Vous êtes sur la page 1sur 62

RESPONSES AND ADAPTATIONS TO

ABIOTIC STRESSES
Plant Physiology, Fifth Edition (2010), L. Taiz & E. Zeiger
Chapter 26 Responses and adapations to abiotic stresses

Biochemistry & Molecular Biology of Plants (2000), B.B. Buchanan, W.
Gruissen, R. J. Jones. Chapter 22 Responses to abiotic stresses
FISIOLOGIA VEGETAL COMPLEMENTAR
(2013/2014)

Abiotic stresses may arise from an excess or deficit in a physical or
chemical component that adversely affect growth, development,
or productivity.

Among the environmental conditions that cause damage are
water-logging, drought, high or low temperatures, excessive soil
salinity, inadequate mineral nutrients in the soil, too much or too
little light, and phytotoxic compounds, like ozone.

Resistance or sensitivity to the stress depends on the species, the
genotype, and the developmental age of the plant.

Many factors determine how plants respond to environmemtal
stresses (Figure 22.1).
The losses of primary productivity associated to environmental
stresses ranges from 51% to 82%, depending on the crop (Table 22.1).
Stress resistance mechanisms can be grouped in two general categories: avoidance
mechanisms, which prevent exposiure to stress, and tolerant mechanisms which
permit the plant to withstand the stress. Other resistance mechanism is achieved
through acclimation (Figure 22.2).
Gene expression patterns often change in response to stress

Stress recognition activates signal transduction pathways that transmit information within
individual cells and throughout the plant (Figure 22.3). Ultimately, changes in gene expression
which occur at the cellular level, are integrated into a response by the whole plant that may
modify growth and development and even influence reproductive capabalities.
STRESSES INVOLVING WATER DEFICIT
Water deficit can be induced by many environmental conditions:
Little rainfall - drought ; Soil salinity; Freezing temperatures
Transpiration rates over water absorption

Two parameters that describe the water status of plants are:
water potential, and relative water content



w - Water potential
s - Solute potential
p - Pressure potential
g - Gravitational potential
m - Matric potential
Turgid weight is determined by
floating tissues on water in an
enclosed chamber, preferably
under light, until the weight of
the tissues stays constant.
The water moves into and out a cell according to its potential gradient (Figure 22.4).
Osmotic adjustment is a
biochemical mechanism
that helps plants acclimate
to dry saline soil.

A plant cannnot extract water from the soil
unless the water potential in the root is
less than the water potential in the
surrounding soil.
Many drought-tolerant plants can regulate
their solute potentials to compensate for
transient or extended periods of water
stress. This process is called osmotic
adjustment (Figure 22.5).
Compatible solutes share specific biochemical attributes

Compatible solutes, also known as compatible osmolytes, are a small group of chemically
diverse organic compounds that are highly diverse organic compounds that are highly soluble
and do not interfere with cellular metabolism, even at high concentrations (Figure 22.6).

Monomeric sugars (e.g. sucrose fructose) can be released from polymeric forms (starch and
fructans, respectively) in response to stress. Once stress is removed, these monomers can be
repolymerised to facilitate rapid and reversible osmotic adjustment.

The organic character of osmolyte molecules probably reflects the
potential toxicity of concentrated inorganic solutes. Many ions found in
cells adversely affect metabolic processes when present at high
concentrations, possibly by binding to and altering the properties of
cofactors, substracts, membranes, and enzymes.

Furthermore, many ions can enter the hydration shells of a protein and
promote its denaturation. In contrast, compatible solues tend to be
neutrally charged at physiological pH, either nonionic or zwitterionic
(dipolar, with spatially separated positive and negative charges), and are
excluded from hydration shells of macromolecules (Figure 22.7).
In addition to stereotypical characteristics, compounds active in osmotic adjustement
demonstrate distribution patterns that support water potential (osmotic) equilibria among
the various membrane-bound compartments of the cell. Vacuoles, which can occupy as
much as 90% of the volume of a mature plant cell, tend to accumulate charged ions and
solutes that would perturb metabolism if present in the cytoplasm.

