arabinan
conjunction with finite element modeling. The Plant Journal 73, 509–520. or galactan during cellulose synthesis is extensive and reversible.
Foster TJ, Ablett S, McCann MC, Gidley MJ. 1996. Mobility-resolved Carbohydrate Polymers 126, 108–121.
C-13-NMR spectroscopy of primary plant cell walls. Biopolymers 39, Lin D, Lopez-Sanchez P, Gidley MJ. 2015b. Interactions of pectins
51–66. with cellulose during its synthesis in the absence of calcium. Food
Fry SC. 1989. Cellulases, hemicelluloses and auxin-stimulated growth—a Hydrocolloids 52, 57–68.
possible relationship. Physiologia Plantarum 75, 532–536. Lintilhac PM, Wei C, Tanguay JJ, Outwater JO. 2000. Ball tonometry:
Gindl W, Schöberl T. 2004. The significance of the elastic modulus a rapid, nondestructive method for measuring cell turgor pressure in thin-
of wood cell walls obtained from nanoindentation measurements. walled plant cells. Journal of Plant Growth Regulation 19, 90–97.
Composites Part A: Applied Science and Manufacturing 35, 1345–1349. Lockhart JA. 1965. An analysis of irreversible plant cell elongation.
Green PB. 1999. Expression of pattern in plants: combining molecular Journal of Theoretical Biology 8, 264–275.
and calculus-based biophysical paradigms. American Journal of Botany Lopez-Sanchez P, Cersosimo J, Wang D, Flanagan B, Stokes JR,
86, 1059–1076. Gidley MJ. 2015. Poroelastic mechanical effects of hemicelluloses on
Green PB, Bauer K, Cummins WR. 1977. Biophysical model for plant cellulosic hydrogels under compression. PLos One 10, e0122132.
cell growth: auxin effects. In: Jungreis AM, Hodges TK, Kleinzeller A, Ma N, Wang Y, Qiu S, Kang Z, Che S, Wang G, Huang J. 2013.
Schultz SG, eds. Water relations in membrane transport in plants . New Overexpression of OsEXPA8, a root-specific gene, improves rice growth and
York: Academic Press, 30–45. root system architecture by facilitating cell extension. PloS One 8, e75997.
Ha MA, Apperley DC, Jarvis MC. 1997. Molecular rigidity in dry and Martinez-Sanz M, Lopez-Sanchez P, Gidley MJ, Gilbert EP. 2015.
hydrated onion cell walls. Plant Physiology 115, 593–598. Evidence for differential interaction mechanism of plant cell wall matrix
Hamann T. 2015. The plant cell wall integrity maintenance mechanism— polysaccharides in hierarchically-structured bacterial cellulose. Cellulose
concepts for organization and mode of action. Plant and Cell Physiology 22, 1541–1563.
56, 215–223. McCann MC, Wells B, Roberts K. 1990. Direct visualization of cross-
Hamant O, Heisler MG, Jonsson H, et al. 2008. Developmental links in the primary plant cell wall. Journal of Cell Science 96, 323–334.
patterning by mechanical signals in Arabidopsis. Science 322, 1650–1655. McKee CT, Last JA, Russell P, Murphy CJ. 2011. Indentation versus
Hamant O, Traas J. 2010. The mechanics behind plant development. tensile measurements of Young’s modulus for soft biological tissues.
New Phytologist 185, 369–385. Tissue Engineering. Part B, Reviews 17, 155–164.
Hansen SL, Ray PM, Karlsson AO, Jorgensen B, Borkhardt B, McQueen-Mason S, Durachko DM, Cosgrove DJ. 1992. Two
Petersen BL, Ulvskov P. 2011. Mechanical properties of plant cell walls endogenous proteins that induce cell wall extension in plants. The Plant
probed by relaxation spectra. Plant Physiology 155, 246–258. Cell 4, 1425–1433.
Hayashi T. 1989. Xyloglucans in the primary cell wall. Annual Review of Meyerowitz EM, Heisler M, Reddy GV, Gordon S, Jonsson H,
Plant Physiology and Plant Molecular Biology 40, 139–168. Mjolsness E. 2005. Dynamic imaging and computational modelling of
Wall extensibility and elastic modulus   |  Page 13 of 14
the Arabidopsis shoot apical meristem. Mechanisms of Development 122, Ray PM, Green PB, Cleland RE. 1972. Role of turgor in plant cell
S2–S2. growth. Nature 239, 163–164.
Milani P, Braybrook SA, Boudaoud A. 2013. Shrinking the hammer: Robinson S, Burian A, Couturier E, Landrein B, Louveaux M,
micromechanical approaches to morphogenesis. Journal of Experimental Neumann ED, Peaucelle A, Weber A, Nakayama N. 2013. Mechanical
Botany 64, 4651–4662. control of morphogenesis at the shoot apex. Journal of Experimental
Milani P, Gholamirad M, Traas J, Arneodo A, Boudaoud A, Argoul F, Botany 64, 4729–4744.
