Sei sulla pagina 1di 36

Critical Reviews in Plant Sciences, 24:2358, 2005

c Taylor & Francis Inc.


Copyright 
ISSN: 0735-2689 print / 1549-7836 online
DOI: 10.1080/07352680590910410

Drought and Salt Tolerance in Plants


Dorothea Bartels and Ramanjulu Sunkar
Institute of Molecular Physiology and Biotechnology of Plants, University of Bonn, Kirschallee 1,
D-53115 Bonn, Germany
Table of Contents
I.

INTRODUCTION .............................................................................................................................................25

II.

OSMOTIC STRESS ..........................................................................................................................................25


A. Osmotic Stress-Induced Growth Arrest ..........................................................................................................25
B. Osmotic Stress Affects Cell Division and Elongation ......................................................................................30

III.

SIGNAL PERCEPTION ...................................................................................................................................30

IV.

SIGNAL TRANSDUCTION ..............................................................................................................................31


A. MAPKinase Pathways ..................................................................................................................................31
B. SNF-1-Like Kinases Are Involved in Osmotic Stress Signalling .......................................................................32
C. Phosphatases ...............................................................................................................................................32
D. Phospholipid Signalling ...............................................................................................................................33
1.
Inositol 1,4,5-Triphosphate (IP3) .........................................................................................................33
2.
Phosphatidic Acid (PA) ......................................................................................................................34
E. Other Signalling Molecules ..........................................................................................................................34
1.
Salicylic Acid ....................................................................................................................................34
2.
Nitric Oxide (NO) .............................................................................................................................34

V.

CALCIUM SIGNALLING DURING DEHYDRATION AND SALT STRESS ....................................................35


A. Calcium-Dependent Protein Kinases (CDPKs) ...............................................................................................35
B. Calcium-Binding Proteins ............................................................................................................................35
C. Ca2+ -Mediated SOS Pathways Are Involved in Ion Homeostasis .....................................................................35
D. Calcineurin B (CBLS) and Osmotic Stress Responses .....................................................................................35
E. Ca2+ ATPases .............................................................................................................................................36

VI.

TRANSCRIPTIONAL REGULATION OF GENE EXPRESSION ....................................................................36


A. ABA Response Elements (ABREs) ...............................................................................................................36
B. The Dehydration Response Element (DRE) ...................................................................................................37
C. Transcription Factors Modulated by Osmotic Stress ........................................................................................37
1.
Basic Region Leucine Zipper (bZIP) Proteins .......................................................................................37
2.
Homeodomain-Leucine Zipper Proteins (HD-ZIP) ................................................................................37
3.
Zn-Finger Proteins .............................................................................................................................37
4.
AP2/ERF-Type Transcription Factors ..................................................................................................38
5.
Myb-Like Proteins .............................................................................................................................38
6.
Myc-Like Proteins .............................................................................................................................39
7.
CDT-1 ..............................................................................................................................................39

Address correspondence to Dorothea Bartels, Institute of Molecular Physiology and Biotechnology of Plants, University of Bonn, Kirschallee
1, D-53115 Bonn, Germany. E-mail: dbartels@uni-bonn.de

23

24
VII.

D. BARTELS AND R. SUNKAR

ACCUMULATION OF SUGARS AND COMPATIBLE SOLUTES ...................................................................39


A. Sugars ........................................................................................................................................................39
B. Cyclitols .....................................................................................................................................................40
C. Proline ........................................................................................................................................................40
D. Glycine Betaine ...........................................................................................................................................40

VIII. PROTECTIVE PROTEINS AND OTHER PATHWAYS INVOLVED IN STRESS ADAPTATION .....................40
A. Late Embryogenesis-Abundant (LEA) Proteins ..............................................................................................40
B. Aquaporins .................................................................................................................................................41
C. Heat Shock Proteins (Hsps) ..........................................................................................................................41
D. Proteases and Proteinase Inhibitors ...............................................................................................................42
E. Polyamines .................................................................................................................................................42
IX.

OXIDATIVE STRESS A CONSEQUENCE OF DEHYDRATION AND SALT STRESS ....................................42


A. Formation of Reactive Molecules ..................................................................................................................42
B. Enzymes That Detoxify Aldehydes ...............................................................................................................43
C. Peroxiredoxins ............................................................................................................................................43
D. Thioredoxins ...............................................................................................................................................43
E. Protein Oxidation ........................................................................................................................................43

X.

IONIC STRESS ................................................................................................................................................44


A. Na+ Toxicity and Homeostasis .....................................................................................................................44
B. Na+ Exclusion ............................................................................................................................................44
C. Na+ Compartmentalization ...........................................................................................................................44
D. Proton Transporters and Salt Tolerance ..........................................................................................................45
E. SOS Pathway and Ion Homeostasis ...............................................................................................................45

XI.

ABSCISIC ACID (ABA) ....................................................................................................................................46


A. Regulation of ABA Levels ............................................................................................................................46
B. The Role of ABA in Stomatal Closure ...........................................................................................................47
C. ABA Signalling Components ........................................................................................................................47

XII.

MONITORING GLOBAL GENE EXPRESSION USING MICROARRAY ANALYSIS .....................................48

XIII. CONCLUSIONS ...............................................................................................................................................48


ACKNOWLEDGMENTS ..................................................................................................................................49
REFERENCES .................................................................................................................................................49

Agricultural productivity worldwide is subject to increasing environmental constraints, particularly to drought and salinity due
to their high magnitude of impact and wide distribution. Traditional breeding programs trying to improve abiotic stress tolerance have had some success, but are limited by the multigenic
nature of the trait. Tolerant plants such as Craterostigma plantagenium, Mesembryanthemum crystallinum, Thellungiella halophila
and other hardy plants could be valuable tools to dissect the extreme tolerance nature. In the last decade, Arabidopsis thaliana, a
genetic model plant, has been extensively used for unravelling the
molecular basis of stress tolerance. Arabidopsis also proved to be

extremely important for assessing functions for individual stressassociated genes due to the availability of knock-out mutants and
its amenability for genetic transformation. In this review, the responses of plants to salt and water stress are described, the regulatory circuits which allow plants to cope with stress are presented,
and how the present knowledge can be applied to obtain tolerant
plants is discussed.
Keywords

dehydration, salinity, ABA, plant stress tolerance, transgenic plants

DROUGHT AND SALT TOLERANCE IN PLANTS

I.

INTRODUCTION
Drought and salinity are two major environmental factors
determining plant productivity and plant distribution. Drought
and salinity affect more than 10 percent of arable land, and desertification and salinization are rapidly increasing on a global
scale declining average yields for most major crop plants by
more than 50 percent (Bray et al., 2000). Understanding plant
tolerance to drought and salinity is therefore of fundamental importance and forms one of the major research topics. Plants can
perceive abiotic stresses and elicit appropriate responses with
altered metabolism, growth and development. The regulatory
circuits include stress sensors, signalling pathways comprising
a network of protein-protein reactions, transcription factors and
promoters, and finally the output proteins or metabolites. Classical breeding approaches revealed that stress tolerance traits
are mainly quantitative trait loci (QTLs), which make genetic
selection of traits difficult. Nevertheless, very respectable stress
tolerant crops have been obtained, mainly by introducing traits
from stress-adapted wild relatives.
As water and salt stresses occur frequently and can affect most
habitats, plants have developed several strategies to cope with
these challenges: either adaptation mechanisms, which allow
them to survive the adverse conditions, or specific growth habits
to avoid stress conditions. Stress-tolerant plants have evolved
certain adaptive mechanisms to display different degrees of tolerance, which are largely determined by genetic plasticity. Differential stress tolerance could be attributed to differences in
plant reactivity in terms of stress perception, signal transduction and appropriate gene expression programs, or other novel
metabolic pathways that are restricted to tolerant plants. The
hypothesis that the genetic program for tolerance is at least to
some extent also present in nontolerant plants is supported by the
observation that gradual acclimation of sensitive plants leads to
acquisition of tolerance to some degree. These plants may need
gradual adaptation for proper expression of genes responsible
for acquisition of tolerance (Zhu, 2001).
Our understanding of how plants respond to water and salt
stress has advanced by analyzing stress-tolerant species like the
desiccation tolerant plant Cratesostigma plantagineum (Bartels
and Salamini, 2001) or the salt-tolerant plant Mesembryanthemum crystallinum (Bohnert and Cushman, 2000). Despite the
fact that research with the desiccation-tolerant plant C. plantagineum revealed additional novel aspects such as specific carbohydrate metabolism and the existence of CDT-1 gene that
were unknown in other nontolerant plant species (Bartels and
Salamini, 2001), the connection between these metabolites and
tolerance is still correlative. Molecular genetic studies have been
performed with Arabidopsis thaliana, which does not display
extreme stress tolerance, but shows many stress responses at
the molecular level and has therefore been successfully used for
a genetic dissection of stress response pathways (Zhu, 2002;
Shinozaki et al., 2003). More importantly, we learned that it is
very likely that the extreme tolerant model plants did not acquire
unique genes since stress-relevant genes are ubiquitously present

25

in the plant kingdom. A particular gene expression pattern is often associated with the tolerant phenotype and it is unknown
to date how this is achieved. This may involve other molecular aspects, like chromatin organization, which have not been
well researched. Recently, the salt-tolerant plant Thellungiella
halophila was introduced as an attractive model plant to study
molecular genetics of salt tolerance, because it is a close relative to Arabidopsis and amenable for transformation unlike
that of other tolerant model plants (Volkov et al., 2003; Bressan
et al., 2001; Zhu, 2001). Thus it may be possible to combine
profound genetic knowledge with the expression of extreme
tolerance.
Exposure to drought or salt stress triggers many common
reactions in plants. Both stresses lead to cellular dehydration,
which causes osmotic stress and removal of water from the cytoplasm into the extracellular space resulting in a reduction of
the cytosolic and vacuolar volumes. Another consequence is the
production of reactive oxygen species which then in turn affects
cellular structures and metabolism negatively. Early responses
to water and salt stress are largely identical except for the ionic
component. These similarities include metabolic processes such
as, for example, a decrease of photosynthesis or hormonal processes like rising levels of the plant hormone ABA. High intracellular concentrations of sodium and chloride ions are an
additional problem of salinity stress.
Adaptation to salinity and drought is undoubtedly one of
the complex processes, involving numerous changes including attenuated growth, the activation/increased expression or
induction of genes, transient increases in ABA levels, accumulation of compatible solutes and protective proteins, increased
levels of antioxidants and suppression of energy-consuming
pathways. However, no consensus has been reached in defining the key processes determining tolerance and the secondary
follow-up processes. With the advancement of high throughput
DNA technologies, several hundred stress-induced or upregulated genes have been identified. The search for stress-associated
genes may have been saturated at least in Arabidopsis. However, the function of only a limited number of gene products are
known (Ingram and Bartels, 1996; Bray, 1997; Shinozaki and
Yamaguchi-Shinozaki, 1997; Hasegawa et al., 2000; Ramanjulu
and Bartels, 2002). Several stress-associated genes have been
evaluated or studies are in progress for their contribution to
drought or salt tolerance in laboratory studies (Table 1). More
rigorous and perhaps field studies are required for possible utilization of these genes for improving stress tolerance in agricultural plants through biotechnological approaches.
II.
A.

OSMOTIC STRESS

Osmotic Stress-Induced Growth Arrest


Arrest of plant growth during stress conditions largely depends on the severity of the stress. Mild osmotic stress leads
rapidly to growth inhibition of leaves and stems, whereas roots
may continue to elongate (Westgate and Boyer, 1985; Sharp

26

D. BARTELS AND R. SUNKAR

TABLE 1
Stress-responsive genes contributing to drought or salt tolerance in transgenic plants
Gene

Plant species

DREB1A (AP2
Arabidopsis
Transcription factor)

OsDREB1A

Arabidopsis

Alfin1 (Transcription
factor)

Alfalfa

Tsi1 (EREBP/AP2
Tobacco
DNA binding motif)

CBF1 (DREB1B)

Tomato

CBF4

Arabidopsis

ABF3/ABF4

Arabidopsis

AtMYC2/AtMYB2

Arabidopsis

ZPT2-3
(Cys2/His2-type
Zincfinger protein)
CpMYB10

Petunia

mt1D (Mannitol-1phosphate
dehydrogenase)

Tobacco

P5CS (Pyrroline-5carboxylate
synthase)

Arabidopsis

Parameters evaluated

Remarks

survived 2 weeks without


improved drought, salt,
watering; activated the
and cold tolerance
expression of genes
involved in stress tolerance
(rd29A)
improved drought, salt,
and freezing
tolerance
increase in root growth under improved salt tolerance
normal and saline
conditions
activated the expression of
improved salt and
several PR genes, PR1,
pathogen tolerance
PR2, PR3, osmotin and
SAR8.2 under unstressed
condition
activated gene expression,
improved drought
catalase1 coupled with
tolerance
decreased accumulation of
H 2 O2
activated the expression of
improved drought
stress responsive genes
tolerance
improved drought
reduced transpiration and
better survival under
tolerance
drought conditions
less electrolyte leakage in
improved drought
transgenic plants
tolerance
better survival rate during
improved drought
drought stress
tolerance
better germination ability
under saline or mannitol
treatments
increased root biomass
accumulation

improved drought and


salt tolerance

Reference
Kasuga et al., 1999;
Liu et al., 1998

Dobouzet et al., 2003

Winicov, 2000

Park et al., 2001

Hsieh et al., 2002

Haake et al., 2002


Kang et al., 2002

Abe et al., 2003


Sugano et al., 2003

Villalobos et al., 2004

improved salt tolerance Tarczynski et al., 1993

Shen et al., 1997


targeted to chloroplasts increased retention of
increased oxidative
of tobacco
chlorophyll against methyl
stress tolerance
viologen treatment
better biomass production
improved salt and
Abede et al., 2003
Wheat
osmotic stress
tolerance
Tobacco
enhanced root biomass and
improved salt tolerance KaviKishor et al., 1995
flower development
Rice
increase in biomass
improved drought and Zhu et al., 1998
accumulation
salt tolerance
(Continued on next page)

27

DROUGHT AND SALT TOLERANCE IN PLANTS

TABLE 1
Stress-responsive genes contributing to drought or salt tolerance in transgenic plants (Continued)
Gene
SacB

Plant species
Tobacco

Reference

improved tolerance to
PEG treatment
improved drought
tolerance
improved drought
tolerance

Pilon-Smits et al., 1995

less inhibition in
photosynthetic rate; better
recovery from stress
better germination and
photosynthetic activity
higher photosynthetic
activity; faster recovery
from stress
antisense plants took longer
duration to lodge the
inflourescence after
subjecting to salinity
increased leaf area, better
photosynthetic activity and
better water retaining
capacity
increased leaf area, better
photosynthetic activity and
better water retaining
capacity
better plant growth and less
photooxidative damage

improved drought and


salt tolerance

Sheveleva et al., 1997

Rice

better growth performance


and photosynthetic
capacity

improved drought and


salt tolerance

Jang et al., 2003

Tobacco

better photosynthetic
capacity under chilling
stress
better photosynthetic
efficiency, yield, and
survival rate

improved cold and


oxidative stress
tolerance
improved water stress
tolerance

Gupta et al., 1993

TPS1 (yeast)
Tobacco
(Trehalose-6phosphate
synthetase) (subunit)
IMT1(myo-Inositol-O- Tobacco
methyltransferase)
Arabidopsis
Rice

ProDH (Proline
dehydrogenase)

antisense-Arabidopsis

OtsA (E. coli)


(Trehalose-6phosphate
synthase)
OtsB (E. coli)
(Trehalose-6phosphate
synthase)
OtsA and OtsB

Tobacco

AtOAT (Ornithine
amino transferase)

Tobacco

BADH1 (Betaine
aldehyde
dehydrogenase)
TPS and TPP
(trehalose-6phosphate
phosphatase)
(E. coli)
Cu/ZnSOD

Tomato

MnSOD (superoxide
dismutase)

Remarks

increased biomass
production
better dry weight
accumulation
delay in withering or
enhanced moisture
retention capacity

Beta vulgaris

CodA (Choline
oxidase)

Parameters evaluated

Tobacco

Rice

Alfalfa

Pilon-Smits et al.,
1999
Holmstrom et al., 1996

improved salt and cold Hayashi et al., 1997


tolerance
improved salt and cold Sakamoto et al., 1998
tolerance
improved salt tolerance Nanjo et al., 1999

improved drought
tolerance

Pilon-Smits et al., 1998

improved drought
tolerance

Pilon-Smits et al., 1998

improved drought, salt Garg et al., 2002


and low-temperature
tolerance
accumulated more proline;
improved NaCl or
Roosens et al., 2002
higher biomass and higher
mannitol tolerance
germination rate under
osmotic stress conditions
better root development and improved salt tolerance Jia et al., 2002
less leakage of electrolytes

McKersie et al., 1996

(Continued on next page)

28

D. BARTELS AND R. SUNKAR

TABLE 1
Stress-responsive genes contributing to drought or salt tolerance in transgenic plants (Continued)
Gene
FeSOD (superoxide
dismutase)

Plant species
Tobacco

Glutathione-STobacco
transferase/Glutathione
peroxidase
MsALR
Alfalfa
(Aldose/aldehyde
reductase)
AtALDH3 (Aldehyde Arabidopsis
dehydrogenase)
Ascorbate peroxidase

Tobacco

OsCDPK
Rice
(Calcium-dependent
protein kinase)
AtGSK1
Arabidopsis

AtNDPK2 (Nucleotide Arabidopsis


diphosphate kinase)
overexpressing and
knock-out mutant

AtNHX1 (Vacuolar
Na+ /H+ antiporter)

Arabidopsis
Tomato

SOS1 (Plasma
membrane
Na+ /H+ antiporter)
AtHAL3a

HAL1 (Yeast)

Brassica napus
Rice
Arabidopsis

Arabidopsis

Watermelon
Arabidopsis

Parameters evaluated
protected the plasmalemma
and PSII against the
damaging effects of
superoxide
decreased oxidative damage

decreased lipid peroxidation


and better photosynthetic
capacity
decreased lipid peroxidation

Remarks
improved salt and
oxidative stress
tolerance

Reference
Van Camp et al., 1996

improved salt and cold Roxas et al., 1997;


tolerance
2000
improved drought and Oberschall et al., 2000
heavy metal
tolerance
improved drought, salt Sunkar et al., 2003
and oxidative stress
tolerance
improved drought and Badawi et al., 2004
salt tolerance

better photosynthetic
capacity under stress
conditions
enhanced levels of
improved drought and Saijo et al., 2000
stress-responsive genes,
salt tolerance
rab16A, SalT, and wsi18
improved salt and
Piao et al., 2001
better root growth,
drought tolerance
expression of
stress-responsive genes in
the absence of NaCl stress;
AtCP1, RD29A, AtCBL1
decreased ROS accumulation enhanced tolerance to Moon et al., 2002
in overexpressing plants,
salt, cold and methyl
while increased ROS
viologen treatments
accumulation in knock-out
mutants
Apse et al., 1999
Na+ compartmentation in
the vacuole

improved salt tolerance Zhang and Blumwald,


2001
Zhang et al., 2001
Ohta et al., 2002
improved salt tolerance Shi et al., 2003

better root growth, PSII


activity and survival under
salt stress conditions
transgenic seedlings
improved tolerance to Espinosa-Ruiz et al.,
developed roots and true
salt and osmotic
1999
leaves but not the WT
stresses
seedlings
better growth performance of improved salt tolerance Yang et al., 2001
transgenic plants;
transgenic plants
Ellul et al., 2003
accumulated less Na+
(Continued on next page)

29

DROUGHT AND SALT TOLERANCE IN PLANTS

TABLE 1
Stress-responsive genes contributing to drought or salt tolerance in transgenic plants (Continued)
Gene
CDT1

Plant species
Craterostigma
plantagenium

Parameters evaluated

desiccation survival of callus


without ABA treatment;
constitutive expression of
stress-induced genes,
CdeT-27-45, CdeT-6-19,
CdeT-11-24
GlyoxylaseI
Brassica juncea
better retention of
chlorophyll
Glyoxylase I and
Tobacco
grew, flower and set viable
Glyoxylase II
seeds under salinity
GS2 (Chloroplastic
Rice
PSII activity retained for two
glutamine
weeks of NaCl salinity,
synthetase)
while the control plant lost
PSII activity within one
week
AtHsp17.6A (Small
Arabidopsis
control plants withered
heat shock protein)
earlier than transgenic
plants under drought
conditions; better survival
rate and fresh weight
accumulation under saline
conditions
()Phospholipase D antisense-Arabidopsis stomatal closure impaired;
more water loss
AtNCED3
Arabidopsis-sense and transpiration rate reduced in
antisense and
sense plants and enhanced
knock-out plants
in antisense and knock-out
mutants, rab18, kin1, and
rd29B gene induction in
sense plants
AVP1
Arabidopsis
reduced stomatal opening
(H+ Pyrophosphatase)
increased salt
accumulation in the
vacuoles
BiP
Tobacco-sense and
sense plants displayed better
antisense plants
survival whereas the
antisense plants are
hypersensitive to drought
stress
Invertase (yeast)
Tobacco
better photosynthetic
efficiency
AtRab7 (Vesicle
Arabidopsis
decreased accumulation of
trafficking protein)
reactive oxygen species
under salinity
ADR1 (CC-NBS-LRR Arabidopsis
increased expression of
protein)
dehydration responsive
genes such as ERD11,
GST, etc.

Remarks
improved dehydration
tolerance

Reference
Furini et al., 1996

improved salt tolerance Veena et al., 1999


improved salt tolerance Singla-Pareek et al.,
2003
improved salt tolerance Hoshida et al., 2000

improved drought and


salt tolerance

Sun et al., 2001

decreased tolerance to
water stress
improved drought
tolerance in sense
and decreased
drought tolerance in
antisense and
knock-out plants
improved salt and
drought tolerance

Sang et al., 2001

improved drought
tolerance

Alvim et al., 2001

Iuchi et al., 2001

Gaxiola et al., 2001

improved salt tolerance Fukushima et al., 2001.


improved salt and
Mazel et al., 2004
osmotic stress
tolerance
improved drought
Chini et al., 2004
tolerance but not to
salinity or heat stress

30

D. BARTELS AND R. SUNKAR

et al., 1988; Nonami and Boyer, 1990; Spollen et al., 1993). The
degree of growth inhibition due to osmotic stress depends on the
time scale of the response, the particular tissue and species in
question, and how the stress treatment was given (rapid or gradual). Growth arrest can be considered as a possibility to preserve
carbohydrates for sustained metabolism, prolonged energy supply, and for better recovery after stress relief. The inhibition
of shoot growth during water deficit is thought to contribute to
solute accumulation and thus eventually to osmotic adjustment
(Osorio et al., 1998). For instance, hexose accumulation accounts for a large proportion of the osmotic potential in the cell
elongation zone in cells of the maize root tip (Sharp et al., 1990).
On the other hand, continuation of root growth under drought
stress is an adaptive mechanism that facilitates water uptake
from deeper soil layers. Similarly, continued root growth under
salt stress may provide additional surfaces for sequestration of
toxic ions, leading to lower salt concentration. For example, salt
tolerance of barley was correlated with the better root growth
rates coupled with fast development and early flowering (Munns
et al., 2000).

B.

Osmotic Stress Affects Cell Division and Elongation


Plants have the unique attribute of modulating their development with the prevailing environmental conditions involving plant hormones, which in turn influence gene expression
programs. Cell division is the principal determinant of meristem activity and determines the overall plant growth rate. It has
been proposed that environmental and developmental controls
of growth rate act by regulating cyclin-dependent kinase (CDK)
activity and cell division (Cockcroft et al., 2000; West et al.,
2004). CDKs are a family of protein kinases, each with a positive regulatory subunit termed a cyclin and the catalytic subunit
CDK (den Boer and Murray, 2000). CDKs are emerging as key
players in regulation of cell division and are likely to be regulated
at both transcriptional and post-translational levels in response
to stress. For example, maize ZmCdc2 (a member of the CDK
family) was shown to be downregulated by water stress leading to a decrease in mitotic cell cycling (Setter and Flanningan,
2001). The decrease in cell division in response to water stress
is characterized by lower CDK activity, which is correlated with
tyrosine phosphorylation (Schuppler et al., 1998). The recent
finding that ABA induces expression of an inhibitor of CDK
(ICK1) links cell division and ABA (Wang et al., 1998). This is
further supported by studies of Kang et al. (2002) who reported
that expression of the ABA inducible cell cycle regulator ICK1
was increased in transgenic plants overexpressing ABF3 and
ABF4 resulting in dwarf phenotypes. These mechanisms could
be responsible for ABA-dependent cell cycle arrest during osmotic stress in plants.
Cell expansion is a coordinately regulated process at the
whole plant level and is influenced by external stimuli including
water availability. The rate of cell expansion is mainly determined by two parameters, cell wall extensibility and cellular

osmotic potential. The enlargement of plant cells involves control of wall synthesis and expansion, solute and water transport,
membrane synthesis, Golgi secretion, ion transport and other
processes (Cosgrove, 1997). Expansins are a family of plant proteins essential for acid-induced cell wall loosening (Cosgrove,
1997).
Expression of three expansin genes Exp1, Exp5, and ExpB8
was upregulated in the apical region of roots after growth at
low water potential leading to higher amounts of expansin protein, which is closely correlated with the root elongation (Wu
et al., 2001). These results are consistent with the hypothesis
that the adaptive wall loosening and growth maintenance in the
apical region of maize roots are partly due to altered expansin
gene expression in the root tip at low water potentials (Wu et al.,
2001). Expansin genes in Craterostigma plantagineum, CpExp1
and CpExp3, are upregulated to different degrees in response to
dehydration. In addition CpExp1 but not CpExp3, is also upregulated in response to rehydration (Jones and McQueen-Mason,
2004). These results implicate a role for expansins during dehydration and rehydration, particularly in increasing wall flexibility. However, the effect of osmotic stress on cell enlargement
is still not clear and may also involve other hormones such as
auxin, cytokinin, or gibberellins.

III. SIGNAL PERCEPTION


It is still an open question how plants sense osmotic stress.
Because no plant molecule has truely been identified as osmosensor, scientists oriented themselves to study how yeast and
microorganisms sense osmotic stress. In yeast, hyperosmotic
stress is sensed by two types of osmosensors, SLN1 and SHO1,
that feed finally into HOG (high-osmolarity glycerol) MAPK
pathway. High osmolarity induces loss of turgor that leads concomitantly to shrinkage of cell volume and an increase in the
distance between plasma membrane and cell wall. SLN1 is likely
to sense the change in turgor pressure (Reiser et al., 2003).
SLN1 is a two-component regulatory system, which is also a
well characterized signal transduction element in prokaryotes.
The two-component regulatory system consists of a phosphorelay between three proteins. The yeast osmosensor SLN1 is a
fused two-component system that autophosphorylates a histidine residue in the N-terminal sensor domain and then transfers
the phosphate group to an aspartate residue in the C-terminallocated response-regulator domain. The phosphate is transferred
to YPD1, which functions as a second histidine phosphorelay
intermediate between SLN1 and the response regulator SSK1.
SSK1 finally feeds into the HOG pathway, which responds to
increased extracellular osmolarity and is responsible for osmolyte (glycerol) accumulation (Posas et al., 1996; WurglerMurphy and Saito, 1997). Another yeast membrane protein involved in osmosensing is SHO1, which also converges with
the MAPK pathway. Which pathway is activated depends on
the osmotic stress level. In Arabidopsis an SLN1 homologue,
AtHK1, was identified; it can function as an osmosensor in yeast

DROUGHT AND SALT TOLERANCE IN PLANTS

and complements SLN1-deficient yeast mutants (Urao et al.,


1999).
Recently, NtC7, which is originally identified as a gene that
is responsive to wounding, has been suggested as another probable candidate for sensing osmotic stress in plants (Tamura et al.,
2003). NtC7 transcript accumulates rapidly and transiently not
only in response to wounding but also to salt and osmotic stress.
The NtC7 gene encodes a receptor-like membrane protein and
overexpression improved osmotic stress tolerance induced by
mannitol but not by NaCl (Tamura et al., 2003). These results
suggest that NtC7 may play a role in sensing specifically osmotic stress. Cre1 (cytokinin response 1) is a cytokinin receptor
in Arabidopsis. Recent experiments implicate Cre1 as another
likely candidate for sensing osmotic stress in plants (Reiser
et al., 2003). Cre1 and Sln1 have similar organizations of the
cytoplasmic histidine kinase and receiver domains. These observations suggest that several candidates for osmosensors have
been proposed, although their role as osmosensors has yet to be
demonstrated.
IV.

