effects of different arbuscular mycorrhizal fungal ...conductivity (lp r) value was achieved by...

17
Original Article Effects of different arbuscular mycorrhizal fungal backgrounds and soils on olive plants growth and water relation properties under well-watered and drought conditions Monica Calvo-Polanco 1,5 , Iván Sánchez-Castro 2 , Manuel Cantos 3 , José Luis García 3 , Rosario Azcón 1 , Juan Manuel Ruiz-Lozano 1 , Carmen R. Beuzón 4 & Ricardo Aroca 1 1 Department of Soil Microbiology and Symbiotic Systems, Estación Experimental del Zaidín (CSIC), C/Profesor Albareda, Granada 18008, Spain, 2 Department of Microbiology, University of Granada, Av. Fuentenueva s/n, Granada 18071, Spain, 3 Department of Plant Biotechnology, Instituto de Recursos Naturales y Agrobiología (CSIC), Av. Reina Mercedes, 10, Sevilla 41012, Spain, 4 Department of Cellular Biology, Genetics and Physiology, Campus de Teatinos, University of Málaga, Málaga 29010, Spain and 5 Biochimie et Physiologie Moléculaire des Plantes, SupAgro/INRA UMR 5004. 2, Place Viala, Montpellier Cedex 2 34060, France ABSTRACT The adaptation capacity of olive trees to different environ- ments is well recognized. However, the presence of microor- ganisms in the soil is also a key factor in the response of these trees to drought. The objective of the present study was to elucidate the effects of different arbuscular mycorrhizal (AM) fungi coming from diverse soils on olive plant growth and water relations. Olive plants were inoculated with native AM fungal populations from two contrasting environments, that is, semi-arid Freila (FL) and humid Grazalema (GZ) regions, and subjected to drought stress. Results showed that plants grew better on GZ soil inoculated with GZ fungi, indicating a preference of AM fungi for their corresponding soil. Furthermore, under these conditions, the highest AM fungal diversity was found. However, the highest root hydraulic conductivity (Lp r ) value was achieved by plants inoculated with GZ fungi and growing in FL soil under drought conditions. So, this AM inoculum also functioned in soils from different origins. Nine novel aquaporin genes were also cloned from olive roots. Diverse correlation and association values were found among different aquaporin expressions and abundances and Lp r , indicating how the interaction of different aquaporins may render diverse Lp r values. Key-words: Olea europaea; aquaporins; arbuscular mycorrhizal fungi; root hydraulic conductivity. Abbreviations: AM, arbuscular mycorrhiza; FL, Freila (arid location); FW, fresh weight;gs, stomatal conductance; GZ, Grazalema (humid location); H, Shannon biodiversity index; HPFM, high-pressure flow meter; IBA, indolbutyric acid; Kr, root hydraulic conductance; Lpr, root hydraulic conductivity; OTU, operational taxonomic unit; PIP, plasma membrane intrinsic protein; RWC, relative water content; S, specific richness; TIP, tonoplast intrinsic protein; ΦPSII, photochemical efficiency of photosystem II. INTRODUCTION Olive trees (Olea europaea L.) are among the most important crops for the Mediterranean area. They grow in very diverse environments (Parodi 1978) and are known for their resistance to water stress, especially during summer, when they face scarce precipitation and high temperatures (Connor 2005). Nevertheless, their response to water stress has been shown to be quite unpredictable, as their intraspecic genetic diversity is high (Guerfel et al. 2009). Even if olive tree is considered as a drought-resistant species, the number of irrigated cultivars is increasing (Carr 2013), and the balance between irrigation and productivity is a major issue as water is scarce in most countries growing this crop. With the current prediction of a decline in water availability and an estimation of crop produc- tion decline for the next years, the selection of olive trees with enhanced drought tolerance is needed to maintain and even improve their productivity. The effects of drought on plants vary among species consid- ered, although one of the rst responses of plants to the lack of water within the soil prole is to adjust their internal water balance by closing stomata (Gómez-del-Campo 2007; Schachtman and Goodger 2008) and by regulating their root hydraulic conductivity (Maurel et al. 2008; Aroca et al. 2012). Even though the regulation of root hydraulic conductivity has been mainly attributed to the action of aquaporins (Luu and Maurel 2005, Maurel et al. 2015), root and leaf osmotic adjustment as well as root architecture may be also playing a very important role in water uptake under prolonged period of drought stress. Three aquaporins have been described in olive trees, that is, OePIP1;1, OePIP2;1 and OeTIP1;1 (Secchi et al. 2007, Lovisolo et al. 2007). As in another tree species, these aquaporins have been reported to be differently regu- lated according to the stress and the intensity of the stresses (Secchi et al. 2007). Osmotic regulation in severely water stressed olive trees has been also described in experiments with Correspondence: R. Aroca. Fax: +34 958 129 600; e-mail: raroca@eez. csic.es © 2016 John Wiley & Sons Ltd 2498 doi: 10.1111/pce.12807 Plant, Cell and Environment (2016) 39, 24982514

Upload: others

Post on 21-Jan-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Original Article

Effects of different arbuscular mycorrhizal fungal backgroundsand soils on olive plants growth and water relation propertiesunder well-watered and drought conditions

Monica Calvo-Polanco1,5, Iván Sánchez-Castro2, Manuel Cantos3, José Luis García3, Rosario Azcón1, Juan Manuel Ruiz-Lozano1,Carmen R. Beuzón4 & Ricardo Aroca1

1Department of Soil Microbiology and Symbiotic Systems, Estación Experimental del Zaidín (CSIC), C/Profesor Albareda, Granada18008, Spain, 2Department ofMicrobiology, University of Granada, Av. Fuentenueva s/n, Granada 18071, Spain, 3Department of PlantBiotechnology, Instituto de Recursos Naturales y Agrobiología (CSIC), Av. Reina Mercedes, 10, Sevilla 41012, Spain, 4Department ofCellular Biology, Genetics and Physiology, Campus de Teatinos, University of Málaga, Málaga 29010, Spain and 5Biochimie etPhysiologie Moléculaire des Plantes, SupAgro/INRA UMR 5004. 2, Place Viala, Montpellier Cedex 2 34060, France

ABSTRACT

The adaptation capacity of olive trees to different environ-ments is well recognized. However, the presence of microor-ganisms in the soil is also a key factor in the response of thesetrees to drought. The objective of the present study was toelucidate the effects of different arbuscular mycorrhizal (AM)fungi coming fromdiverse soils on olive plant growth andwaterrelations. Olive plants were inoculated with native AM fungalpopulations from two contrasting environments, that is,semi-arid – Freila (FL) and humid – Grazalema (GZ) regions,and subjected to drought stress. Results showed that plantsgrew better on GZ soil inoculated with GZ fungi, indicating apreference of AM fungi for their corresponding soil.Furthermore, under these conditions, the highest AM fungaldiversity was found. However, the highest root hydraulicconductivity (Lpr) value was achieved by plants inoculatedwith GZ fungi and growing in FL soil under droughtconditions. So, this AM inoculum also functioned in soils fromdifferent origins. Nine novel aquaporin genes were also clonedfrom olive roots. Diverse correlation and association valueswere found among different aquaporin expressions andabundances and Lpr, indicating how the interaction of differentaquaporins may render diverse Lpr values.