Hamant O. 2011. In vivo analysis of local wall stiffness at the shoot apical Rojas ER, Hotton S, Dumais J. 2011. Chemically mediated mechanical
meristem in Arabidopsis using atomic force microscopy. The Plant Journal expansion of the pollen tube cell wall. Biophysical Journal 101,
67, 1116–1123. 1844–1853.
Nakahori K, Katou K, Okamoto H. 1991. Auxin changes both the Routier-Kierzkowska AL, Smith RS. 2013. Measuring the mechanics of
extensibility and the yield threshold of the cell wall of Vigna hypocotyls. morphogenesis. Current Opinion in Plant Biology 16, 25–32.
Plant and Cell Physiology 32, 121–129.
Ryden P, Selvendran RR. 1990. Structural features of cell-wall
Nakayama N, Smith RS, Mandel T, Robinson S, Kimura S, polysaccharides of potato (Solanum tuberosum). Carbohydrate Research
Boudaoud A, Kuhlemeier C. 2012. Mechanical regulation of auxin- 195, 257–272.
mediated growth. Current Biology 22, 1468–1476.
Sabba RP, Durso NA, Vaughn KC. 1999. Structural and
Niklas KJ. 1992. Plant biomechanics. An engineering approach to plant immunocytochemical characterization of the walls of dichlobenil-
form and function . Chicago: University of Chicago Press. habituated BY-2 tobacco cells. International Journal of Plant Sciences
Nili A, Yi H, Crespi VH, Puri VM. 2015. Examination of biological hotspot 160, 275–290.
hypothesis of primary cell wall using a computational cell wall network Saladie M, Rose JK, Cosgrove DJ, Catala C. 2006. Characterization
model. Cellulose 22, 1027–1038.

Downloaded from http://jxb.oxfordjournals.org/ at Universidad Autonoma de Chiapas on July 11, 2016


of a new xyloglucan endotransglucosylase/hydrolase (XTH) from ripening
Nilsson SB, Hertz CH, Falk S. 1958. On the relation between turgor tomato fruit and implications for the diverse modes of enzymic action. The
pressure and tissue rigidity. 2. Theoretical calculations on model systems. Plant Journal 47, 282–295.
Physiologia Plantarum 11, 818–837. Sampathkumar A, Krupinski P, Wightman R, Milani P, Berquand
Nolte T, Schopfer P. 1997. Viscoelastic versus plastic cell wall A, Boudaoud A, Hamant O, Jonsson H, Meyerowitz EM. 2014a.
extensibility in growing seedling organs: a contribution to avoid some Subcellular and supracellular mechanical stress prescribes cytoskeleton
misconceptions. Journal of Experimental Botany 48, 2103–2107. behavior in Arabidopsis cotyledon pavement cells. Elife 3, e01967.
Ohmiya Y, Takeda T, Nakamura S, Sakai F, Hayashi T. 1995. Sampathkumar A, Yan A, Krupinski P, Meyerowitz EM. 2014b.
Purification and properties of wall-bound endo-1, 4-beta-glucanase from Physical forces regulate plant development and morphogenesis. Current
suspension-cultured poplar cells. Plant and Cell Physiology 36, 607–614. Biology 24, R475–R483.
Palin R, Geitmann A. 2012. The role of pectin in plant morphogenesis. Samuels AL, Giddings TH Jr, Staehelin LA. 1995. Cytokinesis in
BioSystems 109, 397–402. tobacco BY-2 and root tip cells: a new model of cell plate formation in
Park YB, Cosgrove DJ. 2012a. Changes in cell wall biomechanical higher plants. Journal of Cell Biology 130, 1345–1357.
properties in the xyloglucan-deficient xxt1/xxt2 mutant of Arabidopsis. Sanati Nezhad A, Packirisamy M, Geitmann A. 2014. Dynamic, high
Plant Physiology 158, 465–475. precision targeting of growth modulating agents is able to trigger pollen
Park YB, Cosgrove DJ. 2012b. A revised architecture of primary cell tube growth reorientation. The Plant Journal 80, 185–195.
walls based on biomechanical changes induced by substrate-specific Scheller HV, Ulvskov P. 2010. Hemicelluloses. Annual Review of Plant
endoglucanases. Plant Physiology 158, 1933–1943. Biology 61, 263–289.
Park YB, Cosgrove DJ. 2015. Xyloglucan and its interactions with Schopfer P. 2006. Biomechanics of plant growth. American Journal of
other components of the growing cell wall. Plant and Cell Physiology 56, Botany 93, 1415–1425.