SIGNAL TRANSDUCTION
Plants react to external stimuli by initiating signalling cascade which activate the expression of appropriate responses. In
contrast to signal perception various components of the signal
transduction have been identified, although it is largely unknown
how the different molecules interact with each other and where
they are positioned in the complex signalling network. These
signalling pathways comprise a network of protein-protein reactions and signalling molecules (for example, ROS, Ca2+ etc.).
Reversible phosphorylation of proteins is an important mechanism, by which organisms regulate cellular processes in response
to environmental cues. In this review, we will consider several
classes of protein kinases and phosphatases as signal transducers that were shown to be involved in osmotic stress signalling.
This will be followed by a description of the role of calcium as
second messenger molecules.
A.

MAPKinase Pathways
Protein phosphorylation is one of the major mechanisms for
controlling cellular functions in response to external signals. The
mitogen-activated protein kinase (MAPK) cascades are common
signalling modules in eukaryotic cells including plants. A general feature of MAPK cascades is their composition of three
functionally linked protein kinases. MAP kinase activation requires the phosphorylation of conserved threonine and tyrosine
residues in the so-called TEY (Thr, Glu, Tyr) activation loop by
a specific MAPK kinase (MAPKK). A MAPKK kinase (MAPKKK) activates through phosphorylation of conserved threonine
and/or serine residues. At the downstream end of the cascade, activation of the cytoplasmic MAPK module often induces translocation of the MAPK into the nucleus where the kinase is able
to activate genes through phosphorylation of transcription factors (Triesmann, 1996). In some other cases, a given MAPK

31

may translocate to other sites in the cytoplasm to phosphorylate specific enzymes or cytoskeletol components (Robinson and
Cobb, 1997). By tight regulation of the MAPK localization and
through expression of signalling components and substrates in
target cells, tissues or organs, MAPK pathways can mediate
signalling of an extracellular stimulus and bring about specific
responses.
MAPK pathways may integrate a variety of upstream signals
through interaction with other kinases or G proteins (Robinson
and Cobb, 1997). The G proteins often directly serve as coupling
agent between a plasma membranelocated receptor protein that
senses an extracellular stimulus and a cytoplasmic module. G
proteins and kinases in yeast and mammals have been shown to
regulate MAPKKKs (Kyriakis and Avruch, 2001). The actual
mechanisms of MAPKKK activation by osmotic stress in mammalian cells remain largely uncharacterized. Mammalian cells
activate three different MAPKs in response to osmotic stress:
P38, JNK, and ERK5 (extracellular signal regulated protein
kinases) (de Nadal et al., 2002). When compared with mammalian MAPKs, all plant MAPKs have highest homology to the
ERK subfamily. Control of gene expression is a major outcome
of stress-activated MAPK pathways. In mammalian cells, P38
(MAPK) controls the expression of >100 genes, while in yeast,
genome-wide transcription studies revealed that a large number
of genes (7%) show transient changes in their expression levels
after a mild osmotic shock and that the HOG1 MAPK pathway
plays a key role (de Nadal et al., 2002).
At least 20 MAPK, 10 MAPKK and 60 MAPKKK genes
have been identified in Arabidopsis on the basis of sequence similarities (Riechmann et al., 2000; Ichimura et al., 2002). Given
the imbalance in numbers it is likely that the pathways are not
being linear and convergence of pathways is expected. We are
already aware that some MAPKinase-activated pathways apparently overlap, i.e., several genes are induced by more than one
stressor (Knight and Knight, 2001). For instance, AtMPK6 and
AtMPK3 are activated by osmotic stress in Arabidopsis, and the
tobacco orthologs SIPK (salicylic acidinducible protein kinase)
and WIPK (wound-inducible protein kinase) are also activated
by biotic stresses indicating the point of convergence of different signalling cascades and yet leading to appropriate responses
(Singh et al., 2002). Recent studies suggest that different stimuli
activate MAPKs to different levels and with different kinetics,
which may encode signal specificity. Thus, the same MAPKs
may participate in different signalling events (reviewed in Tena
et al., 2001).
Moderate and extreme hyperosmotic stress activated two distinct kinases in alfalfa cells, a 46 kDa MAP kinase identified to
be SIMK under moderate osmotic stress conditions, whereas a
38 kDa protein kinase becomes activated under extreme hyperosmotic stress conditions (Munnik et al., 1999). This situation
resembles the operation of the osmosensing SLN1and SHO1
pathways in yeast, and suggests the possible existence of distinct sensors for moderate and extreme hyperosmotic stress in
plants. To determine the upstream activator of SIMK, Kiegerl

32

D. BARTELS AND R. SUNKAR

et al. (2000) used SIMK as bait in yeast two-hybrid screening


and isolated MAPKK (SIMK kinase). SIMKK encodes a functional protein kinase that specifically activates SIMK in vitro and
in vivo. SIMKK phosphorylates SIMK on the threonine and tyrosine residues of the activation loop, establishing that SIMKK
is a specific dual-specificity protein kinase of SIMK (Kiegerl
et al., 2000). The interaction of SIMK with SIMKK was further
confirmed in an in vivo system. SIMK is only partially activated
by NaCl in transfected parsley protoplasts, whereas it can be
fully activated by co-transfection of SIMKK.
In Arabidopsis, AtMEKK1 (a MAPKinase-kinase-kinase)
and AtMPK3 (a MAPKinase) are activated by dehydration, touch
and cold (Mizoguchi et al., 1996). In accordance with a role for
AtMPK3 during dehydration, the closely related alfalfa homologue SAMK is also transcriptionally upregulated upon drought
stress (Jonak et al., 1996). In Arabidopsis, AtMPK4 and AtMPK6
are post-translationally activated by cold, osmotic stress, and
wounding (Ichimura et al., 2000). Recently, OXI1, a serine/
threonine kinase was identified as a downstream component
of ROS signalling, possibly acting upstream to the AtMPK3
and AtMPK6. OXI1 is transcriptionally upregulated by H2 O2 ,
wounding pathogen attack and osmotic stress (Rentel et al.,
2004). Downstream targets of OXI1 could be AtMPK3 and
AtMPK6, since the OXI1 is required for the activation of MPK3
and MPK6.
ADR1, a CC-NBS-LRR gene, that shows homology with
the serine/threonine protein kinases and that has been originally
implicated as a component of disease resistance turns out to be
important also for abiotic stress responses (Chini et al., 2004).
The ADR1 transgenic plants display dual but opposite responses
to dehydration and other abiotic stresses. These plants exhibit
improved dehydration tolerance but are hypersensitive to salinity
and high temperature stresses. The improved tolerance appears
to be correlated with the dehydration-responsive gene expression
(Chini et al., 2004). These observations implicate overlapping
biotic and abiotic stress signalling pathways using kinases as
converging points of stress signalling pathway and yet capable
of eliciting stress-specific responses.

B.

SNF-1-Like Kinases Are Involved in Osmotic


Stress Signalling
Another family of protein kinases are the SNF1/AMPactivated protein kinases, which were first analysed in yeast from
where the name originated (SNF = sucrose-nonfermenting).
These kinases may sense the ATP/AMP ratio and thus control
fluxes between anabolism and catabolism via transcription of
genes encoding enzymes related to carbohydrate metabolism.
In plants some members of this group of kinases are expressed
in response to dehydration or ABA. The kinases range between
40 and 50 kDa, and are activated by phosphorylation of serine
or threonine. SNF-1-related protein kinases (SnRKs) have been
classified into three families, SnRK1, SnRK2, and SnRK3 with
unknown function (Halford and Hardie, 1998). PKABA1 from

wheat (Anderberg and Walker-Simmons, 1992), ARSK1 (Arabidopsis root specific kinase 1) (Hwang and Goodman, 1995),
two kinases in Dunaliella (Yuasa and Muto, 1996), a maize
45 kDa kinase (Conley et al., 1997), SPK3 and SPK4-kinases
from soybean (Yoon et al., 1997), a 42 kDa kinase from tobacco (Mikolayczyk et al., 2000), a 38 kDa kinase in alfalfa
(Munnik et al., 1999; Munnik and Meijer, 2001), a guard celllocalized ABA-activated protein kinase, AAPK from Vicia faba
(Li et al., 2000), Arabidopsis OST1 protein kinase which is related to AAPK of Vicia faba (Mustilli et al., 2002) are predicted
to belong to this group of kinases and are activated in response
to osmotic stress. The rice genome appears to encode 10 protein kinases belonging to this class and surprisingly all are activated by osmotic stress, and three of them (SAPK8, SAPK9,
and SAPK10) are also activated by ABA (Kobayashi et al.,
2004). Both SAPK1 and SAPK2 from rice are highly homologous to PKABA1. Additionally, these kinases seem to be transcriptionally either upregulated or downregulated by osmotic
stress and ABA. The biological significance of the osmotic- or
ABA-activation of these kinases is largely unknown. However,
their activation in response to different abiotic stresses to different levels implicates a major role for these kinases.

C.

Phosphatases
The action of the protein kinases is counteracted by phosphatases providing modulation and reversibility of the phosphoregulatory mechanism. Phosphatases are classified according to their substrate specificity. There are two major groups
of phosphatases: phosphoprotein (serine/threonine) phophatases
(or PPases) and phosphotyrosine (protein tyrosine phosphatases
or PTPases). PPases are classified into four groups (PP1, PP2A,
PP2B, and PP2C) based on their biochemical and pharmacological properties (Cohen, 1989). The PTPases form three subgroups: receptor-like PTPases, intracellular PTPases, and dualspecific PTPases.
Two major families of phosphatases interact with and inactivate HOG1 in yeast: the serine/threonine protein phosphatase
type 2C (PP2C) and the protein tyrosine phosphatases (PTPases)
(reviewed in de Nadal et al., 2002). Tyrosine specific phosphatases (PTPases) play a major role in the regulation of MAPK
pathway in yeast (Shinozaki and Russel, 1995). PTP2 and PTP3
are major tyrosine specific PTPases responsible for dephosphorylation and inactivation of HOG1 (Wurgler-Murphy et al.,
1997). Expression of genes encoding PTPases is often upregulated by the MAPK pathway, forming a negative feedback loop
for MAPK regulation (Jacoby et al., 1997; Wurgler-Murphy
et al., 1997). Extrapolating from mammals, transient and lowlevel MAPK activation may contribute to stress tolerance in
plants, whereas prolonged and high level activation may be
detrimental to the organism. Salt stress upregulated transiently
the AtPTP1 gene in Arabidopsis (Xu et al., 1998), which is
downregulated by cold stress. This implicates a unique mechanism for AtPTP1 in response to environmental stresses. AtPTP1

DROUGHT AND SALT TOLERANCE IN PLANTS

dephosphorylates AtMPK4 resulting in a complete loss of enzyme activity (Huang et al., 2000). This is important, because
both AtPTP1 and ATMKP4 respond to salt stress (Xu et al.,
1998).
The Arabidopsis mutant Atmkp1 is hypersensitive to genotoxic stress conditions such as UV-C and MMS (Ulm et al.,
2001). AtMKP1 codes for a dual-specificity kinase phosphatase.
Disruption of Atmkp1 has no effect on the plant phenotype under
normal conditions, however, the mutant exhibits hypersensitivity to stress caused by xenotoxic stresses but not to other abiotic
stresses. In wild-type plants, AtMKP1 is specifically activated
in response to UV and MMS. This is the only in vivo analysis
of a phosphatase available up to now that has a role in stress
conditions.
Protein phosphatases 2Cs are serine/threonine phosphatases
and their involvement in stress is well studied in fungal models. In yeast PP2C interacts with a MAP kinase cascade that
controls osmolyte biosynthesis (HOG pathway). Several PP2Cs
(Mpcs) from M. crystallinum have been isolated and studied for
their tissue, developmental stage- and stress-specific responses.
Out of ten PP2Cs, four transcripts, MPC2, MPC3, MPC5, and
MPC8, are transiently increased by salt and dehydration, suggesting a role for these PP2Cs during stress (Miyazaki et al.,
1999). Arabidopsis PP2Cs appear to be the largest protein phosphatase family with 76 genes (reviewed in Schweighofer et al.,
2004). Most of the research is focused on PP2C action in relation to ABA signalling. PP2C has been described as abi1 and
abi2 Arabidopsis mutants defective in a PP2C isoform (Leung
et al., 1994), which seems to function as a negative regulator
in a pathway that mediates responses to environmental stresses
involving ABA. However, the substrates of these PP2C in Arabidopsis are still unknown. Additional studies have shown that
guard cells of abi1-1 and abi2-1 plants are disrupted in ABA activation of hyperpolarization-activated Ca2+ (ICa) channels (Allen
et al., 1999; Murata et al., 2001). Further, experiments on abi11 and abi2-1 mutants revealed both PP2Cs are acting at different levels of the same pathway, i.e., abi1-1 acts upstream and
abi2-1 downstream of ABA-induced ROS production in guard
cells (Murata et al., 2001). It was found that ABI1 can interact
with the ABA-inducible transcription factor ATHB6. ATHB6
promoter-reporter expression was abrogated in abi1-1 mutant,
suggesting that ABI1 acts upstream of the transcription factor
ATHB6 (Himmelbach et al., 2002).
Recently, the involvement of protein phosphatases in stomatal regulation has been demonstrated using pharmacological
approaches. Application of phenylarsine oxide, a specific inhibitor of protein tyrosine phosphatase prevented stomatal closure in response to four stomatal closing signals such as ABA,
H2 O2 , Ca2+ and dark (MacRobbie, 2002). This suggests that
protein tyrosine dephosphorylation is involved mostly downstream of the Ca2+ signalling which is responsible for stomatal
closure. Stomatal aperture regulation is dependent on ion efflux from guard cells. Identification of the target protein whose
dephosphorylation results in activation of ion release from the

33

vacuole will allow us to connect the signalling events in response


to osmotic stress and ABA.
D.

Phospholipid Signalling
The plasma membrane must play an important role in perceiving and transmitting environmental signals. Osmotic stress
often leads to altered membrane fluidity and changes in phospholipids have recently been recognized as important events
mediating osmotic stress signals in plants (Munnik and Meijer,
2001). The current hypothesis is that phopspholipids are cleaved
by phospholipases, which produce phospholipid-derived second
messengers. In plants, like in other organisms four major classes
of phospholipases are distinguished based on their cleavage site:
phospholipase C (PLC), phospholipase D (PLD), and phospholipase A1 and A2 (PL A1 and PL A2) (Wang, 2002). Phospholipid signalling may be regulated through G-proteins and may
be tightly linked with calcium. The major phospholipid-derived
signalling molecules which will be considered in the context of
osmotic stress are inositol 1,4,5-triphosphate (IP3 ), diacylglycerol (DAG) and phosphatidic acid (PA).
1.

Inositol 1,4,5-Triphosphate (IP3)


PLC cleaves the phospholipid phosphatidylinositol 4,5bisphosphate (PIP2) into the soluble IP3 and the membrane
bound DAG. Osmotic stress rapidly increases PIP2 synthesis
in Arabidopsis (Pical et al., 1999; De Wald et al., 2001) and concomitantly the transcript levels increase of PIP5K, a
phosphatidylinositol-kinase, that synthesizes PIP2 (Mikami
et al., 1998). PIP2 is a signal molecule which in animal cells
leads to K+ - desensitization (Kobrinsky et al., 2000). If this
could also be confirmed for plant cells, then ion flux and osmotic
stress can be linked via PIP2. There are reports from several plant
species that osmotic stress leads to increased PLC transcript levels. One of the Arabidopsis phospholipase C genes (AtPLC1) is
induced by dehydration, salinity and low temperature (Hirayama
et al., 1995). Takahashi et al. (2001) have shown that the hyperosmotic stress induces a rapid and transient elevation in IP3
levels due to activation of PI-PLC in Arabidopsis cell cultures.
Two genes, VuPLC1 and VuPLC2, were isolated from droughttolerant and -sensitive varieties of cowpea. Whereas VuPLC1
is constitutively expressed and decreases under drought stress,
VuPLC2 transcript levels accumulate under water stress in the
tolerant plant and progressively declines in the sensitive plant
(El-Maarouf et al., 2001).
Involvement of PLC-genes in dehydration has also been observed in potato (Kopka et al., 1998). PLC activation leads to
the synthesis of IP3 and DAG. IP3 then releases Ca2+ from internal stores (Sanders et al., 1999; Schroeder et al., 2001), while
DAG may be converted to PA or activates a protein kinase C,
which, however, has not been isolated yet in plants. Hyperosmotic stress increased IP3 levels were reported in several studies on osmotic stress (Takahashi et al., 2001; De Wald et al.,
2001; Drbak and Watkins, 2000). PI-PLC was also shown to

34

D. BARTELS AND R. SUNKAR

be involved in the regulation of stomatal movements. Droughtinduced activation of PI-PLC led probably via IP3 to an increase
in cytosolic Ca2+ in guard cells, which triggers stomatal closure
and thus presents a drought avoidance mechanism (Staxen et al.,
1999).
2.

Phosphatidic Acid (PA)


Phosphatidic acid (PA) is another second messenger in animal
cells that can activate PLC and protein kinase C (Munnik and
Meijer, 2001). PA is synthesized through cleavage of membrane
phospholipids by PLD. The predicted structure of plant PLDs
differ from that of yeast and animals in that most plant PLDs
contain not only the conserved catalytic motifs, but also a Ca2+ binding domain (a C2 domain) which is not found in the PLDs
of other organisms (Wang, 2001). This feature indicates a direct
regulation of PLD by calcium, in addition to changes in PLD
gene expression.
Elevated PLD activity was found to be correlated with droughtstress in a drought-sensitive cowpea strain, when it was compared with a drought-tolerant strain (El-Maarouf et al., 1999).
Further evidence was provided by Frank et al. (2000) by isolating
two phospholipase D cDNAs from C. plantagineum: CpPLD1
is constitutively expressed and is likely to be involved in early
responses to dehydration, producing the second messenger phosphotidic acid to amplify the signal after its perception, while the
dehydration-responsive CpPLD2 may be involved in phospholipid metabolism and membrane rearrangements at later stages
of dehydration. Four PLD genes were analyzed for their response to osmotic stress conditions in Arabidopsis (Katagiri
et al., 2001).
Only the transcript levels of PLD were transiently elevated
in response to dehydration, whereas the expression was rapidly
induced by high salt stress. Plants transformed with a PLD
promoter fused to a reporter gene showed strong induction under dehydration, mainly in vascular tissues of leaves, cotyledons and roots. Increased PLD activity coincides with the elevation in phosphatidic acid levels in response to dehydration.
Evidence to support that the elevated phosphatidic acid levels
were linked with stress-increased PLD activity was provided
by antisense PLD transgenic plants, in which the dehydrationinduced production of phosphatidic acid accumulation was substantially reduced (Katagiri et al., 2001). In castor bean leaves
PLD mRNA and enzymatic activity increased by ABA treatment. Further, an increase was observed in the proportion of
enzyme activity associated with microsomal membranes (Ryu
and Wang, 1995) implying both enzyme level and its location
are important factors in the ABA response. Like PLC, PLD has
a role in the regulation of the stomatal aperture during osmotic
stress. ABA-promoted stomatal closure was shown to be mediated by guard cell PLD activity in response to water stress
(Jacob et al., 1999). Antisense suppression of PLD impairs
stomatal closure mediated by ABA or water stress and increases
water loss in Arabidopsis, whereas the overexpression has the
opposite effect and leads to a decreased water loss by enhancing

the sensitivity to ABA (Sang et al., 2001). These experiments


provided direct evidence for PLD in regulating water loss. The
involvement of PLD in ABA responses obtained further support
from experiments using rice protoplasts. The expression of several ABA-related genes was repressed when PLD-derived PA
was blocked, which indicates that PLD activity is important for
ABA-inducible gene expression (Gampala et al., 2001; 2002).
Osmotic stress activated PLDs implies a role for PA that is produced by PLD. The targets of PA in plants are unknown, however,
PIP kinase, PDK (phosphoinositide dependent kinase), MAPK
pathway, MGDG synthetase, K+ channel are possible targets
(Munnik, 2001).

E. Other Signalling Molecules


1. Salicylic Acid
It has been established for quite some time that salicylic acid
(SA) plays an important role in the defense response in many
plant species to pathogen attack. SA mediates the oxidative burst
that leads to cell death in the hypersensitive response, and acts
as a signal for the development of the systemic acquired resistance (Shirasu et al., 1997). Recently the involvement of SA in
osmotic stress was demonstrated by using an SA-deficient transgenic line expressing a salicylate hydroxylase (NahG). Wild type
seeds germinated in the presence of NaCl or mannitol showed
extensive necrosis in the shoot, but not in the NahG mutant. Wild
type and NahG behaved similarly during germination at various
Li+ concentrations, excluding the possible involvement of ionic
components. Greater oxidative damage occurred in wild-type
seedlings compared with NahG seedlings under NaCl stress.
Methyl viologen treatment resulted in necrotic phenotypes only
in wild-type plants. These different observations lend support to
the hypothesis that SA potentiates the effects of salt and osmotic
stress by enhancing the generation of ROS during photosynthesis (Borsani et al., 2001).
2.

Nitric Oxide (NO)


Recent studies revealed that NO is an important signalling
molecule involved in several physiological functions ranging
from plant development to defense responses (Wendehenne et al.,
2001). Nitric oxide is a labile free radical that is produced from
L-arginine by NO synthase in various mammalian cells, where
it has been shown to be protective against damages caused by
oxidative stress conditions. Exogenous application of NO improved water stress tolerance both in wheat seedlings and in
detached leaves (Mata and Lamattina, 2001). Detached leaves
pretreated with the NO releasing agent SNP withstand the imposed stress by retaining a higher water content, lower ion leakage, and greater accumulation of LEA 3 transcripts compared to
control leaves. The drought tolerance of NO could be attributed
to its ability to maintain higher RWC, decreasing the rate of
transpiration by closing stomata and the ability to induce gene
expression involved in stress tolerance.

DROUGHT AND SALT TOLERANCE IN PLANTS

V.

CALCIUM SIGNALLING DURING DEHYDRATION


AND SALT STRESS
In plant cells, calcium functions as a second messenger coupling a wide range of extracellular stimuli to intracellular
responses (Snedden and Fromm, 1998, 2001). Different extracellular stimuli elicit specific calcium signatures: kinetics,
amplitude and duration of Ca2+ transients specify the nature
and the intensity of stimulus. To date, three major classes of
Ca2+ sensors have been characterized in plants. These classes
are calmodulin, CDPKs (calcium-dependent protein kinase) and
CBLs (calcineurin B-like proteins) (Yang and Poovaiah, 2003).
Several lines of evidence suggest that all these three classes of
Ca2+ sensors are involved in stress signal transduction (Snedden
and Fromm, 2001; Luan et al., 2002; Zhu, 2000).
The involvement of Ca2+ signalling in response to osmotic
and ionic stress is well documented. NaCl causes a rapid and
transient increase in cytosolic calcium, that in turn triggers many
signal transduction pathways, including the regulation of enzymatic activity, ion channel activity, and gene expression which
results in diverse cellular responses (Snedden and Fromm, 1998,
2001) and mediates salt adaptation (Bressan et al., 1998; Liu
and Zhu, 1998; Serrano et al., 1999). Recent progress provided
insights into how the response to different osmotic stresses is encoded in the spatial and temporal dynamics of the Ca2+ signal
(Knight and Knight, 2001). The role of calcium and its dynamics were investigated using either pharmacological approaches
or using transgenic plants which express the calcium reporter
protein aequorin in different cellular compartments or under the
control of promoters with different responsiveness to environmental stimuli. This has allowed determining calcium fluxes
within one cell and in different tissues. For instance, mannnitol
and sodium chloride-induced increased cytosolic [Ca2+ ] is due
to release of calcium from the vacuole (Knight et al., 1997).
Cell-type specific changes in cytosolic calcium levels were observed in Arabidopsis root cells in response to drought, salinity,
and low temperature (Kiegle et al., 2000).
A.

Calcium-Dependent Protein Kinases (CDPKs)


Osmotic stress-induced CDPKs have been reported from several plants (Kawasaki et al., 2001; Seki et al., 2002; Ozturk et al.,
2002). Recent experiments on salt-tolerant and salt-sensitive rice
varieties have strengthened the importance of CDPKs in osmotic
stress responses. A specific CDPK is induced earlier and its expression is sustained for longer duration in the tolerant variety
compared to the sensitive variety (Kawasaki et al., 2001). Further
evidence for the involvement of CDPKs in stress was obtained
from studying Arabidopsis CDPKs (Sheen, 1996). Out of several CDPKs tested, only AtCDPK1 and AtCDPK1a were able
to transcriptionally activate selected reporter genes indicating
that specific CDPK isoforms mediate the effects of stress. Interestingly, in the salt-tolerant Mesembryanthemum crystallinum
osmotic stress-induced CDPK (McCDPK1) was shown to interact with CSP1 (calcium-dependent protein kinase Substarte Pro-

35

tein1), which is a transcription factor belonging to a class of twocomponent pseudoresponse regulators (Patharkar and Cushman,
2000). These results establish a role for specific CDPKs in stressinduced gene expression.
B.

Calcium-Binding Proteins
Modulation of intracellular calcium levels is partly regulated
by calcium-binding proteins such as calmodulin, which is activated by increased calcium concentrations and then induces
specific kinases. The importance of the calcium-binding proteins has first been derived from yeast mutant analysis. Mutations in calmodulin genes of yeast render the yeast cells sensitive to high NaCl concentrations and NaCl-induced genes are
no longer induced, suggesting that calmodulin is involved in
the NaCl-stress signal transduction pathway (Cunningham and
Fink, 1996). The regulation of calcium-binding proteins can occur in a cell- or tissue-specific manner as the following examples
illustrate. Ca2+ /calmodulin dependent kinases have been shown
to be regulated by salt stress, e.g., PsCCaMK, a Ca2+ /calmodulin
dependent protein kinase in pea was specifically upregulated by
NaCl in roots, while the shoot kinase was not affected (Pandey
et al., 2002). A family of calmodulin binding transcription activators were first discovered in drought-stressed Brassica napus
and were subsequently shown to be only present in multicellular
organisms (Bouche et al., 2002). Further studies are needed to
determine the importance of this family during osmotic stress.
Other examples for osmotic stress-activated calcium-binding
proteins are the Arabidopsis protein AtCP1, the membraneassociated rice protein OsEFA27, and the Arabidopsis counterpart RD20 (Frandsen et al., 1996; Jang et al., 1998; Takahashi
et al., 2001). A recently identified calmodulin binding protein
in Arabidopsis, AtCaMBP25, has been implicated as a negative
regulator of osmotic stress. AtCaMBP25 is upregulated by osmotic stress and its overexpression leads to sensitivity to osmotic
or salt stresses, whereas the suppression by antisense technology
resulted in improved tolerance (Perruc et al., 2004) suggesting
that AtCaMBP25 is a negative regulator of osmotic and salt
stresses. Further, microarray analysis in these plants will reveal
if there is a correlation between the observed phenotypes and
altered gene expression.
Ca2+ -Mediated SOS Pathways Are Involved
in Ion Homeostasis
Insight into the underlying mechanisms of external calcium
on cellular responses to salt has been provided by the isolation of
salt hypersensitive mutants sos1, sos2 and sos3 from Arabidopsis
thaliana (Zhu, 2003). Cloning and characterization of SOS genes
has led to the discovery of a novel Ca-dependent pathway for
the regulation of ion homeostasis and plant salt tolerance (see
later in Na+ extrusion).
C.

D.

