• Nenhum resultado encontrado

Tese Diego Almeida.pdf

N/A
N/A
Protected

Academic year: 2019

Share "Tese Diego Almeida.pdf"

Copied!
221
0
0

Texto

(1)
(2)
(3)

Regulation of the

OsNHX1

Gene Expression:

Identification and Characterization of Novel

Transcription Factors

Diego Melo Almeida

Dissertation presented to obtain the Ph.D degree in Biochemistry

Instituto de Tecnologia Química e Biológica António Xavier | Universidade Nova de

Lisboa

(4)
(5)

Work performed at:

Supervisors:

Dr. Nelson José Madeira Saibo

Head of the Plant Gene Regulation laboratory (ITQB NOVA) Principal Investigator

Prof. Dr. M. Margarida Oliveira

Head of GPlantS Unit (ITQB NOVA)

Associate Professor with Habilitation (Agregação) at ITQB NOVA

Dr. Glenn B. Gregorio

Crop Breeding Manager at Sweet and Waxy Corn. San Rafael, Philippines. Former senior scientist and plant breeder at International Rice Research Institute (IRRI). Los Baños, Philippines.

Genomics of Plant Stress Unit

Instituto de Tecnologia Química e

Biológica António Xavier

Universidade Nova de Lisboa

Av. Da República

2780-157 Oeiras

Portugal

International Rice Research

Institute

Plant Breeding, Genetics, and

Biotechnology Division

Los Baños

(6)
(7)
(8)
(9)

Equipped with his five senses, man explores the universe around

him and calls the adventure Science.

Edwin Powell Hubble

(10)
(11)

IX

ACKNOWLEDGMENTS

To Nelson Saibo for being my supervisor, for taking care of me during these eight

years, for always pushing me to do and be the best, making me grow. For

teaching me when I was a newly-graduated student. Most importantly, for

open my eyes to science and show what science means.

To Prof. Margarida Oliveira for having accepted me at her lab (already crowded at

that time!), for all precious minutes you devoted me, for always teaching me

more than science. Above all, for show me what it means to love what we

do.

To Glenn Gregorio for having accepted me at IRRI, delighting my world, making

understand that our research produces more than scientific papers. We are

able to shape people´s life.

To Ana Paula Santos for being so positive, for helping me with all micro-world and

for all advices.

To Titi for all great lab conversations, for making our lab a better fun place to

work, for helping me solving the challenges of my work.

To Duarte Figueiredo for all wonderful moments we had inside and outside the

lab, for helping me with everything during my first times in lab, for the great

sense of humor.

To Muffy for all our great weekends/afterhours crazy projects, for all great talks at

23 PM, and for helping me with million things during these years.

To Sónia Negrão for helping me enter in the salt stress world, for all the precious

tips and talks.

To Pedro Barros for all support, wonderful talks, advices, for always ask about my

work, and for the huge help into the bioinformatics world.

To Lena for all support with the physiological work, for always taking care of us,

for being so honest, direct, and for having a big heart.

To Miguel Costa for all enormous support with the physiological work, new ideas,

(12)

X

To Joana Machado for helping me so much during my last months in the lab, and

for all longs days playing with the IRGA.

To Rita Borba, Paulo Gouveia, Sissi, Nuno, Rita Leal, Bruno, Vanessa, Cátia,

Grosa and Inês for making the GPlantS Lab a great place to learn, work and

growth.

To all Forest Biotech group for all patience needed to be our second home.

To Isabel Abreu for supporting me a lot into my trips into the protein-world, for

teaching me how to think as a scientist, for always motivate me to do the best and more, much more…

To Dedas for being a great friend, for all huge support at work, for being such a

great person.

To Gruska for being a great friend, for helping me a lot with my lab work, for being

so patience with my mess, for all the great talks, for teaching me polish. Dziękuję kobietę.

To Tânia Serra for all the tremendous support. I have no words for thank you for

teaching so much. Thank you for listen me, helping me solving challenges,

inspire me.

Aos meus pais e irmã, por estarem sempre comigo. Por me amarem. Obrigado

pela força, equilíbrio e conselhos que dão a cada dia. Por estarmos sempre

fortemente ligados uns aos outros.

A Tânia Lucas, por todo o carinho, amor e paciência durante esta fase. Por

ensinar-me sempre a ser otimista, por acreditar em nós, por levar-nos a

brincar como crianças, por sempre sorrir. Principalmente, por tornar as

(13)

XI

I, Diego Melo Almeida, declare by my honour to have had active participation in

the following papers and book chapter:

- Diego M. Almeida*, M. Cecília Almadanim*, Tiago Lourenço, Isabel A. Abreu,

Nelson J. M. Saibo, M. Margarida Oliveira. “Screening for abiotic stress

tolerance in rice: salt, cold and drought.” Paula Duque (ed). Environmental

Responses in Plants: Methods and protocols, Methods in Molecular Biology,

1398:155-82. DOI: 10.1007/978-1-4939-3356-3_14, Springer Science +

Business Media New York 2016. *First authors

Diego M. Almeida performed part of the review work and manuscript writing.

- Tânia S. Serra, Duarte D. Figueiredo, André M. Cordeiro, Diego M. Almeida,

Tiago Lourenço, Isabel A. Abreu, Alvaro Sebastián, Lisete Fernades, Bruno

Contreras-Moreira, M. Margarida Oliveira, Nelson J.M. Saibo (2013). “OsRMC,

a negative regulator of salt stress response in rice, is regulated by two AP2/ERF transcription factors” Plant Molecular Biology journal. 2013 Jul;82(4-5):439-55. DOI: 10.1007/s11103-013-0073-9.

Diego M. Almeida participated in the experimental work and manuscript writing.

- Diego M. Almeida, Glenn B. Gregorio, M. Margarida Oliveira, Nelson J. M.

Saibo. “Five novel transcriptions factors regulating OsNHX1 expression in a salt tolerant rice genotype” (under re-submission to Plant Molecular Biology journal)

Diego M. Almeida participated in the experimental design, the laboratory work

(14)

XII

- Diego M. Almeida, M. Margarida Oliveira, Nelson J. M. Saibo. “Na+ and K+ homeostasis in plants under salt stress” (Review submitted to Genetics and Molecular Biology journal)

Diego M. Almeida performed the review work and manuscript writing. This

manuscript includes part of the work described in Chapter I.

- Diego M. Almeida, Joana Machado, Pedro M. Barros, Helena Sapeta, M.

Margarida Oliveira, Nelson J. M. Saibo. “OsPCF2, a class I TCP transcription factor involved in stomatal aperture in rice” (In preparation)

Diego M. Almeida performed the review work and manuscript writing. This

(15)

XIII

LIST OF ABBREVIATIONS

3-AT 3-amino-1,2,4-triazole

a.a. Amino acid

ABA Abscisic acid

ABRE ABA-responsiveelement

At Arabidopsis Thaliana

ATP Adenosine triphosphate

bp Base pair

BLAST Basis Local Alignment Search Tool

Bv Beta vulgaris

CaM Calmodulin

CAMV35S Cauliflower Mosaic Virus 35S promote

cDNA Complementary DNA

CM-His Complete Minimal medium lacking Histidine

CM+His Complete Minimal medium supplemented

with Histidine

CPP Cystein-rich Poly comb-like Protein

CPY CarboxypeptidaseY

Cys Cysteine

ºC Degree Celsius

DNA Deoxyribonucleic Acid

dNTPs Deoxynucleotide Triphosphates

dS DeciSiemens

DTT Dithiothreitol

EC Electrical conductivity

EDTA Ethylene Diamine Tetraacetic Acid

EMSA Electrophoretic Mobility Shift Assays

(16)

XIV

g Gram

g Relative centrifugal force

GFP Green fluorescent protein

gs Stomatal Conductance

GUS β-glucuronidase

h Hour

H2O2 Hydrogen peroxide

Hv Hordeum vulgare L.