180–194.
Sneddon IN. 1965. The relation between load and penetration in
Park YW, Tominaga R, Sugiyama J, Furuta Y, Tanimoto E, Samejima the axisymmetric boussinesq problem for a punch of arbitrary profile.
M, Sakai F, Hayashi T. 2003. Enhancement of growth by expression of International Journal of Engineering Science 3, 47–57.
poplar cellulase in Arabidopsis thaliana. The Plant Journal 33, 1099–1106.
Taiz L. 1984. Plant cell expansion—regulation of cell-wall mechanical
Pauly M, Albersheim P, Darvill A, York WS. 1999. Molecular domains properties. Annual Review of Plant Physiology and Plant Molecular Biology
of the cellulose/xyloglucan network in the cell walls of higher plants. The 35, 585–657.
Plant Journal 20, 629–639.
Takahashi K, Hirata S, Kido N, Katou K. 2006. Wall-yielding properties
Peaucelle A, Braybrook S, Hofte H. 2012. Cell wall mechanics and of cell walls from elongating cucumber hypocotyls in relation to the action
growth control in plants: the role of pectins revisited. Frontiers in Plant of expansin. Plant and Cell Physiology 47, 1520–1529.
Science 3, 121.
Tan L, Eberhard S, Pattathil S, et al. 2013. An arabidopsis cell wall
Peaucelle A, Braybrook SA, Le Guillou L, Bron E, Kuhlemeier C, proteoglycan consists of pectin and arabinoxylan covalently linked to an
Hofte H. 2011. Pectin-induced changes in cell wall mechanics underlie arabinogalactan protein. The Plant Cell 25, 270–287.
organ initiation in Arabidopsis. Current Biology 21, 1720–1726.
Tang HR, Belton PS, Ng A, Ryden P. 1999. C-13 MAS NMR studies
Peaucelle A, Louvet R, Johansen JN, Hofte H, Laufs P, Pelloux of the effects of hydration on the cell walls of potatoes and Chinese
J, Mouille G. 2008. Arabidopsis phyllotaxis is controlled by the water chestnuts. Journal of Agricultural and Food Chemistry 47,
methyl-esterification status of cell-wall pectins. Current Biology 18, 510–517.
1943–1948.
Thompson DS. 2005. How do cell walls regulate plant growth? Journal of
Popper ZA, Fry SC. 2008. Xyloglucan–pectin linkages are formed intra- Experimental Botany 56, 2275–2285.
protoplasmically, contribute to wall-assembly, and remain stable in the cell
wall. Planta 227, 781–794. Tomos AD, Leigh RA. 1999. The pressure probe: a versatile tool in plant
cell physiology. Annual Review of Plant Physiology and Plant Molecular
Pritchard J. 1994. The control of cell expansion in roots. New Phytologist
Biology 50, 447–472.
127, 3–26.
Ulvskov P, Wium H, Bruce D, Jorgensen B, Qvist KB,
Radotic K, Roduit C, Simonovic J, Hornitschek P, Fankhauser C,
Skjot M, Hepworth D, Borkhardt B, Sorensen SO. 2005.
Mutavdzic D, Steinbach G, Dietler G, Kasas S. 2012. Atomic force
Biophysical consequences of remodeling the neutral side chains of
microscopy stiffness tomography on living Arabidopsis thaliana cells
rhamnogalacturonan I in tubers of transgenic potatoes. Planta 220,
reveals the mechanical properties of surface and deep cell-wall layers
609–620.
during growth. Biophysical Journal 103, 386–394.
Virgin HI. 1955. A new method for the determination of the turgor of plant
Ray PM. 1987. Principles of plant cell expansion. In: Cosgrove DJ, Knievel
tissues. Physiologia Plantarum 8, 954–962.
DJ, eds. Physiology of cell expansion during plant growth (Symposium on
Plant Physiology, Pennsylvania State University) . Rockville, MD: American Wainwright SA, Biggs WD, Currey JD, Gosline JM. 1976. Mechanical
Society of Plant Physiologists, 1–17. design in organisms . London: Edward Arnold.
Page 14 of 14 | Cosgrove
Wang CX, Wang L, Thomas CR. 2004. Modelling the mechanical Xi X, Kim SH, Tittmann B. 2015. Atomic force microscopy based
properties of single suspension-cultured tomato cells. Annals of Botany nanoindentation study of onion abaxial epidermis walls in aqueous
93, 443–453. environment. Journal of Applied Physics 117, 024703.