Calcineurin B (CBLS) and Osmotic Stress Responses


Despite the genetic analysis of the interaction of the SOS 1
to 3 genes, the complexity of the calcium signals requires a lot

36

D. BARTELS AND R. SUNKAR

more mutant analysis which becomes apparent by the fact that


Arabidopsis thaliana has at least 10 calcineuron B-like (CBL)
genes encoding highly similar but functionally distinct Ca2+ binding proteins (Kudla et al., 1999). Drought, high salt, cold
and wound signals induce AtCBL1 gene transcripts (Kudla et al.,
1999). Both CBL1 and CBL2 respond to light, but CBL2 lacks
the other responses of CBL1 (Nozawa et al., 2001). Knockout
mutant plants of SCaBP5/CBL1 displayed hypersensitivity to
drought and salinity, in contrast to CBL1 overexpressing plants
that showed drought tolerance (Albrecht et al., 2003). Such expression patterns suggest that CBL1 and CBL2 have both overlapping and specific functions.
Ca2+ ATPases
The major physiological role of Ca2+ -ATPases is to restore
and maintain homeostasis by pumping Ca2+ out of the cytosol
to terminate a signalling event, which is critical in all eukaryotic
cells (Sze et al., 2000). Plant and animal cells utilize two distinct
types of Ca2+ pumps, identified as type IIA and type IIB, based
on their protein sequences. NaCl stress has been reported to induce the expression of genes encoding the type IIA Ca2+ pump
in tobacco, tomato, and soybean (Wimmers et al., 1992; Chung
et al., 2000). The consequence of upregulating the Ca2+ pump
in response to NaCl is not known, but it was speculated that it
is likely to provide an adaptive response such that a stimulated
cell would acquire an enhanced efflux capacity capable of subsequently decreasing the magnitude or duration of a calcium release to further exposure to a given stimulus (Chung et al., 2000).
The soybean Ca2+ -ATPase1 was induced by NaCl but not by
KCl and mannitol, indicating that specific calcium signals trigger an increase in transcription (Chung et al., 2000). Geisler et al.
(2000a) have cloned and characterized Ca2+ -ATPase isoform 4
(ACA4), a calmodulin-regulated Ca2+ -ATPase from Arabidopsis. Two lines of evidence suggest that ACA4 might be part of
the Ca2+ -dependent signal transduction pathway linked to salt
stress: (1) Arabidopsis seedlings treated with different concentrations of NaCl for 24 h showed a dose-dependent increase in
ACA4 transcript levels, (2) when N-terminal truncated ACA4
was expressed in yeast, it conferred increased NaCl tolerance to
its host (Geisler et al., 2000b).
E.

VI.

TRANSCRIPTIONAL REGULATION
OF GENE EXPRESSION
Stress responses primarily include transcriptional regulation
of gene expression and this depends on the interaction of transcription factors with cis-regulatory sequences. In several instances, the quantity and availability of regulatory proteins may
depend on their own expression patterns. Such autocatalytic controls may be exerted on the transcriptional, post-transcriptional
or translational level. Phosphorylation of regulatory proteins is
a major event in controlling the gene expression in eukaryotes. Therefore, multiple protein-protein and/or protein-DNA
interactions frequently determine the rate of transcription by

activation/repression of a promoter under given environmental


conditions.
Although the cis-elements of dehydration and ABA-responsive
genes have been intensively studied, our understanding is limited. More efforts are still needed to identify additional novel
cis- and trans-acting elements that function in ABA-dependent
as well as ABA-independent gene expression. Presently, two
classes of drought-responsive DNA elements have been well
characterized: the ABA-responsive element (ABRE) and the
dehydration responsive element (DRE, also referred to as Crepeat; Baker et al., 1994). It is highly likely that yet unidentified
osmotic-stress responsive cis-elements exist in plants.

A.

ABA Response Elements (ABREs)


Most, but not all of the dehydration-induced genes are also induced by the application of the phytohormone ABA
(Chandler and Robertson, 1994; Leung and Giraudat, 1998) (for
details see section XI). Until now, two ABA-dependent pathways
are known to mediate gene expression in plants during osmotic
stress. The distinction is largely based on cis-elements that exist
in the promoters of ABA-inducible genes. The ABA-dependent
pathways are thought to mediate the gene expression through
an ABRE-element and b-ZIP transcription factors (Busk and
Pages, 1998), while the other pathway is through a MYC and
MYB elements and transcription factors (Yamaguchi-Shinozaki
and Shinozaki, 1993).
Many cis-regulatory elements known as ABA-responsive elements (ABREs) have been identified. Among them, (C/T)
ACGTG(G/T)C motifs have been reported to function as ABREs
in many genes (Guiltinan et al., 1990; Mundy et al., 1990; Ingram and Bartels, 1996; Busk and Pages, 1998). The core element of these ABREs is the CACGTG motif also known as
G-box motif, which functions in the regulation of plant genes
stimulated by a variety of environmental signals. Systematic
DNA-binding studies have shown that nucleotides flanking the
ACGT core specify the DNA-binding interactions and subsequent gene activation (Williams et al., 1992; Izawa et al., 1993).
Nevertheless, in promoters such as CDeT27-45 and CDeT6-19,
isolated from C. plantagineum, G-box related ABREs do not appear to be major determinants of the ABA or drought response
(Michel et al., 1993, 1994).
Ho and coworkers identified cis-elements called coupling
elements which are active in combination with an ABRE but
not alone (Shen et al., 1996). In the promoters of the barley
genes HVA22 and HVA1 the coupling elements CE1 and CE3
(ACGCGTGTCCTC) are necessary for activation by ABA. Dissection of these promoters defined ABA response complexes
(ABRC) consisting of a coupling element and an ABRE capable of conferring ABA-inducible transcription. Other ABREs
that do not belong to the above groups have also been reported:
a Sph element-containing sequence (CGTGTCGTCCATGCAT)
of the maize C1gene (Kao et al., 1996), the MYB and the MYC
binding sites of the Arabidopsis rd22 gene (Abe et al., 1997),

DROUGHT AND SALT TOLERANCE IN PLANTS

and a novel element present in the CDeT27-45 gene of C. plantagineum (Michel et al., 1993).
B.

The Dehydration Response Element (DRE)


The investigation of dehydration-induced genes in Arabidopsis has also revealed ABA-independent signal transduction pathways (Shinozaki and Yamaguchi-Shinozaki, 2000). In aba (ABA
deficient) and abi (ABA insensitive) mutants, several genes are
induced by dehydration indicating that these genes do not require ABA for their expression under drought conditions. The
A. thaliana genes rd29A (also described as COR 78 or LTI78) and
rd29B are differentially induced under conditions of dehydration, salt, cold stress, or ABA treatment. This multiple expression
pattern requires at least two cis-acting elements for the rd29A
gene. The 9-base pair direct repeat (TACCGACAT), termed the
dehydration responsive element (DRE), functions in the initial
rapid response of rd29A to dehydration, salt, or low temperature. The slower ABA response is mediated by another promoter fragment that contains an ABRE (Yamaguchi-Shinozaki
and Shinozaki, 1993, 1994). The DRE is an essential cis-acting
element for the regulation of rd29A induction in the ABAindependent response to dehydration in Arabidopsis. DRErelated motifs have been reported in promoters of several genes
regulated by osmotic and low temperature stress, including kin1,
cor6.6/kin2, and rd17/cor47 in Arabidopsis (Wang et al., 1995;
Iwasaki et al., 1997). A similar motif was also reported in the promoter region of cold and dehydration-inducible cor15A (Baker
et al., 1994). In a recent study 16 genes containing DRE (TACCGACAT) or DRE-related core motif (CCGAC) were identified
in the promoters of drought stress inducible Arabidopsis genes,
which are likely targets of DREB1 or DREB2 transcription factors (see below) (Seki et al., 2002). In another microarray experiment which analysed DREB1A overexpressing plants Seki
et al. (2001) have identified 12 target genes of DREB. Among
them, 11 promoters contain the CCGAC sequence motif, and
6 promoters contain ABRE elements (Seki et al., 2001), which
implies additional unidentified stress responsive cis-elements.
C.

Transcription Factors Modulated by Osmotic Stress


Plants need a large number of transcription factors governing
proper and strict transcriptional regulation in response to developmental and environmental cues. Indeed, over 5 percent of the
Arabidopsis genome is devoted to encoding more than 1,500
transcription factors (Riechmann et al., 2000). Despite the fact
that not many stress-specific consensus sequences were identified in promoters of stress specific genes transcription factors
belonging to different families have been shown to be modified by stress. To activate or repress transcription, transcription
factors must be located in the nucleus, bind DNA, and interact
with the basal transcription apparatus. Therefore, environmental
signals that regulate transcription factor activity may affect any
one or a combination of these processes. Regulation of a transcription factor is achieved by reversible phosphorylation or by

37

de novo synthesis of transcription factors. The current analysis


of identified transcription factors suggests that different stress
signalling pathways may overlap or converge at specific points.
1.

Basic Region Leucine Zipper (bZIP) Proteins


Basic region leucine zipper (bZIP) proteins contain a DNAbinding domain rich in basic residues and adjacent to a leucine
zipper dimerization domain. bZIPs are a large family of transcription factors in plants and are represented by 75 members
in Arabidopsis (Jacoby et al., 2002). All plant bZIP proteins
bind to an ACGT core sequence, but the sequences flanking the
core sequence affect the precise DNA binding. Several bZIP
factors that bind ABREs have been cloned as candidates for
ABA-responsive transcription factors that induce gene expression by osmotic stress and/or ABA. One bZIP subfamily has
been linked genetically to an ABA response: ABI5 and its homologs, the ABRE binding factors (ABFs/AREBs). ABRE binding factors (ABFs)/ABA-responsive element binding (AREBs)
proteins respond at the transcriptional and post-transcriptional
level to dehydration and salt stress (Choi et al., 2000; Uno et al.,
2000). ABF3 and ABF4 overexpression in transgenic Arabidopsis plants resulted in enhanced drought tolerance accompanied
by decreased transpiration, suggesting that ABF3 and ABF4 are
involved in stomatal closure mediated by ABA (Kang et al.,
2002). Promoters of both ABF3 and ABF4 were found to be
most active in roots and guard cells, consistent with their roles
in stomatal regulation and water stress response.
2.

Homeodomain-Leucine Zipper Proteins (HD-ZIP)


HD-ZIP genes encode proteins that have only been identified in plants so far and are thought to regulate developmental
processes and responses to environmental cues. HD-ZIP proteins are characterized by the presence of a DNA-binding homeodomain with a closely linked leucine zipper motif. The activity
of HD-ZIP resides primarily in the specific DNA-binding property of the homeodomain and the ability of the leucine zipper
to mediate protein-protein interaction with other HD-ZIPs. HDZIPs have been described for several plant species such as Arabidopsis, carrot, tomato, rice, sunflower, and C. plantagineum.
Seven HD-ZIP are affected by dehydration in C. plantagineum,
which are placed in different branches of the signalling network
(Deng et al., 2002).
In Arabidopsis three HD-ZIP genes (ATHB-6, ATHB-7, and
ATHB-12) are shown to be responsive to dehydration (Soderman
et al., 1996, 1999; Lee and Chun, 1998). Hahb-4 is upregulated by drought and ABA in sunflower roots, stems, and leaves
(Gago et al., 2002). HD-ZIP proteins can form both homo- and
heterodimers, which gives many degrees of freedom in regulation processes. Direct targets of HD-ZIP genes have still to be
identified.
3.

Zn-Finger Proteins
A zinc-finger motif represents the sequence in which cysteines and/or histidines coordinate a zinc atom(s) to form local

38

D. BARTELS AND R. SUNKAR

peptide structures that are required for their specific functions.


The zinc-finger motifs are classified based on the arrangement
of the zinc-binding amino acids. They play critical roles in interactions with other molecules. The cDNA clone for Alfin1 was
isolated by differential screening of salt-tolerant alfalfa cells
grown on NaCl (Winicov, 1993). Alfin1 cDNA encodes a novel
member of putative Zn-finger proteins, which binds to promoter
sequences of MsPRP2 in vitro (Bastola et al., 1998). Overexpression of the Alfin1 gene in transgenic plants increased the
MsPRP2 transcript levels in root and enhanced NaCl tolerance,
indicating that the Alfin1 gene product regulates MsPRP2 expression in vivo (Winicov and Bastola, 1999). The function of
MSPRP2 is yet unknown, however, it was predicted that it is
likely to be a cell wall protein with potential as an anchor to the
membrane.
4.

AP2/ERF-Type Transcription Factors


The group of AP2/ERF DNA binding proteins were identified in different scientific contexts and the first proteins described gave this group its name. AP2/ERF domain proteins
include the DREB or CBF proteins binding to dehydration response elements (DRE) or C-repeats. The Arabidopsis genome
encodes 145 DREB/ERF-related proteins (Sakuma et al., 2002).
The Arabidopsis AP2 proteins have been classified into five
classes based on similarities in their DNA-binding domains:
AP2 subfamily (14 genes), RAV (related to ABI3/VP1) subfamily (6 genes), DREB subfamily (55 genes; Group A), ERF
subfamily (65 genes, Group B), and others (4 genes) (Sakuma
et al., 2002). A major transcriptional regulatory system is represented by DRE/C-repeat promoter sequences in stress-activated
genes and DREBs/CBF factors that control stress gene expression (Stockinger et al., 1997; Liu et al., 1998). A C-repeat
binding factor, CBF1, was isolated by screening an Arabidopsis
cDNA expression library using the C-repeat of the cold stress
COR15a gene as target (Stockinger et al., 1997). The same
factor and other CBF family members, referred to as DREBs
(Dehydration Responsive Element Binding proteins) have been
reported from studies of dehydration stress genes (Liu et al.,
1998). A detailed expression analysis showed that these factors
may be associated with different physiological conditions. For
example, expression of DREB1A, B, C/CBF1, 2, 3 is induced
by low temperature, DREB1A, DREB1D/CBF4, DREB2A and
DREB2B are induced by salt and dehydration, DREB1F is induced by salt, whereas, DREB1E is induced by ABA (Liu et al.,
1998; Shinwari et al., 1998; Nakashima et al., 2000; Haake
et al., 2002; Sakuma et al., 2002). Several stress-inducible genes
such as rd29A, Cor6.6, Cor15a and Kin1 are target genes of
DREBs/CBFs in Arabidopsis and contain DRE/C-repeat sequences in their promoters. Seki et al. (2001) have identified six
new genes containing both DRE/C-repeat and ABRE motifs in
their promoters, implying complex regulation of stress-induced
genes by ABA-dependent and ABA-independent pathways. The
emerging picture from functional studies of CBFs/DREBs indicates that CBF genes coordinate both, activation or repression

of stress responsive genes. Ectopic overexpression of CBF1/


DREB1B and CBF3/DREB1A in Arabidopsis results in the constitutive expression of downstream stress-inducible genes and
enhances freezing tolerance and dehydration/salt tolerances, respectively (Kasuga et al., 1999; Gilmour et al., 2000). On the
other hand, the functional analysis of cbf2/dreb1c knock-out mutant plants revealed somewhat surprising results. Cbf2 mutant
plants displayed tolerant phenotype to dehydration and salinity (Novillo et al., 2004). Another interesting feature of these
mutant plants is enhanced expression of CBF1/DREB1B and
CBF3/DREB1A and the downstream genes such as LTI78, KIN1,
COR47, and COR15A which are known to impart stress tolerance. These results imply that CBF2 is a negative regulator of
CBF1 and CBF3 transcription factors and of the corresponding
downstream genes.
Rice AP2 transcription activators, OsDREB1A, OsDREB1B,
OsDREB1C, OsDREB1D and OsDREB2A have been isolated.
OsDREB2A was induced by dehydration and high salinity stress.
Overexpression of OsDREB1A in transgenic Arabidopsis improved stress tolerance (Dubouzet et al., 2003). Two novel DRE
binding factors (DBF1 and DBF2) were isolated from maize
(Kizis and Pages, 2002). DBF1 and DBF2 are capable of binding DRE motifs and are not homologues of DREBs from Arabidopsis. DBF1 is induced by dehydration, salinity, and ABA
whereas DBF2 is constitutively expressed at low levels and not
responsive to stress.
The Tsi1 (tobacco stress-inducible gene 1) from tobacco belongs to the ERF subfamily (Sakuma et al., 2002) and is induced
by NaCl, salicyclic acid, and ethylene. Tsi1 was able to bind GCC
motifs in both ethylene response elements and DRE/CRT sequences. Tobacco plants overexpressing Tsi1 showed enhanced
resistance to osmotic stress and pathogen attack (Tsugane et al.,
1999). This demonstrates that overexpression of a single transcription factor effects both biotic and abiotic stress responses
suggesting convergence of both signalling cascades.
5.

Myb-Like Proteins
The Myb-motif comprises three imperfect repeats forming a
helix-turn-helixrelated motif. Each repeat contains three conserved tryptophan residues every 18 to 19 amino acids, which
promotes a secondary structure with a functional Myb-domain.
In plants, the first tryptophan of R3 is substituted by phenylalanine or isoleucine. This latter amino acid is present in Atmyb2 in
Arabidopsis and myb-related genes from C. plantagineum cpMYB7 and cpMYB10 whose expression is upregulated by dehydration and ABA (Urao et al., 1993; Iturriga et al., 1996). Ectopic
expression of CpMYB10 in transgenic Arabidopsis plants resulted in drought and salinity tolerance (Villalobos et al., 2004).
The Myb gene family is represented by 190 genes in Arabidopsis (Riechmann et al., 2000). Atmyb-2 has been identified as a
positive regulator of stress-induced rd22 gene expression (Abe
et al., 1997). Consistent with the role, AtMYB-2 overexpressing transgenic plants exhibited hypersensitivity whereas knockout mutants showed insensitivity to ABA (Abe et al., 2003). In

DROUGHT AND SALT TOLERANCE IN PLANTS

addition, three additional Arabidopsis Myb-like transcription


factors were identified as upregulated by dehydration, high salinity, or cold stress in a microarray study (Seki et al., 2002).
6.

Myc-Like Proteins
Myc-like proteins contain the basic helix-loop-helix (bHLH)
domain, which is composed of two subdomains: the basic region (as found in bZIPs) responsible for DNA binding and the
HLH (helix-loop-helix) region for dimerization with interacting
proteins. The Arabidopsis rd22BP1 gene encodes a Myc-like
transcription factor and is induced by dehydration, high salt
conditions, and ABA. It activates the downstream rd22 gene
by binding to the myc promoter motif (Abe et al., 1997). The
expression of rd22BP1 (AtMYC-2) precedes the rd22 gene expression, which underlines the hierarchic relationship. Overexpression and knock-out mutants displayed contrasting phenotypes with respect to ABA sensitivity, similar to AtMYb-2 (Abe
et al., 2003). Hypersensitivity to ABA was enhanced in transgenics when both genes are overexpressed together (Abe et al.,
2003), confirming the cooperation of AtMYB-2 and AtMYC-2
interaction in vivo in stress-activated gene expression.
7.

CDT-1
A novel gene CDT-1 was isolated from C. plantagineum using a T-DNA activation tagging approach. A gain-of-function
phenotype in C. plantagineum calli was due to the activation
of the CDT-1 gene that confers desiccation tolerance to callus
cells and activates stress genes independent of exogenous ABA
(Furini et al., 1997). However, it is not known whether CDT-1
functions as RNA or as a peptide; recent experiments favor RNA
as functional molecule (Furini and Bartels, unpublished). There
is no direct sequence homologue to CDT-1 in Arabidopsis.
VII.

ACCUMULATION OF SUGARS
AND COMPATIBLE SOLUTES
Almost all organisms, ranging from microbes to animals
and plants, synthesize compatible solutes in response to osmotic stress (Burg et al., 1996). Compatible solutes are nontoxic molecules such as amino acids, glycine betaine, sugars, or
sugar alcohols. They do not interfere with normal metabolism
and accumulate predominantly in the cytoplasm at high concentrations under osmotic stress (reviewed in Chen and Murata,
2002). These molecules may have a primary role of turgor maintenance but they may also be involved in stabilizing proteins
and cell structures (Yancey et al., 1982). Initially it was thought
that compatible solutes have their main role in osmotic adjustment, but there is increasing discussion of other roles (Serraj and
Sinclair, 2002). The accumulation of these solutes per se may
not be important for osmotic stress tolerance but the metabolic
pathways may have adaptive value (Hasegawa et al., 2000). A
further hypothesis is that compatible solutes are also involved in
scavenging reactive oxygen species (Shen et al., 1997a, b; Hong
et al., 2000; Akashi et al., 2001; Chen and Murata, 2002). Engi-

39

neering the synthesis of compatible solutes has been a relatively


successful approach to obtain stress tolerant plants (see Table 1).
A.

Sugars
Several physiological studies suggested that under stress conditions nonstructural carbohydrates (sucrose, hexoses, and sugar
alcohols) accumulate although to varying degree in different
plant species. A strong correlation between sugar accumulation
and osmotic stress tolerance has been widely reported, including
transgenic experiments (Ramanjulu et al., 1994a; Abd-El Baki
et al., 2000; Gilmour et al., 2000; Streeter et al., 2001; Taji et al.,
2002).
The increase in sugars mostly results in increased starch hydrolysis, which requires activities of hydrolytic enzymes. Resurrection plants and seeds of many higher plants are good examples
for a link of carbohydrate (in particular sucrose) accumulation
and the acquisition of stress tolerance (Hoekstra et al., 2001;
Phillips et al., 2002). This is illustrated with the example of
the resurrection plant C. plantagineum, which contains a high
amount of the unusual sugar octulose (an 8-carbon sugar), which
is rapidly converted into sucrose during dehydration (Bianchi
et al., 1991). This sugar conversion is coupled with the increased expression of sucrose synthase and sucrose phosphate
synthase genes (Ingram et al., 1997). The currant hypothesis is
that sugars either act as osmotica and/or protect specific macromolecules and contribute to the stabilization of membrane structures. Sugars may protect cells during desiccation by forming
glasses (Black and Pritchard, 2002). Sugars are also thought to
interact with polar headgroups of phospholipids in membranes
so that membrane fusion is prevented. It is unknown whether
sugars fulfil this function on their own or in conjunction with
other molecules such as LEA proteins. Many seeds accumulate
considerable amounts of raffinoase-type oligosaccarides (RFOs)
such as raffinose and stachyose which are thought to play a role
in the acquisition of desiccation tolerance. Despite many studies
the link between the presence of these carbohydrates and desiccation tolerance has not always been confirmed. Galactinol
synthase catalyzes the first committed step in the biosynthesis
of RFOs. Overexpression of galactinol synthese resulted in accumulation of galactinol and raffinose under controlled conditions
and improved drought tolerance (Taji et al., 2002), confirming
the importance of these sugars under stress conditions.
A sugar which has been shown to contribute to desiccation
tolerance in yeast and some nematodes is trehalose. In higher
plants substantial amounts of trehalose were identified in two
resurrection plants Myrothamnus flabellifolia and Sporobolus
stapfianus (Phillips et al., 2002). It has been reported that many
higher plants possess trehalase activity, which is perhaps responsible for rapid degradation of any trehalose synthesized.
Arabidopsis thaliana has at least one gene which encodes
trehalose-6-phosphate phosphatase which is required for trehalose synthesis, but the physiological role of this enzyme is not
clear (Vogel et al., 1998).

40

D. BARTELS AND R. SUNKAR

B.

Cyclitols
Accumulation of cyclic polyols such as D-pinitol (1D-3O-methyl-chiro-inositol) or D-ononitol (1D-4-O-methyl-myoinositol) has frequently been reported in response to drought
and salinity (Vernon and Bohnert, 1992; Streeter et al., 2001).
Pinitol may accumulate in chloroplasts, which is consistent with
the positive correlation between cyclitol accumulation and CO2
assimilation under drought (Sheveleva et al., 1997). Direct evidence for a role of cyclitols has been provided by transgenic tobacco plants that accumulated ononitol which resulted in drought
and salt stress tolerance (Sheveleva et al., 1997).

which accumulated proline but showed no change in osmotic


stress tolerance (Mani et al., 2002).

Glycine Betaine
Glycine betaine is another extensively studied compatible solute. Glycine betaine is thought to protect the plant by maintaning
the water balance between the plant cell and the environment and
by stabilizing macromolecules (reviewed in Chen and Murata,
2002; Rontein et al., 2002). Plants synthesize glycine betaine via
a two-step oxidation of choline: Choline betaine aldehyde
glycine betaine (Rhodes and Hanson, 1993). The first reaction
is catalyzed by a ferredoxin-dependent choline monooxygenase
(CMO) and the second step by a NAD+ -dependent betaine aldeC. Proline
Proline is probably the most widely distributed osmolyte, hyde dehydrogenase (BADH) (Chen and Murata, 2002; Rontein
and it occurs not only in plants but also in many other organisms et al., 2002). Glycine betaine accumulation is associated with
(McCue and Hanson, 1990; Delauney, 1993). Besides osmotic upregulated CMO and BADH gene expression concomitantly
adjustment other roles have been proposed for proline in os- leading to elevated enzymatic activity.
Glycine betaine accumulation marginally improves osmotic
motically stressed plant tissues: protection of plasma membrane
stress
tolerance in transgenic plants (Hayashi et al., 1997). The
integrity (Mansour et al., 1998), a sink of energy or reducing
levels
of glycine betaine thus far obtained by engineering are
power (Verbruggen et al., 1996), a source for carbon and nitrogen
low,
and
the increments in stress tolerance are small (Nuccio
(Ahmad and Hellebust, 1988; Peng et al., 1996), or hydroxyl radet
al.,
1999).
The major factors that limit the accumulation of
ical scavenger (Smirnoff and Cumbes, 1989; Hong et al., 2000).
glycine
betaine
are the available choline as the substrate for
Proline accumulation can occur via two biosynthetic paththe
reaction
and
its transport from the chloroplast (where it is
ways in plants: the ornithine-dependent pathway and the glutamatesynthesized)
to
the
cytosol (Nuccio et al., 1998, 2000; McNeil
dependent pathway. Proline biosynthesis from glutamate apet
al.,
2000;
Huang
et
al., 2000; Chen and Murata, 2002; Rontein
pears to be the predominant pathway, especially under stress
et
al.,
2002).
conditions (Delauney and Verma, 1993; Delauney et al., 1993).
L-proline is synthesized from L-glutamic acid via 1 -pyrroline5-carboxylate (P5C). This reaction is catalyzed by two enzymes,
P5C synthetase (P5CS) and P5C reductase (P5CR). Genes encoding the enzymes involved in proline metabolism have been
isolated from several plant species (Yoshiba et al., 1997). The
second proline biosynthesis pathway involves transamination
of ornithine and is catalyzed by ornithine--aminotransferase
(OAT) yielding two possible intermediates P5C (-pyrroline-2carboxylate) and P2C (-pyrroline-2-carboxylate), which can
both be reduced to proline (Mestichelli et al., 1979). There are
indications from Arabidopsis that the ornithine pathway operates mainly in young seedlings (Roosens et al., 1998). The other
important process that controls proline levels is oxidation of
L-proline by proline dehydrogenase (ProDH) to P5C, which is
converted to L-glutamic acid by P5C dehydrogenase.
Direct evidence for the role of proline during osmotic stress
has been provided by transgenic approaches. Different strategies
were used to manipulate proline biosynthesis including overexpression of P5CS in tobacco, rice and Arabidopsis plants,
overexpression of OAT, expression of a feedback inhibitioninsensitive form of P5CS, and antisense suppression of proline
oxidation by ProDH (Kavi Kishor et al., 1995; Zhu et al., 1998;
Nanjo et al., 1999; Hong et al., 2000; Roosens et al., 2002).
All approaches resulted in elevated proline pools and improved
osmotic stress tolerance. In contrast to these observations is a
report about antisense ProDH transgenic Arabidopsis plants,

D.

VIII.
A.

PROTECTIVE PROTEINS AND OTHER PATHWAYS


INVOLVED IN STRESS ADAPTATION

Late Embryogenesis-Abundant (LEA) Proteins


Lea genes encode a diverse group of stress-protection proteins expressed during embryo maturation in all angiosperms
including orthodox and recalcitrant seeds. LEA proteins were
first identified and characterized in cotton and represents the
dominant protein and mRNA species in mature embryos (Dure
et al., 1981; Galau et al., 1986). Accumulation of LEA proteins during embryogenesis correlates with increased levels of
ABA and with acquisition of desiccation tolerance (Galau et al.,
1986). ABA treatment of embryos isolated from an early developmental stage results in a precocious accumulation of LEA
proteins and the acquisition of desiccation tolerance indicating
a connection between these parameters (Bartels et al., 1988).
LEA proteins are not normally expressed in vegetative tissues,
but they are induced by osmotic stress or exogenous application
of ABA (Ingram and Bartels, 1996). LEA proteins comprise the
vast majority of stress-responsive proteins. Evidence derived
from expression profiles strongly supports a role for LEA proteins as protective molecules, which enable the cells to survive
protoplasmic water depletion (Ingram and Bartels, 1996). This
is corroborated by the finding that a lea transcript is expressed
in response to dehydration in the nematode Aphelenchus avenae

DROUGHT AND SALT TOLERANCE IN PLANTS

(Browne et al., 2002). This nematode belongs to the group of


animals that enter a state of suspended animation known as anhydrobiosis (Crowe et al., 1992), surviving for indefinite periods
until rehydration allows them to resume normal metabolism.
The genomes of some microorganisms also contain sequences
that encode LEA-like proteins (Dure, 2001). This suggests that
plants, animals, and microorganisms may use common protective strategies against dehydration, although the biochemical
function of the LEA proteins has not been proven. Some positive
evidence for LEA proteins as protective molecules has been derived from overexpression studies in plants and yeast (Table 1).
LEA proteins have been divided into different groups based
on conserved structural features (Dure et al., 1989; Dure, 1993),
although there is some debate about the classification and better
systems may evolve. Group 1 LEA proteins are characterized
by high glycine contents (ca. 20%), amino acids with charged
R-groups (ca. 40%) and the presence of a stretch of 20 hydrophilic amino acids. The high hydrophilicity of these proteins
renders them soluble after boiling suggesting that these proteins are highly hydrated and do not assume a globular tertiary
structure (Dure, 1993). The group 1 proteins may be involved
in binding or replacement of water. The group 2 LEA proteins
(dehydrins) are characterized by a tract of serine residues, a conserved motif containing the consensus sequence DEYGNP near
the N-terminus and a lysine-rich 15 amino acid sequence motif
termed K-segment (EKKGIMDKIKEKLGP), which is present
in most cases at or near the carboxy terminus and is predicted
to form an amphipathic -helix. Group 2 and Group 4 proteins
may contribute to maintenance of protein and membrane structures (Dure, 1993; Ingram and Bartels, 1996). The main characteristic of group 3 LEA proteins is the presence of several
copies of an 11-amino acid peptide predicted to form an amphipathic -helix with possibilities for intra-and intermolecular
interactions (Dure, 1993). Group 4 is characterized by a conserved N-terminus predicted to form an -helix and a diverse
C-terminal part with a random coil structure. Group 5 LEA proteins contain a higher proportion of hydrophobic residues than
the other four groups and probably adopt a globular conformation. Proteins of group 3 and 5 have been suggested to form
dimers with a coiled-coil structure capable of sequestering ions,
which accumulate due to water depletion (Dure et al., 1989).