In Ipomea nil

IRRI International Rice Research Institute

KDa KiloDalton

Kg Kilogram

Km Michaelis–Menten constant

L Litre

LUC Luciferase

M Molarity

m Mass

Mb Mega base pair

µCi MicroCurie

µg Microgram

µL Microliter

µM Micromolar

min Minute

mL Mililitre

mM Milimolar

mRNA Message Ribonucleic Acid

MYC Myelocytomatosis Oncogene

MYB Myeloblastosis Oncogene

(17)

XV

NIN Nodule Inception

ng Nanogram

NLS Nuclear Localization Signal

NSCC Non Selective Cation Channels

Os Oryza sativa

PCNA Proliferating Cell Nuclear Antigen

PCR Polymerase Chain Reaction

PEG Polyethylene Glycol

PPM Parts per Million

PM Plasma Membrane

PVC Prevacuolar compartment

QTL Quantitative Trait Locus

RIL Recombinant Inbred Line

RNA Ribonucleic Acid

ROS Reactive Oxygen Species

rpm Rotations per minute

RT Room temperature

RT-PCR Reverse Transcription – PCR

RT-qPCR Quantitative Real-Time RT-PCR

RWC Relative Water Content

s Second

Sc Saccharomyces cerevisiae

SD Standard Deviation

SDS-PAGE Sodium Dodecyl Sulfate –

Polyacrylamide Gel Electrophoresis

SE Standard Erro

SES Standard Evaluation Score

SOS Salt Overly Sensitive

(18)

XVI

T-DNA Transfer-DNA

TF Transcription Factor

TGN Trans-Golgi Network

TRX Thioredoxin

Vv Vitis vinifera

(19)

XVII

SUMMARY

For half of the world´s population, rice is life. This cereal crop is

considered an important staple food worldwide, and more than three billion people

count on it for 50-80% of their daily calorie intake. Soil salinity is a major

environmental constraint to crop production, resulting in considerable yield losses

around the globe every year. According to the Food and Agriculture Organization

(FAO), in 2008 over 6% of world's total landand over 20% of irrigated land were

affected by high levels of salt. Irrigated land is only 15%of cultivated land, but it produces one third of the world’s food, raising awareness about salinity as a serious problem for crop productivity. Rice like as most crops is very sensitive to

salt, showing salt stress symptoms and reduced yield at relatively low soil salinity

levels (≈ 40 mM NaCl). Among the agronomically important cereals, rice shows

the highest sensitivity to salt. However, some degree of genotype tolerance for

salt stress is available in rice germplasm. To cope with salt stress conditions,

plants evolved several and diverse response mechanisms. One of these

mechanisms is tissue tolerance, in which high salt concentration is found in leaves

but is compartmentalized, especially in the vacuole, reducing the deleterious

effect of Na+ in the cytosol and driving water uptake to cells. Cation/H+ antiporters

mediate the transport of Na+ into the vacuole. This Na+/H+ exchange is mediated

by members of a family of transporters referred to as K+,Na+/H+ antiporters

(NHX-type). Among them, NHX1 is the most abundant and the best characterized

member. Several studies have shown that NHX1 overexpression leads to

improved salt and drought stress tolerance in various plant species. Given that

transcription factors (TFs) can act as master regulators of different cellular

processes, they are promising candidates for modifying complex traits in crop

plants, such as salt stress tolerance. Nevertheless, NHX1 transcriptional

regulation under salt stress is poorly understood.

The main objective of our study was the identification and functional

characterization of TFs regulating OsNHX1 expression under salt stress in a salt

(20)

XVIII

tolerant rice genotype in which the regulation of the OsNHX1 gene expression in

response to salt stress could be relevant for the salt stress tolerance at seedling

stage. Among the seven rice genotypes analyzed, we have selected Hasawi,

which showed a strong salt stress tolerance and high OsNHX1 responsiveness to

salt stress. Using the Yeast-One-Hybrid (Y1H) system to screen a salt-induced

rice cDNA expression library from Hasawi, five TFs belonging to three distinct

families were identified as binding to OsNHX1 promoter: one TCP (OsPCF2), one

CPP (OsCPP5) and three NIN-likes (OsNIN-like 2, OsNIN-like 3 and OsNIN-like

4). Transactivation activity assays performed in Arabidopsis and rice protoplasts

showed that OsPCF2 and OsNIN-like 4 are activators of the OsNHX1 gene

expression, while OsCPP5 and OsNIN-like 2 act as repressors. The

transactivation activity of OsNIN-like 3 needs to be further investigated.

When we analyzed the transcript levels of these TFs in rice seedlings

subjected to abiotic stress conditions, it was observed that all of them are early

regulated by both salt stress and PEG-simulated drought, especially in roots. The

expression of OsPCF2 in roots under salt stress and the OsNIN-like 4 in roots

subjected to PEG were mainly up-regulated in Hasawi, indicating that these TFs

may be associated with the salt and drought stress tolerance observed in Hasawi.

Analyses of the rice NHX-type gene promoters showed that OsPCF1 and

OsPCF2 (both TFs are TCP class I) binding motifs were over-represented in the

promoter of all OsNHX genes. Using an Electrophoretic Mobility Shift Assay

(EMSA), we showed that both OsPCF1 and OsPCF2 proteins bind to all OsNHX

gene promoters. In addition, a genome-wide search identified TCP class I binding

motifs in the promoter region of 3.089 rice genes. Among these genes, ten

(OsAKT2, OsKAT2, OsKAT3, OsKC1.2, OsALMT1, OsVHA-a1, OsVHA-a2,

OsVHA-a3, OsVHA-F, and OsPIP1;1) are somehow related to stomatal aperture.

We showed that OsPCF2 binds to the respective cis-regulatory elements present

in the promoters of all these genes. In addition, a rice T-DNA insertion line for

OsPCF2 (gene knockout) revealed a general down-regulation for most of the

(21)

XIX

conductance to water vapour under light conditions as well as reduced salt stress

tolerance. Moreover, we observed that OsPCF2 seems to be posttranslationally

regulated by H2O2, thus modulating its binding to the OsNHX1 promoter.

This work allowed the identification of five novel TFs binding to the

promoter of OsNHX1, which is known to have a role controlling plant cell turgor

and expansion, thus mediating abiotic stress effects on plant development.

Further characterization of these TFs will help us to better understand their

function and it will unveil novel targets for improvement of plant abiotic stress

(22)
(23)

XXI

SUMÁRIO

Para metade da população mundial, o arroz é vida. Este cereal é

considerado um alimento essencial em todo o mundo. Mais de três mil milhões de

pessoas dependem diretamente do arroz para satisfazer cerca de 50-80% da sua

ingestão diária de calorias. A salinização dos solos é um dos maiores obstáculos

ambientais que limita a produção agrícola, resultando em perdas significativas na

produtividade a nível mundial. De acordo com a Organização das Nações Unidas

para Alimentação e Agricultura (FAO), em 2008 mais de 6% da área cultivada

mundial total e mais de 20% das superfícies irrigadas estavam afetadas pela

salinidade. As superfícies irrigadas representam apenas 15% das terras

cultivadas, mas produzem um terço dos alimentos a nível mundial. Estes

números despertam uma maior sensibilização para o grave problema da

salinidade dos solos na produtividade de diversas culturas. O arroz, assim como

a maioria das culturas cerealíferas, é sensível à salinidade e apresenta sintomas

de stress salino e redução na produção em solos com relativamente baixa salinidade (≈ 40 mM NaCl). Entre os cereais de maior interesse agronómico, o arroz é o mais sensível à salinidade. No entanto, alguns genótipos de arroz

apresentam um maior grau de tolerância à salinidade. Para lidar com a elevada

salinidade, as plantas desenvolveram muitos e variados mecanismos, sendo a

tolerância ao nível dos tecidos um deles. Neste caso, as folhas apresentam uma

elevada concentração de sal, mas este está compartimentado no vacúolo,

reduzindo os efeitos nefastos do Na+ no citosol e promovendo a absorção de

água para as células. Os anti-portadores catião/H+ medeiam o transporte de Na+

para o vacúolo. O transporte Na+/H+ entre o citosole o vacúolo é mediado por

membros da família de transportadores referidos como anti-portadores K+,Na+/H+

(tipo NHX). Entre estes transportadores, o NHX1 é o membro mais abundante e

melhor caracterizado. Além disso, vários estudos demostraram que a

sobre-expressão do NHX1 conduz a um aumento da tolerância aos stresses salino e

hídrico, em várias espécies de plantas. O fato dos fatores de transcrição (FT)

(24)

XXII

bons candidatos para regular características complexas em plantas, como por

exemplo a tolerância ao stress salino. No entanto, a regulação transcricional do

NHX1 em resposta ao stress salino está longe de ser bem conhecida.