Compatible solutes in the citoplasm, however, allow the cytosol to achieve an osmotic
balance with the vacuole. For example, little or no glycine betaine is associated with
vacuolar sap from salt-stressed spinach leaves, but concentrations of this osmolyte in the
cytosol and chloroplasts can exceed 250 mM (Figure 22.8).
Glycine betaine accumulation is regulated by rates of its
synthesis and transport

Glycine betaine is synthesized and accumulated by many algae and higher plants.
Its distribution among plants is sporadic. All Chenopodiaceae species examined
to data accumulates glycine betaine. In other families this trait is found in only
some species.

Glycine betaine is synthesized from choline in a two-step pathway (Figure 22.9).

Evidence from subcellular localization studies with spinach places this pathway in the
chloroplast.

The first enzyme, choline monooxigenase, catalyzes the oxidation of choline to betaine
aldehyde, using photosynthetically reduced ferredoxin and molecular oxigen.

The second enzyme, betaine aldehyde dehydrogenase, catalyzes the oxidation of betaine
aldehyde to glycine betaine.

Plants that accumulate glycine betaine may use different pathways to generate the
choline precursor (Figure 22.9)
( A )
( B )
(A) - spinach, sugar beet
(B) - barley
In some plant species, salt stress inhibits sucrose synthesis
and promotes accumulation of Manitol

The manitol is the reduced form of the sugar mannose (Figure 22.10)
Taxonomically diverse plants
accumulate pinitol in response to
salt stress.

The biosynthesis of pinitol is believed to proceed
from the O-meythylation of myo-inositol, which
forms ononitol in angiosperms (Figure 22.11) and
the onotitol isomer, sequoyitol, in gimnosperms.
Impact of water deficit and salinity on transport across plant
membranes



Drought and salt stress both entail acclimation to low water potentials.
However, plants growing under saline conditions must also cope with
potentially toxic amounts of specific ions. In fact high cytosolic
concentrations of some ions, such as Na, perturb metabolism.


Under such conditions, regulating the concentration, composition and
distribution of ions within the plant cells can be viewed as an essential
feature of tolerance to osmotic stress.
Carriers, pumps, and channels opperate to minimize the
impact of perturbing ions on cell metabolism.


Due to K- Na cotransport system, cells subjected to high
concentrations of Na tends to accumulate this ion and to block the
absortion of K (this is one of the mechanisms of toxicity due to high concentrations of Na).


A Na toxic preventing mechanism involves the active transport of Na
across the plasma membrane and out of the cell. Another mechanism
is the cytosolic Na transport across the tonoplast and sequestration
on vacuole through Na/H antiporters energized by vacuolar H-
ATPase pumps as it was demonstrated for Mesembryanthemum
crystalinum and Arabidopsis.
Synthesis and activity of aquaporin may be up-regulated in
response to drought
Plasma membranes and tonoplasts can be rendered more permeable to water by
proteinaceous transmembrane water chanels called aquaporins.
Water movement through aquaporins can be
modulated rapidly. These channels may facilitate
water movement in drought stressed tissues and
promote recovery of turgor on watering.

In Arabidopsis, water deficit strongly induces
expression of the Rd28 gene which encode for
the aquaporin RD28 (a MIP - Major Intrinsed Protein).

In Mesembryanthemum crystalinum plants
subjected to a 400 mM NaCl shock treatment,
the amount of transcrips of this type of genes
first decrease after the initial shock, as does
turgor. Transcript concentrations then increase
as the turgor is restored (Figure 22.12).


Some seed proteins (LEA proteins) may protect
vegetative tissues from stress

LEA (Late Embryogenesis Abundant) proteins are overwhelmingly hydrophilic
cytoplasmic proteins sharing a biased amino acid composition, rich
in alanine and glycine, and lacking cysteine and tryptophan,
dependent of Lea genes firstly identified as genes induced in seeds
during maturation and desiccation.