Key-words:Olea europaea; aquaporins; arbuscular mycorrhizalfungi; root hydraulic conductivity.

Abbreviations: AM, arbuscular mycorrhiza; FL, Freila (aridlocation); FW, fresh weight;gs, stomatal conductance; GZ,Grazalema (humid location); H′, Shannon biodiversity index;HPFM, high-pressure flow meter; IBA, indolbutyric acid; Kr,root hydraulic conductance; Lpr, root hydraulic conductivity;OTU, operational taxonomic unit; PIP, plasma membraneintrinsic protein; RWC, relative water content; S, specific

richness; TIP, tonoplast intrinsic protein;ΦPSII, photochemicalefficiency of photosystem II.

INTRODUCTION

Olive trees (Olea europaea L.) are among the most importantcrops for the Mediterranean area. They grow in very diverseenvironments (Parodi 1978) and are known for their resistanceto water stress, especially during summer, when they facescarce precipitation and high temperatures (Connor 2005).Nevertheless, their response to water stress has been shownto be quite unpredictable, as their intraspecific genetic diversityis high (Guerfel et al. 2009). Even if olive tree is considered as adrought-resistant species, the number of irrigated cultivars isincreasing (Carr 2013), and the balance between irrigationand productivity is a major issue as water is scarce in mostcountries growing this crop. With the current prediction of adecline in water availability and an estimation of crop produc-tion decline for the next years, the selection of olive trees withenhanced drought tolerance is needed to maintain and evenimprove their productivity.

The effects of drought on plants vary among species consid-ered, although one of the first responses of plants to the lack ofwater within the soil profile is to adjust their internal waterbalance by closing stomata (Gómez-del-Campo 2007;Schachtman and Goodger 2008) and by regulating their roothydraulic conductivity (Maurel et al. 2008; Aroca et al. 2012).Even though the regulation of root hydraulic conductivity hasbeen mainly attributed to the action of aquaporins (Luu andMaurel 2005, Maurel et al. 2015), root and leaf osmoticadjustment as well as root architecture may be also playing avery important role in water uptake under prolonged periodof drought stress. Three aquaporins have been described inolive trees, that is, OePIP1;1,OePIP2;1 andOeTIP1;1 (Secchiet al. 2007, Lovisolo et al. 2007). As in another tree species,these aquaporins have been reported to be differently regu-lated according to the stress and the intensity of the stresses(Secchi et al. 2007). Osmotic regulation in severely waterstressed olive trees has been also described in experiments with

Correspondence: R. Aroca. Fax: +34 958 129 600; e-mail: [email protected]

© 2016 John Wiley & Sons Ltd2498

doi: 10.1111/pce.12807Plant, Cell and Environment (2016) 39, 2498–2514

these trees and may be a complementary strategy to aquapo-rins in order to resist long-term exposure to soil water deficit(Chartzoulakis et al. 2000; Ennajeh et al. 2008). At the sametime, Tataranni et al. (2015) found a close correlation betweenmorpho-anatomical traits and Lpr values in olive treessubjected to drought.The use of AM fungi as relievers of the drought stress effects

on plants has been studied for many years. AM fungi arebelieved to have a great capacity to resist fast environmentalchanges under drought conditions and long-term stresses,allowing the plants to have a wider range of possibilities toadapt and survive (Al-Karaki 2006; Smith and Read 2008;Ruiz-Lozano et al. 2012). Olive trees appear to be highlydependent on AM fungi under arid conditions (Mekahliaet al. 2013), although the effect of the AM symbiosis on plantsvaries according to the AM fungal strain used and the plantcultivars (Binet et al. 2007), as well as to the soil chemicalcomposition (Burns et al. 2015). AM fungi establishes anextensive hyphal network in the soil, mobilizing soil nutrients,mainly phosphorus and nitrogen (Bonfante and Genre 2010,Bompadre et al. 2013) that would play a key role in plantsurvival under stress conditions. In general, AM fungi havebeen shown to increase plant resistance to drought (Azconet al. 1996; Porcel et al. 2004) and to alleviate water stress(Sheng et al. 2008; Barzana et al. 2012), while improving plantproductivity (Navarro-Fernández et al. 2011; Abbaspour et al.2012) and nutrient status (Farzaneh et al. 2011; Lee et al.2012). Furthermore, AM symbiosis causes significant changesin aquaporin abundance and activity in host plants (Arocaet al. 2007, Jahromi et al. 2008, Barzana et al. 2014). Forexample, a consistent diminution of PIP2 aquaporinphosphorylation has been observed in bean roots colonizedby the AM fungus Rhizophagus irregularis under optimalconditions (Aroca et al. 2007; Benabdellah et al. 2009;Sanchez-Romera et al. 2016), suggesting less activity of suchaquaporins (Maurel et al. 1995). These results support the ideathat each aquaporin has a specific function under different en-vironmental conditions (Aroca et al. 2007; Jang et al. 2007;Calvo-Polanco et al. 2014a) and that each plant will responddifferently to each AM colonizing fungus. Therefore, the regu-lation of root hydraulic conductivity (Lpr) and aquaporin ex-pression and abundance by AM symbiosis is far from beingunderstood. In studies using the same plant and AM fungalspecies, the response of aquaporin expression to AM inocula-tion was different (Aroca et al. 2007; Sanchez-Romera et al.2016), probably because the soil used was not the same. Fur-thermore, El-Mesbahi et al. (2012) found that the response ofLpr and aquaporin expression to AM symbiosis depended onthe soil K+ content. Furthermore, different AM fungal specieshad different effects on aquaporin expression in soybean andlettuce plants (Porcel et al. 2006).In the present study, we analysed the response to drought

stress of 7-month-old olive plants (O.europaea cv. Picual)grown in natural soils from two locations with contrastingclimatology: Freila (FL), aMediterranean location in the southof Spain with low-average annual precipitation (380mm), andGrazalema (GZ), whose average annual precipitation is high(2223mm). Plants were inoculated with the native AM fungal

population of each soil, including crossing of soils withnon-autochthonous AM fungi. The plants were subjected todrought stress for 4weeks. The objectives of the study were(1) to study how different combinations of soils, AM fungalcommunities and water regime modify root hydraulicproperties of olive trees including aquaporin expression andabundance, (2) to identify novel aquaporin genes in oliveplants taking advantage of the OLEAGEN database(Muñoz-Mérida et al. 2013) and (3) to find out whichphysiological or molecular traits are potential moredeterminants of Lpr regulation.