O objetivo principal deste estudo foi a identificação e caracterização

funcional de FT que se ligam ao promotor do OsNHX1 proveniente de um

genótipo de arroz tolerante à elevada salinidade. Inicialmente, este projecto visou

a identificação de um genótipo de arroz tolerante ao stress salino em que a

resposta do OsNHX1 ao stress salino pudesse ser relevante para a tolerância da

planta. Entre os sete genótipos de arroz analisados selecionámos o genótipo

Hasawi, pois apresentou uma acentuada tolerância ao stress salino assim como

uma elevada indução do gene OsNHX1 pelo mesmo stress. Utilizando o sistema

Yeast-One-Hybrid (Y1H) para fazer a triagem de uma biblioteca de expressão de

cDNA de arroz, genótipo Hasawi, induzido pelo stress salino, identificámos cinco

FT, pertencentes a três famílias distintas, que se ligam ao promotor do OsNHX1

de Hasawi: um TCP (OsPCF2), um CPP (OsCPP5) e três NIN-Like (OsNIN-like 2,

OsNIN-like 3 e OsNIN-like 4). Observámos que alguns destes FT funcionam

como repressores (OsCPP5, OsNIN-like 2) e outros como ativadores da

transcrição (OsPCF2, OsNIN-like 4). A atividade transcricional do OsNIN-like 3

necessita ser mais investigada.

Quando analisámos o nível de transcrição dos FTs em plântulas de arroz

submetidas a condições de stress abiótico, observou-se que a expressão génica

de todos os FT é rapidamente modulada pelo stress salino e hídrico (induzido por

tratamento com PEG), especialmente nas raízes. A expressão de OsPCF2 pelo

stress salino, nas raízes, e do OsNIN-like 4 pelo PEG, nas raízes, foram

principalmente sobre induzidos em Hasawi (tolerante ao stress salino e hídrico),

indicando que estes FTs podem estar associada na tolerância ao sal e seca

observados em Hasawi.

A análise da região promotora dos genes NHX de arroz revelou que os

motivos de ligação para o OsPCF1 e OsPCF2 (FTs TCP classe I), encontram-se

(25)

XXIII

Electrophoretic Mobility Shift Assay (EMSA) mostrámos que as proteínas

OsPCF1 e OsPCF2 interagem com todos os motivos identificados. Além disso,

uma análise in silico a todo o genoma do arroz permitiu identificar locais de

ligação para FT da família TCP classe I na região promotora de 3.089 genes.

Entre estes genes, dez (OsAKT2, OsKAT2, OsKAT3, OsKC1.2, OsALMT1,

OsVHA-a1, a2-OsVHA, OsVHA-A3, OsVHA-F e OsPIP1; 1) estão, de alguma

maneira, relacionados com a abertura dos estomas. A ligação da proteína

OsPCF2 aos promotores destes genes foi demostrada por EMSA. A análise de

uma linha mutante de arroz com inserção de T-DNA (knockout para o gene

OsPCF2) revelou uma redução generalizada da expressão génica dos alvos do

OsPCF2, redução do teor de K+ nas partes aéreas e raízes, redução na

condutância estomática em condições de luz e redução da tolerância ao stresse

salino. Para além disso, verificámos que a ligação do OsPCF2 ao promotor do

OsNHX1 parece ser regulada por modificações pós-traducionais induzidas pelo

H2O2.

Este trabalho permitiu a identificação de cinco FT que se ligam ao

promotor do OsNHX1, o qual tem sido descrito como tendo uma função na

regulação da turgescência e expansão celular das plantas, mediando assim o

desenvolvimento das plantas em resposta aos stresses abióticos. Estudos

adicionais de caracterização funcional destes FT, irão revelar possivelmente

novos alvos para o aumento da tolerância das plantas aos stresses abióticos e

(26)
(27)

XXV

TABLE OF CONTENTS

Acknowledgments ... IX

List of Abbreviations ... XIII

Summary ... XVII

Sumário ... XXI

Chapter I ... 1

General introduction and Research Objective

Chapter II ... 57

Physiological and Molecular Responses of Seven Rice Genotypes to Salt Stress

Chapter III ... 85

Identification and Characterization of Five Novel Transcriptions Factors

Regulating OsNHX1 Expression in a Salt Tolerant Rice Genotype

Chapter IV ... 131

OsPCF2, a class I TCP Transcription Factor Involved in Stomatal Aperture in Rice

Chapter V ... 183

(28)
(29)

1

Chapter I

(30)

2

TABLE OF CONTENTS – CHAPTER I

Importance of rice ... 3

Salt stress effects on plant growth and yield ... 4

Sodium uptake from soil, sensing and signaling mechanism ... 7

Mechanisms of salt tolerance ... 9

Sodium transporters and plants salt stress tolerance ... 9

H+-Pumps and the plant salt stress response ... 11

Plasma membrane H+-ATPase ... 12

Vacuolar H+-ATPase... 13

Plasma membrane and vacuolar H+-PPase ... 16

SOS1 and the plant salt stress response ... 17

HKTs and the plant salt stress response... 21

HKT1 family ... 22

HKT2 family ... 25

NHX1 and the plant salt stress response ... 27

Diversity of plant NHX antiporters ... 28

Plant tissue localization ... 28

NHX gene expression under stress conditions ... 29

NHX structural organization and regulatory properties ... 30

Topology ... 30

Post-translational modifications ... 31

Function of NHX antiporters ... 32

Salt tolerance ... 32

K+ homeostasis ... 33

pH homeostasis ... 34

Vesicular trafficking ... 35

Transcriptional regulation ... 37

Research Objectives ... 38

(31)

3

IMPORTANCE OF RICE

For half of the world´s population, rice is life. Globally, this cereal crop is

considered a staple food for more than three billion people, accounting for 50-80%

of their daily calorie intake. Asia produced and consumed more than 90% of the world’s rice, and in many Asian countries rice represents a significant portion of its economy, labor force and even cultural heritage (Papademetriou, 2000). Rice is

cultivated in more than a hundred countries, as far north as (53º N) on the border

between Russia and China, and as far south as central Argentina (40º S) (IRRI,

2013; Papademetriou, 2000); clearly showing rice adaptability to diverse growing

conditions. There are two cultivated rice species: Oryza sativa, grown worldwide,

and Oryza glaberrima, grown in West and Central Africa. Based on the

morphology, physiology, agronomy, genetics and biochemistry features, Oryza

sativa can be divided in two main subspecies adapted to various environmental

conditions, indica and japonica. Indica includes the genotypes grown in tropical

and subtropical regions, whereas japonica genotypes grown in the cooler zones of

the subtropics and in the temperate environments (IRRI, 2013; Papademetriou,

2000).