Overexpression of transgenic LEA proteins in rice and yeast has been
shown to enhance resistance to specific water-deficit stresses


LEA proteins include five distinct families, designated Groups 1,
through 5 (Table 22.2)


Osmotin, a tobacco protein with antifungal activity, accumulates
during the water deficit

Osmotin is an abundant alkaline protein, classified as a pathogenic-related (PR)
protein, because in early studies it was found to accumulate after pathogenic
infection. The postulated sequence of events associated with its antifungal
properties has been described according to the model shown in Figure 22.13.

Transcription of an osmotin gene is induced by at least 10 signals: ABA, ethylene,
auxin, infection by tobacco mosaic virus (TMV), salinity, lack of water, cold, UV
light, wounding, and fungal infection.
Freezing stress

Some plants can acclimate to subfreezing temperatures
At a given temperature, the chemical
potential of ice is less than that of liquid
water. Thus, as ice formation os initiated in
the intercellular spaces, cellular water
moves down the water potential gradient,
across the plasma membrane, and toward
the extracellular ice. Therefore, awater
deficit develops within the cell in response
to freezing (Figure 22.15).

The details of the mecanisms that permit
freezing tolerance are not well understood.
Several processes that can occur in the
development of freezing tolerance are:
(a) stabilization of membranes;
(b) accumulation of sugars, other
osmolytes, and antifreeze proteins; and
(c) multiple changes in gene expression.


Stabilization of membranes
Multiple mechanisms are involved in the development of freezing tolerance. One
mechanism involves changes in membrane lipid composition, including enhanced fatty
acid desaturation in membrane phospholipids and changes in the abundance of
various membrane sterols and cerebrosides.

Accumulation of osmolytes and antifreezing proteins
Constitutive freeze-tolerant Arabidopsis plants and other ones, overexpress sugars and
proline. During periods of cold acclimation, some plants accumulate apoplastic PR
proteins that can retard growth of ice crystals. In winter rye, these antifreezing
proteins include endochitinases, (13)-endoglucanases, and osmotin (thaumatin)-
like proteins. Cold-induced PR proteins form oligomeric complexes that interact with
ice, inhibiting its growth and recrystallization more efectively than the individual
polypeptides can.

Changes in gene expression
Group 5 LEA proteins and the producs of other unic genes, have been found to
accumulate during freezing stress.



Flooding and oxygen deficit

Plants ca be damaged not only by the absence of water but also by too much water,
which blocks entry of O into the soil so that roots and other organs cannot carry out
respiration.

The supply of O to root cells is influenced by several factors, including soil porosity,
water content, temperature, root density, and the presence of competing algae and
aerobic microoganisms.

Oxygen concentration in root tissues also vary according to root deph, root
thickness, the volume of intercellular gaseous spaces, and cellular metabolic activity.

Plant or cellular oxygen status can be defined as NORMOXIC, HYPOXIC, or ANOXIC
(Table 22.3).

To survive short-term flooding, plants must generate NADP and NAD, and avoid
accumulation of toxic metabolites.

Periods of oxygen deficit can trigger developmental responses that promote
acclimation to hypoxic or anoxic conditions.
Plants vary in ability to tolerate flooding

Plant species generally can be classified as wetland, flood-tolerant,
or flood sensitive, according to their ability to withstand periods of
oxygen deficit (Table 22.4).
Wetland species possess diverse anatomical, morphological, and
physiological features that permit survival in aquatic environments or
waterlogged soils.

Growth in a wetland environment promotes formation of a thick root
hypodermis to reduce O loss to anaerobic soil.