MATERIAL AND METHODS

Olive plants production and growth

We used olive (Olea europaea L. cv. Picual) plants to test theeffect of two contrasting natural soils from the Mediterra-nean area in Spain combined with the natural AM popula-tion from those two soils on plant tolerance to severedrought stress. This particular olive cultivar is known for itsmoderate drought tolerance, because it has been selectedfor rain fed conditions (Tugendhaft et al. 2016). Oliveexplants were produced from mature olive trees growing atFL (37°31′43″N, 2°54′34″W), a Mediterranean location inthe south of Spain. Olive branches were removed andtransported in a humid piece of cloth to a greenhouse.Cuttings were produced as explained by Suárez et al. (1999).Briefly, 18 cm length and 4–6 cmdiameter cuttings were treatedwith indolbutyric acid (IBA) 3500ppm by immersing thecutting base for 10 s in a 1:1 (v:v) IBA hydro-alcoholic solution.The explants were immediately transferred to a mistpropagation system for 90d on a perlite substrate at 25 °C basalheating for rooting.

Once the explants were rooted, they were transferred into2L pots using two different natural soils; one originating fromFL area and another one from the GZ area (36°46′4″N, 5°21′57″ W). These areas were chosen for their high olive oilproduction and their contrasting climatology. FL is a typicalsemi-arid Mediterranean location, with dry and hot summers,mild winters and low-annual total precipitation (380mm),while the GZ area has more a continental climate with coldwinters and hot summers and much higher total annualprecipitation (2223mm). Soil was collected from four differentpoints (20kg at each point), spaced 10m each other in eacholive field. The soil from the four points was mixed later. OnlyHorizon-A was collected. The soils were sieved (50mm) andsterilized by steaming for 1 h, for three consecutive days at100 °C. Plants were then transferred to a greenhouse at22/18 °C, 65% relative humidity, and were grown for 7months.Plants were irrigated with 50% modified Hoagland’s solutiononce per week (Epstein (1972)): 2.5mM KNO3, 0.5mM

KH2PO4, 2.5mM Ca(NO3)2, 1mM MgSO4, 23μM H3BO3,5μM MnCl2, 0.3μM ZnSO4, 0.2μM CuSO4, 0.01μM (NH4)6Mo7O24 and 90μM EDTA-Fe.

Inoculation was carried out at the time of planting. Theplants growing in each soil were divided into three differentgroups and the inoculation treatments were applied as follows:

Effects of AM fungal backgrounds on olive plants 2499

© 2016 John Wiley & Sons Ltd, Plant, Cell and Environment, 39, 2498–2514

(1) control, non-AM inocula; (2) naturalAM fungal populationfrom FL soil; and (3) natural AM fungal population from GZsoil. The AM fungal inocula from the natural soils wereobtained by using the wet sieving and decanting technique(Navarro-Fernández et al. 2011). Briefly, 310 g of thecorresponding 2mm sieved soil (which is equivalent to500mL of each soil with the original texture) was suspendedin 2L of water. Then, the suspension was strongly stirredfor sample homogenization and left for soil decantation.The supernatant was poured out through coupled sieves of700, 500, 250 and 50μm. This procedure was carried outtwice with the same soil sample. Finally, the material fromall sieves except from that of 700μm was added to eachpot in the planting hole, according to the correspondingtreatment. Thus, the experiment consisted of 12 treatmentsarising from the combination of two different soils (FLand GZ), three AM inoculum (non-inoculated, native AMfungi from FL soil and native AM fungi from GZ soils)and two water regime treatments (well-watered and droughtstress conditions). Each treatment was composed of sixplants.

Soil analyses and watering regime

Elemental analyses of the natural soils from FL and GZ andfrom olive leaves were carried out by an inductively coupledplasma analyser at the Instrumental Service of the EstaciónExperimental del Zaidín. The results of soil analyses arepresented in Table 1.

Six months after planting, plants from each fungal treat-ment were separated into two groups: half of them were wellwatered (95% field capacity) and the other half submitted todrought stress (55% field capacity). Soil moisture wascontrolled using a ML2 ThetaProbe (AT Delta-T DevicesLtd., Cambridge, UK), and after that, the water content ofthe soil was maintained by weighing the pots every day andreplacing the water lost to recover the desired level of soilwater content (Porcel et al. 2004). The drought treatmentlasted 4weeks.

Arbuscular mycorrhizal colonization rates andfungal mycelial length

Mycorrhizal root colonization was estimated by analysing fourroots (n=4) per treatment combination. Approximately 0.5 gof root tissues were cleared in 10% KOH and stained with0.05% trypan blue in lactic acid (v/v). The extent ofmycorrhizal colonization was calculated according to thegridline intersect method (Giovannetti and Mosse 1980).

Fungal mycelial length was determined as explained inAbbott et al. (1984). For the calculation of the total length ofthe mycelia, the following equation was used as R= (πAN/2H)×Fd ×Fps, were R is the hyphal length, A is the area ofthe filter (m2), N is the number of intersections, H the totallength of the reticulum used, Fd is the dilution factor and Fpsis the dry-weight factor. Tab

le1.

Elemen

tala

nalysesof

Grazalemaan

dFreila

soils

Location

pH(extract

1:2.5)

EC(1:5)

(mScm

�1)

P(m

gkg

�1 )

K(m

gkg

�1)

Ca

(mgkg

�1 )

Mg

(mgkg

�1)

CaC

O3

(%)

AC(%

CaC

O3)

C (%)

OM

(%)

N (%)

C/N

N-N

O3

(mgL�1 )

N-N

H4

(mgL�1)

Amon

ium

(extract

KCl2

N;

mgkg

�1)

Cl

(mgL�1)

Grazalema

8.33

0.08

4.65

221.50

4785

450

53.30

8.55

0.83

1.44

0.12

6.99

1.41

0.47

4.65

2.30

Freila

8.98

0.06

<1

45.50

3490

317

34.40

8.10

0.41

0.71

0.05

8.39

0.57

0.35

3.50

2.25

EC,electricalcon

ductivity

;AC,activecarbon

ates;O

M,o

rgan

icmaterial.