In Europe, rice has an important sociocultural and ecological role in

several Mediterranean countries. Milled rice consumption in Europe ranges from

3.5-5.5 kg/person/year in northern countries to 6–14.8 kg/person/year in southern

countries. About 80% of the European rice production takes place in Italy and

Spain, and 12% in Portugal and Greece. The period of introduction of rice in

Europe is not concerted among academics, varying from 8th century up to late

10th century. Certainly, the Moors introduced rice to the south of the Iberian

Peninsula. Portugal is by far the largest rice consumer in Europe, consuming 14.8

Kg/person/year. Portuguese rice production is concentrated in three regions:

Mondego, Tejo and Sado rivers beds (IRRI, 2013).

The current world population of 7.3 billion is expected to reach 8.8 billion

in 2035. Most people will be living in Asia and Africa (UN, 2015). Combined, these

(32)

4

than 96% of the world’s rice consumption (IRRI, 2013; USDA, 2012). For the majority of the developing countries, rice availability means food security which is

closely connected to political stability (Bradsher, 2008; FAO, 2011). The task of

producing additional rice to meet the expected demands of people poses a major

challenge; for every one billion people added to the world’s population, it is

estimated that more 100 million tons of rice (paddy) have to be produced annually

(IRRI, 2013). This means an effort for overall increase in rice production of 26% in

the next 20 years, which must be achieved in a more efficient and

environmental-friendly system, using fewer resources (land, water, labor, etc.). To meet this goal,

high yield genotypes better adapted to adverse environmental conditions are

needed, while limiting yield losses. This is not possible without a comprehensive

understanding of the mechanisms controlling plant growth, development and

environmental stress adaptation (Bradsher, 2008; FAO, 2011; IRRI, 2013;

Papademetriou, 2000).

Many plant biology studies use Arabidopsis thaliana as model system.

However, Arabidopsis is not the best model for monocots, and our main staple

food crops, such as wheat, rice, and maize, are all monocots. In addition,

dicotyledons (dicots) and monocots are significantly distinct in many aspects of

their development (Izawa and Shimamoto, 1996). In spite of being a crop species,

rice has also emerged as a model organism for plant molecular biology studies,

and the main reasons for this are: it is relatively small, compared to other

monocots, and it has a fully sequenced genome (390 Mb); tools for functional

genomic analysis, T-DNA insertional mutant libraries are available, and the

production of transgenic plants is relatively easy, as compared to other cereals,

due to highly efficient transformation protocols (Nishimura et al., 2006; Shimamoto

and Kyozuka, 2002).

SALT STRESS EFFECTS ON PLANT GROWTH AND YIELD

Soil salinity is a major environmental constrain to crop production,

(33)

5

(Munns, 2005; Munns and Tester, 2008; Shabala and Cuin, 2008). Salt stress affects over 6% of the world’s total land area, most of this salt affected land has arisen from natural causes, including rainfall, windblown salt from ocean,

tsunamis, and rock weathering. Apart from natural causes, soil salinization is

commonly associated to irrigation practices, such as the use of water with high

salt concentration, or land cleaning by removal of deep rooted vegetation or

replaced with shallow-rooted plants that use less water, leaving more water to

pass through soil to groundwater, raising the water table and bringing salt to the

surface where it can be left behind as the water evaporates (Abrol et al., 1988).

These Man-made actions led to a significantly increase in salt affected agriculture

cultivated land. Currently it is estimated that 20% of the total irrigated land is

salt-affected. Given that irrigated land produces at least twice more than rain-fed land and is responsible for one third of the world’s food production, it raises awareness for salinity as a serious problem for crop productivity (Munns, 2005; Munns and

Tester, 2008).

High soil salinity is a condition characterized by a high concentration of

soluble salts, in which NaCl is the most soluble and widespread salt. Soils are

classified as saline when the electrical conductivity (EC) is 4 dS/m (≈ 40 mM

NaCl) or more, which significantly reduces growth and yield of most crops. Rice as

well as most crop plants are glycophytes and show salt stress symptoms and

reduced yield even when the EC is lower than 4.0 dS/m. Among cereal crops, rice

is the most salt sensitive one (Munns and Tester, 2008). The salinity threshold for

rice is 3.0 dS/m with a 12% reduction in yield, per dS/m, beyond this threshold

(Gao et al., 2007). However, some degree of genotype tolerance for salt stress

tolerance is available in rice germplasm. Among 180.000 rice genotypes screened

by the International Rice Research Institute (IRRI), 17% had acceptable tolerance

at an EC of 10 dS/m at seedling stage (Gregorio et al., 2002).

Salt stress affects plants in two distinct phases. The first phase is the

osmotic effect; independent of the accumulation of salt in the shoot. Salts

(34)

6

from roots thermodynamically unfavorable, which induces water deficit (Pardo,

2010; Roy et al., 2014). Water deficit is rapidly transmitted (within minutes) from

roots to shoots causing intracellular turgor reduction and decreased cell

expansion (Munns, 2005; Munns and Tester, 2008). This signal also promotes the

biosynthesis of abscisic acid (ABA), which will induce stomatal closure and

consequently reduction in transpirational water loss (Munns, 2005; Munns and

Tester, 2008; Roy et al., 2014). Lower stomatal conductance leads to a lower

carbon assimilation, biomass production and decreased yield. The second phase

of salinity is ionic specific; this is due to the accumulation to toxic concentrations

of sodium (Na+) and/or chloride (Cl-) ions, especially in the older leaves, inducing

tissue necrosis and early leaf senescence (Roy et al., 2014). For most plant

species Na+ appears to reach a toxic concentration earlier than Cl- (Tester and

Davenport, 2003), and for rice (Chi Lin and Huei Kao, 2001; Tsai et al., 2004) Na+

has been shown to be the primary toxic ion. Furthermore, osmotic and ionic stress

disturb aerobic metabolism and induce the accumulation of reactive oxygen species (ROS) beyond the plant’s capacity for cellular oxidant detoxification, which in turn negatively affects cellular structures and metabolism (Chaves and

Oliveira, 2004; Chaves et al., 2009).

A deleterious effect imposed by salt stress, during the second phase, is

ions imbalance (Munns and Tester, 2008). Potassium (K+) is an essential

macronutrient that plays important functions related to enzyme activation, osmotic

adjustment and turgor generation, regulation of membrane potential, and

cytoplasmatic pH homeostasis (Barragán et al., 2012; PPI, 1998). Due to

similarity in physicochemical properties between Na+ and K+ (i.e., ionic radius and

ion hydration energy), the former competes with K+ for major binding sites in key

metabolic processes in the cytoplasm, such as enzymatic reactions, protein

synthesis and ribosome functions (PPI, 1998; Marschner, 1995). Na+ inhibits

enzyme activity of many of these enzymes that require K+ for functioning

(Duggleby and Dennis, 1973). With over 50 different cytoplasmic enzymes being

(35)

7

impairment, both in root and leaf tissues (PPI, 1998; Marschner, 1995). It has

been suggested that for plant survival under salt stress, it is essential to maintain

a high K+ concentration while keeping a low concentration of Na+ in the cytosol,

resulting in a high cytosolic K+/Na+ ratio. The restriction of Na+ accumulation in

shoots under salt stress has been reported as correlating with the salt stress

tolerance of rice (Lutts et al., 1996) and maize (Zea mays L.) (Tester and

Davenport, 2003).