To facilitate transport of O from aerial structures to submerged roots
and thereby maintain aerobic metabolism and growth, some plants
develop specific structures:
(a) aerenchyma - continuos, colunar intracellular spaces formed in
root cortical tissues (Figure 22.18);
(b) adventitious roots from hypocotyl or stem (Figure 22.19),
(c) lenticels openings in the periderm that allow gas exchange
(Figure 22.19):
(d) Pneumatophores shallow roots that grow with negative
geotrophy out of the aquatic environment (Figure 22.20).
Submergence of deepwater rice seedlings promotes formation of adventitious roots and
accelerates the internodal stem elongation, which enables stems and leaves to be
established above the aquatic environment (Figure 22.21).

Mitochondrial morphology is altered, but functional electron transport chain complexes of
the inner membrane and enzymes of the cytric acid cycle in the matrix are maintained
(Figure 22.22).
During short-term acclimation to anoxic conditions, plants
generate ATP through glycolysis and fermentation


In tomato (flood-sensitive), maize (flood-tolerant), and rice (wetland),
flooding stimulates an increase in glycolytic flux known as the Pasteur
effect. Sucrose or glucose from the floem is directed toward glycolysis in
flooded organs. Flooded rice coleoptiles also hydrolyze stored starch to
obtain additional sugars.

The energy-requiring steps of glycolysis generate low amounts of ATP while
reducing NAD to NADH. To support ongoing glycolysis in the absence of
mitochondrial respiration, the glycolytic substrate NAD must be
regenerated through fermentative reactions.

The principal end products of glycolysis in oxygen-deprived plant tissues
are lactate and ethanol (Figure 22.24); alanine, succinate, and y-
aminobutyrate may also be formed (Figure 22.25).

Besides lactate and ethanol
(Figure 22.24), alanine, succinate,
and y-aminobutyrate may also be
formed as end products of
glycolysis in oxygen-deprived plant
tissues (Figure 22.25).
Both lactate- and ethanol- producing fermentations yield NAD. However,
lactate lowers cytosolic pH, whereas ethanol does not. According to the
Davis-Roberts lactate dehydrogenase (LDH)/pyruvate decarboxylase (PDC)
pH-stat hypothesis, anaerobic metabolism is regulated by the activities of
pH-sensitive enzymes (Figure 22.26).

According to this model, the pyruvate produced initially by glycolysis is
converted to lactate in a reaction catalyzed by LDH, an enzyme with an
optimum at physiological pH. Lactate production reoxidizes NADH but also
lowers the cytoplasmic pH. Therefore the accumulation of lactate ultimely
stimulates the conversion of pyruvate to acetaldehyde.

Alcohol dehydrogenase (ADH) subsequently reduces the acetaldehyde
to ethanol while oxidizing NADH to NAD. Unlike lactate, ethanol is an
uncharged molecule at cellular pH and can diffuse across plasma
membranes. As a result, the switch to ethanol production stabilizes
cytoplasmic pH at a slightly acidic value.

In Nature, hypoxia frequently precedes anoxia in flooded roots. If
maize or rice seedlings are transfered to hypoxia conditions (3 kPa
oxygen) before transfer to anoxia (0 kPa oxygen), their survival rates
increase considerably and the ability to continue cell elongation is
greatly improved. This hypoxic pretreatment promotes acclimation by
increasing glycolytic flux and ATP production.

Hypoxia pretreatment results in greater specific activity of
hexokinase, fructokinase, pyruvate kinase, and other enzymes
involved in glycolysis and ethanolic fermentation.

Another important feature of this acclimation is development of the
capacity to export lactate from the cytoplasm to the surrounding
medium a further indication that avoidance of cytoplasmic acidosis
which is a major factor in survival of low-oxygen conditions (Figure
22.28) .
Shifting from aerobic metabolism to glycolytic metabolism
involves changes in gene expression.

Although most gene expression is repressed in response to oxygen deprivation (70% reduction
in total protein synthesis, in anoxia; in lesser extent under hypoxia), an important subset of
genes are upregulated. Most of the known proteins synthesized in high amounts in anoxic
roots are enzymes involved in sucrose and starch degradation, glycolysis, and ethanol
fermentation (Figure 22.29).
Several genes that encode isoforms
of the same enzyme, for example
ADH (Alcohol DeHydrogenase) and
glyceraldehyde-3-phosphate
dehydrogenase, are expressed
differentially in aerobic and anoxic
cells because of sequence
differences in their promotors.
Many of the genes
expressed in aerobic roots
are transcribed during
anoxia in approximately the
same amounts as during
aerobic conditions; their
mRNAs, however, are very
poorly translated.