2500 M. Calvo-Polanco et al.

© 2016 John Wiley & Sons Ltd, Plant, Cell and Environment, 39, 2498–2514

Molecular identification of arbuscular mycorrhizalfungi colonizing the roots

For AM fungal identification, genomic DNA was extractedfrom 120 to 140mg of fine roots from each treatmentcombination by using the DNeasy Plant Mini Extraction Kit(Qiagen Inc., Mississauga, ON, Canada) following manufac-turer’s instructions. A nested polymerase chain reaction(PCR) approachwas used to amplify a partial LSU rRNAgeneregion (approx. 370bp) of the AM fungal DNA from the rootsamples. The primer combinations LR1/FLR2 and FLR3/FLR4 were employed as explained in Gollotte et al (2004).All amplifications were performed on a Mastercycler NexusPCR cycler (Eppendorf, Hamburg,Germany) by using CanvaxBiotech Taq polymerase (Canvax Biotech, Cordoba, Spain).PCR products were separated by gel electrophoresis on a 1%agarose gel in TAE buffer, and DNA was visualized underUV light after being stained with ethidium bromide.Four to five replicates of both PCR amplifications were

performed for each sample and the resulting amplicons pooledto yield a composite sample (Renker et al 2006). The mixedproducts from the second PCR were cloned into thep-GEM® T-Easy Vector (Promega, Madison, USA) and usedto transform commercial DH5α competent cells (Invitrogen,Carlsbad, USA), following the instructions of the manufac-turer. Up to 40 clones were screened from each cloningreaction to check the correct length of plasmid inserts bycolony PCR using vector primers T7 (5′-TAATACGACT-CACTATAGGG-3′) and M13r (5′-GGATAACAATTTCA-CACAGG-3′). At least 30 clones, containing inserts with thecorrect size, from each LSU rDNA library were sent forsequencing to GATC (GATC Biotech, Konstanz, Germany),using the vector primer M13-FP (5′-TGTAAAACGAC-GGCCAGT-3′).All sequences obtained were aligned and adjusted manually

using the programme BioEditversion 7.1.3 (`http://www.mbio.ncsu.edu/bioedit). Sequence types or operational taxonomicunits (OTUs) were defined in a conservativemanner as clustersof closely related sequences with a level of pairwise similarityhigher than 98%, except if different OTUs containedsequences from the same deposited isolate. Through BasicLocal Alignment Search Tool (BLAST) online tool, a prelimi-nary taxonomic classification was obtained for all definedOTUs. Clones of non-AM fungal origin were excluded fromfurther analyses.The abundance of each OTU in the different treatment

combinations was represented as relative abundance inrelation to the total number of AM fungal clones consideredin each clone library. It is assumed that each AM fungal typeis amplified and cloned proportionally, being the abundanceof each OTU as an approximate estimation of their proportionin the root. The Shannon biodiversity index (H′) was used toevaluate the genetic diversity (hereafter AM fungal diversity)and calculated by the formula H′=�Σ(ni/N)ln(ni/N), whereni represents the number of sequences belonging to eachphylotype and N the total number of phylotypes (Shannon &Weaver 1963). The phylotypes’ specific richness (S) was alsocalculated for each plant.

Physiological parameters

Root and leaf fresh weight (FW) were determined at the timeof harvest in six (n=6) plants per treatment combination.Relative height increments and relative number of leaves weredetermined from themeasurements taken at the beginning andat the end of drought treatment.

Stomatal conductance (gs) was measured before harvesting,in the last fully developed mature leaves with a portable AP4Porometer (Delta-TDevices Ltd, Cambridge) 4 h after sunrise.The photochemical efficiency of photosystem II (ΦPSII) andleaf relative water content (RWC) were determined in sixmature leaves per treatment combination (n=6), using thesame leaves as for gs. For ΦPSII, we used a FluorPen FP100(Photon Systems Instruments, Brno, Czech Republic), whichallows the measurement of chlorophyll a fluorescence andhence plant photosynthetic performance. To determinedRWC, mature, fully developed leaves were excised from themain shoot, weighed (W0) and introduced into 15mLcentrifuge tubes (BD Falcon, Fisher Scientific) with a piece ofwet cotton for 24h at 4 °C. Leaves were weighed again (Wh),dried at 75 °C for 2 d and then weighed (Wd) a third time,and the relative water content was calculated as RWC=(W0�Wd)/(Wh�Wd)×100. Leaf chlorophyll contents wereextracted in 100% methanol and concentration calculatedusing the coefficients and equations reported in Lichtenthaler(1987).

Root hydraulic conductance (Kr) was determined for sixcomplete roots (n=6), per each of the 12 treatments. We useda high-pressure flow meter (Dynamax, Inc., Houston), and themeasurements were taken in the same plants used for gs,between 3 and 4h after sunrise. Detached roots wereconnected to the high-pressure flow meter and water waspressurized into the roots from 0 to 0.5MPa in the transientmode to calculate root hydraulic conductance (Kr;Calvo-Polanco et al. 2012). Root hydraulic conductivity (Lpr)was determined by dividing Kr by the root volume(Calvo-Polanco et al. 2012).

Olive aquaporins identification

For olive aquaporins identification, total RNA was isolatedfrom roots by a phenol/chloroform extractionmethod followedby LiCl precipitation (Kay et al. 1987). cDNAwas synthesizedfrom 2μg of total RNA using oligo(dT)12–18 as a primer andM-MLV as reverse transcriptase (Invitrogen). For theidentification of aquaporins, cDNA was amplified by PCRusing the degenerate primers from Park et al. (2010). PCRreactions were performed as described for AM fungalidentification, except that the annealing temperature waschanged to 60 °C. After gel electrophoresis of the PCRproducts, visible bands of the expected size (approximately400kb) were recovered from the gel, eluted with a QIAquickGel Extraction Kit (Qiagen) and cloned as explainedpreviously. To determine the sequences of the full-lengthcDNAs, we used the SRS database from the OLEAGENconsortium (Muñoz-Mérida et al. 2013).

Effects of AM fungal backgrounds on olive plants 2501

© 2016 John Wiley & Sons Ltd, Plant, Cell and Environment, 39, 2498–2514

Amino acid sequences predicted for the olive aquaporingenes were used for phylogenetic analysis along with thosepredicted for other plant aquaporin genes using MEGA4software (Tamura et al. 2007). The reliability of the branchesin each resulting tree was supported with 1000 bootstrapresampling. Sequences generated in this study have beendeposited at the European Molecular Biology Laboratorydatabase under the accession numbers KT380900 toKT380908.