SODIUM UPTAKE FROM SOIL, SENSING AND SIGNALING MECHANISMS

The enormous negative membrane potential across the plasma

membrane of plants root cells (negative inside) favor the passive transport of Na+

into the cells, and especially so when the sodium concentration increases in the

soil solution. In contrast, Na+ efflux (i.e., removal from the cell) is not passive and

requires energy expenditure (Maathuis et al., 2014). The major pathway for

passive Na+ entry into root cells at high soil salinity is mediated by a family of Non

Selective Cation Channels (NSCCs family), but their molecular identity remains

unknown (Blumwald et al., 2000; Kronzucker and Britto, 2011) (Fig. 1). In addition

to the Na+ flow across cellular membranes to enter the roots (symplast flow), it has

been reported that, at least in some species, interruptions in the endodermis

(passage cells) allow the movement of water and solutes (i.e., Na+) through the

cell wall and intercellular spaces. This type of transport, to the xylem stream,

without crossing the plasma membrane is referred as “apoplast flow” (Kronzucker

and Britto, 2011; Yeo et al., 1987) (Fig. 1). Casparian strips and suberine layers in

the root endoderm and exodermal layers provide some barrier to apoplast flow

(Yeo et al., 1987). In many plant species, such as rice, the apoplast flow is

considered to be the major port of Na+ entry ( 50% of total Na+ uptake) (Yeo et

al., 1987), especially at high salinity levels, and is responsible for a significant

amount of Na+ transported to the shoot (Kronzucker and Britto, 2011; Yeo et al.,

1987). Na+ ions taken up by the roots are then transported to shoots via xylem

(36)

8

creating tensions in the root xylem, which provides the major force to move water

from roots to up to the shoots (Nobel, 2009) (Fig. 1).

Sodium has also a strong inhibitory effect on K+ uptake by cells, probably

by inhibiting K+ transporters, such as AKT1 (hyperpolarization-activated

inward‐rectifying K+ channel), a major player in K+ acquisition by plants (Fuchs et

al., 2005; Hirsch et al., 1998), and HAK5 (carrier-type HUP/HAK/KT transport)

(Nieves-Cordones et al., 2010), both present in the plasma membrane of root

cells. Additionally, membrane depolarization caused by large cytosolic Na+ influx

results in increased K+ efflux possible through depolarization-activated

outwardrectifying K+ channels (e.g.,GORK) (Adams and Shin, 2014) and NSCCs

(Sun et al., 2009).

Very little is known about how Na+ is sensed in most cellular systems. In

theory, Na+ can be sensed either before or after entry the cell, or both.

Extracellular Na+ may be sensed by a membrane receptor, whereas intracellular

Na+ may be sensed either by membrane proteins or by any of the many Na+

sensitive enzymes in the cytoplasm (Conde et al., 2011). The plasma membrane

Na+/H+ antiporter SOS1 (Salt Overly Sensitive 1) has been described as a

possible Na+ sensor (Shi et al., 2000). Its transport activity is essential for Na+

efflux from cells (Quintero et al., 2002), but its unusually long cytoplasmatic tail is

thought to be involved in Na+ sensing (Shi et al., 2000) (Fig. 3). However, this

mechanism it is not fully clear.

In plant cells, Ca2+ acts as a second messenger connecting a wide range

of extracellular stimuli with various intracellular responses (Conde et al., 2011).

Salt stress originates a fast and transient increase in free cytosolic Ca2+, likely

released from the vacuole (Pottosin et al., 2009), that is decoded by Ca2+ sensors

such as calmodulin (CaM), calcineurin B-like proteins (CBLs) and CBL-interacting

protein kinases (CIPKs). When acting as a CBL-CIPK complex, these Ca+ sensors

are often designed as calcium-dependent protein kinases (CDPKs) (Conde et al.,

2011; Yang and Poovaiah, 2003). Cytosolic Ca2+ sensors in turn trigger many

(37)

9

(e.g,. NSCCs are strongly blocked by external Ca+2), as well as enzymatic activity

and gene transcription, ending up in ion homeostasis (Adams and Shin, 2014;

Conde et al., 2011; Martinez-Atienza et al., 2007; Pardo and Quintero, 2002;

Yamaguchi et al., 2005).

MECHANISMS OF SALT TOLERANCE

Salt stress frequently affects plant habitats and many species evolved

varied mechanisms for dealing with it. These mechanisms for salt tolerance can

be classified into three main categories. The first one is osmotic stress tolerance,

which is regulated by is regulated by long distance signals that reduce shoot

growth (Roy et al., 2014) and involves biosynthesis and accumulation of

compatible solutes to maintain water uptake (Peleg et al., 2011). Another

mechanism is ion exclusion, in which Na+ transport reduces the accumulation of

toxic Na+ within leaves. This system operates by controlling the Na+ loading to the

xylem and Na+ retrieval from the xylem, before reaching the shoot photosynthetic

tissues (Fig. 1). Finally, the third mechanism is tissue tolerance, in which high salt

concentration is found in leaves, but Na+ is compartmentalized at the cellular and

intracellular level(especially in the vacuole) reducing the deleterious effect of Na+

in the cytosol and driving water uptake to cells (Fig. 1) (Munns and Tester, 2008).

In most cases, the plant salt stress tolerance relies on the three mechanisms

together, rather than only one mechanism is particular (Munns and Tester, 2008;

Pires et al., 2015; Roy et al., 2014).

SODIUM TRANSPORTERS AND PLANTS SALT STRESS TOLERANCE

The study of salt stress tolerance in plants usually focuses on the control

of Na+ movement, namely on: Na+ exclusion in roots, Na+ long distance transport,

and Na+ compartmentalization at both cellular and tissue level (Conde et al., 2011;

Munns, 2005; Roy et al., 2014). These processes are mediated by membrane

transporters reason why the manipulation of their activity has an enormous

(38)

10

Here, we focus on the specific membrane transporters described as involved in

the above outlined tolerance processes. In contrast to animal cells, higher plants

do not have Na+-ATPases or Na+/K+-ATPases and rely on H+-ATPases and H+

-pyrophosphatases (PPases) to create a proton-motive force necessary to drive

Na+ transport across membranes (Conde et al., 2011). The plasma membrane

localized SOS1 (Ji et al., 2013; Martinez-Atienza et al., 2007) and the vacuole

membrane (tonoplast) localized NHX1 (Fukuda et al., 2011; Jiang et al., 2010) are

two Cation/H+ antiporters involved in Na+ exclusion back to the soil and in K+-Na+

compartmentalization in the vacuole. In addition, members of the HKT1 family of

HKTs (high affinity potassium transporters) are involved in the control of Na+ long

distance transport by reabsorption of Na+ from the xylem sap into the root cells,

preventing the large accumulation of Na+ in the above-ground tissues (Rus et al.,

2004) (Fig. 1). It is noteworthy that HKT1 Na+ exclusion mechanism from the

transpiration stream has been frequently indicated as a strong trait in salt

tolerance of different cereals, such as rice (Ren et al., 2005) and durum wheat

(Triticumturgidum L. subsp. durum) (James et al., 2006).

In the following sections, the role that different Na+ transporters and H+

(39)

11

Figure 1. Summary diagram showing key plasma and tonoplast membrane transporters, channels and pumps mediating Na+ and K+ homeostasis in plants under salt stress (adapted

from Roy et al. 2014). Na+ ions enter the cells via Non Selective Cation Channels (NSCCs),

likely other cation transporters (not shown) and through the cell wall and intercellular spaces (apoplast flow – red arrow). The Na+/H+ antiporter SOS1 extrudes Na+ at the root soil interface,

thus reducing the Na+ net influx of Na+. At the xylem parenchyma cells, HKT1-like proteins

retrieve Na+ from the xylem sap thereby restricting the amount of Na+ reaching the

photosynthetic tissues. To translocate Na+ back to the root, ions unloaded from xylem may be

transported into phloem via additional HKT1-like protein. In addition, HKT1-like proteins also load Na+ into shoot phloem and then Na+ is transferred into roots via downstream of phloem,

preventing Na+ accumulation in shoots. SOS1, xylem parenchyma cells localized, is also

suggested to mediate Na+ efflux from xylem vessels under high salinity. Incoming Na+, in root

and shoots, is stored in the large central vacuole by tonoplast localized NHX exchangers (NHX1-4). Plasma membrane (PM) H+-ATPase (P-ATPase), PM H+-PPase (PM-PPase),

tonoplast H+-ATPase (V-ATPase) and tonoplast H+-PPase (V-PPase) generate electrochemical

potential gradient for secondary active transport.