B.

Aquaporins
Dehydration and salt stress require changes in water flow
to allow cells and tissue to adapt to the stress situation. The
rate of water flux into or out of cells is determined by the water potential gradient that acts as the driving force for transport
and by the water permeability of the membrane. Evidence has
been accumulating that aquaporins are central components in
plant water relations; this subject has recently been reviewed in
detail by Tyerman et al. (2002). Aquaporin proteins facilitate
osmosis by forming water-specific pores as an alternative to water diffusion through the lipid bilayer, thus increasing the water

41

permeability of the membrane (Schaffner, 1998; Kjellbom et al.,


1999). Aquaporins are members of a large super-family of membrane spanning proteins, the major intrinsic proteins (MIPs). In
plants, aquaporins localized in the tonoplast are called tonoplast
intrinsic proteins (TIPs), while those in the plasma membrane
are PIPs. Across different plant species aquaporins have very
conserved structures and hence are encoded by similar DNA sequences. In A. thaliana, the MIPs are forming a family of at least
30 genes that display differential patterns of expression. In general, MIPs are most abundantly expressed in rapidly growing tissues and in cells involved in high-volume water flux (Weig et al.,
1997; Kjellbom et al., 1999). There are several reports that aquaporin genes are induced by dehydration and salt stress; this may
trigger greater osmotic water permeability and facilitate water
flux (Guerrero et al., 1990; Yamaguchi-Shinozaki et al., 1992;
Fray et al., 1994; Yamada et al., 1997; Sarda et al., 1999). On the
other hand, there are also examples of downregulated aquaporins
under dehydration or salt stress, which should result in decreased
membrane permeability and may allow cellular water conservation (Yamada et al., 1997; Johansson et al., 1998). Guard cell
MIP genes (TIPs) NgMIP2, NgMIP3 and (PIP) NgMIP4 are
downregulated by drought stress (Smart et al., 2001). A role for
aquaporins in fine tuning of water availability is supported by a
recent example from the resurrection plant C. plantagineum. It is
shown that a specific aquaporin gene is upregulated by dehydration but downregulated by salt stress, which coincides with the
tolerance behavior of this plant (Smith-Espinoza et al., 2003).
Direct evidence for a role of aquaporins in maintaining the water status of plants has been derived from aquaporin antisense
experiments in Arabidopsis or tobacco (Kaldenhoff et al., 1998;
Siefritz et al., 2002). Kaldenhoff et al. (1998) showed that Arabidopsis plants, in which the expression of the plasma membrane
aquaporin PIP1b is suppressed, performed better in a hypotonic
solution than wild-type plants, and these plants compensated
for the reduced water flow by increasing the root system. Arabidopsis plants with reduced aquaporin1 gene expression had
reduced hydraulic conductivity and decreased drought tolerance
(Siefritz et al., 2002).

C.

Heat Shock Proteins (Hsps)


Synthesis of correctly structured proteins and their maintenance is important for efficient cellular functions. The heat shock
protein family encompasses many chaperones, which have an
important role in the folding and assembly of proteins during
synthesis, and in the removal and disposal of nonfunctional and
degraded proteins. Heat shock proteins (Hsps) are usually undetectable in vegetative tissues under normal growth conditions,
but can be induced by environmental stress and developmental stimuli. Accumulation of Hsps coincides with acquisition
of stress tolerance. In desiccation sensitive Arabidopsis mutant
seeds Hsp expression is reduced (Wehmeyer and Vierling, 2000).
Hsps are induced by water stress in several plants (Alamillo et al.,
1995; Coca et al., 1996; Campalans et al., 2001). Transgenic

42

D. BARTELS AND R. SUNKAR

Arabidopsis plants overexpressing AtHSP17.7 accumulate high


levels of AtHSP17.7 protein and show enhanced tolerance to
drought and salinity (Sun et al., 2001). This provides in vivo evidence for a protective role of Hsps in plants. Experiments in vitro
suggest that cytosolic Hsps function as molecular chaperones
by preventing the thermal aggregation of substrate proteins and
facilitating the subsequent reactivation (Lee et al., 1995). The
concept that low RWC impairs protein structure explains the necessity of molecular chaperones to accumulate under a range of
stresses. The abundance of small heat shock proteins (sHsps) in
plants and their functional characteristics of binding and stabilizing denatured proteins suggest that sHsps play an important
role in plant stress tolerance (reviewed in Wang et al., 2004).
Targeting is controlled by molecular chaperones of which one
of the major players is the chaperone binding protein (BiP), a
member of the HSP70 protein family. Synthesis of BiP is induced
by physiological stress conditions that promote accumulation
of mis-folded proteins. Plant BiPs were shown to be induced
by a variety of environmental stresses, including water stress
(Cascardo et al., 2000). Overexpression of BiP-enhanced water stress tolerance in transgenic tobacco plants, whereas the
antisense transgenic plants showed hypersensitivity to water
stress. The water-stress tolerant phenotype of BiP overexpressing plants is likely to be due to better osmotic adjustment which
prevents turgor loss (Alvim et al., 2001).
D.

Proteases and Proteinase Inhibitors


Proteolysis is an important cellular activity to maintain protein homeostasis. Increased proteolysis in response to stress is
frequently observed and can be interpreted as a way to get rid
of damaged proteins or to mobilize nitrogen (Vierstra, 1996).
Increased proteolysis during drought/salt stress conditions has
been reported (Guerrero et al., 1990; Ramanjulu et al., 1994b;
Ramanjulu and Sudhakar, 1997; de Carvalho et al., 2001). Cysteine proteases were shown to be induced during drought and
salinity (Koizumi et al., 1993; Forsthoefel et al., 1998; KhannaChopra et al., 1999; Campalans et al., 2001; Seki et al., 2002).
The water and salinity stress-induced Arabidopsis ERD1 gene
encodes a chloroplast-localized protein with sequence homology to the regulatory ATPase subunit ClpA of the ATP-dependent
Clp protease from E. coli (Nakashima et al., 1997). The ERD1
gene product may interact with other subunits such as ClpP (a
chloroplast gene) of the ATP-dependent protease and may function in the degradation of chloroplast proteins. An Arabidopsis
DegP2 gene encoding a novel chloroplast homologue of the
prokaryotic trypsin was isolated and was shown to increase to
high concentrations of NaCl, desiccation and high light (Hausuhl
et al., 2001). AtDegP2 is likely to have a role in primary cleavage
of the photo-damaged D1 protein of PSII prior to its removal by
secondary proteolysis.
E.

Polyamines
The polyamines spermidine, spermine, and putrescine are
polycationic small aliphatic amines that have been implicated

in a variety of physiological processes such as growth and development in plants. A role for polyamines has been proposed
in stress responses. Salt-tolerant Pokkali rice plants accumulate higher levels of polyamines compared to the salt-sensitive
rice plants in response to salinity stress (Chattopadhyay et al.,
2002). Exogenously supplied putrescine prevented stress damage and increased stress tolerance in Conyza bonariensis and
maize (Ye et al., 1998). Biosynthesis of polyamines in plants
is controlled by the enzymes ornithine decarboxylase and arginine decarboxylase which are responsible for the production
of putrescine, and S-adenosyl-L-methionine (SAM) decarboxylase that is necessary for the formation of spermidine and spermine. Plants subjected to osmotic stress show a rapid increase
in putrescine levels due to transcription and activation of arginine decarboxylase (Borrell et al., 1996; Flores and Galston,
1982). Arabidopsis mutants spe1-1 and spe2-1 with reduced activity of arginine decarboxylase are more sensitive to salt stress
than wild-type plants (Kasinathan and Wingler, 2002). Furthermore, AtADC2 (arginine decarboxylase) expression correlated
with free putrescine accumulation under salinity and dehydration (Urano et al., 2003). Ds insertion mutant adc2-1 displayed
a salt-sensitive phenotype coupled with reduced accumulation
of putrescine (Urano et al., 2004). Transgenic rice plants with
enhanced level of ADC gene expression increased biomass compared to control plants under saline conditions (Roy and Wu,
2001). These observations lend support for the protective function of polyamines. Polyamines may possibly exert their protective function by scavanging ROS, which may occur as a consequence of stress (Tiburcio et al., 1994). However, accumulation
of polyamines seems to be toxic to the plants under normal conditions and therefore constitutive overexpression may not be the
appropriate way to obtain stress tolerance.
IX.
A.

OXIDATIVE STRESS A CONSEQUENCE


OF DEHYDRATION AND SALT STRESS

Formation of Reactive Molecules


A secondary effect of dehydration and salt stress is the increase of reactive oxygen species (ROS), which include singlet
oxygen, superoxide anion radicals, hydroxyl radicals, and hydrogen peroxide (Smirnoff, 1998; Bartels, 2001; Apel and Hirt,
2004). ROS are predominantly generated in the chloroplast by
direct transfer of excitation energy from chlorophyll to produce
singlet oxygen, or by univalent oxygen reduction at photosystem I, in the Mehler reaction (Foyer et al., 1994; Allen, 1995)
and to some extent in mitochondria. Chloroplasts are the first
targets in plant cells since this is the major site of ROS production. The increased concentration of ROS inhibits the ability to
repair damage to photosystem II and inhibits the synthesis of
the D1 protein. Stress-enhanced photorespiration and NADPH
activity also contributes to the increased H2 O2 accumulation,
which may inactivate enzymes by oxidizing their thiol groups.
The toxicity of H2 O2 is not due to its reactivity per se, but
requires the presence of a metal reductant to form the highly

DROUGHT AND SALT TOLERANCE IN PLANTS

reactive hydroxyl radical, which potentially reacts with all biological molecules. Transition metals such as cuprous and ferrous
ions may be released from enzymes and electron carriers during stress and promote the Fenton reaction to produce highly
reactive hydroxyl radicals, which extensively oxidizes proteins,
lipids and nucleic acids (Halliwell and Gutteridge, 1999).
The alleviation of oxidative damage and increased resistance
to environmental stresses is often correlated with an efficient antioxidative system (Smirnoff, 1998; Shalata et al., 2001; Kranner
et al., 2002). Overproduction of SOD, APX and catalase have
been shown to improve oxidative stress tolerance in transgenic
plants (Allen, 1995; Roxas et al., 1997). The importance of antioxidant systems has been demonstrated in the resurrection
plant Myrothamnus flabellifolia (Kranner et al., 2002). This
plant was able to recover after four months of desiccation, but
not after eight months when the antioxidant defense mechanisms were broken down. The ability to recover after desiccation has been correlated with the ability to maintain/resynthesize
antioxidants such as -tocopherol, ascorbate, and glutathione.
Strategies have been developed to keep the concentrations of
ROS under tight control by detoxification. ROS detoxification
mechanisms can be broadly divided into nonenzymatic and enzymatic mechanisms. Major nonenzymatic antioxidants include
ascorbate (vitamin C) and glutathione in plants, although tocopherol (vitamin E), flavonoids, alkaloids, and carotenoids can
also act as antioxidants. Enzymatic mechanisms include superoxide dismutase, peroxidases, and catalase. These aspects have
been discused extensively in recent reviews (Mittler, 2002; Apel
and Hirt, 2004). Here we would like to highlight additional components that aid in detoxification of ROS and secondary products
that are derived from ROS interaction with biomolecules.

B.

Enzymes That Detoxify Aldehydes


Free radical-mediated lipid peroxidation in biological systems is always accompanied by the generation of complex patterns of aldehydes. Aldehydes are highly reactive molecules
and toxic at low concentrations. Aldehyde dehydrogenases have
been emerging as crucial enzymes in detoxification processes in
a wide range of organisms. These enzymes are critical in the conversion of toxic aldehydes to the less reactive carboxylic acid
forms.
In plants, not much is known about aldehyde and detoxifying aldehydes. An aldehyde dehydrogenase from C. plantagenium (Cp-ALDH) and its Arabidopsis homologue Ath-ALDH3
were induced by diverse abiotic stressors that induce oxidative
stress (Kirch et al., 2001; Sunkar et al., 2003). Osmotic stressinduced aldehyde dehydrogenases have also been described in
other plants (Seki et al., 2002; Ozturk et al., 2002; Chen et al.,
2002). Another group of enzymes are aldose/aldehyde reductases which reduce a wide range of aldehydes and ketones to
alcohols. Osmotic stress-inducible aldose/aldehyde reductases
were isolated from plants (Oberschall et al., 2000; Mundree
et al., 2000). Functional analyses of aldose/aldehyde reductase

43

and aldehyde dehydrogenases point to a role of detoxification of


aldehydes in vitro. The function is also supported by transgenic
plants overexpressing these enzymes (Oberschall et al., 2000;
Sunkar et al., 2003).
C.

Peroxiredoxins
Peroxiredoxins are a group of enzymes with a catalytic function in the detoxification of cellular-toxic peroxides (Dietz et al.,
2002). Peroxiredoxins reduce peroxide to the corresponding alcohol. Peroxiredoxins are able to protect DNA, membranes and
certain enzymes in vitro against damage by ROS and to remove
H2 O2 , alkyl hydroperoxides and hydroxyl radicals (Lim et al.,
1993). Stress-inducible peroxiredoxins have been identified in
plants (Seki et al., 2001; Mowla et al., 2002). The peroxiredoxin transcript is constitutively expressed at high levels in
35S:DREB1A overexpressing plants (Kasuga et al., 1999; see
under section VI C AP2/ERF-type-transcription factors) suggesting that it is one of the targets of DREB1A.
D.

Thioredoxins
Thioredoxins are small proteins functioning as hydrogen
donor and thereby reducing disulfide bridges in proteins. Their
involvement in the response to oxidative stress is well documented in bacteria, yeast, and animal cells (Arner and Holmgren,
2000). The role of thioredoxins in oxidative stress responses
in plants is largely unexplored. Drought- and high light stressinducible chloroplast localized protein CDSP32 has been identified in plants (Pruvot et al., 1996; Rey et al., 1998; Broin
et al., 2000). The mature CDSP32 protein exhibits structural
features similar to those described for bacterial thioredoxins. A
decreased photochemical efficiency by suppression of CDSP32
in antisense plants suggested its role in oxidative stress. However, overexpression of CDSP32 in transgenic plants leads to
susceptibility to oxidative stress and it was hypothesized that
overproduction of thioredoxin could disturb the plastidic redox
state (Broin et al., 2002).
E.

Protein Oxidation
ROS can lead to oxidation of amino acid residue side chains,
formation of protein-protein cross-linkages and oxidation of the
protein backbone, resulting in protein fragmentation (Berlett and
Stadtman, 1997). Such modified proteins are more susceptible to
degradation by intracellular proteases. The sulphur-containing
amino acids methionine (Met) and cysteine (Cys) in proteins are
more susceptible than others. Oxidation of Met leads to the formation of the methionine sulfoxide [Met(O)], a process which
leads to the loss of function (Rodrigo et al., 2002). However, enzymes such as peptide-methionine sulfoxide reductases (MsrA)
found in several organisms can compensate the damage by catalyzing the reduction of free and protein-bound (Met(O) residues
to Met (Moskovitz et al., 1999). The Arabidopsis gene AtSXL3
is induced by dehydration and oxidative stress. Recently Rodrigo et al. (2002) have analyzed SXL genes in knock-out yeast

44

D. BARTELS AND R. SUNKAR

and Arabidopsis mutants. Knock-out mutants have significantly


greater quantities of Met(O) compared to wild type and revertants under oxidative stress.
X.

IONIC STRESS
Similar to osmotic stress, high concentrations of Na+ in the
soil/increased Na+ accumulation in the plant system may be recognized by extracellular and intracellular sensors such that the
effective counteracting mechanisms will be initiated. As yet, the
molecular identity of Na+ sensor(s) is unknown, but the long Cterminal cytoplasmic domain of SOS1 (see below) is predicted
as a possible cytosolic Na+ sensor (Zhu, 2003). The identification of Na+ sensor is as important as that of osmosensor.

Na+ Toxicity and Homeostasis


The complexity of the plant response to salt stress can be
partially explained by the fact that salinity imposes salt toxicity in addition to osmotic stress (Niu et al., 1995; Hasegawa
et al., 2000). Sodium is toxic to many organisms, except for
halotolerant organisms like halobacteria and halophytes, which
possess specific mechanisms that keep intracellular sodium concentrations low. Accumulation of sodium in the cytoplasm is
prevented by restricting its uptake across the plasma membrane
and by promoting its extrusion or sequestration in halophytes
(Hasegawa et al., 2000). High salt concentrations (>400 mM)
inhibit the activities of most enzymes because of perturbation of
the hydrophobic-electrostatic balance between the forces maintaining protein structure. However, toxic effects on cells occur
at much lower salt concentrations (about 100 mM), pointing to
specific salt toxicity targets (Serrano, 1996). Surprisingly, the
cytosolic enzymes from glycophytes are as sensitive as those of
halophytes to Na+ (Blumwald et al., 2000). On the other hand,
extracellular (apoplastic) enzymes from glycophytes and halophytes have been shown in vitro to be remarkably salt-insensitive
tolerating concentrations up to 500 mM NaCl (Thiyagarajah
et al., 1996).
Many physiological studies have demonstrated that Na+ toxicity is not only due to toxic effects of Na+ in the cytosol, but
also because K+ homeostasis is disrupted possibly due to the
ability of Na+ competing for K+ binding sites.
Considerable progress has been made in understanding ion
homeostasis (Hasegawa et al., 2000; Blumwald et al., 2000;
Apse and Blumwald, 2002; Zhu, 2003). Ion transporters are
considered to play an important role in salt tolerance. In principle, three mechanisms exist to prevent excess Na+ accumulation
in the symplast of plant cells:
A.

1. Restricting the Na+ permeation and entry into plants by Na+


transporters, whose molecular identity is unknown.
2. Compartmentalizing the Na+ in the vacuole.
3. Extruding Na+ : cytosolic Na+ can be transported back to
the external medium or the apoplast via plasma membrane
Na+ /H+ antiporter activity.

Na+ Exclusion
Na+ levels should be controlled at the entry point (Munns,
2002). The Na+ uptake across the plasma membrane has been
attributed to low Na+ permeability properties of systems that
transport essential K+ (Blumwald et al., 2000; Hasegawa et al.,
2000). The Arabidopsis transporter AtHKT1 has been identified
as one of the pathways of Na+ entry into plants. Mutation in
AtHKT1 suppresses the hypersensitivity of sos3 mutants (Ruis
et al., 2001) suggesting that the wild-type SOS3 may inhibit
the activity of AtHKT1 as a Na+ influx transporter. In addition,
nonselective cation channels present in the root plasma membrane may also contribute to Na+ entry into plants (Amtmann
and Sanders, 1999), although the molecular identity of such
transporters is still unknown.

B.

Na+ Compartmentalization
In plants, the central vacuole plays a vital role in regulation of
cytoplasmic ion homeostasis. Exclusion of excess Na+ from the
cytoplasm and the accumulation in the vacuole represents one
of the adaptive mechanisms during salt stress. Halophytes have
the ability to use Na+ as an osmoticum by compartmentalizing
it into vacuoles. The vacuolar sodium sequestration is mediated by a Na+ /H+ antiport at the tonoplast (Apse et al., 1999).
Sequestration or compartmentalization of sodium into the vacuole through vacuolar Na+ /H+ antiporters uses the proton motive force generated by the vacuolar H+ -translocating enzymes,
H+ -adenosine triphosphatase (ATPase), and H+ -inorganic pyrophosphatase (PPiase), to couple the downhill movement of
H+ with the uphill movement of Na+ against the electrochemical potential (Blumwald and Gelli, 1997).
The presence of Na+ /H+ antiporter activities has been physiologically characterized in tonoplast vesicles and it is molecularly represented by six Arabidopsis genes AtNHX1-6 (Blumwald
et al., 2000; Yokoi et al., 2002). AtNHX1 steady-state transcript
levels were increased in response to NaCl, KCl, sorbitol, and
ABA suggesting that the AtNHX1 transcript upregulation is not
specific to ionic stress but common to osmotic stress (Gaxiola
et al., 1999; Quintero et al., 2002; Shi and Zhu, 2002; Yokoi
et al., 2002). AtNHX1 was able to suppress the salt-sensitive phenotype of the nhx1 yeast mutant (Gaxiola et al., 1999; Quintero
et al., 2002). Evidence for the in vivo function was derived
from ectopically expressing AtNHX1 in Arabidopsis, tomato and
canola (Apse et al., 1999; Zhang and Blumwald, 2001; Zhang
et al., 2001). This was further supported by the overexpression
of the rice homologue in transgenic rice plants (Fukuda et al.,
2004). Tonoplast vesicles isolated from transgenic plants displayed significantly higher Na+ /H+ exchange rates than those of
wild-type plants.
Expression of both AtNHX2 and 5 also suppressed the alkali cation-sensitive phenotype of an nhx1 mutant yeast, like
AtNHX1 (Yokoi et al., 2002). AtNHX2 mediated relatively higher
tolerance than AtNHX1 and 5 in yeast, which may suggest that
NHX proteins perform similar functions but with varying
C.

DROUGHT AND SALT TOLERANCE IN PLANTS

efficiency. AtNHX1 and 2 are responsive to NaCl, sorbitol and


ABA suggesting that AtNHX1 and 2 are responsive to hyperosmotic stress rather than to Na+ toxicity, while AtNHX5 was
specifically increased in response to NaCl but not to sorbitol,
an indication for its involvement in Na+ exchange (Yokoi et al.,
2002).
AtNHX1 is constitutively expressed and thus it is likely to
have physiological roles under nonstress conditions. One such
function seems to be pH regulation (Fukuda-Tanaka et al., 2000).
Venema et al. (2002) showed that AtNHX1 was able to mediate
low affinity Na+ as well as K+ transport in artificial, reconstituted liposomes and therefore AtNHX1 may also have a role of
K+ accumulation in vacuoles. These observations implicate that
NHX proteins may have additional roles such as K+ exchange
activity and pH regulation.
In yeast the protein phosphatase PPZ1 and its regulatory subunit HAL3 act as important determinants of salt tolerance by
regulating the expression of the ENA1 gene encoding the major sodium extrusion pump (de Nadal et al., 1998). Transgenic
Arabidopsis plants constitutively overexpressing AtHAL3a, the
HAL3 Arabidopsis homologue, showed improved growth and
osmotic stress tolerance (Espinosa-Ruiz et al., 1999). This possibly suggests a similar role for AtHAL3 in plants as for its yeast
counterpart HAL3.

subunits of V-ATPases accumulate in response to salt stress


leading to increased enzyme activity (Ratajczak et al., 1994).
Salt induced the transcription of several V-ATPase subunits (A,
B, C, and E) in the common ice plant (Dietz and Arbinger,
1996), subunit A in salt-adapted tobacco cell suspension culture
(Narasimhan et al., 1991) and subunits A and C in sugar beet
(Kirsch et al., 1996; Lehr et al., 1999). A close correlation between the expression of subunit E of the V-ATPase and salt
tolerance was reported for the ice plant, i.e., juvenile non-salttolerant plants did not alter subunit E; whereas the transcript levels increased in salt-tolerant mature leaves (Golldack and Dietz,
2001). Coupled with ATPase activity is the vacuolar H+ -PPase
(pyrophosphatase) activity, which is encoded by Avp1in Arabidopsis. The activity of Avp1 is correlated with increased ion
content of the plants. Overexpression of Avp1 leads to enhanced
salt and drought tolerance in transgenic plants (Gaxiola et al.,
2001). This suggests that pumping of H+ across the vacuolar
membrane is an additional driving force for vacuolar sodium
accumulation. Therefore saltadaptation is correlated with VATPase in many plants. In summary, salt adaptation is correlated
with the tonoplast Na+ /H+ antiporter and with increased activity of V-H+ -ATPase submits. The coordination of both systems
is crucial for sequestering Na+ in cellular compartments.

E.

D.

Proton Transporters and Salt Tolerance


Regulation of ion transport is one of the important factors for
salt tolerance of plants. Protons are used as coupling ions for ion
transport systems, and the proton gradient, generated by proton
pumps in the membrane systems, is the driving force for Na+
transport across membranes. The plasma membrane H+ -ATPase
(P-H+ -ATPases) and vacuolar ATPases (V-ATPases) coupled
with vacuolar H+ -PPase (pyrophosphatase) (V-H+ -PPases) act
at the plasma membrane and tonoplast, respectively. The tonoplast H+ -ATPase and H+ -PPase play important roles in transporting H+ into the vacuole and generation of proton motive
force across the tonoplast, which operates Na+ /H+ antiporters
(Hasegawa et al., 2000; Blumwald et al., 2000).
The P-H+ -ATPases (proton pumps) are encoded by at least
12 genes in Arabidopsis. P-H+ -ATPase from Arabidopsis was
identified as one of the important components of salt tolerance in
plants (Vitart et al., 2001). A mutation in the ATPase gene AHA4
causes a reduction in root and shoot growth under NaCl stress.
Salt stress resulted in more Na+ and less K+ accumulation in
the leaves of aha4-1 mutant plants compared to wild type. It is
hypothesized that AHA4 functions in the control of Na+ flux
across the root endodermis, possibly by partially disrupting the
activity of other pumps (Vitart et al., 2001).
The V-type ATPase is a multimeric enzyme localized in endomembranes of eukaryotic cells that establishes an electrochemical H+ -gradient (Luttge and Ratajczak, 1997). The
V-ATPases are of prime importance in energizing sodium sequestration into the central vacuole. Transcripts encoding

45

SOS Pathway and Ion Homeostasis


Genetic approaches and yeast complementation assays have
identified individual components of sodium homeostasis, but
only the systematic genetic approach in Arabidopsis by Zhu
and colleagues have led to the discovery of a linear salt-stressinduced Ca2+ -regulated pathway involved in Na+ homeostasis.
The molecular basis for this was the characterization of the SOS
(Salt Overly Sensitive) mutants, which render the SOS plants
hypersensitive to NaCl. The identification of SOS genes has uncovered a novel Ca2+ -dependent pathway determining ion (Na+
and K+ ) homeostasis and salt tolerance. Molecular interaction
and complementation analysis indicate that SOS3 is required
for activation of SOS2 that regulates SOS1 transcription. Double
mutant analysis showed that three SOS genes function in a linear
pathway (Zhu, 2001a). The SOS3 gene encodes a Ca2+ sensor
protein similar to the calcineurin B subunit from yeast, a Ca2+
dependent protein phosphatase, and neuronal calcium sensors
from animals (Liu and Zhu, 1998). Calcineurin B and neuronal
calcium sensors are both activated by Ca2+ , which raises the
possibility that the SOS3 gene product might control K+ /Na+
transport via a Ca2+ -regulated pathway (Liu and Zhu, 1998).
Calcineurin is a protein phosphatase type 2B that plays a vital
role in signalling pathways in regulating ion homeostasis and
salt tolerance of yeast (Mendoza et al., 1994) and plants (Pardo
et al., 1998; Liu and Zhu, 1998). Evidence suggests that SOS3
is involved in regulating the SOS2 protein kinase (Halfter et al.,
2000). SOS2 encodes a serine/threonine protein kinase with an
N-terminal kinase domain and a unique C-terminal domain with
FISL motif that is sufficient for interaction with SOS3 (Liu et al.,

46

D. BARTELS AND R. SUNKAR

2000). Removal of the FISL motif renders the SOS2 kinase constitutively active (Guo et al., 2001). SOS1 mutant plants are hypersensitive to salinity and accumulate high amounts of Na+
compared to wild-type plants (Liu and Zhu, 1997). The SOS1
gene encodes a plasma membrane Na+ /H+ antiporter, with a
long cytoplasmic C-terminal tail (Shi et al., 2000). The upregulation of SOS1 in response to NaCl is reduced in sos2 and sos3
mutant plants, suggesting that SOS1 expression is controlled by
SOS2/SOS3, which is further evidence for the interaction of the
three SOS genes. SOS3 forms a complex with SOS2, and this
complex is necessary for the phosphorylation and subsequent
activation of SOS1, the Na+ /H+ antiporter (Qiu et al., 2002).
Overexpression of SOS1 in Arabidopsis plants improved salt
tolerance with reduced Na+ accumulation in shoots. SOS1 probably retrieves Na+ from the xylem (Shi et al., 2003). It seems
that the SOS pathway also regulates the vacuolar Na+ /H+ exchange activity and contributes to Na+ compartmentalization
(Qiu et al., 2004). The emerging picture from these studies is
that the SOS pathway coordinately regulates plasma membrane
and tonoplast Na+ /H+ antiporter activity which leads to Na+
homeostasis and thus salt tolerance.
XI. ABSCISIC ACID (ABA)
The plant hormone abscisic acid (ABA) plays a central role
in many aspects of stress responses as well as seed development, dormancy, and germination. Different aspects have recently been covered in excellent reviews (Leung and
Giraudat, 1998; Rock, 2000; Finkelstein et al., 2002). Here we
will focus on the role of ABA in osmotic stress. During vegetative growth, ABA-mediated adaptive responses are critical to
plant survival during drought, salt, and cold stress. These stressors serve as a trigger for the accumulation of ABA, which in
turn activates various stress-associated genes that are thought
to function in the accumulation of osmoprotectants, (LEA) proteins, signalling, transcriptional regulation etc. Exposure to exogenous ABA mimics the induction of genes similar to stress
treatment. This gene induction can be correlated with acquisition of desiccation tolerance similar to what has been shown
for callus derived from the resurrection plant C. plantagineum
(Bartels and Salamini, 2001). ABA-insensitive and biosynthetic
mutants have confirmed the role of ABA as an intermediate
molecule between stress perception and cellular stress
response.
A.