Root aquaporin expression analysis

Aquaporins expression was determined in the roots of threeplants of each treatment combination (n=3). Total RNAwasisolated as described earlier. DNase treatment of total RNAand reverse transcription were performed following theinstructions provided by themanufacturer (Quantitect ReverseTranscription KIT Cat#205311, Qiagen, CA). We used, for theroot aquaporin expression analyses, the three known O.europaea aquaporins (OePIP1;1 GeneBank accession no.DQ202708, OePIP2;1 GeneBank accession no. DQ202709and OeTIP1;1 GeneBank accession no. DQ202710), plus thenew ones described in this study (OePIP1;2 GeneBankaccession no. KT380904; OePIP1;3 GeneBank accession no.KT380905; OePIP2;2 GeneBank accession no. KT380900;OePIP2;3 GeneBank accession no. KT380903; OePIP2;4GeneBank accession no. KT380908; OePIP2;5 GeneBankaccession no. KT380901; OePIP2;6 GeneBank accession no.KT380906; OeTIP1;2 GeneBank accession no. KT3809902and OeTIP1;3 GeneBank accession no. KT380907). Theexpression of the different aquaporins was determined usinga real time quantitative PCR (iCycler-Bio-Rad, Hercules,CA) as explained in Calvo-Polanco et al. (2014a, 2014b). Wecould not detect any expression fromOePIP1;1 andOePIP2;1in the roots of our olive plants. For the aquaporins analysed,the annealing temperature was 58 °C, and the primers usedare described in the Supporting Information. The specificityof the PCR amplification procedure was confirmed using a heatdissociation protocol (from 60 to100 °C) after the final cycle ofthe PCR. The relative abundance of transcription wascalculated using the 2-ΔΔCt method (Livak and Schmittgen2001). To normalize aquaporin expression, we tested differentolive housekeeping genes: actin, ubiquitin and elongationfactor. The elongation factor was the one chosen afterRT-qPCR as it displayed stable mRNA levels throughout alltreatments. Elongation factor-specific primers were used forstandardization by measuring the expression of elongationfactor gene in each sample. Negative controls without cDNAwere used in all the PCR reactions.

Proteins isolation and enzyme-linkedimmunosorbent assay analysis

Microsomes were isolated as described in Hachez et al (2006).For enzyme-linked immunosorbent assay analysis, twomicrograms of the protein extracts were processed as describedin Calvo-Polanco et al. (2014a, 2014b). We used, as primary

antibodies (at a dilution of 1:1000), the two antibodies that rec-ognize several PIP1 and PIP2 and three antibodies that recog-nize the phosphorylation of PIP2 proteins at their C-terminalregion at Serine 280 (PIP2280), Serine 283 (PIP2283) and bothat Serine 280 and 283 (PIP2280/283) as described in Calvo-Polanco et al. (2014a, 2014b). A goat anti-rat IgG coupled tohorseradish peroxidase (Sigma-Aldrich Co., USA) was usedas secondary antibody at 1:10 000 for PIP1. Goat anti-rabbitIgG coupled to horseradish peroxidase (Sigma-Aldrich Co.,USA) was used as secondary antibody at 1:10 000 for PIP2and PIP2280, PIP2283 and PIP2280/283. Protein quantificationwas carried out in three different independent root samplesper treatment (n=3), replicated three times each. PIP2 anti-bodies antigens were aligned to see whichO. europaea aquapo-rins could be recognized by each antibody (SupportingInformation). The specificity of the PIP2 and phosphorylatedantibodies PIP2280, PIP2283 and PIP2280/283 is described inCalvo-Polanco et al. (2014b).

Statistics

Data were analysed using analysis of variance (ANOVA) withthe Proc MIXED procedure in SAS (version 9.2, SAS instituteInc., NC, USA) together with the post hoc Tukey’s test todetect significant differences among all treatment means.When the ANOVA P-values for the three-way interaction werenot significant, we proceed to run t-test for the significantinteractions. The aquaporin expression and abundance data,mycorrhizal colonization rates and length, stomatal conduc-tance, together with the Lpr data were firstly sorted out usinga principal components analyses and posterior Pearson’s corre-lations to determinewhich of the aquaporins may have contrib-uted to Lpr and if gs was correlated with Lpr. Pearsoncorrelations were also calculated among Lpr and stomatalconductance and leaf N, P and K contents.

RESULTS

Soil conditions and plant growth

Grazalema and FL soil analyses showed high pH values,between 8 and 9, and low-electrical conductivity (Table 1).GZ soil was richer in nutrients (higher N, P, K and Ca)and also in organic matter, which should favour plantgrowth. Significant lower values of N, P, K and S contentswere observed in leaves of plants growing in FL soil, mostlyin non-inoculated plants (Table 2). Inoculation with AMfungi from FL soil increased P and K contents of leaves ofplants growing in FL soil (Table 2). The difference in Kcontent between leaves from plants growing in both kindsof soils was only significant under well-watered conditions(Table 2).

When plant growth was analysed using either leaf or rootFW as a parameter, the impact of the soil origin becomeclear as GZ soils supported larger growth (P< 0.0001 andP=0.044, respectively) (Table 3). However, drought stressinduced a significant reduction of the leaf FW (P=0.0008),in both non-inoculated plants and plants inoculated with

2502 M. Calvo-Polanco et al.

© 2016 John Wiley & Sons Ltd, Plant, Cell and Environment, 39, 2498–2514

Table 2. Nutrient content of leaves ofOlea europaea plants growing into two different types of soils (GZ and FL), inoculated with differentmycorrhizal communities from GZ and FL soils and cultivated under well-watered conditions or subjected to drought stress for 4weeks

Grazalema soil Freila soil

No Myc Myco GZ Myco FL No Myc Myco GZ Myco FL

N (mg g�1) Well-watered 1.09 ± 0.12 1.21 ± 0.03* 1.10 ± 0.12 0.83 ± 0.01 0.80 ± 0.03 0.94 ± 0.03Drought 0.92 ± 0.07 1.02 ± 0.02 1.03 ± 0.03 0.85 ± 0.02 0.85 ± 0.04 0.88 ± 0.03

P (mg g�1) Well-watered 1.09 ± 0.17* 2.12 ± 0.11* 1.12 ± 0.18 0.53 ± 0.01 0.87 ± 0.08 1.16 ± 0.05Drought 0.65 ± 0.12 1.23 ± 0.04 1.03 ± 0.11 0.57 ± 0.06 0.73 ± 0.08 1.01 ± 0.16

K (mg g�1) Well-watered 7.37 ± 0.68ab 7.99 ± 0.36a 8.03 ± 0.46a 6.02 ± 0.25cde 6.89 ± 0.43abc 7.86 ± 0.11aDrought 6.03 ± 0.32cde 6.28 ± 0.34bcd 5.71 ± 0.51de 5.02 ± 0.19e 4.99 ± 0.25e 5.63 ± 0.57de

Mg (mg g�1) Well-watered 1.22 ± 0.18 1.33 ± 0.19 1.14 ± 0.12 1.30 ± 0.08 1.28 ± 0.12 1.15 ± 0.05Drought 1.07 ± 0.12 1.01 ± 0.05 1.16 ± 0.03 1.36 ± 0.26 1.26 ± 0.15 1.20 ± 0.16

S (mg g�1) Well-watered 0.96 ± 0.07 1.08 ± 0.05 1.12 ± 0.18 0.87 ± 0.05 0.92 ± 0.06 0.92 ± 0.06Drought 0.81 ± 0.04 0.92 ± 0.04 1.03 ± 0.11 0.70 ± 0.02 0.92 ± 0.03 0.88 ± 0.03