H+-Pumps and the plant salt stress response

Proton gradients are crucial for the transport of ions and solutes across

the different plant cell membranes. Three primary proton transport proteins are

found in plant cells: (1) plasma membrane (PM) and (2) vacuolar H+-ATPases,

which couple ATP hydrolysis with proton transport, and (3) PM and vacuolarH+

(40)

12

al., 2010; Gaxiola et al., 2007). In plant cells, H+-ATPase and H+-PPase are major

components of the vacuole membrane (Silva and Gerós, 2009). The H+-Pumps

generate an electrochemical potential gradient across membranes, which is the

motive force for a large set of secondary transports.

1- Plasma membrane H+-ATPase

The PM H+-ATPase belongs to a class known as P-type ATPases

(P-ATPases), and is encoded by a large gene family (Fuglsang et al., 2010; Gaxiola

et al., 2007). The pump is formed by a single subunit protein, which contains ten

trans-membrane helices and a large cytoplasmatic domain (Fuglsang et al.,

2010). Arabidopsis and rice genomes encode eleven and ten P-ATPases,

respectively (Arango et al., 2003; Axelsen and Palmgren, 2001).

The proton motive force created by P-ATPases is largely responsible for

an inside negative potential across the plasma membrane, which is essential for

root nutrient uptake, stomatal aperture, phloem loading, and cell growth

(Blumwald et al., 2000; Gaxiola et al., 2007; Mansour, 2014). Besides regulation

of many physiological processes, the P-ATPases have a critical role in plant

adaptation to salt stress conditions. Higher P-ATPases activity under salt stress

conditions repolarizes the NaCl-induced depolarization of PM. This response has

been strongly associated with salt stress tolerance (Mansour, 2014). The

maintenance of the PM potential under salt stress through P-ATPases activity has

a great effect on reduction of Na+ influx via depolarization-activated NSCCs and

K+ efflux via KORs and NSCCs, which help to restore higher K+/Na+ levels (Sun et

al., 2009). Also, P-ATPases higher activity under stress energizes the active

transport that exclude Na+ from root cells, a process dependent on the SOS1

Na+/H+ antiporter (Gaxiola et al., 2007). Furthermore, it was reported that higher

activation of P-ATPases is often found in halophytes and salt tolerant genotypes,

which may correlate with salt stress tolerance (Mansour, 2014). For instance, in

rice callus lines, a higher activation of P-ATPases occurred in salt-tolerant lines as

(41)

13

The salt-dependent activation of PM H+-pump is associated with

increased levels of gene expression as well as post-translational modifications of

the enzyme present in a preexisting pool (Gaxiola et al., 2007; Mansour, 2014).

However, it is likely that most regulation of the pump activity occurs at the

post-translational level (Fuglsang et al., 2010; Gaxiola et al., 2007). The pump activity

can be modulated by phosphorylation/dephosphorylation of the penultimate a.a.

residue of the cytoplasmatic C-terminus domain, a threonine residue. The

phosphorylated threonine residue promotes binding of the activating 14-3-3

protein (Fuglsang et al., 2010).

Stomatal aperture involves regulation of osmotic pressure within the

guard cells, a process powered by P-ATPases activity and responsive to a wide

variety of external signals (Gaxiola et al., 2007). Blue light perception in guard

cells is mediated by phototropins, which intitiate a signal transduction signal

pathway that involves an upstream protein phosphatase I and a downstream

protein kinase that phosphorylates the penultimate C-terminus a.a. residue of the

P-ATPase (Gaxiola et al., 2007; Takemiya et al., 2006). Under drought and salt

stress conditions, stomatal closure is induced by ABA through a mechanism that

involves production of hydrogen peroxide (H2O2) and dephosphorylation of the

P-ATPases (Gaxiola et al., 2007; McAinsh et al., 1996; Zhang et al., 2001).

2- Vacuolar H+-ATPase

Among the three proton-pumps found in plant cells, the vacuolar H+

-ATPase (-ATPase) is the most complex one (Gaxiola et al., 2007). The

V-ATPase was first found associated with the endomembrane system where it

acidifies and generates a proton force motive within diverse cell compartments

(e.g.,vacuole, endoplasmic reticulum and trans-Golgi network) (Ratajczak, 2000).

However, V-ATPases have also been associated with cell plasma membrane

(Hanitzsch et al., 2007). The ability of the V-ATPase to maintain the cytosolic pH

homeostasis and to acidify the endomembrane compartments is crucial during

(42)

14

Vacuolar H+-ATPases are multisubunit enzymes composed of two

subcomplexes: Three peripheral V1 complexes consisting each of eight subunits

(A, B, C, D, E, F, G and H) responsible for ATP hydrolysis, and the

membrane-integral V0 complex comprising up to six subunits (a, c, c´, c´´, d and e)

responsible for proton translocation (Gaxiola et al., 2007) (Fig. 2). In plants, the

sub-unit c´ is not found and many of the V-ATPase subunits are encoded by gene

families. In Arabidopsis andrice, the 13 subunits which compose the vacuolar H+

-ATPases (A, B, C, D, E, F, G, H, a, c, c´´, d and e) are encoded by a total of 27

genes and 22 genes respectively (known as VHA genes). If all possible isoform

combinations are used, we will have hundreds of different V-ATPase complexes

(Hanitzsch et al., 2007; Sze et al., 2002).

By convention, the subunits of V1 and V0 complexes are distinguished

with capital and case letters, respectively. The V1 complex consists of: (1) a

globular hexameric head with three alternating copies of subunits A and B forming

a ring, (2) a central rotational stalk composed of single copies of subunits D and F,

and (3) a outer stalk made of subunits C, E, G and H. Subunits A and B mediate

the hydrolyses of ATP at three reaction sites associated with subunit A. Both the

central rotational stalk and fixed outer stalk connect the V1 complex to the

membrane inserted V0 complex. The proton transporting V0 complex consists of

six or more c subunits, also forming a ring structure. In addition, each V0 complex

contains one copy of subunits a, d and e (Beyenbach and Wieczorek, 2006;

Hanitzsch et al., 2007) (Fig. 2). It has been reported that structural changes of the

V-ATPase complex or presence/absence of individual protein isoforms could be

correlated with differences in V-ATPase localization and activity between plant

(43)

15

The plant vacuole plays a very important role in the maintenance of

cellular metabolism due to its role in long term storage of toxic ions, long or short

term storage of mineral and/or organic acids and in pH and Ca2+ cytoplasmatic

homeostasis. Furthermore, the V-ATPase is the most abundant H+-pump in the

tonoplast, and it has been shown that its activity is modulated to cope with

environmental and metabolic changes (Ratajczak, 2000). For instance, under salt

stress, a general increase of V-ATPase activity has been reported in many plant

species (Matsumoto and Chung, 1988; Silva and Gerós, 2009). The V-ATPase

provides the driving force necessary for K+-Na+ vacuole compartmentalization, a

process linked on the NHX1 antiporter activity (Bassil and Blumwald, 2014; Jiang

et al., 2010).