Regulation of ABA Levels


Mutant plants with altered biosynthesis, perception, or response have been crucial in identification of various components involved in ABA biosynthesis and signalling. The genetic
screens and selections that have been used include production
of nondormant seeds, loss or gain of sensitivity to ABA during germination, seedling or root growth, and altered expression of reporter genes. These approaches have yielded three
classes of ABA mutants: ABA-deficient, hypersensitive, and

insensitive. ABA-deficient mutants (impaired ABA biosynthesis) exhibited a wilty phenotype due to impaired stomatal
closure.
Significant progress has been made in recent years with cloning
genes encoding enzymes involved in ABA biosynthetis. This
revealed the critical steps in the regulation of ABA levels. The
main ABA biosynthetic pathway starts from carotenoids C-40
(Zeevart and Creelman, 1988). Biochemical studies have indicated that 9-cis-epoxycarotenoid dioxygenase (NCED) is a
key enzyme in ABA biosynthesis (Kende and Zeevaart, 1997).
NCED catalyzes the cleavage reaction of epoxycarotenoids which
produces xanthoxin (the first C15 intermediate). The ABA accumulation is regulated at the transcriptional level in response to
osmotic stress (Guerro and Mullett, 1986; Stewart et al., 1986).
Accordingly, the induction of the NCED gene by drought stress
has been reported in maize, tomato, Arabidopsis, bean, and cow
pea (Burbridge et al., 1997; Tan et al., 1997; Neill et al., 1998;
Qin and Zeevaart et al., 1999; Iuchi et al., 2000; Thompson
et al., 2000a). The timing of the induction of the VuNCED1 gene
was shown to be slightly earlier than that of ABA accumulation
under drought stress suggesting that the transcriptional regulation of the genes involved in the ABA biosynthesis pathway
are responsible for drought-induced ABA accumulation (Iuchi
et al., 2000). The important role of NCED was confirmed by
transgenic and mutant studies which confirmed for Arabidopsis and tomato that the expression of the NCED gene caused
an increase in ABA levels (Iuchi et al., 2001; Thompson et al.,
2000b). Plants overexpressing AtNCED3 showed reduced transpiration rates and improved drought tolerance, whereas plants
with nonfunctional NCED genes had a drought-sensitive phenotype (Iuchi et al., 2001). These experiments provide convincing
evidence that modulating of endogenous ABA levels is possible
via engineering NCED expression, which subsequently changes
drought tolerance and most likely other responses to osmotic
stress. The Arabidopsis genome contains at least seven homologous NCED genes, but there is hardly any knowledge of their
individual functions.
Recent findings confirmed that ZEP and NCED are located
in plastids, and that the product of the ABA levels are modulated not only by the rate of synthesis, but also by the activity
of degradation enzymes. The hydroxylation at the 8 -position
of ABA is thought to be the key step of ABA catabolism, and
this reaction is catalyzed by ABA 8 -hydroxylase, a cytochrome
P450. The gene responsible for this process in Arabidopsis is
CYP707A3. ABA treatment resulted in increased transcript accumulation of CYP707A3 (Saito et al., 2004). Further, plants
that accumulated ABA due to overexpression of NCED also
hyperaccumulated phaseic acid (Qin and Zeevaart, 2002), suggesting that the ABA catabolism has been increased in these
plants. These observations support the notion that ABA might
restrict its own accumulation by activating its degradation. The
findings suggest that both ABA biosynthesis and catabolism are
activated by stress/ABA in fine tuning the levels of ABA. Therefore, it is possible that ABA negatively regulates the biosynthetic

DROUGHT AND SALT TOLERANCE IN PLANTS

genes and positively regulates the catabolic genes in feed-back


and feed-forward mechanisms.

47

hypersensitive mutants display enhanced sensitivity to ABA,


resulting in diminished germination rates at low ABA concentrations and reduced water loss due to enhanced ABA-induced
B. The Role of ABA in Stomatal Closure
stomatal closure (Arabidopsis era1, sad1, abh1, hyl1, rop10 and
Transpiration through stomatal pores is a crucial response of fiery 1). The ERA1 gene encodes a protein farnesyltransferase,
the plant under osmotic stress regulated by ABA. The closure HYL1, SAD1, and ABH1 genes encode different types of RNA
of stomatal pores in aerial tissues is an important mechanism by binding proteins. Farnesyltransferases influence protein strucwhich higher plants regulate their water balance. Guard cells, ture or localization through mechanisms other than phosphorywhich flank stomatal pores integrate and respond appropriately lation. Protein farnesylation mediates the COOH-terminal lipito changes in water levels with ABA as a major signal. ABA- dation of specific cellular proteins such as Ras and G-proteins.
deficient mutants are prone to wilting and cannot withstand wa- Pei et al. (1998) reported that deletion of the Arabidopsis farneter deficit conditions due to unregulated stomatal control. There syltransferase gene ERA1 or application of farnesyltransferase
have been major advances in our understanding of the cellular inhibitors resulted in ABA hypersensitivity of guard cell anionevents that underlie the regulation of stomatal aperture in re- channel activation and of stomatal closure implicating a role
sponse to osmotic stress (reviewed in Schroeder et al., 2001). for farnesyltransferases in stomatal regulation. The era1 mutant
Different signalling components in this process such as protein plants exhibited a reduction of transpirational water loss during
kinases and phosphatses, Ca2+ , ROS, IP3, IP6, NO, syntaxin, dehydration suggesting ERA1 is a negative regulator in this prosphingosine-1-phosphate, phosphatidic acid, G proteins, G pro- cess (Pei et al., 1998). FRY1 encodes an inositolpolyphosphatetein coupled receptor (GCR1) and cADPR have been identi- 1-phosphatase. Fry1 mutation results in elevated levels of inositolfied recently (Leung et al., 1994; Leymzann et al., 1999; Pei 1, 4, 5-triphosphate, super-induction of ABA and stress-inducible
et al. 2000; Li et al., 2000; Ng et al., 2001; Wang, 2001; Mata genes in response to ABA, salt, and dehydration. Therefore,
and Lamattina, 2001; Zhu et al., 2002; Coursol et al., 2003; FRY1 is a negative regulator of ABA and stress signalling and
Pandey and Assmann, 2004). ABA-induced stomatal closure establishes a connection between phosphoinositols and ABA
is mediated by a reduction in the turgor pressure of guard cells, and stress signalling in plants (Xiong et al., 2001c).
which requires an efflux of K+ and Cl , sucrose removal and the
Recent identification of RNA-binding proteins ABH1 (ABA
conversion of malate to osmotically inactive starch (Schroeder hypersensitive), SAD1 (supersensitive to ABA and drought 1)
et al., 2001). ABA triggers an increase in cytosolic calcium lev- and HYL1 (hyponastic leaves 1), whose mutation confers an
els in guard cells which regulate ion channels hat control ion ABA-hypersensitive phenotype (Lu and Federoff, 2000;
efflux and stomatal closure. These aspects have been covered Hugouvieux et al., 2001; Xiong et al., 2001b) and AKIP1 (an hnin some recent excellent reviews by Assmann and Wang (2001) RNP binding protein) (Li et al., 2002) implicates that ABA may
and Schroeder et al. (2001).
also play a role in post-transcriptional RNA processing. HYL1
ABA-induced stomatal closure requires a reorganization of (hyponastic leaves 1) is a double strandedRNA binding protein
the actin cytoskeleton of guard cells (Eun and Lee, 1997), sug- and it was shown recently that HYL1 is involved in microRNA
gesting that the ABA-triggered cytoskeleton reorganization is production. Another modulator of ABA signalling Abh1 encodimportant for the process of stomatal closure. A link for such ing a mRNA cap binding protein was identified (Hugouvieux
an action has been demonstrated between ABA and actin through et al., 2001). Abh1 plants exhibited drought tolerance compared
AtRac1. AtRac1 (Arabidopsis Rho-related small guanosine triphos-to wild-type due to increased stomatal closure coupled with elphatase, GTPase) has been identified as a central component in evations in cytosolic calcium levels.
the ABA-mediated stomatal closure. In animals and yeast Rho
DNA chip analysis indicated only a few of the genes imGTPases are key regulators of the actin cytoskeleton. GTPases plicated in ABA signalling are downregulated in the mutant.
are inactivated by ABA treatment leading to the disruption of the One important gene is type 2C phosphoprotein phosphatase that
guard cell actin in wild-type but not in abi1-1 mutant. Further has been proposed to function as a negative regulator of ABA
expression of a dominant-negative mutant resulted in stomatal signal transduction (Sheen, 1998). It seems that ABH1 influclosure, while the dominant-positive mutant inhibited the induc- ences ABA signalling pathways upstream of [Ca2+ ], since abh1
tion of stomatal closure by ABA (Lemichez et al., 2001). A link mutant guard cells showed hypersensitive ABA-induced Ca2+
between ABA-induced increases in [Ca2+ ] and the cytoskeleton elevations when compared with wild-type. Sad1 (supersensitivreorganization has been also established (Hwang and Lee, 2001) ity to ABA and drought 1) mutant is impaired in the last step
suggesting that the ABA-induced increases in [Ca2+ ] is a com- of drought-induced ABA biosynthesis, i.e., conversion of ABA
ponent in the pathways leading to both alteration in the activity aldehyde to ABA (Xiong et al., 2001). SAD1 encodes a Smlike small ribonucleoprotein that is predicted to be involved in
of ion channels and the control of cytoskeletal reorganization.
mRNA splicing, export, and degradation. Further studies are reC. ABA Signalling Components
quired to establish the link between SAD1 expression and ABA
Characterization of ABA hypersensitive mutants led to the biosynthesis (Xiong and Zhu, 2003). Another RNA binding proidentification of several components of ABA signalling. ABA tein, AKIP1 was specifically phosphorylated by ABA-activated

48

D. BARTELS AND R. SUNKAR

AAPK that increases the affinity of AKIP1 for the dehydrin transcript (Li et al., 2002). How these proteins execute their role in
the ABA signalling pathway is unknown but the findings implicate a connection between RNA processing and ABA signalling.

XII.

MONITORING GLOBAL GENE EXPRESSION


USING MICROARRAY ANALYSIS
The field of plant stress tolerance has recently devoted considerable research effort towards identifying stress response genes.
This approach is largely boosted by the availability of Arabidopsis and rice genome sequences. Genomics technologies
now provide high-throughput integrated approaches for transcript and protein analysis, which should complete our view of
stress-activated gene expression and may also lead to novel gene
discovery. Although it is believed that high-throughput experimentation will be highly useful in finding the whole set of genes
that are involved in the process, single-gene studies remain essential to verify and evaluate high-throughput data sets. The first
available results on microarray data in relation to abiotic stresses
in Arabidopsis and rice will be summarized here. These studies
provide important information, given the complexity of plant
response to abiotic stresses. Seki et al. (2001) have constructed
Arabidopsis full-length cDNA microarrays using about 1300
full-length cDNAs. Forty-four genes were identified as drought
inducible, of which 30 genes are novel drought-inducible genes.
In a subsequent experiment, Seki et al. (2002) have analyzed
the expression of 7000 Arabidopsis full-length cDNAs in response to dehydration, high-salinity, and cold stress treatment.
Dehydration and high-salinity treatments resulted in increased
expression of 277 and 194 genes, respectively. Nearly 70 percent (141 genes) of the high-salinityinducible genes are also
induced by dehydration, emphasizing an extensive overlap of dehydration and salt stress. The same array experiment identified
40 transcription factor genes as dehydration, high-salinity, or
cold-inducible genes. These include six members of the DREB
family, two members of the ERF family, ten members of the Znfinger family, four members of the WRKY family, three members of the Myb family, two members of the bHLH family, four
members of the bZIP family, five members of the NAC family
and three members of the HDZIP.
In another study, a comparative analysis was made between
salt-tolerant and salt-sensitive rice varieties in terms of salt
shockinduced gene expression profiles (Kawasaki et al., 2001).
Salt-tolerant Pokkali rice plants survived a 150 mM salt shock
treatment, whereas salt-sensitive IR29 plants collapsed within
24 h of the treatment. Thirtythree percent of the Pokkali transcripts were upregulated after 1 h treatment with 150 mM NaCl
shock, whereas only 7 percent of the salt-sensitive IR29 transcripts were upregulated under similar conditions. Mostly, transcripts homologous with ribosomal proteins were upregulated
in Pokkali but not in IR29. After 3 h 38 percent of the transcripts
were altered in both varieties not showing significant differences
in the two lines. After 6 h of treatment, 38 percent of the tran-

scripts were downregulated in IR29, but only 13 percent were


downregulated in Pokkali. The differences in expression behavior between the two lines include a delay in the initial response by
IR29. For instance, one of the early responsive genes, CDPK-like
sequence was threefold upregulated in Pokkali but not in IR29
after 1 h of salt shock treatment, suggesting a difference in signal
transduction at the early stages of stress (Kawasaki et al., 2001).
Global expression profiles also provide an overview of those
genes, which are downregulated by dehydration or salt stress.
A significant percent of genes are downregulated by dehydration or salinity (Bockel et al., 1998; Kawasaki et al., 2001; Seki
et al., 2002; Oztur et al., 2002). For example, Seki et al. (2002)
have identified 79 and 89 Arabidopsis genes downregulated by
drought and high-salinity, respectively. Mostly, photosynthesisrelated genes and the components of PSI and PSII belong to
this group of genes (Seki et al., 2002; Bockel et al., 1998). A
logical hypothesis for such a response is that photosynthesis is
decreased as a consequence of stress imposition, which in turn
downregulates the components involved in this process. This
represents an adaptive mechanism to decrease the formation of
ROS. On the other hand, some of the downregulated genes may
be of direct adaptive value. For example, downregulation of proline dehydrogenase is important for proline accumulation under
osmotic stress conditions. In addition, some of the downregulated genes may include negative regulators of stress responses.
Microarray studies show that at least several hundred genes
are involved in response to drought or salt stress in plants. A careful analysis of the transcription profiles should reveal not only individual stress-activated genes but pathways. With the advances
in proteomics we should be able to simultaneously quantify the
levels of individual proteins or follow post-translational modifications that occur in response to abiotic stresses. Unfortunately,
the global expression analysis will at present be restricted to
a few species, in particular Arabidopsis, rice, and maize, from
which large EST collections are available. Interesting stresstolerant species are not amenable to this analysis because of
lack of tools. It remains to be seen to what extent the data can
be extrapolated from those species, which can undergo global
expression analysis.

XIII. CONCLUSIONS
Despite the fact that research efforts have produced an enormous amount of information, we are far from understanding the
complete circuits of stress reactions. Only a few components
of many pathways have been the subject of investigations. Future priorities should be aimed at the identification of molecules
connecting pathways and of key components in each pathway.
It is proposed that our understanding of plant stress tolerance
can be greatly refined by thorough characterization of individual genes and assessing their contribution to stress tolerance.
Such experiments indicated that many individual genes appear
to have some positive impact on stress tolerance; mainly master switches such as transcription factor or upstream signalling

DROUGHT AND SALT TOLERANCE IN PLANTS

molecules are promising candidate genes for biotechnological


approaches. However, a combination of such master genes that
act in different pathways, e.g., ROS scavenging and osmotic
adaptation, may prove even more beneficial for improving stress
tolerance.
ACKNOWLEDGMENTS
The work in the laboratory of D. Bartels was supported by
grants from the European Commission and the German research
Council (DFG). S. Ramanjulu was supported by a fellowship
from the Alexander von Humboldt Foundation. We appologize
to all colleagues whose work has not been cited due to restrictions in length.
REFERENCES
Abd-El Baki, G. K., Siefritz, F., Man, H. M., Weiner, H., Kaldenhoff, R., and
Kaiser, W. 2000. Nitrate reductase in zea mays L. under salinity. Plant Cell
Environ. 23: 515521.
Abe, H., Urao, T., Ito, T., Seki, M., Shinozaki, K., and Yamaguchi-Shinozaki,
K. 2003. Arabidopsis AtMYC2 (b-HLH) and AtMYB2 (MYB) function as
transcriptional activators in abscisic acid signalling. Plant Cell 15: 6378.
Abe, H., Yamaguchi-Shinozaki, K., Urao, T., Iwasaki, T., Hosakawa, D., and
Shinozaki, K. 1997. Role of Arabidopsis MYC and MYB homologs in
drought- and abscisic acid-regulated gene expression. Plant Cell 9: 1859
1868.
Abede, T., Guenzi, A. C., Martin, B., and Cushman, J. C. 2003. Tolerance of
mannitol-accumulating transgenic wheat to water stress and salinity. Plant
Physiol. 131: 17481755.
Ahmad, I. and Hellebust, J. A. 1988. The relationship between inorganic nitrogen
metabolism and proline accumulation in osmoregulatory responses of two
euryhaline microalgae. Plant Physiol. 88: 348354.
Akashi, K., Miyake, C., and Yokota, A. 2001. Citrulline, a novel compatible
solute in drought-tolerant wild watermelon leaves, is an efficient hydroxyl
radical scavenger. FEBS Lett. 508: 438442.
Alamillo, J., Almogura, C., Bartels, D., and Jordano, J. 1995. Constitutive expression of small heat shock proteins in vegetative tissues of the resurrection
plant Craterostigma planatgenium. Plant Mol. Biol. 29: 10931099.
Albrecht, V., Wein, S., Blazevic, D., DAngelo, C., Batistic, O., Kolukisaoglu,
U., Bock, R., Schulz, B., Harter, K., and Kudla, J. 2003. The calcium sensor
CBL1 integrates plant responses to abiotic stresses. Plant J. 36: 457470.
Allen, G. J., Kuchitsu, K., Chu, S. P., Murata, Y., and Schroeder, J. I. 1999.
Arabidopsis abi1-1 and abi2-1 phosphatase mutations reduce abscisic acidinduced cytosolic calcium rises in guard cells. Plant Cell 11: 17851798.
Allen, R. 1995. Dissection of oxidative stress tolerance using transgenic plants.
Plant Physiol. 107: 10491054.
Alvim, F. C., Carolino, S. M. B., Cascardo, J. C. M., Nunes, C. C., Martinez,
C. A., Otoni, W. C., and Fontes, E. P. B. 2001. Enhanced accumulation of
BiP in transgenic plants confers tolerance to water stress. Plant Physiol. 126:
10421054.
Amtmann, A. and Sanders, D. 1999. Mechanisms of Na+ uptake by plant cells.
Advances in Botanical Research 29: 75112.
Apel, K. and Hirt, H. 2004. Reactive oxygen species: Metabolism, oxidative
stress, and signal transduction. Annu. Rev. Plant. Biol. 55: 373399.
Apse, M. P., Aharon, G. S., Snedden, W. A., and Blumwald, E. 1999. Salt
tolerance conferred by overexpression of a vacuolar Na+/H+ antiport in
Arabidopsis. Science 285: 12561258.
Apse, M. P. and Blumwald, E. 2002. Engineering salt tolerance in plants. Curr.
Opin. Biotechnol. 13: 146150.
Arner, E. S. J. and Holmgren, A. 2000. Physiological functions of thioredoxin
and thioredoxin reductase. Eur. J. Biochem. 267: 61026109.

49

Assmann, S. M. and Wang, X. O. 2001. From milliseconds to millions of years:


Guard cells and environmental responses. Curr. Opin. Plant Biol. 4: 421428.
Badawi, G. H., Kawano, N., Yamauchi, Y., Shimada, E., Sasaki, R., Kubo, A.,
and Tanaka, K. 2004. Over-expression of ascorbate peroxidase in tobacco
chloroplasts enhances the tolerance to salt stress and water deficit. Physiol.
Plant 121: 231237.
Baker, S. S., Wilhelm, K. S., and Thomashow, M. F. 1994. The 5 -region of Arabidopsis thaliana cor15a has cis-acting elements that confer cold-, droughtand ABA-regulated gene expression. Plant Mol. Biol. 24: 701713.
Bartels, D. 2001. Targeting detoxification pathways: An efficient approach to
obtain plants with multiple stress tolerance? Trends Plant Sci. 6: 284286.
Bartels, D. and Salamini, F. 2001. Desiccation tolerance in the resurrection plant
Craterostigma plantagineum. A contribution to the study of drought tolerance
at the molecular level. Plant Physiol. 127: 13461353.
Bartels, D., Singh, M., and Salamini, F. 1988. Onset of desiccation tolerance
during development of the barley embryo. Planta 175: 485492.
Bastola, D. R., Pethe, V. V., and Winicov, I. 1998. Alfin1, a novel zinc finger
protein in alfalfa roots that binds to promoter in the salt-inducible MSPRP2
gene. Plant Mol. Biol. 38: 11231135.
Berlett, B. S. and Stadtman, E. R. 1997. Protein oxidation in aging, disease and
oxidative stress. J. Biol. Chem. 272: 2031320316.
Bianchi, G., Gamba, A., Murelli, C., Salamini, F., and Bartels, D. 1991.
Novel carbohydrate metabolism in the resurrection plant Craterostigma plantagineum. Plant J. 1: 355359.
Blumwald, E., Aharon, G. S., and Apse, M. P. 2000. Sodium transport in plant
cells. Biochimica et Biophysica Acta 1465: 140151.
Blumwald, E. and Gelli A. 1997. Secondary inorganic ion transport in plant
vacuoles. Adv. Bot. Res. 25: 401407.
Bockel, C., Salamini, F., and Bartels, D. 1998. Isolation and characterization
of genes expressed during early events of the dehydration process in the
resurrection plant Craterostigma plantagineum. J. Plant Physiol. 25: 158
166.
Bohnert, H. J. and Cushman, J. C. 2000. The ice plant cometh: Lessons in abiotic
stress tolerance. J. Plant Growth Regul. 19: 334346.
Borrell, A., Besford, R. T., Altabella, T., Masgrau, C., and Tiburcio, A. F. 1996.
Regulation of arginine decarboxylase by spermine in osmotically-stressed oat
leaves. Physiol. Plant 98: 105110.
Borsani, O., Valupesta, V., and Botella, M. A. 2001. Evidence for a role of
salicylic acid in the oxidative damage generated by NaCl and osmotic stress
in Arabidopsis seedlings. Plant Physiol. 126: 10241030.
Bouche, N., Scharlat, A., Snedden, W., Bouchez, D., and Fromm, H. 2002. A
novel family of calmodulin-binding transcription activators in multicellular
organisms. J. Biol. Chem. 277: 2185121861.
Bray, E. A. 1997. Plant responses to water deficit. Trends Plant Sci. 2: 4854.
Bray, E. A., Bailey-Serres J., and Weretilnyk, E. 2000. Responses to abiotic
stresses. In: Biochemistry and Molecular Biology of Plants. Gruissem, W.,
Buchnnan, B., and Jones, R. eds. American Society of Plant Physiologists,
Rockville, MD, 11581249.
Bressan, R. A., Hasegawa, P. M., and Pardo, J. M. 1998. Plants use calcium to
resolve salt stress. Trends Plant Sci. 3: 411412.
Bressan, R. A., Zhang, C. Q., Zhang, H., Hasegawa, P. M., Bohnert, H. J., and
Zhu, J. K. 2001. Learning from the Arabidopsis experience. The next gene
search paradigm. Plant Physiol. 127: 13541360.
Broin, M., Cuine, S., Eymery, F., and Rey, P. 2002. The plastidic 2-cystein
peroxiredoxin is a target for a thioredoxin involved in the protection of the
photosynthetic apparatus against oxidative damage. Plant Cell 14: 1417
1432.
Broin, M., Cuine, S., Peltier, G., and Rey, P. 2000. Involvement of CDSP32
a drought-induced thioredoxin in the response to oxidative stress in potato
plants. FEBS Lett. 467: 245248.
Browne, J., Tunnacliffe, A., and Burnell, A. 2002. Plant desiccation gene found
in a nematode. Nature 416: 38.
Burbridge, A., Grieve, T. M., Jackson, A., Thompson, A., and Taylor, I. B. 1997.
Structure and expression of a cDNA encoding a putative neoxanthin cleavage

50

D. BARTELS AND R. SUNKAR

enzyme (NCE) isolated from a wilt-related tomato (Lycopersicon esculentum


Mill) library. J. Exp. Bot. 47: 21112112.
Burg, M. B., Kwon, E. D., and Kultz, D. 1996. Osmotic regulation of gene
expression. FASEB J. 10: 15981606.
Busk, P. K. and Pages, M. 1998. Regulation of abscisic acid-induced transcription. Plant Mol. Biol. 37: 425435.
Campalans, A., Pages, M., and Messeguer, R. 2001. Identification of differentially expressed genes by the cDNA-AFLP technique during dehydration of
almond (Prunus amygdalus). Tree Physiol. 21: 633643.
Cascardo, J. C. M., Almeida, R. S., Buzeli, R. A. A., Carolino, S. M. B., Otoni,
W. C., and Fontes, E. P. B. 2000. The phosphorylation state and expression of
soybean BiP isoforms are differentially regulated following abiotic stresses.
J. Biol. Chem. 275: 1449414500.
Chandler, P. and Robertson, M. 1994. Gene expression regulated by abscisic
acid and its relation to stress tolerance. Annu. Rev. Plant Physiol. Plant Mol.
Biol. 25: 113141.
Chattopadhyay, M. K., Tiwari, B. S., Chattopadhyay, G., Bose, A., Sengupta, D.
N., and Ghosh, B. 2002. Protective role of exogenous polyamines on salinitystressed rice (Oryza sativa) plants. Physiol. Plant 116: 192199.
Chen, T. H. H. and Murata, N. 2002. Enhancement of tolerance of abiotic stress
by metabolic engineering of betaines and other compatible solutes. Curr.
Opin. Plant Biol. 5: 250257.
Chini, A., Grant, J. J., Seki, M., Shinozaki, K., and Loake, G. J. 2004. Drought tolerance established by enhanced expression of the CC-NBS-LRR gene, ADR1,
requires salicylic acid, EDS1 and ABI1. Plant J. 38: 810822.
Choi, H. I., Hong, J. H., Ha, J. O., Kang, J. Y., and Kim, S. Y. 2000. ABFs,
a family of ABA-responsive element binding factors. J. Biol. Chem. 275:
17231730.
Chung, W. S., Lee, S. H., Kim, J. C., Heo, W. D., Kim, M. C., Park, C. Y., Park,
H. C., Lim, C. O., Kim, W. B., Harper, J. F,. and Cho, M. J. 2000. Identification
of a calmodulin-regulated Ca2+-ATPase (SCA1) that is located in the plasma
membrane. Plant Cell 12: 13931407.
Coca, M., Almoguera, C., Thomas, T., and Jordano, J. 1996. Differentail regulation of small heat-shock genes in plants: Analysis of a water stress-inducible
and developmentally activated sunflower promoter. Plant Mol. Biol. 31: 863
876.
Cockcroft, C. E., den Boer, B. G. W., Healyy, J. M. S., and Murray, J. A. H.
2000. Cyclin D control of plant growth rate in plants. Nature 405: 575579.
Cohen, P. 1989. The structure and regulation of protein phosphatases. Annu. Rev.
Biochem. 58: 453508.
Conley, T. R., Sharp, R. E., and Walker, J. C. 1997. Water deficit rapidly stimulates the activity of a protein kinase in the elongation zone of the maize
primary root. Plant Physiol. 113: 219226.
Cosgrove, D. J. 1997. Relaxation in a high stress environment: The molecuklar
basis of extensible cell walls and cell enlargement. Plant Cell 9: 10311041.
Coursol, S., Fan, L. M., Le Stunff, H., Spiegel, S., Gilroy, S., and Assmann,
S. 2003. Sphingolipid signalling in Arabidopsis guard cells involves heterotrimeric G proteins. Nature 423: 651654.
Crowe, J. H., Hoekstra, F. A., and Crowe, L. M. 1992. Anhydrobiosis. Annu Rev
Plant Physiol. 54: 579599.
Cunningham, K. W. and Fink, G. R. 1996. Calcineurin inhibits VCX1-dependent
H+ /Ca2+ exchange and induced Ca2+ ATPases in Saccharomyces cerevisiae.
Mol. Cell Biol. 16: 22262237.