Ca (mg g�1) Well-watered 10.11 ± 0.95 11.32 ± 1.48 10.95 ± 1.49 8.70 ± 1.76 10.50 ± 0.96 8.54 ± 0.25Drought 9.15 ± 1.1 9.07 ± 0.42 10.37 ± 0.18 7.75 ± 0.43 11.20 ± 0.51 10.22 ± 1.22

Soil Myc Soil ×Myc Drought Soil × drought Myc × drought Soil ×Myc × drought

N <0.0001 0.8919 0.0290 0.2833 0.0019 0.3662 0.0744P <0.0001 <0.0001 0.0098 0.0002 0.0034 0.2691 0.4605K 0.0010 0.0613 0.1147 <0.0001 0.1150 0.0019 0.0455Mg 0.1992 0.7224 0.7163 0.4618 0.2672 0.5663 0.7945S 0.0027 0.0216 0.7647 0.1420 0.1032 0.4277 0.3421Ca 0.2627 0.0951 0.2972 0.8970 0.0535 0.6630 0.9349

ANOVAs plus post hoc Tukey’s test were run when significant three-way interaction P-values allowed it (only in the case of K contents), and then dif-ferent lettersmean significant differences (P< 0.05) among treatments in the same row.When the three-way interactionwas non-significant, t-test wasused to indicate significant differences (P< 0.05) between treatment means at the same column indicated by an asterisk.ANOVA P-values for the different parameters measured: soil – soil type; Myc – non-inoculated, inoculated with AM from GZ or inoculated with AMfrom FL; and well-watered and drought stress.ANOVAs, analyses of variance; GZ, Grazalema; FL, Freila.

Table 3. Leaf and root FW, leaf RWC, photochemical efficiency ofΦPSII and chlorophyll content inOlea europaea plants growing into two differenttypes of soils (GZ and FL), inoculated with different mycorrhizal communities fromGZ and FL soils and cultivated under well-watered conditions orsubjected to drought stress for 4weeks

Grazalema soil Freila soil

No Myc MycoGZ Myco FL No Myc MycoGZ Myco FL

Leaf FW (g) Well-watered 5.6 ± 1.0* 7.1 ± 0.8 5.9 ± 0.7* 2.7 ± 0.4 3.6 ± 0.3 3.9 ± 0.6Drought 3.4 ± 0.4 5.9 ± 0.8 3.1 ± 0.8 2.6 ± 0.5 3.0 ± 0.6 2.9 ± 0.5

Root FW (g) Well-watered 4.7 ± 0.9 4.5 ± 0.3 5.0 ± 1.1 3.0 ± 0.7* 5.1 ± 0.8* 4.8 ± 0.7*Drought 3.3 ± 0.8 4.0 ± 0.5 3.8 ± 1.34 1.5 ± 0.4 3.3 ± 0.5 2.5 ± 0.8

ΦPSII (r.u) Well-watered 0.61 ± 0.01* 0.61 ± 0.01* 0.60 ± 0.01* 0.54 ± 0.10* 0.57 ± 0.05* 0.58 ± 0.06*Drought 0.16 ± 0.07 0.18 ± 0.04 0.19 ± 0.07 0.25 ± 0.02 0.30 ± 0.07 0.31 ± 0.05

RWC (%) Well-watered 94± 1* 92 ± 1* 94± 1* 94± 1* 94± 1* 91 ± 1*Drought 85 ± 1 85 ± 2 90± 1 90± 1 85± 3 86 ± 3

Chlorophyll content(mg g�1)

Well-watered 7.4 ± 1.1* 7.3 ± 0.2* 6.4 ± 0.5 5.3 ± 0.7* 6.1 ± 0.7 6.5 ± 0.3Drought 4.7 ± 0.2 5.3 ± 0.4 4.8 ± 1.6 3.8 ± 0.2 5.0 ± 0.6 6.3 ± 0.3

Soil Myc Soil ×Myc Drought Soil × drought Myc × drought Soil ×Myc× drought

Leaf FW <0.0001 0.0100 0.0630 0.0008 0.0534 0.5122 0.6709Root FW 0.0441 0.0604 0.2186 0.0010 0.3005 0.8372 0.7938ΦPSII 0.4322 0.5239 0.9134 <0.0001 0.0296 0.7099 0.8641RWC 0.7804 0.3389 0.1711 <0.0001 0.9200 0.3983 0.3298Chlorophyll 0.0346 0.0377 0.0015 <0.0001 0.0154 0.1724 0.8686

As the three-way interaction was non-significant for any parameter, t-test was used to indicate significant differences (P< 0.05) between treatmentmeans at the same column indicated by an asterisk (n= 6).ANOVA P-values for the different parameters measured: soil – soil type; Myc – non-inoculated, inoculated with AM from GZ or inoculated with AMfrom FL; and drought- Well-watered and drought stress.ANOVA, analysis of variance; GZ, Grazalema; FL, Freila; FW, fresh weight; RWC, relative water content; ΦPSII, photosystem II.

Effects of AM fungal backgrounds on olive plants 2503

© 2016 John Wiley & Sons Ltd, Plant, Cell and Environment, 39, 2498–2514

the FL-inocula growing in GZ soil (Table 3). On the otherhand, root FW was significantly reduced by drought in allplants growing in FL soil, while no effect could be detectedon GZ soil (Table 3). There was also a significant positiveeffect of the mycorrhizal inoculation in leaf FW (P=0.01;Table 3).

Drought treatment also affected relative plant height(P=0.04) (Fig. 1b), as well as the relative number of leavesproduced during the drought treatment (P< 0.0001)(Fig. 1a). Plants growing in GZ soil with the GZ inoculawere the only ones that displayed a significant increase intheir relative heights under well-watered conditions, whilethe relative height did not change in any of the othertreatments considered (Fig. 1b). For the relative numberof leaves, drought caused a significant loss of leaves in theGZ soil plants in either non-inoculated plants or thoseinoculated with FL-AM fungi (Fig. 1a). It is noteworthy that

under these conditions, inoculation with GL-AM fungiprevented such decrease.

Fungal mycelium length, root mycorrhization ratesand fungal diversity

The highest extension of external mycelia under well-wateredconditions was found in plants growing in FL soil and inocu-lated with AM fungi from GZ (Fig. 2a, P< 0.0001). Whendrought was applied, the highest external fungal developmentwas found in FL soils with FL AM fungi. In GZ soils, hyphaewithin the soil were longer in the presence of GZ fungi, despitethe water regime (Fig. 2a). However, the fungal rootcolonization within the roots only varied when the plants weresubjected to drought stress and the FL inoculum was applied(Fig. 2b).