The ability to respond salt stress via changes in the expression of the

V-ATPase subunits encoding genes might be a prerequisite and a characteristic of

salt stress tolerance in plants. It has been reported that the transcript levels of

some subunits are up-regulated in response to salt stress (Kirsch et al., 1996;

Narasimhan et al., 1991; Silva and Gerós, 2009). However, the expression of

V-ATPase genes does not always involve a fixed stoichiometry of mRNAs for the

different subunits (Silva and Gerós, 2009). Other factors may also account for the Figure 2. Structural model of the plant V-ATPase adapted from Gaxiola et al. (2007). The peripheral V1 complex (blue)

and the membrane integral V0

complex (orange) are joined through a peripheral stalk formed by subunits a, C, E, G and H. Hydrolysis of ATP is couple with H+ transport to the

(44)

16

regulation of V-ATPase activity. Indeed, VHA-A subunit from barley (Hordeum

vulgare L.) was shown to interact to 14-3-3 proteins, well known activators of PM

ATPases, in a phosphorylation-dependent way. That interaction was suggested to

activate V-ATPase activity (Klychnikov et al., 2007).

3- Plasma membrane and vacuolar H+-PPase

H+-pyrophosphatases (H+-PPase) are highly hydrophobic single subunit

proteins that generate proton gradients across the vacuole, Golgi and plasma

membranes using the energy of hydrolysis of pyrophosphate (PPi) molecules

(Gaxiola et al., 2007). Plants have two phylogenetically distinct types of H+

-PPases: type I and type II. Type I H+-PPases depend on cytosolic K+ for their

activity and are moderately sensitive to inhibition by Ca2+, and type II H+-PPases

are K+ insensitive but extremely Ca2+ sensitive.

The Arabidopsis genome encodes two H+-PPases: one type I H+-PPase

(AVP1) and another type II H+-PPase (AVP2) (Drozdowicz et al., 2000). Rice

genome also encodes two H+-PPases: OVP1 and OVP2 (Sakakibara et al., 1996).

However, more isoforms have been proposed (Choura and Rebai, 2005).

Phylogenetic analysis of V-PPase sequences showed that rice H+-PPases are

likely to be type I H+-PPases (Drozdowicz et al., 2000). Type I H+-PPases are

mainly suggested to acidify the vacuole (Gaxiola et al., 2007). However these H+

-pumps were also found in the plasma membrane (Alexandersson et al., 2004;

Ratajczak et al., 1999). Arabidopsis type II H+-PPases, AVP2, has been shown to

localize exclusively to Golgi apparatus (Mitsudaa et al., 2001).

The expression levels of the H+-PPases are strictly regulated at

transcriptional level in response to various environmental conditions or

developmental stages (Fuglsang et al., 2010; Silva and Gerós, 2009). Under salt

stress an increased H+-PPase activity has been reported (Maeshima, 2000). It

has also been shown that the pollen-specific cis-acting region of the AVP1 gene is

involved in the expression of the AVP1 gene during pollen development.

AtCAM15, AtCAMTA 1 (calmoduline-binding transcription factors) (Mitsuda et al.,

(45)

17

(Mitsuda et al., 2004) were identified as binding to the cis-acting region of the

AVP1 gene. However, a comprehensive mechanism of H+-PPase gene

expression and post-translational regulation is still needed. It is likely that the

protein C-terminus plays an essential role in supporting the physiological function

of H+-PPase activity (Fuglsang et al., 2010).

Given the importance of the pH homeostasis in the cytosol for cell

metabolism, it is likely that the activity of all three H+-pumps (P-ATPase,

V-ATPase and H+-PPase) is regulated by common regulatory mechanisms (e.g.,

14-3-3 proteins).

SOS1 and the plant salt stress response

Comparisons of unidirectional Na+ fluxes and rates of net accumulation of

Na+ in root indicate that 70-99% of the Na+ fluxed into the root is extruded back to

the apoplast (Munns, 2005; Tester and Davenport, 2003). For rice, that value is

indicated as 96% (Munns, 2005), meaning that over time Na+ will be accumulated

into roots and transferred to shoot via the transpiration stream, later accumulating

there. Since it is important to maintain low cytoplasmatic Na+ concentrations for

growth and survival under saline conditions, plants have developed a direct

mechanism to extrude Na+ from cells across the plasma membrane to the soil or

apopoplast. Small differences in Na+ exclusion capacity creates major changes in

Na+ net accumulation (Brini and Khaled., 2012; Munns, 2005; Tester and

Davenport, 2003). However, the role of cellular Na+ efflux is not intuitive in

multicellular plants, as Na+ transport out of one cell would negatively impact the

surrounding neighbor cells. So, the role of Na+ efflux has to be considered in

specific tissues and in the context of whole plants (Zhu, 2003). Sodium efflux is

catalyzed by the plasma membrane Na+/H+ antiporter encoded by SOS1 (Salt

Overly Sensitive1 = AtNHX7) gene, identified in several plants including

Arabidopsis (Wu et al., 1996), rice (Martinez-Atienza et al., 2007), wheat (Xu et

al., 2008), and tomato (Xu et al., 2008). SOS1 uses the proton gradient

(46)

18

for H+ across the membrane (Ji et al., 2013; Qiu et al., 2004; Shi et al., 2002).

Activity of the ArabidopsisSOS1 promoter is detected ubiquitously in virtually all

tissues, but it appears to be more active in: (1) root epidermal cells (particularly at

the root tip), suggesting that meristem requires special protection, since the root

tip cells have very small vacuoles and thus incapable of vacuolar Na+

compartmentalization, and (2) root parenchyma cells lining the vasculature

(Kronzucker and Britto, 2011; Shi et al., 2002). The SOS1 gene expression

pattern, together with the results of ion analysis in sos1 mutant plants, suggests

that SOS1 has several roles: (1) Na+ efflux from roots; (2) slow down Na+

accumulation in the cytoplasm in order to gain time for Na+ storage in the vacuole;

and (3) control of long-distance Na+ transport between roots and leaves by loading

and unloading Na+ into and from the xylem (Conde et al., 2011; Zhu, 2003). SOS1

may mediate active loading of Na+ to the xylem under mild salinity (25mM NaCl).

However, at salt stress (100 mM NaCl), expression of SOS1 is induced and SOS1

may function in Na+ retrieval from the xylem (Shi et al., 2002). Such SOS1 role in

long-distance transport is important for coordination between transpiration Na+

flow and Na+ vacuolar sequestration in leaves. However, thermodynamic analysis

by Munns and Tester (2008) indicates the Na+ removal from the xylem is unlikely

to be mediated by a Na+/H+ antiporter such as SOS1, because its operation “in reverse” under high Na+ conditions is thermodynamically unfavorable. Instead, class I HKTs have been shown to be involved in xylem unloading of Na+

(Davenport et al., 2007; James et al., 2006; Ren et al., 2005). Thus, the role of

SOS1 in long-distance Na+ transport remains unclear. Nevertheless, many reports

suggest that SOS1 plays a critical role in Na+ exclusion, thus maintaining cellular

ion homeostasis and allowing plants to survive and grow under salt stress

conditions (Cuin et al., 2011; Shi et al., 2003).

The transcript level of SOS1 is upregulated by salt stress (Shi et al.,

2000). Analysis of the 2 Kb upstream of the SOS1, CIPK24/SOS2 and

CBL4/SOS3 transcription initiation site revealed that the promoter of these genes

(47)

19

WRKY, and TCP classes (Ji et al., 2013). However, TF(s) mediating promoter

activity of SOS genes have not yet been identified. Upregulation of SOS1

transcripts under salt stress is suggested to be regulated at post-transcriptional

levels, as SOS1 promoter activity is not upregulated by salt stress but the SOS1

gene expression driven by the constitutive Cauliflower mosaic virus 35S promoter

is (Shi et al., 2003). This may indicate that the SOS1 transcript is unstable in the

absence of salt stress and the salt stress causes a post-transcriptional

stabilization of the transcript (Shi et al., 2003). More recently, it was suggested

that the Na+ stress induced SOS1 mRNA stability is mediated by reactive oxygen

species (ROS) (Chung et al., 2008). In addition, SOS1 upregulation by salt stress

is partly under the control of SOS2 and SOS3 (Shi et al., 2000). CIPK24/SOS2 is

a protein kinase and CBL4/SOS3 is a calcium sensor, that together with SOS1 are

the three key components comprising the Salt Overlay Sensitive (SOS) signaling

pathway identified in Arabidopsis (Wu et al., 1996) and rice (Martinez-Atienza et

al., 2007). At the cellular level, the SOS signaling pathway has been proposed to

mediate cellular signaling under salt stress to maintain the ion homeostasis (Ji et

al., 2013).