de Carvalho, M. H., dArcy-Lameta,


A., Roy-Macauley, H., Garcil, M., El
Maarouf, H., Pham-Thi, A.-T., and Zuily-Fodil, Y. 2001. Aspartic protease in
leaves of common bean (Phaseolus vulgaris L.) and cowpea (Vigna unguiculata L. Wasp): Enzymatic activity, gene expression and relation to drought
susceptibility. FEBS Lett. 492: 242246.
Delauney, A. J., Hu, C.-A. A., Kavi Kishor, P. B., and Verma, D. P. S. 1993.
Cloning of ornithine -aminotransferase cDNA from Vigna aconitifolia by
trans-complementation in Escherichia coli and regulation of proline biosynthesis. J. Biol. Chem. 268: 1867318678.
Delauney, A. and Verma, D. P. S. 1993. Proline biosynthesis and osmoregulation
in plants. Plant J. 4: 215223.

de Nadal, E., Alepuz, P. A., and Posas, F. 2002. Dealing with osmostress through
MAP kinase activation. EMBO Reports 3: 735740.
de Nadal, E., Clotet, J., Posas, F., Serrano, R., Gomez, N., and Arino, J. 1998.
The yeast halotolerance determinant Hal3p is an inhibitory subunit of the
Ppz1p Ser/Thr protein phosphatase. Proc. Natl. Acad. Sci. U S A 95: 7357
7362.
den Boer, B. G. W. and Murray, J. A. H. 2000. Triggering the cell cycle in plants.
Trends Cell Biol. 10: 245250.
Deng, X., Philips, J., Meijer, A., Salamini, F., and Bartels, D. 2002. Characterization of five novel dehydration-responsive homeodomain leucine zipper
genes from the resurrection plant Craterostigma plantagenium. Plant Mol.
Biol. 49: 601610.
De Wald, D. B., Torabinejad, J., Jones, C. A., Shope, J. C., Cangelosi, A. R.,
Thompson, J. E., Prestwich, G. D., and Hama, H. 2001. Rapid accumulation
of phosphatidylinositol 4,5-bisphosphate and inositol 1,4,5-trisphosphate correlates with calcium mobilization in salt-stressed Arabidopsis. Plant Physiol.
126: 759769.
Dietz, K.-J. and Arbinger, B. 1996. cDNA sequence and expression of subunit
E of the vacuolar H+ -ATPase in the inducible crassulacean acid metabolism
plant Mesembryanthemum crystallinum. Biochim. Biophys. Acta 1281: 134
138.
Dietz, K.-J., Horhing, F., Konig, J., and Baier, M. 2002. The function of chloroplast 2-cysteine peroxiredoxin in peroxide detoxification and its regulation.
J. Exp. Bot. 53: 13211329.
Drbak, B. K. and Watkins, P. A. C. 2000. Inositol(1,4,5)trisphosphate production in plant cells: an early response to salinity and hyperosmotic stress. FEBS
Lett. 481: 240244.
Dubouzet, J. G., Sakuma, Y., Ito, Y., Kasuga, M., Dubouzet, E., Miura, S., Seki,
M., Shinozaki, K., and Yamaguchi-Shinozaki. K. (2003). OsDREB genes in
rice, Oryza sativa L., encode transcription activators that function in drought-,
high-salt- and cold-responsive gene expression. Plant J. 33: 751763.
Dure, III L. 1993. Structural motifs in LEA proteins. In: Close T. J. Ed. Plant
Responses to cellular dehydration during environmental stress: pp 91103.
American Society of Plant Physiologists, Rockville, MN.
Dure, L. 2001. Occurrence of a repeating Il-mer amino acid sequence motif in
diverse organisms. Protein Pept. Lett. 8: 115122.
Dure, III L., Crouch, M., Harada, J., Ho, T. H. D., Mundy, J., Quatrano, R. S.,
Thomas, T., and Sung, Z. R. 1989. Common amino acid sequence domains
among the LEA proteins of higher plants. Plant Mol. Biol. 12: 475486.
Dure, III L., Greenway, S., and Galau, G. A. 1981. Developmental biochemistry of cotton seed embryogenesis and germination XIV. Changing mRNA
populations as shown in vitro and in vivo protein synthesis. Biochemistry 20:
41624168.
Ellul, P., Rios, G., Atares, A., Roig, L. A., Serrano, R., and Moreno, V. 2003.
The expression of the Saccharomyces cerevisiae HAL1 gene increases salt
tolerance in transgenic watermelon [Citrullus lanatus (Thunb.) Matsun. &
Nakai. ]. Theor. Appl. Genet. 107: 462469.
El-Maarouf, H., Arcy-Lameta, A., Gareil, M., Zuily-Fodil, Y., and Pham-Thi,
A.-T. 2001. Cloning and expression under drought of cDNAs coding for two
PI-PLCs in cowpea leaves. Plant Physiol. Biochem. 39: 167172.
El-Maarouf, H., Zuily-Fodil, Y., Gareil, M., Arcy-Lameta, A., and Pham-Thi,
A.-T. 1999. Enzymatic activity and gene expression under water stress of
phospholipase D in two cultivars of Vigna unguiculata L. Walp. differing in
drought tolerance. Plant Mol. Biol. 39: 12571265.
Espinosa-Ruiz, A., Belles, J. M., Serrano, R., and Culianez-Macia, F. A. 1999.
Arabidopsis thaliana AtHAL3: A flavoprotein related to salt and osmotic
tolerance and plant growth. Plant J. 20: 529539.
Eun, S. O. and Lee, Y. 1997. Actin filaments of guard cells are reorganized in
response to light and abscisic acid. Plant Physiol. 115: 14911498.
Finkelstein, R. R., Gampala, S. S., and Rock, C. D. 2002. Abscisic acid signaling
in seeds and seedlings. Plant Cell 14 Suppl: S15S45.
Flores, H. E. and Galston, A. W. 1982. Osmotic stress-induced polyamine accumulation in cereal leaves. I. Physiological parameters of the response. Plant
Physiol. 75: 102109.

DROUGHT AND SALT TOLERANCE IN PLANTS


Forsthoefel, N. R., Cushman, M. A. F., Ostrem, J. A., and Cushman, J. C.
1998. Induction of a cysteine protease cDNA from Mesembryanthemum crystallinum leaves by environmental stress and plant growth regulators. Plant Sci.
136: 195206.
Foyer, C. H., Descourvieres, P., and Kunert, K. J. 1994. Protection against oxygen
radicals, an important defense mechanism studied in transgenic plants. Plant
Cell Environ. 17: 507523.
Frandsen, G., Muller-Uri, F., Nielsen, M., Mundy, J., and Skriver, K. 1996.
Novel plant Ca2+ -binding protein expressed in response to abscisic acid and
osmotic stress. J. Biol. Chem. 271: 343348.
Frank, W., Munnik, T., Kerkmann, K., Salamini, F., and Bartels, D. 2000.
Water deficit triggers phospholipase D activity in the resurrection plant
Craterostigma plantagineum. Plant Cell 12: 111123.
Fray, R. G., Wallace, A., Grierson, D., and Lycett, G. W. 1994. Nucleotide
sequence and expression of a ripening and water stress-related cDNA from
tomato with homology to the MIP class of membrane channel proteins. Plant
Mol. Biol. 24: 539543.
Fukuda, A., Nakamura, A., Tagiri, A., Tanaka, H., Miyao, A., Hirochika, H., and
Tanaka, Y. 2004. Function, intracellular localization and the importance in salt
tolerance of a vacuolar Na+ /H+ antiporter from rice. Plant Cell Physiol. 45:
146159.
Fukada-Tanaka, S., Inagaki, Y., Yamaguchi, T., Saito, N., and Iida, S. 2000.
Colour-enhancing protein in blue petals. Nature 407: 581.
Fukushima, E., Arata, Y., Endo, T., Sonnewald, U., and Sato, F. 2001. Improved
salt tolerance of transgenic tobacco expressing apoplastic yeast-derived invertase. Plant Cell Physiol. 42: 245249.
Furini, A., Koncz, C., Salamini, F., and Bartels, D. 1997. High level transcription
of a member of a repeated gene family confers dehydration tolerance to callus
tissue of Craterostigma plantagineum. EMBO J. 16: 35993608.
Gago, G. M., Almoguera, C., Jordano, J., Gonzalez, D. H., and Chan, R. L. 2002.
Hahb-4, a homeobox-leucine zipper gene potentially involved in abscisic
acid-dependent responses to water stress in sunflower. Plant Cell Environ.
25: 633640.
Galau, G. A, Hughes, D. W., and Dure, III L. 1986. Abscisic acid induction of
cloned cotton late embryogenesis-abundant (Lea) mRNAs. Plant Mol. Biol.
7: 155170.
Gampala, S. S., Finkelstein, R. R., Sun. S. S., and Rock, C. D. 2002. ABI5
interacts with abscisic acid signalling effectors in rice protoplasts. J. Biol.
Chem. 277: 16891694.
Gampala, S. S., Hagenbeek, D., and Rock, C. D. 2001. Functional interactions
of lanthanium and phospholipase D with the abscisic acid signalling effectors
VP1 and ABI1-1 in rice protoplasts. J. Biol. Chem. 276: 98559860.
Garg, A. K., Kim, J-K., Owens, T. G., Ranwala, A. P., Choi, Y. D., Kochian, L.
V., and Wu, R. J. 2002. Trehalose accumulation in rice plants confers high
tolerance levels to different abiotic stresses. Proc. Nat. Acad. Sci USA 99:
1589815903.
Gaxiola, R. A., Rao, R., Sherman, A., Grisafi, P., Alper, S. L., and Fink, G. R.
1999. The Arabidopsis thaliana proton transporters, Atnhx1 and Avp1, can
function in cation detoxification in yeast. Proc. Natl. Acad. Sci. USA 96:
14801485.
Gaxiola, R., Li, J., Undurraga, S., Dang, L. M., Allen, G. J., Alper, S. L., and
Fink, G. R. 2001. Drought- and salt-tolerant plants result from overexpression
of the AVP1 H+ pump. Proc. Nat. Acad. Sci. USA 98: 1144411449.
Geisler, M., Axelsen, K. B., Harper, J. F., and Palmgren, M. G. (2000a). Molecular aspects of higher plant P-type Ca(2+)-ATPases. Biochim. Biophys. Acta
1465: 5278.
Geisler, M., Frangne, N., Gomes, E., Matinoia, E., and Palmgreen, M. G.
(2000b). The ACA4 gene of Arabidopsis encodes a vacuolar membrane calcium pump that improves salt tolerance in yeast. Plant Physiol. 124: 1814
1827.
Gilmour, S. J., Seblot, A. M., Salazar, M. P., Everard, J. D., and Thomashow,
M. F. 2000. Overexpression of the Arabidopsis CBF3 transcriptional activator
mimics multiple biochemical changes associated with cold acclimation. Plant
Physiol. 124: 18541865.

51

Golldack, D. and Dietz, K-J. 2001. Salt-induced expression of the vacuolar H+ATPase in the common ice plant is developmentally controlled and tissue
specific. Plant Physiol. 125: 16431654
Guerroro, F. and Mullet, J. E. 1986. Increased abscisic acid biosynthesis during
plant dehydration requires transcription. Plant Physiol. 80: 588591.
Guerrero, F. D., Jones, J. T., and Mullet, J. E. 1990. Turgor responsive gene
transcription and RNA levels increase rapidly when pea shoots are wilted:
sequence and expression of three inducible genes. Plant Mol. Biol. 15: 11
26.
Guiltinan, M. J., Marcotte, W. R., and Quatrano, R. S. 1990. A plant leucine
zipper protein recognizes an abscisic acid responsive element. Science 25:
267271.
Guo, Y., Halfter, U., Ishitani, M., and Zhu, J.-K. 2001. Molecular characterization
of functional domains in the protein kinase SOS2 that is required for plant
salt tolerance. Plant Cell 13: 13831400.
Gupta, A. S., Heinen, J. L., Holaday, A. S., Burke, J. J., and Allen, R. D. 1993.
Increased resistance to oxidative stress in transgenic plants that overexpress
chloroplastic Cu/Zn superoxide dismutase. Proc. Natl. Acad. Sci. USA 90:
16291633.
Haake, V., Cook, D. Riechmann, J. L., Pineda, O., Thomashow, M. F., and Zhang,
J. Z. 2002. Transcription factor CBF4 is a regulator of drought adaptation in
Arabidopsis. Plant Physiol. 130: 639648.
Halford, N. G. and Hardie, D. G. 1998. SNF-1 related protein kinases: Global
regulators of carbon metabolism in plants? Plant Mol. Biol. 37: 735748.
Halfter, U., Ishitani, M., and Zhu, J. -K. 2000. The Arabidopsis SOS2 protein
kinase physically interacts with and is activated by the calcium-binding protein
SOS3. Proc. Natl. Acad. Sci. USA 97: 37353740.
Halliwell, B. and Gutteridge, J. M. C. 1999. Free Radicals in Biology and
Medicine: 3rd ed., Oxford University Press, New York.
Hasegawa, P. M., Bressan, R., Zhu, J.-K., and Bohnert, H.-J. 2000. Plant cellular
and molecular responses to high salinity. Annu. Rev. Plant Physiol. Plant Mol.
Biol. 51: 463499.
Haushul, K., Andersson, B., and Adamska, I. 2001. A chloroplast DegP2 protease performs the primary cleavage of the photodamaged D1 protein in plant
photosystem II. EMBO J. 20: 713722.
Hayashi, H., Alia, Mustardy, L., Deshnium, P., Ida, M., and Murata, N. 1997.
Transformation of Arabidopsis thaliana with the codA gene for choline oxidase; accumulation of glycinebetaine and enhanced tolerance to salt and cold
stress. Plant J. 12: 133142.
Himmelbach, A., Hoffmann, T., Leube, M., Hohener, B., and Grill, E. 2002.
Homeodomain protein ATHB6 is a target of protein phosphatase ABI1
and regulates hormone responses in Arabidopsis. EMBO J. 21: 3029
3038.
Hirayama, T., Ohto, C., Mizoguchi, T., and Shinozaki, K. 1995. A gene encoding
a phosphotidylinositol-specific phospholipase C is induced by dehydration
and salt stress in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 92: 3903
3907.
Hoekstra, F. A., Golovina, E. A., and Buitink, J. 2001. Mechanisms of plant
desiccation tolerance. Trends Plant Sci. 6: 431438.
Hong, Z., Lakkineni, K., Zhang, Z., and Verma, D. P. S. 2000. Removal of
feedback inhibition of 1-pyrroline-5-carboxylate synthetase results in increased proline accumulation and protection of plants from osmotic stress.
Plant Physiol. 122: 11291136.
Hoshida, H., Tanaka, Y., Hibino, T., Hayashi, Y., Tanaka, A., Takabe, T., and
Takabe, T. 2000. Enhanced tolerance to salt stress in transgenic rice that
overexpresses chloroplast glutamine synthetase. Plant Mol. Biol. 43: 103
111.
Hsieh, T.-H., Lee, J.-T., Charng, Y.-Y., and Chan, M.-T. 2002. Tomato plants
ectopically expressing Arabidopsis CBF1 show enhanced resistance to water
deficit stress. Plant Physiol. 130: 618626.
Huang, J., Hirji, R., Adam, L., Rozwadowski, K. L., Hammerlindl, J. K., Keller,
W. A., and Selvaraj, G. 2000. Genetic engineering of glycinebetaine production toward enhancing stress tolerance in plants: metabolic limitations. Plant
Physiol. 122: 747756.

52

D. BARTELS AND R. SUNKAR

Huang, Y., Li, H., Gupta, R., Morris, P. C., Luan, S., and Kieber, J. J. 2000.
ATMPK4, an Arabidopsis homolog of mitogen-activated protein kinase, is activated in vitro by AtMEK1 through threonine phosphorylation. Plant Physiol.
122: 13011310.
Hugovieux, V., Kwak, J. M., and Schroeder, J. I. 2001. An mRNA cap binding
protein, ABH1, modulates early abscisic acid signal transduction. Cell 106:
477487.
Hwang, I. and Goodman, H. M. 1995. An Arabidopsis thaliana root-specific
kinase homolog is induced by dehydration, ABA, and NaCl. Plant J. 8: 37
43.
Hwang, J. and Lee, Y. 2001. Abscisic acid-induced actin reorganization in guard
cells of dayflower is mediated by cytosolic calcium levels and by protein
kinase and protein phosphatase activities. Plant Physiol. 125: 21202128.
Ichimura, K., Mizoguchi, T., Yoshida, R., Yuasa, T., and Shinozaki, K. 2000.
Various abiotic stresses rapidly activate Arabidopsis MAP kinases AtMAPK4
and AtMPK6. Plant J. 24: 655665.
Ichimura, K., Shinozaki, K., Tina, G., Sheen, J., Henry, Y. et al. 2002. Mitogenactivated protein kinase cascades in plants: A new nomenclature. Trends Plant
Sci. 7: 301308.
Ingram, J. and Bartels, D. 1996. The molecular basis of dehydration tolerance
in plants. Annu. Rev. Plant Physiol. Plant Mol. Biol. 47: 377403.
Ingram, J., Chandler, J. W., Gallagher, L., Salamini, F., and Bartels, D. 1997.
Analysis of cDNA clones encoding sucrose-phosphate synthase in relation to
sugar interconversions associated with dehydration in the resurrection plant
Craterostigma plantagineum Hochst. Plant Physiol. 115: 113121.
Iturriga, G., Leyns, L., Villegas, A., Gharabeih, R., Salamini, F., and Bartels,
D. 1996. A family of novel myb-related genes from the resurrection plant
Craterostigma plantagineum are specifically expressed in callus and roots in
response to ABA or desiccation. Plant Mol. Biol. 25: 707716.
Iuchi, S., Kobayashi, M., Taji, T., Naramoto, M., Seki, M., Kato, T., Tabata, S.,
Kakubari, Y., Yamaguchi-Shinozaki, K., and Shinozaki, K. 2001. Regulation
of drought tolerance by gene manipulation of 9-cis-epoxycarotenoid dioxygenase, a key enzyme in abscisic acid biosynthesis in Arabidopsis. Plant J.
27: 325333.
Iuchi, S., Kobayashi, M., Yamaguchi-Shinozaki, K., and Shinozaki, K. 2000.
A stress-inducible gene for 9-cis-epoxycarotenoid dioxygenase involved in
abscisic acid biosynthesis under water stress in drought tolerant cow pea.
Plant Physiol. 123: 553562.
Iwasaki, T., Kiyosue, T., Yamaguchi-Shinozaki, K., and Shinozaki, K. 1997.
The dehydration-inducible rd17 (cor47) gene and its promoter region in Arabidopsis thaliana. Plant Physiol. 115: 12871294.
Izawa, T., Foster, R., and Chua, N.-H. 1993. Plant bZIP Protein DNA Binding
Specificity J. Mol. Biol. 230: 11311144.
Jacob, T., Ritchie, S., Assmann, S. M., and Gilroy, S. 1999. Abscisic acid signal
transduction in guard cells is mediated by phospholipase D activity. Proc.
Natl. Acad. Sci. USA 95: 26972702.
Jacoby, M., Weisshaar, B., Droge-Laser, W., Vicente-Carbajosa, J., Tiedemann,
J., Kroj, T., and Parcy, F. 2002. bZIP transcription factors in Arabidopsis.
Trends Plant Sci. 7: 106111.
Jacoby, T., Flanagan, H., Faykin, A., Seto, A. G., Mattison, C., and Ota, I.
1997. Two protein-tyrosine phosphatases inactivate the osmotic stress response pathway in yeast by targeting the mitogen activated protein kinase,
HOG1. J. Biol. Chem. 272: 1774917755.
Jang, H. J., Pih, K. T., Kang, S. G., Lim, J. H., Jin, J. B., Piao, H. L., and Hwang,
I. 1998. Molecular cloning of a novel Ca2+ binding protein that is induced
by NaCl stress. Plant Mol. Biol. 37: 839847.
Jang, I.-C., Oh, S.-J., Seo, J.-S., Choi, W.-B., Song, S. I., Kim, C. H., Kim, Y.
S., Seo, H.-S., Choi, Y. D., Nahm, B. H., and Kim, J.-K. 2003. Expression of
a bifunctional fusion of the Escherichia coli genes for trehalose-6-phosphate
synthase and trehalose-6-phosphate phosphatase in transgenic rice plants increases trehalose accumulation and abiotic stress tolerance without stunting
growth. Plant Physiol. 131: 516524.
Johansson, I., Karlsson, M., Shukla, V. K., Chrispeels, M. J., Larsson, C.,
and Kjellbom, P. 1998. Water transport activity of the plasma membrane

aquaporin PM28A is regulated by phosphorylation. Plant Cell 25: 451


460.
Jonak, C., Kiegerl, S., Ligterink, W., Barker, P. J., Huskisson, N. S., and Hirt, H.
1996. Stress signaling in plants: A mitogen-activated protein kinase pathway
is activated by cold and drought. Proc. Natl. Acad. Sci. USA 93: 1127411279.
Jones, L. and McQueen-Mason, S. 2004. A role for expansins in dehydration
and rehydration of the resurrection plant Craterostigma plantagineum. FEBS
Lett. 559: 6165.
Kaldenhoff, R., Grote, K., Zhu, J. J., and Zimmermann, U. 1998. Significance of
plasmalemma aquaporins for water-transport in Arabidopsis thaliana. Plant
J. 14: 121128.
Kang, J., Choi, H., Im, M., and Kim, S. Y. 2002. Arabidopsis basic leucine zipper
proteins that mediate stress-responsive abscisic acid signaling. Plant Cell 14:
343357.
Kao, C. Y., Cocciolone, S. M., Vasil, I. K., and McCarty, D. R. 1996. Localization
and interaction of the cis-acting elements for abscisic acid, VIVIPAROUS1,
and light activation of the C1 gene of maize. Plant Cell 8: 11711179.
Kasinathan, V. and Wingler, A. 2002. Effect of reduced arginine decarboxylase
activity on salt tolerance and on polyamine formation during salt stress in
Arabidopsis thaliana. Physiol. Plant 121: 101107.
Kasuga, M., Liu, Q., Miura, S., YamaguchiShinozaki, K., and Shinozaki, K.
1999. Improving plant drought, salt, and freezing tolerance by gene transfer
of a single stress-inducible transcription factor. Nat. Biotechnol. 17: 287291.
Katagiri, T., Takahashi, S., and Shinozaki, K. 2001. Involvement of a novel
Arabidopsis phospholipase D, AtPLD in dehydration-inducible accumulation
of phosphatidic acid in stress signalling. Plant J. 26: 595605.
KaviKishor, P. B., Hong, Z., Miao, G.-H., Hu, C.-A. A., and Verma, D. P. S. 1995.
Over-expression of -pyrroline-5-carboxylate synthetase increases proline
production and confers osmotolerance in transgenic plants. Plant Physiol.
108: 13871394.
Kawasaki, S., Borchert, C., Deyholos, M., Wang, H., Brazille, S., Kawai, K.,
Galbraith, D., and Bohnert, H. 2001. Gene expression profiles during the
initial phase of salt stress in rice. Plant Cell 13: 889905.
Kende, H. and Zeevaart, J. 1997. The five classical plant hormones. Plant Cell
9: 11971210.
Khanna-Chopra, R., Srivalli, B., and Ahlawat, Y. 1999. Drought induces many
forms of cysteine proteases not observed during natural senescence. Biochem.
Biophys. Res. Commun. 255: 324327.
Kiegerl, S., Cardinale, F., Siligan, C., Gross, A., Baudouin, E., Liwosz, A.,
Ekloff, S., Till, S., Bogre, L., Hirt, H., and Meskiene, I. 2000. SIMKK, a
mitogen-activated protein kinase (MAPK) kinase is a specific activator of the
salt stress-induced MAPK, SIMK. Plant Cell 12: 22472258.
Kiegle, E., Moore, C. A., Haseloff, J., Tester, M. A., and Knight, M. R. 2000. Celltype-specific calcium responses to drought, salt and cold in the Arabidopsis
root. Plant J. 23: 267278.
Kirch, H.-H., Nair, A., and Bartels, D. 2001. Novel ABA- and dehydrationinducible aldehyde dehydrogenase genes isolated from the resurrection plant
Craterostigma plantagineum and Arabidopsis thaliana. Plant J. 28: 555567.
Kirsch, M., An, Z., Viereck, R., Low, R., and Rausch, T. 1996. Salt stress induces
an increased expression of V-type H(+)-ATPase in mature sugar beet leaves.
Plant Mol. Biol. 32: 543547.
Kizis, D. and Pages, M. 2002. Maize DRE-binding proteins DBF1 and DBF2
are involved in rab17 regulation through the drought-responsive element in
an ABA-dependent pathway. Plant J. 30: 679689.
Kjellbom, P., Larsson, C., Johansson, I., Karlsson, M., and Johanson, U. 1999.
Aquaporins and water homeostasis in plants. Trends Plant Sci. 25: 308314.
Knight, H. and Knight, M. R. 2001. Abiotic stress signalling pathways: Specificity and cross-talk. Trends Plant Sci. 6: 262267.
Knight, H., Trewavas, A. J., and Knight, M. R. 1997. Calcium signalling in
Arabidopsis thaliana responding to drought and salinity. Plant J. 12: 1067
1078.
Kobayashi, Y., Yamamoto, S., Minami, H., Kagaya, Y., and Hattori, T. 2004. Differential activation of the rice sucrose nonfermenting1-related protein kinase2
family by hyperosmotic stress and abscisic acid. Plant Cell 16: 11631177.