In contrast, Adh1 mRNA is
translated efficiently in
anoxic cells (Figure 22.32).
The plant hormone ethylene promotes long-term
acclimative responses, including formation of aerenchyma
and stem elongation, in wetland and flood-tolerant plants.

Aerenchyma provide a conduit for gas diffusion between roots and aerial organs.
They can form by cell death and dissolution (lysigeny), by separation of cells without
collapse (schizogeny), or by a combination of lysigeny and schizogeny
(schizolysigeny).

The presence of low amounts of oxygen (less than 12.5 kPa to 3 kPa) stimulates
production of the hormone ethylene, which promotes the formation of
aerenchyma in central portion of the root cortex.

Anoxic roots develop fewer aerenchyma than hypoxic roots because oxygen is
required for ethylene synthesis (Figure 22.33A).

The abundance of 1-aminocyclopropane-1-carboxylic acid (ACC) synthase and ACC
oxidase, enzymes in the ethylene biosynthesis pathway, increases considerably in
response to hypoxia in maize roots.
Development of aerenchyma also involves Ca-mediated signal triggered by
ethylene, the ramifications of which are not fully understood.

Ethylene production and the Ca signal stimulate the death of cells withing the central
portion of the root cortex. At least two cell wall-degrading enzymes, cellulase and
xyloglucanase, are present or synthesized in large amounts in hypoxic roots; they do, most
likely, play a role in aerenchyma formation (Figure 22.33B).
S-Anenosyl-Methionine
The role of ethylene in aerenchyma development was confirmed by
using inhibitors of ethylene synthesis and antagonists of ethylene action
and by exposing aerobic roots to exogenous ethylene (Table 22.5)
Ethylene triggers epinasty i some flood-sensitive species.

Ethylene is responsible for transient changes in the morphology of aerial tissues in
flood-sensitive plants. Leaf epinasty curvature caused by cell expansion of the adaxial
cells of the petiole, is a common response to waterlogging associated with flood-
sensitive plants. Epinasty reduces the foliar absorption of light, thereby slowing
transpirational water loss in plants for which water absorption by roots is limited by
anoxia.

Due to absence of O in root ACC (1-aminocyclopropane-1-carboxylic acid)cannot, there, be converted in
ethylene. However some of this metabolite is transported to leaves where ACC oxidase
concentrations subsequently increase and ethylene is produced (Figure 22.34B).
Oxidative stress
Oxidative stress results from conditions promoting the formation
of active oxygen species that damage or kill cells. Several
environmental factors cause oxidative stress, (Figure 22.35).
Reactive oxygen species (ROS) are
formed during certain redox
reactions and during incomplete
reduction of oxygen or oxidation
of water by the mitochondrial or
chloroplast electron transfer
chains (Figure 22.36).
The antioxidant defense systems
include nonenzymatic and enzymatic
antioxidants.

The major antioxidant species in plants
are ascorbate (Vitamin C), reduced
glutathione (GSH), -tocopherol (vitamin
E), and carotenoids; polyamines and
flavonoids also may provide some
protection from the free radical injury.
These compounds and enzymes are not
distributed uniformely, so defense
systems vary among specific subcelular
compartments (Tables 22.6 and 22.7).
The ascorbate-glutathione cycle is the major antioxidant pathway in
plastids, where ROS are generated during normal biochemical
processes that include photosynthetic transfer of electrons.