Figure 1. Relative plant height (a), relative number of leaves (b) and ANOVA P-values table in Olea europaea cv. Picual plants growing into twodifferent types of soils (Grazalema – GZ and Freila – FL) and inoculated with different mycorrhizal communities from GZ and FL soils. The plantswere either cultivated under well-watered conditions or subjected to drought stress for 4 weeks. Asterisks above bars indicate significant differences(P< 0.05) between well-watered and drought treatments after t-student test (n= 6).

2504 M. Calvo-Polanco et al.

© 2016 John Wiley & Sons Ltd, Plant, Cell and Environment, 39, 2498–2514

A total of nine OTUs in the roots of all analysedsamples were found (Fig. 2c). Only one of them, OTU1,likely affiliated to an undescribed species from the genusDominikia, was found in all treatments. In addition, thisOTU1 resulted to be dominant in all samples except inroots grown on GZ soil with the GZ inoculum and culti-vated under well-watered conditions. Certain OTUs weredetected exclusively in one of the root treatments relatedto a specific water regime or soil-inoculum association.Maximum AM fungal diversity values corresponded towell-watered treatments and if AM fungal inoculum wasapplied to its original soil, while the lowest diversity (0)was found in roots involving FL inoculum on GZ soilunder well-watered conditions or in FL soil under droughtconditions (Fig. 2c).

Leaf stomatal conductance, relative water content,photochemical efficiency of photosystem II,chlorophyll content and root hydraulic conductivity

Drought treatment caused a masive reduction of gs in alltreatments except in non-inoculated plants growing in FL soil,as their initial values were already very low (Fig. 3a). In FL soil,AM inoculation significantly increased gs under well-wateredconditions, especially with the GZ inoculum (Fig. 3a). Thismassive reduction of gs by drought could indicate a strongdrought stress.

Leaf RWC and ΦPSII were reduced by the drought treat-ment (P< 0.0001, Table 3), with no effect of the differentinoculation treatments or soil (Table 3). Leaf chlorophyllcontents were significantly reduced by drought in GZ soils in

Figure 2. Soil hyphal length (a), percentage of mycorrhizal root length (b) and root mycorrhizal diversity (c) in Olea europaea cv. Picual plantsgrowing into two different types of soils (Grazalema –GZ and Freila – FL), and inoculated with different mycorrhizal communities fromGZ and FLsoils. The plants were either cultivated under well-watered conditions or subjected to drought stress for 4weeks Different letters means significantdifferences (P< 0.05) among treatments after ANOVA and Tuckey’s test (n= 4). The numbers above the columns in panel C are the Shannonbiodiversity index and the numbers under brackets are the specific richness.

Effects of AM fungal backgrounds on olive plants 2505

© 2016 John Wiley & Sons Ltd, Plant, Cell and Environment, 39, 2498–2514

non-inoculated plants or those inoculated with GZ-AM fungi,as well as in non-inoculated plants growing in FL soil(Table 3). Under drought condition, AM inoculation (withboth GZ and FL) significantly increased chlorophyll contentsof plants growing in FL soil (Table 3).

The trends observed in gs were not the same as the trendsfound in root hydraulic conductivity (Lpr) (Fig. 3), in fact asignificant negative correlation was found between gs and Lpr(R=�0.320 P=0.0249). Our study showed that the plantshaving higher Lpr values were found in FL soil, with asignificant increase under drought in plants inoculated withGZ-AM fungi (Fig. 3b). However, in FL soil both AMinocula reducedLpr under well-watered conditions. InGZ soil,we did not observed any effects on Lpr by any treatment(Fig. 3b)

Molecular identification, expression andabundance of olive aquaporins

From the molecular approach used, nine new olive aquaporinswere described: OePIP1;2, OePIP1;3, OePIP2;2, OePIP2;3,OePIP2;4, OePIP2;5, OePIP2;6, OeTIP1;2 and OeTIP1;3. Aphylogenetic tree was built with the currently known PIPsand TIPs aquaporins (OePIP1;1, OePIP2;1 and OeTIP1;1)from O. europaea and Poplar (Fig. 4).

We next proceeded to analyse expression in the roots of the12 aquaporins (the three already described and the nine newones). RNA accumulation could not be detected for the previ-ously known aquaporins OePIP1;1 and OePIP2;1 using qRT-PCR analyses. To analyse aquaporin expression, we firstlyran an examination of principal components to elucidate which

Figure 3. Stomatal conductance (A) and root hydraulic conductivity (B) inOlea europaea cv. Picual plants growing into two different types of soils(Grazalema-GZand Freila-FL), and inoculatedwith differentmycorrhizal communities fromGZandFL soils. The plants were either cultivated underwell-watered conditions or subjected to drought stress for fourweeks. Different letters means significant differences (P< 0.05) among treatments afterANOVA and Tukey’s test (n= 6).

2506 M. Calvo-Polanco et al.

© 2016 John Wiley & Sons Ltd, Plant, Cell and Environment, 39, 2498–2514

aquaporins weremore relatedwith theLpr and contributed themost to water transport (Fig. 5a). We found that expression ofOePIP1;2, OePIP2;2, OePIP2;3, OePIP2;5, OePIP2;6 andOeTIP1;1 were related to whole Lpr, while the expression ofaquaporins OePIP1;3, OePIP2;4, OeTIP1;2 and OeTIP1;3 ex-plained the higher sources of variation within the data. At thesame time, soil mycelial hyphal length and proportion of rootlength colonized also are closely related to Lpr. Stomatal con-ductance (gs) also explained the variation observed inLpr data.We ran Pearson correlations with the different aquaporins andLpr, and found that Lpr was correlated positively withOePIP1;2 and OeTIP1;2; and negatively correlated withOePIP1;3, OePIP2;4 and OeTIP1;3 (Table 3).Analysis of variance analyses were run to detect significant

differences between treatment means at the different treat-ments considered. The P-values showed, as in previous studies(Barzana et al. 2014), that the expression of all the aquaporins

did not follow the same trend at the different treatmentsconsidered (Table 4 – ANOVA P-values, and SupportingInformation). However, plants growing in FL soil showed apositive increase in OeTIP1;3 mRNA accumulation underdrought conditions (Fig. 5b).

Protein analyses

The PIP1 proteins were undetected with the methodologyapplied. Amounts of the PIP2 proteins, as well as the differentphosphorylated proteins changed as with the different treat-ments. The ANOVA P-values showed significant differencesbetween treatmentmeans for all the proteins studied (Table 4).We also ran Pearson’s correlations to test possible linksbetween our Lpr data and the PIP2 protein abundance andphosphorylation state. It was found that a general positive

Figure 4. Phylogenetic tree of the PIP and TIP aquaporins from Populus trichocarpa (Gupta and Sankararamakrishnan 2009) and Olea europaea(Secchi et al. 2007; present study) species. The aquaporins of the two plant species were analysed with MEGA4 software using neighbour-joiningmethod. The distance scale represents evolutionary distance, expressed in the number of substitutions per amino acid.