DROUGHT AND SALT TOLERANCE IN PLANTS


Kobrinsky, E., Mirshahi, T., Zhang, H, Jin, T., and Logothetis, D. E. 2000.
Receptor-mediated hydrolysis of plasma membrane messenger PIP2 leads to
K+-current desensitization. Nat. Cell. Bio. 2: 507514.
Koizumi, M., Yamaguchi-Shinozaki, K., Tsuji, H., and Shinozaki, K. 1993.
Structure and expression of two genes that encode distinct drought-inducible
cysteine proteinases in Arabidopsis thaliana. Gene 129: 175182.
Kopka, J., Pical, C., Gray, J. E., and Muller-Rober, B. 1998. Molecular and
enzymatic characterization of three phosphoinositide-specific phospholipase
C isoforms from potato. Plant Physiol. 116: 239250.
Kranner, I., Beckett, R. P., Wornik, S., Zorn, M., and Pfeifhofer, H. W. 2002.
Revival of resurrection plant correlates with its antioxidant status. Plant J.
31: 1324.
Kudla, J., Xu, Q., Harter, K., Gruissem, W., and Luan, S. 1999. Genes for
calcineurin B-like proteins in Arabidopsis are differentially regulated by stress
signals. Proc. Natl. Acad. Sci. USA 96: 47184723.
Kyriakis, J. M. and Avruch, J. 2001. Mammalian mitogen-activated protein
kinase signal transduction pathways activated by stress and inflammation.
Physiol. Rev. 81: 807869.
Lee, G., Pokala, N., and Vierling, E. 1995. Structure and in vitro molecular
chaperone activity of cytosolic small heat shock proteins from pea. J. Biol.
Chem. 270: 1043210438.
Lee, Y. H. and Chun, J. Y. 1998. A new homeodomain-leucin zipper gene from
Arabidopsis thaliana induced by water stress and abscisic acid. Plant Mol.
Biol. 37: 377384.
Lehr, A., Kirsch, M., Viereck, R., Schiemann, J., and Rausch, T. 1999. cDNA and
genomic cloning of sugar beet V-type H+ -ATPase subunit A and c isoforms:
Evidence for coordinate expression during plant development and coordinate
induction in response to high salinity. Plant Mol. Biol. 39: 46375.
Lemichez, E., Wu, Y., Sanchez, J.-P., Mettouchi, A., Mathur, J., and Chua, N.-H.
2001. Inactivation of AtRac1 by abscisic acid is essential for stomatal closure.
Genes Dev. 15: 18081816.
Leung, J., Bouvier-Durand, M., Morris, P. C., Guerrier, D., Chedfor, F., and
Giraudat, J. 1994. Arabidopsis ABA-response gene ABI1: Features of a
calcium-modulated protein phosphatase. Science 264: 14481452.
Leung, J. and Giraudat, J. 1998. Abscisic acid signal transduction. Annu. Rev.
Plant Physiol. Plant Mol. Biol. 25: 199221.
Leymann, B., Geelen, D., Quintero, F. J., and Blatt, M. R. 1999. A tobacco
syntaxin with a role in hormonal control of guard cell ion channels. Science
283: 537540.
Li, J., Kinoshita, T., Pandey, S., Ng, C. K-Y., Gygl, S. P., Shimazaki, K-I., and
Assmann, S. M. 2002. Modulation of an RNA-binding protein by abscisicacid-activated protein kinase. Nature 418: 793797.
Li, J., Wang, X. -Q., Watson, M. B., and Assmann, S. M. 2000. Regulation
of abscisic acid-induced stomatal closure and anion channels by guard cell
AAPK kinase. Science 287: 300303.
Lim, Y. S., Cha, M-K., Kim, H. K., Uhm, T. B., Park, J. W., Kim, K., and Kim, I.
H. 1993. Removal of hydrogen peroxide and hydroxyl radical by thiol-specific
antioxidant protein as a possible role in vivo. Biochem. Biophy. Res. Comm.
192: 273280.
Liu, J., Ishitani, M., Halfter, U., Kim, C.-S., and Zhu, J.-K. 2000. The Arabidopsis
thaliana SOS2 gene encodes a protein kinase that is required for salt tolerance.
Proc. Natl. Acad. Sci. USA 97: 37303734.
Liu, J. and Zhu, J.-K. 1997. An Arabidopsis mutant that requires increased
calcium for potassium nutrition and salt tolerance. Proc. Natl. Acad. Sci. USA
94: 1496014964.
Liu, J. and Zhu J.-K. 1998. A calcium sensor homolog required for plant salt
tolerance. Science 280: 19431945.
Liu, Q., Kasuga, M., Sakuma, Y., Abe, H., Miura, S., Yamaguchi-Shinozaki, K.,
and Shinozaki, K. 1998. Two transcription factors, DREB1 and DREB2, with
an EREBP/AP2 DNA binding domain separate two cellular signal transduction pathways in drought- and low-temperature-responsive gene expression,
respectively, in Arabidopsis. Plant Cell 10: 13911406.
Lu, C. and Fedoroff, N. 2000. A mutation in the Arabidopsis HYL1 gene encoding a dsRNA binding protein affects responses to abscisic acid, auxin, and
cytokinin. Plant Cell 12: 23512366.

53

Luan, S., Kudla, J., Rodriguez-Concepcion, M., Yalovsky, S., and Gruissem,
W. 2002. Calmodulins and calcineurin-B like proteins: Calcium sensors for
specific signal response coupling in plants. Plant Cell 14: S389S400.
Luttge, U. and Ratajczak, R. 1997. The physiology, biochemistry and molecular
biology of the plant vacuolar ATPase. Adv. Bot. Res. 25: 253296.
MacRobbie, E. A. C. 2002. Evidence for a role for protein tyrosine phosphatase
in the cytosol of ion release from the guard cell vacuole in stomatal closure.
Proc. Natl. Acad. Sci. USA 99: 1196311968.
Mani, S., van de Cotte, B., Van Montagu, M., and Verbruggen, N. 2002. Altered levels of proline dehydrogenase cause hypersensitivity to proline and
its analogs in Arabidopsis. Plant Physiol. 128: 7383.
Mansour, M. M. F. 1998. Protection of plasma membrane of onion epidermal
cells by glycinebetaine and proline against NaCl stress. Plant Physio. Biochem
36: 767772.
Mata, C. G. and Lamattina, L. 2001. Nitric oxide induces stomatal closure and
enhances the adaptive plant responses against drought stress. Plant Physiol.
126: 11961204.
Mazel, A., Leshem, Y., Tiwari, B. S., and Levin, A. 2004. Induction of salt
and osmotic stress tolerance by overexpression of an intracellular vesicle
trafficking protein AtRab7 (AtRabG3e). Plant Physiol. 134: 118128.
McCue, K. F., and Hanson, A. D. 1990. Drought and salt tolerance: towards
understanding and application. Trends Biotech. 8: 358362.
McKersie, B. D., Bowley, S. R., Harjanto, E., and Leprince, O. 1996. Waterdeficit tolerance and field performance of transgenic alfalfa overexpressing
superoxide dismutase. Plant Physiol. 111: 11771181.
McNeil, S. D., Rhodes, D., Russell, B. L., Nuccio, M. L., Shachar-Hill, Y.,
and Hanson A. D. 2000. Metabolic modeling identifies key constraints on an
engineered glycine betaine synthesis pathway in tobacco. Plant Physiol. 124:
153162.
Mendoza, I., Rubio, F., Rodriguez-Navarro, A., and Pardo, J. M. 1994. The protein phosphatase calcineurin is essential for NaCl tolerance of Saccharomyces
cerevisiae. J. Biol. Chem. 269: 87928796.
Mestichelli, L. J., Gupta, R. N., and Spenser, I. D. 1979. The biosynthetic route
from ornithine to proline. J. Biol. Chem. 254: 640647.
Michel, E., Salamini, R., Barels, E., Dale, P., Baga, M., and Szalay, A. 1993.
Analysis of a desiccation and ABA-responsive promoter isolated from the
resurrection plant Craterostigma plantagineum. Plant J. 4: 2940.
Michel, D. Furini, A., Salamini, F., and Bartels, D. 1994. Structure and regulation of an ABA- and desiccation-responsive gene from the resurrection plant
Craterostigma plantagineum. Plant Mol. Biol. 24: 549560.
Mikami, K., Katagiri, T., Luchi, S., Yamaguchi-Shinozaki, K., and Shinozaki, K.
1998. A gene encoding phosphatidylinositol 4-phosphate 5-kinase is induced
by water stress and abscisic acid in Arabidopsis thaliana. Plant J. 15: 563
568.
Mikolajczyk, M., Awotunde, O. S., Muszynska, G., Klessig, D. F., and
Dobrowolska, G. 2000. Osmotic stress induced rapid activation of a salicylic acid-induced protein kinase and a homolog of protein kinase ASK1 in
tobacco cells. Plant Cell 12: 165178.
Mittler, R. 2002. Oxidative stress, antioxidants and stress tolerance. Trends Plant
Sci. 7: 405410.
Miyazaki, S., Koga, R., Bohnert, H. J., and Fukuhara, T. 1999. Tissue- and
environmental response-specific expression of 10PP2C transcripts in Mesembryanthemum crystallinum. Mol. Gen. Genet. 261: 307316.
Mizoguchi, T., Irie, K., Hirayama, T., Hayashida, N., Yamaguchi-Shinozaki,
K., Matsumoto, K., and Shinozaki, K. 1996. A gene encoding a mitogenactivated protein kinase kinase kinase is induced simultaneously with genes
for a mitogen-activated protein kinase and an S6 ribosomal protein kinase by
touch, cold, and water stress in Arabidopsis thaliana. Proc. Natl. Acad. Sci.
USA 93: 765769.
Moskovitz, J., Berlett, B. S., Poston, J. M., and Stadtman, E. R. 1999. Methionine sulfoxide reductase in antioxidant defense. Methods Enzymol. 300: 239
244.
Mowla, S. B., Thomson, J., Farrant, J. M., and Mundree, S. G. 2002. A novel
stress-inducible antioxidant enzyme identified from the resurrection plant
Xerophyta viscosa baker. Planta 215: 716726.

54

D. BARTELS AND R. SUNKAR

Mundree, S. G., Whittaker, A., Thomson, J., and Farrant, J. M. 2000. An aldose reductase homolog from the resurrection plant Xerophyta viscosa Baker.
Planta 211: 693700.
Mundy, J., Yamaguchi-Shinozaki, K., and Chua, N. H. 1990. Nuclear proteins
bind conserved elements in the abscisic acid-responsive promoter of a rice
rab gene. Proc. Natl. Acad. Sci. USA 87: 14061410.
Munnik, T. 2001. Phosphatidic acid: an emerging plant lipid second messenger.
Trends Plant Sci. 6: 227233.
Munnik, T., Ligternick, W., Meskiene, I., Calderini, O., Beyerly, J., Musgrave,
A., and Hirt, H. 1999. Distinct osmo-sensing protein kinase pathways are
involved in signalling moderate and severe hyper-osmotic stress. Plant J. 20:
381388.
Munnik, T. and Meijer, H. J. G. 2001. Osmotic stress activates distinct lipid and
MAPK signalling pathways in plants. FEBS Lett. 498: 172178.
Munns, R. 2002. Comparative physiology of salt and water stress. Plant Cell
Environ. 25: 239250.
Munns, R., Passioura, J. B., Guo, J., Ehazen, O., and Cramer, G. R. 2000. Water
relations and leaf expansion: Importance of timing. J. Exp. Bot. 51: 1495
1504.
Murata, Y., Pei, Z. M., Mori, I. C., and Schroeder, J. I. 2001. Abscisic acid activation of plasma membrane Ca2+ channels in guard cells require cytosolic
NAD(P)H and is differentially disrupted upstream and downstream of reactive oxygen species production in abi1-1 and abi2-1 protein phosphatase 2C
mutants. Plant Cell 13: 25132523.
Mustilli, A.-C., Merlot, S., Vavasseur, A., Fenzi, F., and Giraudat, J. 2002. Arabidopsis OST1 protein kinase mediates the regulation of stomatal aperture by
abscisic acid and acts upstream of reactive oxygen species production. Plant
Cell 14: 30893099.
Nakashima, K., Kiyosue, T., Yamaguchi-Shinozaki, K., and Shinozaki, K. 1997.
A nuclear gene, erd1, encoding a chloroplast-targeted Clp protease regulatory
subunit homolog is not only induced by water stress but also developmentally
up-regulated during senescence in Arabidopsis thaliana. Plant J. 12: 851861.
Nakashima, K., Shinwari, Z. K., Sakuma, Y., Seki, M., Miura, S., Shinozaki,
K., and Yamaguchi-Shinozaki, K. 2000. Organization and expression of
two Arabidopsis DREB2 genes encoding DRE-binding proteins involved in
dehydration- and high-salinity-responsive gene expression. Plant Mol. Biol.
42: 657665.
Nanjo, T., Kobayashi, M., Yoshiba, Y., Kakubari, Y., Yamaguchi-Shinozaki,
K., and Shinozaki, K. 1999. Antisense suppression of proline degradation
improves tolerance to freezing and salinity in Arabidopsis thaliana. FEBS
Lett. 461: 205210.
Narasimhan, M. L., Binzel, M. L., Perez-Prat, E., Chen, Z., Nelson, D. E., Singh,
N. K., Bressan, R. A., and Hasegawa, P. M. 1991. NaCl regulation of tonoplast
ATPase 70-kilodalton subunit mRNA in tobacco cells. Plant Physiol. 97: 562
568.
Neill, S. J., Burnett, E. C., Desikan, R., and Hancock, J. T. 1998. Cloning of
a wilt-responsive cDNA from an Arabidopsis thaliana suspension culture
cDNA library that encodes a putative 9-cis-epoxycarotenoid dioxygenase. J.
Exp. Bot. 49: 18931894.
Ng, C. K.-Y., Carr, K., McAinsh, M. R., Powell, B., and Hetherington, A. M.
2001. Drought-induced guard cell signal transduction involves sphingosine1-phosphate. Nature 410: 596599.
Niu, X., Bressan, R. A., Hasegawa, P. M., and Pardo, J. M. 1995. Ion homeostasis
in NaCl stress environments. Plant Physiol. 109: 735742.
Nonami, H. and Boyer, J. S. 1990. Primary events regulating stem growth at low
water potentials. Plant Physiol. 94: 16011609.
Novillo, F., Alonso, J. M., Ecker, J. R., and Salinas, J. 2004. CBF2/DREB1C is
a negative regulator of CBF1/DREB1B and CBF3/DREB1A expression and
plays a central role in stress tolerance in Arabidopsis. Proc. Natl. Acad. Sci.
USA 101: 39853990.
Nozawa, A., Koizumi, N., and Sano, H. 2001. An Arabidopsis SNF-1related
protein kinase, atsr1, interacts with a calcium-binding protein, AtCBL2, of
which transcripts respond to light. Plant Cell Physiol. 42: 976981.
Nuccio, M. L., McNeil, S. D., Ziemak, M. J., Hanson, A. D., Jain, R. K., and
Selvaraj, G. 2000. Choline import into chloroplasts limits glycine betaine

synthesis in tobacco: analysis of plants engineered with a chloroplastic or a


cytosolic pathway. Metab Eng. 2: 300311.
Nuccio, M. L., Rhodes, D., McNeil, S. D., and Hanson, A. D. 1999. Metabolic
engineering of plants for osmotic stress resistance. Curr. Opin. Plant Biol. 2:
128134.
Nuccio, M. L., Russell, B. L., Nolte, K. D., Rathinasabapathi, B., Gage, D. A.,
and Hanson, A. D. 1998. The endogenous choline supply limits glycine betaine synthesis in transgenic tobacco expressing choline monooxygenase.
Plant J. 16: 101110.
Oberschall, A., Deak, M., Torok, K., Saa, L., Vass, I., Kovacs, I., Feher, A.,
Dudits, D., and Horvath, G. V. 2000. A novel aldose/aldehyde reductase protects transgenic plants against lipid peroxidation under chemical and drought
stresses. Plant J. 24: 437446.
Ohta, M., Hayashi, Y., Nakashima, A., Hamada, A., Tanaka, A., Nakamura,
T., and Hayakawa, T. 2002. Introduction of a Na+ /H+ antiporter gene from
Atriplex gmelini confers salt tolerance to rice. FEBS Lett. 532: 279282.
Osorio, J., Osorio, M. L., Chaves, M. M., and Pereira, J. S. 1998. Water deficits
are more important in delaying growth than in changing patterns of carbon
allocation in Eucalyptus globulus. Tree Physiol. 18: 363373.
Oztur, Z. N., Talame, V., Deyholds, M., Michalowski, C. B., Galbraith, D. W.,
Gozukirmizi, N., Tubrosa, R., and Bohnert, H. J. 2002. Monitoring largescale changes in transcript abundance in drought- and salt-stressed barley.
Plant Mol. Biol. 48: 551573.
Pandey, S. and Assmann, S. 2004. The Arabidopsis putative G protein-coupled
receptor GCR1 interacts with the G protein subunit GPA1 and regulates
abscisic acid signalling. Plant Cell 16: 16161632.
Pandey, S., Tiwari, S. B., Tyagi, W., Reddy, M. K., Upadhyaya, K. C., and Sopory,
S. K. 2002. A Ca2+ /CaM-dependent kinase from pea is stress regulated and
in vitro phosphorylates a protein that binds to AtCaM5 promoter. Eur. J.
Biochem. 269: 31933204.
Pardo, J. M., Reddy, M. P., Yang, S., Maggio, A., Huh, G. H., Matsumoto, T.,
Coca, M. A., Paino-DUrzo, M., Koiwa, H., Yun, D. J., Watad, A. A., Bressan,
R. A., and Hasegawa, P. M. 1998. Stress signaling through Ca2+/calmodulindependent protein phosphatase calcineurin mediates salt adaptation in plants.
Proc. Natl. Acad. Sci. USA 95: 96819686.
Park, J. M., Park, C.-J., Lee, S.-B., Ham, B.-K., Shin, R., and Paek, K.-H.
2001. Overexpression of the tobacco Tsi1 gene encoding an EREBP/AP2-type
transcription factor enhances resistance against pathogen attack and osmotic
stress in tobacco. Plant Cell 13: 10351046.
Patharkar, O. R. and Cushman, J. C. 2000. A stress-induced calcium-dependent
protein kinase from Mesembryanthemum crystallium phosphorylates a twocomponent pseudo-response regulator. Plant J. 24: 679691.
Pei, Z. M., Ghassemian, Kwak, C. M., McCourt, P., and Schroeder, J. I. 1998.
Role of farnesyltransferase in ABA regulation of guard cell anion channels
and plant water loss. Science 282: 287290.
Pei, Z. M., Murata, Y., Benning, G., Thomine, S., Klusener, B., Allen, G. J.,
Grill, E., and Schroeder, J. I. 2000. Calcium channels activated by hydrogen
peroxide mediate abscisic acid signaling in guard cells. Nature 406: 731734.
Peng, Z., Lu, Q., and Verma, D. P. 1996. Reciprocal regulation of 1-pyrroline5-carboxylate synthetase and proline dehydrogenase genes controls proline
levels during and after osmotic stress in plants. Mol. Gen. Genet. 253: 334
341.
Perruc, E., Charpenteau, M., Ramirez, B. C., Jauneau, A., Galaug, J.-P., Ranjeva,
R., and Ranty, B. 2004. A novel calmodulin-binding protein functions as a negative regulator of osmotic stress tolerance in Arabidopsis thaliana seedlings.
Plant J. 38: 410420.
Phillips, J. R., Oliver, M. J., and Bartels, D. 2002. Molecular genetics of desiccation and tolerant systems. In CAB International Desiccation and Survival in
Plants: Drying without dying, M. Black and H. Pritchard. (Eds.), pp. 319341.
Piao, H. L., Lim, J. H., Kim, S. J., Cheong, G. W., and Hwang, I. 2001. Constitutive over-expression of AtGSK1 induces NaCl stress responses in the absence
of NaCl stress and results in enhanced NaCl tolerance in Arabidopsis. Plant
J. 27: 305314.
Pical, C., Westergren, T., Dove, S. K., Larsson, C., and Sommarin,
M. 1999. Salinity and hyperosmotic stress induce rapid increases in

DROUGHT AND SALT TOLERANCE IN PLANTS


phosphotidylinositol-4,5-bisphosphate, diacylglycerol pyrophosphate, and
phosphotidylcholine in Arabidopsis thaliana cells. J. Biol. Chem. 274: 38232
38240.
Pilon-Smits, E. A. H., Ebskamp, M. J. M., Paul, M. J., Jeuken, M. J. W.,
Weisbeek, P. J., and Smeekens, S. C. M. 1995. Improved performance of
transgenic fructan-accumulating tobacco under drought stress. Plant Physiol.
25: 125130.
Pilon-Smits, E. A. H., Terry, N., Sears Tobin, K. H., and Van Dun, K. 1999.
Enhanced drought resistance in fructan-producing sugar beet. Plant Physiol.
Biochem 25: 313317.
Pilon-Smits, E. A. H., Terry, N., Sears Tobin, K. H. et al. 1998. Trehaloseproducing transgenic tobacco plants show improved growth performance under drought stress. J Plant Physiol. 25: 525532.
Posas, F., Wurgler-Murphy, S. M., Maeda, T., Witten, E. A., Thai, T. C., and
Saito, H. 1996. Yeast HOG1 MAPkinase cascade is regulated by a multistep phosphorelay mechanism in the SLN1-YPD1-SSK1 two-component
osmosensor. Cell 86: 865875.
Pruvot, G., Cuine, S., Peltier, G., and Rey, P. 1996. Characterization of a novel
drought-induced 34-kDa protein located in the thylakoids of Solanum tuberosum L. plants. Planta 25: 471479.
Qin, X. and Zeevaart, J. A. D. 1999. The 9-cis-epoxycarotenoid cleavage reaction
is the key regulatory step of abscisic acid biosynthesis in water-stressed bean.
Proc. Natl. Acad. Sci. USA 96: 1535415361.
Qin, X. and Zeevaart, J. A. D. 2002. Overexpression of a 9-cis-epoxycarotenoid
dioxygenase gene in Nicotiana plumbaginifolia increases abscisic acid and
phaseic acid levels and enhances drought tolerance. Plant Physiol. 128: 544
551.
Qiu, Q. S., Guo, Y., Dietrich, M., Schumaker, K. S., and Zhu, J.-K. 2002. Regulation of SOS1, a plasma membrane Na+ /H+ exchanger in Arabidopsis thaliana
by SOS2 and SOS3. Proc. Natl. Acad. Sci. USA 99: 84368441.
Qiu, Q.-S., Guo, Y., Quintero, F. J., Pardo, J. M., Schumaker, K. S., and Zhu,
J.-K. 2004. Regulation of vacuolar Na+ /H+ exchange in Arabidopsis thaliana
by the salt-overly-sensitive (SOS) pathway. J. Biol. Chem. 279: 207215.
Quintero, F. J., Ohta, M., Shi, H., Zhu, J.-K., and Pardo, J. M. 2002. Reconstitution in yeast of the Arabidopsis SOS signaling pathway for Na+ homeostasis.
Proc. Natl. Acad. Sci. USA 99: 90619066.
Ramanjulu, S. and Bartels, D. 2002. Drought- and desiccation-induced modulation of gene expression in plants. Plant Cell Environ. 25: 141151.
Ramanjulu, S. and Sudhakar, C. 1997. Drought tolerance is partly related to
amino acid accumulation and ammonia assimilation: A comparative study in
two mulberry genotypes differing in drought sensitivity. J Plant Physiol. 150:
345350.
Ramanjulu, S., Veeranjaneyulu, K., and Sudhakar, C. 1994a. Relative tolerance
of certain mulberry (Morus alba L.) varieties to NaCl salinity. Serecologia
34: 695702.
Ramanjulu, S., Veeranjaneyulu, K., and Sudhakar, C. 1994b. Short-term shifts
in nitrogen metabolism in mulberry, Morus alba under salt shock. Phytochem.
37: 991995.
Ratajczak, R., Richter, J., and Luttge, U. 1994. Adaptation of the tonoplast Vtype H+ -ATPase of Mesembryanthemum crystallinum to salt stress, C3-CAM
transition and plant age. Plant Cell Environ. 17: 11011112.
Reiser, V., Raitt, D. C., and Saito, H. 2003. Yeast osmosensor Sln1 and plant
cytokinin receptor Cre1 respond to changes in turgor pressure. J. Cell. Biol.
161: 10351040.
Rentel, M. C., Lecourieux, D., Ouaked, F., Usher, S. L., Petersen, L., Okamoto,
H., Knight, H., Peck, S. C., Grierson, C. S., Hirt, H., and Knight, M. R.
2004. OXI1 kinase is necessary for oxidative burst-mediated signalling in
Arabidopsis. Nature 427: 858861.
Rey, P., Pruvot, G., Becuwe, N., Eymery, F., Rumeau, D., and Peltier, G. A.
1998. A novel thioredoxin-like protein located in the chloroplast is induced
by water deficit in Solanum tuberosum L. plants. Plant J. 25: 97107.
Rhodes, D. and Hanson, A. D. 1993. Quaternary ammonium and tertiary sulfonium compounds in higher plants. Annu Rev Plant Physiol Plant Mol Biol
44: 357384.

55

Riechmann, J., Heard, G., Martin, L., Reuber, C.-Z., Jiang, J., Keddie, L., Adam,
O., Pineda, O. J., Ratcliffe, R. R., Samaha, R., Creelman, M., Pilgrim, P.,
Broun, J. Z., Zhang, D., Ghandehari, B. K., Sherman, G., and Yu L. 2000.
Arabidopsis transcription factors: Genome-wide comparative analysis among
eukaryotes. Science 290: 21052110.
Robinson, M. J. and Cobb, M. H. 1997. Mitogen-activated protein kinase pathways. Curr. Opin. Cell Biol. 9: 180186.
Rock, C. 2000. Pathways to abscisic acid-regulated gene expression. New Phytol.
148: 357396.
Rodrigo, M.-J., Moskovitz, J., Salamini, F., and Bartels, D. 2002. Reverse genetic
approaches in plants and yeast suggest a role for novel, evolutionarily conserved, selenoprotein-related genes in oxidative stress defense. Mol. Genet.
Genomics 267: 613621.
Rontein, D., Basset, G., and Hanson, A. D. 2002. Metabolic engineering of
osmoprotectant accumulation in plants. Metab. Eng. 4: 4956.
Roosens, N., Hal Bitar, F., Loenders, K., Angenon, G., and Jacobs, M. 2002.
Overexpression of ornithine--aminotransferase increases proline biosynthesis and confers osmotolerance in transgenic plants. Mol. Breed. 9: 7380.
Roosens, N. H. J., Thu, T. T., Iskandar, H. M., and Jacobs, M. 1998. Isolation of
ornithine--aminotransferase cDNA and effect of salt stress on its expression
in Arabidopsis. Plant Physiol. 117: 263271.
Roxas, V. P., Smith, Jr, R. K., Allen, E. R., and Allen, R. D. 1997. Overexpression of glutathione-S-transferase/glutathione peroxidase enhances the growth
of transgenic tobacco seedlings during stress. Nat. Biotechnol. 15: 988
991.
Roy, M. and Wu, R. 2001. Arginine decarboxylase transgene expression and
analysis of environmental stress tolerance in transgenic rice. Plant Sci. 160:
869875.
Rus, A., Yokoi, S., Sharkhuu, A., Reddy, M., Lee, B. H., Matsumoto, T. K.,
Koiwa, H., Zhu, J. K., Bressan, R. A., and Hasegawa, P. M. 2001. AtHKT1 is
a salt tolerance determinant that controls Na(+) entry into plant roots. Proc
Natl Acad Sci USA. 98: 141501455.
Ryu, S. B. and Wang, X. 1995. Expression of phospholipase D during castor
bean senescence. Plant Physiol. 108: 713719.
Saijo, Y., Hata, S., Kyozuka, J., Shimamoto, K., and Izui, K. 2000. Overexpression of a single Ca2+ -dependent protein kinase confers both cold and
salt/drought tolerance on rice plants. Plant J. 23: 319327.
Saito, S., Hirai, N., Matsumoto, C., Ohigashi, H., Ohta, D., Sakata, K., and
Mizutani, M. 2004. Arabidopsis CYP707As encode (+)-abscisic acid 8 hydroxylase, a key enzyme in the oxidative catabolism of abscisic acid. Plant
Physiol. 134: 14391449.
Sakuma, Y., Liu, Q., Dubouzet, J. G., Abe, H., Shinozaki, K., and YamaguchiShinozaki, K. 2002. DNA-binding specificity of the ERF/AP2 domain of
Arabidopsis DREBs, transcription factors involved in dehydration- and coldinducible gene expression. Biochem. Biophys. Res. Commun. 290: 998
1009.
Sanders, D., Brownlee, C., and Harper, J. F. 1999. Communicating with calcium.
Plant Cell 11: 691706.
Sang, Y., Zhang, S., Li, W., Huang, and Wang, X. 2001. Regulation of plant
water loss by manipulating the expression of phospholipase D. Plant J. 28:
135144.
Sarda, X., Tousch, D., Ferrare, K., Cellier, F., Alcon, C., Dupuis, J. M., Casse, F.,
and Lamaze, T. 1999. Characterization of closely related delta-TIP genes encoding aquaporins which are differentially expressed in sunflower roots upon
water deprivation through exposure to air. Plant Mol. Biol. 25: 179191.
Schaffner, A. R. 1998. Aquaporin function, structure, and expression: Are there
more surprises to surface in water relations? Planta 25: 131139.
Schroeder, J. I., Kwak, J. M., and Allen, G. J. 2001. Guard cell abscisic acid
signalling and engineering drought hardiness in plants. Nature 410: 327330.
Schuppler, U., He, P.-H., John, P. C. L., and Munns, R. 1998. Effect of water
stress on cell division and cell-division-cycle 2-like cell-cycle kinase activity
in wheat leaves. Plant Physiol. 117: 667678.
Schweighofer, A., Hirt, H., and Meskiene, I. 2004. Plant PP2C phosphatases:
Emerging functions in stress signalling. Trends Plant Sci. 9: 236243.