Wang T, Park YB, Caporini MA, Rosay M, Zhong L, Cosgrove Yamamoto R, Shinozak K, Masuda Y. 1970. Stress–relaxation
DJ, Hong M. 2013. Sensitivity-enhanced solid-state NMR detection of properties of plant cell walls with special reference to auxin action. Plant
expansin’s target in plant cell walls. Proceedings of the National Academy and Cell Physiology 11, 947– 956.
of Sciences, USA 110, 16444–16449. Yanagisawa M, Desyatova AS, Belteton SA, Mallery EL, Turner JA,
Wang T, Park YB, Cosgrove DJ, Hong M. 2015. Cellulose–pectin Szymanski DB. 2015. Patterning mechanisms of cytoskeletal and cell
spatial contacts are inherent to never-dried Arabidopsis primary cell walls: wall systems during leaf trichome morphogenesis. Nature Plants 1, 15014.
evidence from solid-state nuclear magnetic resonance. Plant Physiology Yoshida K, Komae K. 2006. A rice family 9 glycoside hydrolase isozyme
168, 871–884. with broad substrate specificity for hemicelluloses in type II cell walls. Plant
Wang T, Zabotina O, Hong M. 2012. Pectin–cellulose interactions in the and Cell Physiology 47, 1541–1554.
Arabidopsis primary cell wall from two-dimensional magic-angle-spinning Yuan S, Wu Y, Cosgrove DJ. 2001. A fungal endoglucanase with plant
solid-state nuclear magnetic resonance. Biochemistry 51, 9846–9856. cell wall extension activity. Plant Physiology 127, 324–333.
Wang Y, Ma NN, Qiu SC, Zou HY, Zang GC, Kang ZH, Wang GX, Zamil MS, Yi H, Puri VM. 2014. Mechanical characterization of outer
Huang JL. 2014. Regulation of the alpha-expansin gene OsEXPA8 epidermal middle lamella of onion under tensile loading. American Journal
expression affects root system architecture in transgenic rice plants. of Botany 101, 778–787.
Molecular Breeding 34, 47–57.
Zamil MS, Yi H, Puri VM. 2015. The mechanical properties of plant cell
Weber A, Braybrook S, Huflejt M, Mosca G, Routier-Kierzkowska walls soft material at the subcellular scale: the implications of water and of
AL, Smith RS. 2015. Measuring the mechanical properties of plant cells the intercellular boundaries. Journal of Materials Science 50, 6608–6623.

Downloaded from http://jxb.oxfordjournals.org/ at Universidad Autonoma de Chiapas on July 11, 2016


by combining micro-indentation with osmotic treatments. Journal of Zdunek A, Kurenda A. 2013. Determination of the elastic properties of
Experimental Botany 66, 3229–3241. tomato fruit cells with an atomic force microscope. Sensors (Basel) 13,
White PB, Wang T, Park YB, Cosgrove DJ, Hong M. 2014. Water– 12175–12191.
polysaccharide interactions in the primary cell wall of Arabidopsis thaliana Zhang T, Mahgsoudy-Louyeh S, Tittmann B, Cosgrove DJ. 2014.
from polarization transfer solid-state NMR. Journal of the American Visualization of the nanoscale pattern of recently-deposited cellulose
Chemical Society 136, 10399–10409. microfibrils and matrix materials in never-dried primary walls of the onion
Whitney SE, Gothard MG, Mitchell JT, Gidley MJ. 1999. Roles of epidermis. Cellulose 21, 853–862.
cellulose and xyloglucan in determining the mechanical properties of Zhao Z, Crespi VH, Kubicki JD, Cosgrove DJ, Zhong L. 2014.
primary plant cell walls. Plant Physiology 121, 657–664. Molecular dynamics simulation study of xyloglucan adsorption on cellulose
Wilson RH, Smith AC, Kacurakova M, Saunders PK, Wellner N, surfaces: effects of surface hydrophobicity and side-chain variation.
Waldron KW. 2000. The mechanical properties and molecular dynamics Cellulose 21, 1025–1039.
of plant cell wall polysaccharides studied by Fourier-transform infrared Zimmermann U, Steudle E. 1978. Physical aspects of water relations of
spectroscopy. Plant Physiology 124, 397–405. plant cells. Advances in Botanical Research 6, 45–117.
Wolf S, Greiner S. 2012. Growth control by cell wall pectins. Protoplasma Zykwinska A, Thibault JF, Ralet MC. 2008. Competitive binding of
249 Suppl 2, S169–S175. pectin and xyloglucan with primary cell wall cellulose. Carbohydrate
Wolf S, van der Does D, Ladwig F, et al. 2014. A receptor-like protein Polymers 74, 957–961.
mediates the response to pectin modification by activating brassinosteroid Zykwinska AW, Ralet MC, Garnier CD, Thibault JF. 2005. Evidence
signaling. Proceedings of the National Academy of Sciences, USA 111, for in vitro binding of pectin side chains to cellulose. Plant Physiology 139,
15261–15266. 397–407.

You might also like