Activation of the Na+/H+ antiport activity of SOS1 by salt stress is

controlled by SOS3 and SOS2 (Ji et al., 2013; Zhu, 2003). In response to an

external stimulus, such as high Na+ concentration, transient increases in

cytoplasmatic Ca2+ occur and that is decoded by the calcineurin B and neuronal

Ca2+ sensor-like protein SOS3. Activation of SOS3 requires N-myristoylation and

Ca2+ bound on EF-hand Ca2+ binding sites. Activated SOS3 then physically

interacts with the auto-inhibitory domain of the SOS2, a member of the SnRK

(sucrose non-fermenting-related serine/threonine kinase) family, which activates

the kinase and facilitates the localization of the SOS2-SOS3 complex to the

plasma membrane. The SOS2-SOS3 complex then associates with the Na+/H+

antiporter SOS1, phosphorylating its C-terminal auto-inhibitory domain, which

becomes activated and pumps Na+ out of the cell (Brini and Khaled., 2012;

(48)

20

The SOS pathway is not limited to the three main proteins, as it interacts

with other stress related proteins. A SOS3 homolog SOS3-LIKE Calcium Binding

Protein8 (SCABP8/CBL10) interacts with SOS2 to form an alternative protein

kinase complex that regulates SOS1 activity in the plasma membrane in response

to salt stress, mainly in shoots; while SOS3 functions primarily in the root (Quan et

al., 2007) (Fig. 3). SOS2 phosphorylates CBL10 in a Ca2+ independent manner

upon salt stress, and this phosphorylation stabilizes the SOS2-CBL10 complex

association with the plasma membrane and increases SOS1 antiporter activity

(Hasegawa, 2013; Kim et al., 2007; Quan et al., 2007). Abscisic acid insensitive

(ABI2) interacts with SOS2 to prevent SOS3 binding to SOS2 and kinase

activation. Such ABI2-SOS2 interaction may represent an integrating node

between salt stress and ABA signaling (Hasegawa, 2013; Ohta et al., 2003).

The SOS pathway may also regulate the Na+ vacuolar

compartmentalization. Interaction of SOS2-CBL10 may result in localization of the

kinase complex at the vacuolar membrane where it is possibly involved in the

regulation of Na+/H+ exchange at the tonoplast, presumably by regulation of NHX

antiporter activity (Kim et al., 2007; Qiu et al., 2004). However, to date no NHX

antiporter has been shown to be directly regulated by SOS2 and/or by the

SOS2-complex. In addition, SOS2 has been suggested to regulate the V-ATPase

activity. SOS2 was found to interact with the B1 and B2 subunits of the V-ATPase

in the absence of CBL proteins, and tonoplast vesicles from the Arabidopsis

sos2-2 mutant showed reduced ATPase and H+-translocation activities (Batelli et al.,

2007).

Potassium homeostasis has also been shown to be modulated by the

SOS signaling pathway. The CIPK23 directly phosphorylates and activates the

AKT1 K+ channel at the plasma membrane, significantly increasing the K+ uptake

under low-K+ stress (Pardo, 2010; Ren et al., 2013). Furthermore, the protein

CBL10 has been indicated to directly interact with AKT1 channel, negatively

regulating its activity in roots in a CIPK-independent way (Ren et al., 2013). It is

(49)

21

high K+/Na+ cytosolic ratio under stress (Tester and Davenport, 2003). The

possibility that CBL10 functions as an interconnecting regulator of SOS1 and

AKT1 may indicate that CBL10 plays a crucial role in ion homeostasis (K+/Na+)

under salt stress by regulating both Na+ and K+ uptake/exclusion (Ren et al.,

2013).

HKTs and the plant salt stress response

Another important determinant of salt stress tolerance in plants is the

activity of the HKT (high affinity potassium transporter) proteins (Munns and

Tester, 2008; Roy et al., 2014). The HKT family is quite diverse, and this diversity

reflects their large amplitude of functions (Almeida et al., 2013; Munns and Tester,

2008; Roy et al., 2014). The HKT family is divided in two distinct classes

according to their transport characteristics. The main distinguishing feature is the

a.a. sequence that constitutes the first pore domain (PD) (Platten et al., 2006).

Members of class I transporters (HKT1) have a serine (S), forming an S-G-G-G

motif, where most of the members of class II (HKT2) have a G in the position

occupied by the S in class I transporters, forming a G-G-G-G motif (Maser et al.,

2002). The presence of either S or G at this position is critical for K+ specificity of

the transporter. The presence of a S is characterized by a preference for Na+

conductance over other cations (HKT1), whereas the presence of a G is

characterized by transport of Na+ and/or K+ depending on the external

concentrations of these two ions (HKT2) (Kronzucker and Britto, 2011; Platten et

al., 2006). However, there are notable exceptions, in particular HKT2;1 from

cereals, in which the G has reverted to S (Kronzucker and Britto, 2011), but it has

been clearly shown to be involved in mediating Na+ and K+ entry into roots

(Kronzucker and Britto, 2011; Munns and Tester, 2008). The main role of HKT1 is

believed to be Na+ retrieval from the transpiration stream avoiding the over

Imagem

Figure  1.  Summary  diagram  showing  key  plasma  and  tonoplast  membrane  transporters,  channels and pumps mediating Na +  and K +  homeostasis in plants under salt stress (adapted  from  Roy  et  al
Figure  2.  Structural  model  of  the  plant  V-ATPase  adapted  from Gaxiola et al. (2007)
Figure  3.  Summary  diagram  of  an  hypothetical  Arabidopsis  cell  indicating  subcellular  localizations, functions, and regulations of NHXs antiporters (NHX1-6), plasma membrane H +  -ATPase  (P--ATPase),  tonoplast  membrane  H + -ATPase  (V-ATPase)
Figure 1 shows salt stress visual injuries effects on selected genotypes.
+7

Referências

Documentos relacionados

Through computational analysis of the spatial distribution of the GR consensus binding motif in upstream promoter regions, and its relationship with GR occupancy as measured

We studied the polymorphism of MHC class I ( HLA-A , -B , -C ), and class II ( HLA-DRB1 , -DQB1 ) genes using high-resolution sequence based typing (SBT) method and we structured

In our study, a variety of results indicated that CKS2 is a direct target of miR-26a in TPC-1 cells: (i) a seed binding site sequence for miR-26a was identified in the 3 9 UTR of

Although 69 nucleotides upstream of the annotated ATG start site is conserved sequence of Delta class, this slightly longer protein would still be too short to be a viable GST enzyme

A 3D model of the target sequence was constructed with the automodel class of SPDBV to generate five similar iterative models of the target sequence based on its template structure

lation (positive or negative) of the expression pattern of genes which possess an ApiAP2 domain binding motif in their upstream region with the profile displayed by the gene

A crystal structure of a class I MHC protein with a pentapeptide epitope occupying only the C- terminal end of the binding site [54] and molecular dynamics studies of peptide-free

Figure 3. Discovery of a motif involved in early zygotic gene transcription. Analysis of 400 bp upstream sequence from 61 EZGs by SCOPE. A) Snapshot of SCOPE output showing the top