The photosynthetic apparatus receives additional protection from
oxidative damage by the exothermic production of the xanthophyll
zeaxanthin.
Tropospheric ozone is linked to oxidative stress in plants.
Anthropogenic hydrocarbons and oxides of nitrogen (NO, NO) and sulfur (SOx) react with
solar UV radiation to generate ozone (O). Stractosphere ozone is benefical because it shields
the earth from UV irradiation, but tropospheric ozone is harmful to life because it is a highly
reactive oxidant (Figure 22.38).
Ozone causes oxidative damage to biomolecules

The negative effects of ozone on plants include decreased rates of
photosynthesis, reduced growth of shoots and roots, acelerated
senescence, reduced crop yield, and leaf injury (Figure 22.39).
The mechanism of ozone toxicity is not completely understood (Figure 22.40).
Most likely, damage occurs after stomatal
uptake of the ozone, which results in oxidative
destruction of lipids and proteins of the plasma
membrane and production of free radicals or
other reactive intermediates.
O may react with ethylene and other alkenes in
the apoplastic fluid to form HO

, O
-
and HO.

O and ROS not decomposed in the apoplast will
react with membrane lipids to form reactive
lipid peroxidases , which will perpetuate ROS
formation.
Some ozone and ROS may succed in entering
the cytoplasm and membrane damage. Those
that promote further production of radicals
inside cell.
Ozone-stimulated damage to the plasma membrane permeability, inhibits H-
pump activity, collapses membrane potential, and increases Ca uptake from
apoplast. Within the cell, the presence of ROS and damaged biomolecules may
trigger the antioxidant defense system. Ozone stimulates the wound-induced
production of ethylene as well as the accumulation of salicylic acid (SA). Both of
these hormones operate in different signal transduction pathways that induce
specific changes in gene expression and metabolism.

The mechanism of ozone action can differ depending of whether the exposure is
acute, chronic, or repeated (Table 22.8).
Increased synthesis of antioxidants and antioxidant enzymes
can improve tolerance to oxidative stress.

In many plants, ozone exposure and other oxidative stresses can stimulate synthesis of
antioxidant metabolites and enhance antioxidant enzyme activities (Table 22.9).

Tolerant plants or acclimated plants may contain higher concentrations of antioxidant
factors to minimize damage from ROS.

To study the role of antioxidant enzymes in oxidative stress tolerance, researchers
have overexpressed these proteins in transgenic plants (Table 22.10).
Heat Stress
Plants exposed to excess heat exhibit a characteristic set of cellular and metabolic
responses, many of which are conserved in all organisms. The signature response to
heat stress is a decrease in the synthesis of normal proteins, acompanied by an
accelerated transcription and translation of a set of proteins known as heat shock
proteins (HSPs).
This response is observed when plants are
exposed to temperatures at least 5C above
their optimal growing conditions.

HSPs can be visualised easily on two-dimensional
electrophoretic gels (Fig. 22.41).
Plants can acclimate to heat stress
Plants can acquire thermotolerance if subjected to a nonlethal (permissive) high
temperature for a few hours before encounting heat shock conditions. An acclimated
plant survive exposure to a temperature that would otherwise be lethal (Fig. 22.42).
Five classes of HSPs are defined according to size (Table 22.11)
Expression of many HSPs is controlled by a transcription
factor that recognizes a conserved promoter sequence.
The heat shock transcription factor (HSF) is expressed constititutively but must be
activated during heat stress to recognize its DNA target, the heat shock element (HSE).

The HSE is made up of 5-bp repeats in alternating orientations with consensus nGAAn. An HSF-regulated promoter may
contain five to seven of these repeats close to TATA box. Many HSEs contain the DNA element 5-CTnGAAnnTTCnAG-3.
Like other HSFs, the HSF of
Arabidopsis (ATHSF1; Figure 22.43)
only can bind DNA as trimers.

Heat stress is required for
trimerization (Figure 22.44).

Trimerization depends on the presence of
a leucine zipper configuration of
hydrophobic heptad repeats located
adjacent to the DNA-binding domain.
DNA binding and transcriptional activity
are repressed in the absence of heat
stress.
END

Vous aimerez peut-être aussi