56

D. BARTELS AND R. SUNKAR

Seki, M., Narusaka, M., Abe, H., Kasuga, M., Yamaguchi-Shinozaki, K.,
Carninic, P., Hayashizaki, Y., and Shinozaki, K. 2001. Monitoring the
expression pattern of 1300 Arabidopsis genes under drought and cold stresses
by using a full-length cDNA microarray. Plant Cell 13: 6172.
Seki, M., Narusaka, M., Ishida, J., Nanjo, T., Fujita, M., Oono, Y., Kamiya,
A., Nakajima, M., Enju, A., Sakurai, T., Satou, M., Akiyama, K., Taji, T.,
Yamaguchi-Shinozaki, K., Carninci, P., Kawai, J., Hayashizaki, Y., and Shinozaki, K. 2002. Monitoring the expression profiles of ca. 7000 Arabidopsis
genes under drought, cold, and high-salinity stresses using a full-length cDNA
microarray. Plant J. 31: 279292.
Serraj, R. and Sinclair, T. R. 2002. Osmolyte accumulation: Can it really help
increase in crop yield under drought conditions? Plant Cell Environ. 25: 333
341.
Serrano, R. 1996. Salt tolerance in plants and microorganisms: Toxicity targets
and defense responses. Int. Rev. Cytol. 165: 152.
Serrano, R., Mulet, J. M., Rios. G., Marquez. J. A., de Larrinoa, I., Leube, M.
P., Mendizabal, I., Pascual-Ahuir, A., Proft, M., Ros, R., and Montesinos, C.
1999. A glimpse of the mechanisms of ion homeostasis during salt stress. J.
Exp. Bot. 50: 10231036.
Setter, T. L. and Flannigan, B. A. 2001. Water deficit inhibits cell division and
expression of transcripts involved in cell proliferation and endoreduplication
in maize endosperm. J. Exp. Bot. 52: 14011408.
Shalata, A., Mittova, V., Volokita, M., Guy, M., and Tal, M. 2001. Response of
the cultivated tomato and its wild salt-tolerant relative Lycopersicon pennellii
to salt-dependent oxidative stress: The root antioxidative system. Physiol.
Plant 112: 487494.
Sharp, R. E., Hsiao, T. C., and Silk, W. K. 1988. Growth of the maize primary
root at low water potentials. I. Spatial distribution of expansive growth. Plant
Physiol. 87: 5057.
Sharp, R. E., Hsiao, T. C., and Silk, W. K. 1990. Growth of the maize primary
root at low water potentials. II. Role of growth and deposition of hexose and
potassium in osmotic adjustment. Plant Physiol. 93: 13371346.
Sheen, J. 1996. Ca2+ -dependent protein kinases and stress signal transduction
in plants. Science 274: 19001902.
Sheen, J. 1998. Mutational analysis of protein phosphatase 2C involved in abscisic acid signal transduction in higher plants. Proc. Natl. Acad. Sci. USA
95: 97580.
Shen, B., Jensen, R. G., and Bohnert, H. J. 1997a. Mannitol protects against
oxidation by hyderoxyl radicals. Plant Physiol. 115: 527532.
Shen, B., Jensen, R. G., and Bohnert, H. J. 1997b. Increased resistance to oxidative stress in transgenic plants by targeting mannitol biosynthesis to chloroplasts. Plant Physiol. 113: 11771183.
Shen, Q., Zhang, P., and Ho, T. -H. D. 1996. Modular nature of abscisic acid
(ABA) response complexes: Composite promoter units that are necessary
and sufficient for ABA induction of gene expression in barley. Plant Cell 8:
11071119.
Sheveleva, E., Chmara, W., Bohnert, H. J., and Jensen, R. G. 1997. Increased
salt and drought tolerance by D-ononitol production in transgenic Nicotiana
tabacum L. Plant Physiol. 115: 12111219.
Shi, H., Ishitani, M., Kim, C., and Zhu, J.-K. 2000. The Arabidopsis thaliana
salt tolerance gene SOS1 encodes a putative Na+/H+ antiporter. Proc. Natl.
Acad. Sci. USA 97: 68966901.
Shi, H., Lee, B.-H., Wu, S.-J., and Zhu, J.-K. 2003. Overexpression of a plasma
membrane Na+ /H+ antiporter gene improves salt tolerance in Arabidopsis
thaliana. Nature Biotechnol. 21: 8185.
Shi, H. and Zhu, J.-K. 2002. Regulation of expression of the vacuolar Na+ /H+
antiporter gene AtNHX1 by salt stress and ABA. Plant Mol. Biol. 50: 543
550.
Shinozaki, K. and Russel, P. 1995. Cell-cycle control linked to extracellular
environment by MAP kinase pathway in fission yeast. Nature 378: 739743.
Shinozaki, K. and Yamaguchi-Shinozaki, K. 1997. Gene expression and signal
transduction in water-stress response. Plant Physiol. 25: 327334.
Shinozaki, K. and Yamaguchi-Shinozaki, K. 2000. Molecular response to dehydration and low temperature: Differences and cross-talk between two stress
signaling pathways. Curr. Opin. Plant Biol. 3: 217223.

Shinozaki, K., Yamaguchi-Shinozaki, K., and Seki, M. 2003. Regulatory network of gene expression in the drought and cold stress responses. Curr. Opin.
Plant Biol. 6: 410417.
Shinwari, Z. K., Nakashima, K., Miura, S., Kasuga, M., Seki, M., YamaguchiShinozaki K., and Shinozaki, K. 1998. An Arabidopsis gene family encoding
DRE/CRT binding proteins involved in low-temperature-responsive gene expression. Biochem. Biophys. Res. Commun. 250: 161170.
Shirasu, K., Nakajima, H., Rajasekhar, K., Dixon, R. A., and Lamb, C. 1997.
Salicylic acid potentiates an agonist-dependent gain control that amplifies
pathogen signal in the activation of defense mechanisms. Plant Cell 9: 261
270.
Siefritz, F. Tyree, M. T., Lovisolo, C., Schubert, A., and Kaldenhoff, R. 2002.
PIP1 plasma membrane aquaporins in tobacco: From cellular effects to function in plants. Plant Cell 14: 869876.
Singh, K. B., Foley, R. C., and Onate-Sanchez, L. 2002. Transcription factors in plant defense and stress response. Curr. Opin. Plant Biol. 5: 430
436.
Singla-Pareek, S. L., Reddy, M. K., and Sopory, S. K. 2003. Genetic engineering
of the glyoxalase pathway in tobacco leads to enhanced salinity tolerance.
Proc. Natl. Acad. Sci. USA 100: 1467214677.
Smart, L. B., Moskal, W. A., Cameron, K. D., and Bennett, A. 2001. MIP
genes are down-regulated under drought stress in Nicotiana glauca. Plant
Cell Physiol. 42: 686693.
Smirnoff, N. 1998. Plant resistance to environmental stress. Curr. Opin. Plant
Biol. 9: 214219.
Smirnoff, N., and Cumbes, Q. J. 1989. Hydroxyl radical scavenging activity of
compatible solutes. Phytochem. 28: 10571060.
Smith-Espinoza, C. J., Richter, A., Salamini, F., and Bartels, D. 2003. Dissecting the response to dehydration and salt (NaCl) in the resurrection plant
Craterostigma Plantagenium. Plant Cell Environ. 26: 13071315.
Snedden, W. A. and Fromm, H. 1998. Calmodulin, calmodulin-related proteins
and plant responses to the environment. Trends Plant Sci. 3: 299304.
Snedden, W. A. and Fromm, H. 2001. Calmodulin as a versatile calcium signal
transducer in plants. New Phytol. 151: 3566.
Soderman, E., Hjellstrom, M., Fahleson, J., and Engstrom, P. 1999. The HDZIP gene ATHB6 in Arabidopsis is expressed in developing leaves, roots
and carpels and up-regulated by water-deficit conditions. Plant Mol. Biol. 40:
10731083.
Soderman, E., Mattsson, J., and Engstrom, P. 1996. The Arabidopsis homeobox
gene ATHB-7 is induced by water deficit and by abscisic acid. Plant J. 10:
375381.
Spollen, W. G., Sharp, R. E., Saab, I. N., and Wu, Y. 1993. Regulation of cell
expansion in roots and shoots at low water potentials. In: Water Deficits: Plant
Responses from Cell to Community. Smith J. A. C., Griffiths H., Eds. BIOS
Scientific Publishers, Oxford, pp. 3752.
Staxen, I., Pical, C., Montgomery, L. T., Gray, J. E., Hetherington, A. M., and
McAinsh, M. R. 1999. Abscisic acid induces oscillations in guard-cell cytosolic free calcium that involve phosphoinositide-specific phospholipase C.
Proc. Natl. Acad. Sci. USA 96: 17791784.
Stewart, C. R., Voetberg, G., and Rayapati, P. J. 1986. The effects of benzyladenine, cycloheximide and cordycepin on wilting-induced abscisic acid and
praline accumulations and abscisic acid- and salt-induced praline accumulation in barley leaves. Plant Physiol. 82: 703707.
Stockinger, E. J., Gilmour, S. J., and Thomashow, M. F. 1997. Arabidopsis
thaliana CBF1 encodes an AP2 domain-containing transcriptional activator
that binds to the C-repeat/DRE, a cis-acting DNA regulatory element that
stimulates transcription in response to low temperature and water deficit.
Proc. Natl. Acad. Sci. USA 94: 10351040.
Streeter, J. G., Lohnes, D. G., and Fioritto, R. J. 2001. Pattern of pinitol accumulation in soybean plants and relationships to drought tolerance. Plant Cell
Environ. 24: 429438.
Sugano, S. Kaminaka, H. Rybka, Z. Catala, R. Salinas, J., Matsui, K. OhmeTkagi, M., and Takatsuji, H. 2003. Stress-responsive zinc finger gene
ZPT2-3 plays a role in drought tolerance in petunia. Plant J. 36: 830
841.

DROUGHT AND SALT TOLERANCE IN PLANTS


Sun, W., Bernard, C., van de Cotte, Van Montagu, M., and Verbruggen, N.
2001. At-HSP17. 6A, encoding a small heat-shock protein in Arabidopsis,
can enhance osmotolerance upon overexpression. Plant J. 27: 407415.
Sunkar, R., Bartels, D., and Kirch, H.-H. 2003. Overexpression of a stressinducible aldehyde dehydrogenase gene from Arabidopsis thaliana in transgenic plants improves stress tolerance. Plant J. 35: 452464.
Sze, H., Liang, F., Hwang, I., Curran, A. C., and Harper, J. F. 2000. Diversity
and regulation of plant Ca2+ pumps: Insights from expression in yeast. Annu.
Rev. Plant Physiol. Plant Mol. Biol. 51: 433462.
Taji, T., Ohsumi, C., Iuchi, S., Seki, M., Kasuga, M., Kobayashi, M., YamaguchiShinozaki, K., and Shinozaki, K. 2002. Important roles of drought- and coldinducible genes for galactinol synthase in stress tolerance in Arabidopsis
thaliana. Plant J. 29: 417426.
Takahashi, S., Katagiri, T., Hirayama, T., Yamaguchi-Shinozaki, K., and
Shinozaki, K. 2001. Hyperosmotic stress induces a rapid and transient increase
in Inositol1,4,5-triphosphate independent of abscisic acid in Arabidopsis cell
cultures. Plant Cell Physiol. 42: 214222.
Tamura, T., Hra, K., Yamaguchi, Y., Koizumi, N., and Sano, H. 2003. osmotic stress tolerance of transgenic tobacco expressing a gene encoding a
membrane-located receptor-like protein from tobacco plants. Plant Physiol.
131: 454462.
Tan, B. C., Schwartz, S. H., Zeevart, J. A., and McCarty, D. R. 1997. Genetic
control of abscisic acid biosynthesis in maize. Proc. Natl. Acad. Sci. USA 94:
1223512240.
Tena, G., Asai, T., Chiu, W.-L., and Sheen, J. 2001. Plant mitogen-activated
protein kinase signaling cascades. Curr. Opin. Plant Biol. 4: 392400.
Thiyagarajah, M., Fry, S., and Yeo, A. 1996. In vitro salt tolerance of cell wall
enzymes from halophytes and glycophytes. J. Exp. Bot. 47: 17171724.
Thompson, A. J., Jackson, A. C., Parker, R. A., Morpeth, D. R., Burbidge, A.,
and Taylor, I. B. 2000a. Abscisic acid biosynthesis in tomato: Regulation
of zeaxanthin epoxidase and 9-cis-epoxycarotenoid dioxygenase mRNAs by
light/dark cycles, water stress and abscisic acid. Plant Mol. Biol. 42: 833845.
Thompson, A. J., Jackson, A. C., Symonds, R. C., Mulholland, B. J., Dadswell,
A. R., Blake, P. S., Burbidge, A., and Taylor, I. B. 2000b. Ectopic expression
of a tomato 9-cis-epoxycarotenoid dioxygenase gene causes over-production
of abscisic acid. Plant J. 23: 363374.
Tiburcio, A. F., Besford, R. T., Capell, T., Borrell, A., Testillano, P. S., and
Risueno, M. C. 1994. Mechanisms of polyamine action during senescence
responses induced by osmotic stress. J. Exp. Bot. 45: 17891800.
Triesmann, R. 1996. Regulation of transcription by MAP kinase cascades. Curr.
Opin. Cell Biol. 8: 205215.
Tsugane, K., Kobayashi, K., Niwa, Y., Ohba, Y., Wada, K., and Kobayashi, H.
1999. A recessive Arabidopsis mutant that grows photoautotrophically under
salt stress shows enhanced active oxygen detoxification. Plant Cell 11: 1195
1206.
Tyerman, S. D., Niemietz, C. M., and Bramley, H. 2002. Plant aquaporins:
Multifunctional water and solute channels with expanding roles. Plant Cell
Environ. 25: 173194.
Ulm, R., Rekenkova, E., Sansebastiano, G. P., Bechtold, N., and Paszkowski, J.
2001. Mitogen-activated protein kinase phosphatse is required for genotoxic
stress releif in Arabidopsis. Genes Dev. 15: 699709.
Uno, Y., Furihata, T., Abe, H., Yoshida, R., Shinozaki, K., and YamaguchiShinozaki, K. 2000. Arabidopsis basic leucine zipper transcription factors
involved in an abscisic acid-dependent signal transduction pathway under
drought and high-salinity conditions. Proc. Natl. Acad. Sci. USA 97: 11632
11637.
Urano, K., Yoshiba, Y., Nanjo, T., Igarashi, Y., Seki, M., Sekiguchi, K.,
Yamaguchi-Shinozaki, K., and Shinozaki, K. 2003. Characterization of Arabidopsis genes involved in biosynthesis of polyamines in abiotic stress responses and developmental stages. Plant Cell Environ. 26: 19171926.
Urano, K., Yoshiba, Y., Nanjo, T., Ito, T., Yamaguchi-Shinozaki, K., and
Shinozaki, K. 2004. Arabidopsis stress-inducible gene for arginine decarboxylase AtADC2 is required for accumulation of putrescine in salt tolerance.
Biochem. Biophys. Res. Commun. 313: 369375.

57

Urao, T., Yakubov, B., Satoh, R., Yamaguchi-Shinozaki, K., Seki, B., Hirayama,
T., and Shinozaki, K. 1999. A transmembrane hybrid-type histidine kinase in
Arabidopsis functions as an osmosensor. Plant Cell 11: 17431754.
Urao, T., Yamaguchi-Shinozaki, K., Urao, S., and Shinozaki, K. 1993. An Arabidopsis myb homolog is induced by dehydration stress and its gene product binds to the conserved MYB recognition sequence. Plant Cell 5: 1529
1539.
Van Camp, W., Capiau, K., Van Montagu, M., Inze, D., and Slooten, L. 1996.
Enhancement of oxidative stress tolerance in transgenic tobacco plants overproducing Fe-superoxide dismutase in chloroplasts. Plant Physiol. 112: 1703
1714.
Veena, Reddy, V. S., and Sopory, S. K. 1999. Glyoxylase I from Brassica juncea:
Molecular cloning, regulation and its overexpression confer tolerance in transgenic tobacco under stress. Plant J. 17: 385395.
Venema, K., Quintero, F. J., Pardo, J. M., and Donaire, J. P. 2002. The Arabidopsis Na+ /H+ exchanger AtNHX1 catalyses low affinity Na+ and K+ transport
in reconstituted liposomes. J. Biol. Chem. 277: 24132418.
Verbruggen, N., Hua, X. J., May, M., and VanMontagu, M. 1996. Environmental and developmental signals modulate proline homeostasis: evidence for
a negative transcriptional regulator. Proc. Natl. Acad. Sci. USA, 93: 8787
8791.
Vernon, D. M., and Bohnert, H. J. 1992. A novel methyl transferase induced by
osmotic stress in the facultative halophyte, Mesembryanthemum crystallinum.
EMBO J. 11: 20772085.
Vierstra, R. D. 1996. Proteolysis in plant: mechanisms and functions. Plant Mol.
Biol. 32: 275302.
Villalobos, M. A., Bartels, D., and Iturriga, G. 2004. Stress tolerance and glucose
insensitive phenotypes in Arabidopsis overexpressing the CpMYB10 transcription factor gene. Plant Physiol. 135: 309324.
Vitart, V., Baxter, I., Doerner, P., and Harper, J. F. 2001. Evidence for a role
in growth and salt resistance of a plasma membrane H+-ATPase in the root
endodermis. Plant J. 27: 191201.
Vogel, G., Aeschbacher, R. A., Muller, J., Boller, T., and Wiemken, A. 1998.
Trehalose-6-phosphate phosphatases from Arabidopsis thaliana: identification by functional complementation of the yeast tps2 mutant. Plant J. 13:
673683.
Volkov, V., Wang, B., Dominy, P. J., Fricke, W., and Amtmann, A. 2003. Thellungiella halophila, a salt-tolerant relative of Arabidopsis thaliana, possesses
effective mechanisms to discriminate between potassium and sodium. Plant,
Cell Environ. 27: 114.
Wang, H., Datla, R., Georges, F., Loewen, M., and Cuter, A. J. 1995. Promoters from kin1 and cor6. 6, two homologues of Arabidopsis thaliana genes:
Transcriptional regulation and gene expression induced by low temperature,
ABA, osmoticum and dehydration. Plant Mol. Biol. 28: 605617.
Wang, H., Qi, Q., Schorr, P., Cutler, A. J., Crosby, W. L., and Fowke, L. C.
1998. ICK1, a cyclin-dependent protein kinase inhibitor from Arabidopsis
thaliana interacts with both Cdc2a and CycD3, and its expression is induced
by abscisic acid. Plant J. 15: 501510.
Wang, W., Vinocur, B., Shoseyov, O., and Altman, A. 2004. Role of plant heatshock proteins and molecular chaperones in the abiotic stress response. Trends
Plant Sci. 9: 244252.
Wang, X. 2001. Plant phospholipases. Annu. Rev. Plant Physiol. Plant Mol. Biol.
52: 211231.
Wang, X. 2002. Phospholipase D in hormonal and stress signalling. Curr. Opin.
Plant Biol. 5: 408414.
Wehmeyer, N. and Vierling, E. 2000. The expression of small heat shock proteins
in seeds responds to discrete developmental signals and suggests a general
protective role in desiccation tolerance. Plant Physiol. 122: 10991108.
Weig, A., Deswarte, C., and Chrispeels, M. J. 1997. The major intrinsic protein
family of Arabidopsis has 23 members that form three distinct groups with
functional aquaporins in each group. Plant Physiol. 25: 13471357.
Wendehenne, D., Pugin, A., Klessig, D. F., and Durner, J. 2001. Nitric oxide:
Comparative synthesis and signalling in animal and plant cells. Trends Plant
Sci. 6: 177183.

58

D. BARTELS AND R. SUNKAR

West, G., Inze, D., and Beemster, G. T. S. 2004. Cell cycle modulation in the
response of the primary root of Arabidopsis to salt stress. Plant Physiol. 135:
10501058.
Westgate, M. E. and Boyer, J. S. 1985. Osmotic adjustment and the inhibition
of leaf, root, stem and silk growth at low water potentials in maize. Planta
164: 540549.
Williams, M. E., Foster, R., and Chua, N.-H. 1992. Sequences flanking the
hexameric G-box core CACGTG affect the specificity of protein binding.
Plant Cell 4: 485496.
Wimmers, L. E., Ewing, N. N., and Bennett, A. B. 1992. Higher plant Ca2+ ATPase: Primary structure and regulation of mRNA abundance by salt. Proc.
Natl. Acad. Sci. USA 89: 92059209
Winicov, I. 1993. cDNA encoding zinc finger motifs from salt-tolerant alfalfa
(Medicago sativa L.) cells. Plant Physiol. 102: 681682.
Winicov, I. 2000. Alfin1 transcription factor overexpression enhances plant root
growth under normal and saline conditions and improves salt tolerance in
alfalfa. Planta 210: 416422.
Winicov, I. and Bastola, D. R. 1999. Transgenic overexpression of the transcription factor Alfin1 enhances expression of the endogenous MSPRP2 gene
in alfalfa and improves salinity tolerance of the plants. Plant Physiol. 120:
473480.
Wu, Y., Thorne, E. T., Sharp, R. E., and Cosgrove, D. J. 2001. Modification of
expansin transcript levels in the maize primary root at low water potentials.
Plant Physiol. 126: 14711479.
Wurgler-Murphy, S. M., Maeda, T., Witten, E. A., and Saito, H. 1997. Regulation
of the Saccharomyces cerevisiae HOG1 mitogen-activated protein kinase by
the PTP2 and PTP3 protein tyrosine phosphatases. Mol. Cell Biol. 17: 1289
1297.
Wurgler-Murphy, S. M., and Saito, S. 1997. Two-component signal transducers
and MAPK cascades. Trends Biochem. Sci. 25: 172176.
Xiong, L., Gong, Z., Rock, C. D., Subramanian, S., Guo, Y., Xu, W., Galbraith,
D., and Zhu, J. K. 2001. Modulation of abscisic acid signal transduction and
biosynthesis by an Sm-like protein in Arabidopsis. Dev. Cell 1: 771778.
Xiong, L., Ishitani, M., Lee, H., and Zhu, J.-K. 2001. The Arabidopsis
LOS5/ABA3 locus encodes a molybdenum cofactor sulfurylase and modulates cold stress and osmotic stress-responsive gene expression. Plant Cell
13: 20632083.
Xiong, L., Lee, B. H., Ishitani, M., Lee, H., Zhang, C., and Zhu, J. K. 2001c.
FIERY1 encoding an inositol polyphosphate 1-phosphatase is a negative regulator of abscisic acid and stress signalling in Arabidopsis. Genes Dev. 15:
19711984.
Xiong, L. and Zhu, J. K. 2003. Regulation of abscisic acid biosynthesis. Plant
Physiol. 133: 2936.
Xu, Q., Fu, H.-H., Gupta, R., and Luan, S. 1998. Molecular characterization of
a tyrosine-specific protein phosphatase encoded by a stress-responsive gene
in Arabidopsis. Plant Cell 10: 849857.
Yamada, S., Komori, T., Myers, P. N., Kuwata, S., Kubo, T., and Imaseki, H.
1997. Expression of plama membrane water channel genes under water stress
in Nicotiana excelsior. Plant Cell Physiol. 25: 12261231.
Yamaguchi-Shinozaki, K., Koizumi, M., Urao, S., and Shinozaki, K. 1992.
Molecular cloning and characterization of 9 cDNAs for genes that are responsive to desiccation in Arabidopsis thaliana: Sequence analysis of one
cDNA that encodes a putative transmembrane channel protein. Plant Cell
Physiol. 25: 217224.
Yamaguchi-Shinozaki, K. and Shinozaki, K. 1993. The plant hormone abscisic
acid mediates the drought-induced expression but not the seed-specific expres-

sion of rd22, a gene responsive to dehydration stress in Arabidopsis thaliana.


Mol. Gen. Genet. 238: 1725.
Yamaguchi-Shinozaki, K. and Shinozaki, K. 1994. A novel cis-acting element in an Arabidopsis gene is involved in responsiveness to drought, lowtemperature, or high-salt stress. Plant Cell 6: 251264.
Yancey, P. H., Clark, M. E., Hand, S. C., Bowlus, R. D., and Somero, G. N.
1982. Living with water stress: Evolution of osmolyte systems. Science 217:
12141222.
Yang, S. X., Zhao, Y. X., Zhang, Q., He, Y. K., Zhang, H., and Luo. 2001. HAL1
mediate salt adaptation in Arabidopsis thaliana. Cell Res. 11: 142148.
Yang, T. and Poovaiah, B. W. 2003. Calcium/Calmodulin-mediated signal network in plants. Trends Plant Sci. 8: 505512.
Ye, B., Muller, H. H., Zhang, J., and Gressel, J. 1998. Constitutively elevated
levels of putrescine and putrescine generating enzymes correlated with oxidant stress resistance in Conyza bonariensis and wheat. Plant Physiol. 115:
14431451.
Yokoi, S., Quintero, F. J., Cubero, B., Ruiz, M. T., Bressan, R. A., Hasegawa, P.
M., and Pardo, J. M. 2002. Differential expression and function of Arabidopsis
thaliana NHX Na+ /H+ antiporters in the salt stress response. Plant J. 30:
529539.
Yoon, H. W., Kim, M. C., Shin, P. G., Kim, J-S., Lee, S. Y., Hwang, I., Bahk, J.
D., Hong, J. C., Han, C., and Cho, M. J. 1997. Differential expression of two
functional serine/threonine protein kinases from soybean that have an unusual
acidic domains at the carboxy terminus. Mol. Gen. Genet. 255: 359371.
Yoshiba, Y., Kiyosue, T., Nakashima, K., Yamaguchi-Shinozaki, K., and
Shinozaki, K. 1997. Regulation of levels of proline as an osmolyte in plants
under water stress. Plant Cell Physiol. 38: 10951102.
Yuasa, T., Ichimura, K., Mizoguchi, T., and Shinozaki, K. 2001. Oxidative stress
activates ATMPK6, an Arabidopsis homologue of MAP kinase. Plant Cell
Physiol. 42: 10121016.
Yuasa, T. and Muto, S. 1996. Activation of 40-kDa protein kinases in response
to hypo- and hyperosmotic shocks in a halotolerant green alga Dunaliella
tertiolecta. Plant Cell Physiol. 37: 3542.
Zeevaart, J. A. D. and Creelman, R. A. 1988. Metabolism and physiology of
abscisic acid. Annu. Rev. Plant Physiol. Plant Mol. Biol. 39: 439473.
Zhang, H.-X. and Blumwald, E. 2001. Transgenic salt-tolerant tomato plants
accumulate salt in foliage but not in fruit. Nature Biotechnol. 19: 765768.
Zhang, H.-X., Hodson, J. N., Williams, J. P., and Blumwald, E. 2001. Engineering salt-tolerant Brassica plants: Characterization of yield and seed oil
quality in transgenic plants with increased vacuolar sodium accumulation.
Proc. Natl. Acad. Sci. USA 98: 1283212836
Zhu, B. C., Su, J., Chan, M. C., Verma, D. P. S., Fan, Y. L., and Wu, R. 1998.
Over-expression of a -pyrroline-5-carboxylate synthetase gene and analysis
of tolerance to water-stress and salt-stress in transgenic rice. Plant Sci. 139:
4148.
Zhu, J., Gong, Z., Zhang, C., Song, C.-P., Damsz, B., Inan, G., Koiwa, H., Zhu,
J.-K., Hasegawa, P. M., and Bressan, R. A. 2002. OSM1/SYP61: A syntaxin
protein in Arabidopsis controls abscisic acid-mediated and non-abscisic acidmediated responses to abiotic stress. Plant Cell 14: 30093028.
Zhu, J.-K. 2000. Genetic analysis of plant salt tolerance using Arabidopsis. Plant
Physiol. 124: 941948.
Zhu, J.-K. 2001. Plant salt tolerance. Trends Plant Sci. 6: 6671.
Zhu, J.-K. 2002. Salt and drought stress signal transduction in plants. Annu. Rev.
Plant Biol. 53: 247273.
Zhu, J. K. 2003. Regulation of ion homeostasis under salt stress. Curr. Opin.
Plant Biol. 6: 441445.

Potrebbero piacerti anche