Effects of AM fungal backgrounds on olive plants 2507

© 2016 John Wiley & Sons Ltd, Plant, Cell and Environment, 39, 2498–2514

Figure 5. Principal component analyses (A) for the expression of 10 aquaporin genes, PIP1 and PIP2 abundances, PIP2 phoshorylation state, roothydraulic conductivity (Lpr), stomatal conductance (gs), root colonization rate (MycP) and soil hyphal lengh (MycL). Relative mRNA expression ofOeTIP1;3 (B) in roots ofOlea europaea cv. Picual plants growing into two different types of soils (Grazalema-GZ and Freila-FL), and inoculated withdifferent mycorrhizal communities from GZ and FL soils. The plants were either cultivated under well-watered conditions or subjected to droughtstress for four weeks Asterisks above bars indicate significant differences (P< 0.05) between well-watered and drought treatments after t-student test(n = 3).

Table 4. ANOVA P-values for expression of 10Olea europaea aquaporins and the PIP2 root protein abundance and phosphorylation state (PIP2280 –proteins phosphorylated at Ser-280, PIP2283 – proteins phosphorylated at Ser-283 and PIP2280–283 – proteins phosphorylated at both Ser-280 andSer-283)

Soil Myc Soil ×Myc Drought Soil × drought Myc × drought Soil ×Myc × drought

OePIP1;2 0.6948 0.6960 0.2279 0.0099* 0.7861 0.2295 0.5441OePIP1;3 0.0247* 0.0028* 0.0002* 0.1258 0.1597 0.0746 0.0482*OePIP2;2 0.9796 0.0077* 0.0026* <0.0001* 0.4775 0.0075* 0.1077OePIP2;3 0.0369* 0.1614 0.4516 0.5105 0.9305 0.1303 0.0443*OePIP2;4 0.5710 0.0143* 0.4131 0.0005* 0.9357 0.1137 0.0606OePIP2;5 0.0243* 0.1782 0.0133* 0.7510 0.0898 0.1810 0.1267OePIP2;6 0.6627 0.6126 0.3996 0.3079 0.0066* 0.0271* 0.0167*OeTIP1;1 0.5439 0.6830 0.4257 0.6290 0.8697 0.5195 0.6225OeTIP1;2 0.9458 0.6177 0.0060* 0.0037* 0.8828 0.8179 0.0203*OeTIP1;3 0.9647 0.1395 0.0536 <0.0001* 0.0015* 0.3041 0.5517PIP2 0.0358 0.5679 0.0453 0.9705 0.0002* 0.7366 0.0061*PIP2280 0.0007* <0.0001* 0.0104* <0.0001* 0.0271* <0.0001* <0.0001*PIP2283 0.6411 0.1466 0.0887 0.5913 0.0380* 0.5233 0.0162*PIP2280/283 0.6581 0.0067* 0.0112* <0.0001* 0.0061* <0.0001* <0.0001*

2508 M. Calvo-Polanco et al.

© 2016 John Wiley & Sons Ltd, Plant, Cell and Environment, 39, 2498–2514

correlation between Lpr and the abundance of PIP2 proteins,but a negative correlation with the abundance of phosphory-lated PIP2Ser280 and PIP2 Ser280–283 proteins (Table 5). How-ever, at the treatment level, the higher values of Lpr indrought plants in FL soil inoculated with the native mycorrhi-zas from GZ was not correlated with the abundance of PIP2proteins (Fig. 6a) or the phosphorylation of these proteins atPIP2Ser280 (Fig. 6b), but with the increase on the abundanceof phosphorylated proteins PIP2Ser283 and PIP2Ser280–283(Fig. 6c,d).

DISCUSSION

Olive plants were generated from explants of mature olivetrees growing in a typically Mediterranean area. The plantswere grown for 7months into two contrasting Mediterraneansoils and were inoculated with fungal communities obtainedfrom those soils. In addition, plants were subjected to well-watered and drought stress conditions for 4weeks. Underwell-watered conditions, the plants growing on GZ soils andinoculated with the GZ fungal community were the onlyones which increased their relative height during the 4weeksof the stress period. It is well known that AM fungi canexhibit a considerable level of selectivity in their associationwith different plants species or plant ecological groups(Öpik et al. 2009, Varela-Cevero et al. 2015). However, thesestudies have not related the AM associations with thephysical and chemical characteristics of the soil, only withthe presence of certain AM fungi species. In our study, thehigher relative heights of the plants were obtained with fungifrom GZ and soils also from GZ, which have the highestcontent in organic matter, K, P and N. It has been previouslyfound that soil chemistry influences soil microbial communitycomposition, diversity and activity (Ehrenfeld et al. 2005).Greater soil fertility increases bacterial biomass and activityin grassland mesocosms, and these effects interact with plantspecies identity in some systems (Innes et al. 2004). Thus,effects of soil chemistry are often system-dependent.Whether or how these soil chemistry effects may interactwith plant species identity or plant relatedness is notgenerally known. In our study, it seems to be crucial forthe development of the AM fungi within the roots (thatwas similar in both well-watered and drought stressedplants), as well as, with the development of the externalmycelia within the soil. This combination had also the higherdiversity in the OTUs found within the roots under well-watered conditions.

It is also remarkable that plants growing in FL soil withoutinoculation had the lowest leaf FW, root FW and chlorophyllcontent, but these values were increased by AM inoculation.These results confirm that under some environmental condi-tionsAM fungi are essential for plant growth and development(Estrada et al. 2013). Moreover, these plants, even under well-watered conditions, had gs values similar to plants underdrought stress conditions. The cause could be the lower valuesof leaf P and K content. It is known that nutrient deficiency,including P and K, causes reduction of stomatal conductance(Flores et al. 2015). In fact, we found a positive correlationT

able

5.Pearson

correlations

betw

eenroot

hydrau

liccond

uctiv

ity(L

p r)an

dthedifferen

taqu

aporin

expression

determ

ined

aswella

sthePIP2root

proteinab

unda

ncean

dph

osph

orylationstate

(PIP2 2

80–proteins

phosph

orylated

atSe

r-280,PIP2 2

83–proteins

phosph

orylated

atSe

r-283an

dPIP2 2

80–283–proteins

phosph

orylated

atbo

thSe

r-280an

dSe

r-283)

OeP

IP1;2

OeP

IP1;3

OeP

IP2;2

OeP

IP23

OeP

IP2;4

OeP

IP2;5

OeP

IP2;6

OeT

IP1;1

OeT

IP1;2

OeT

IP1;3

PIP2

PIP2 2

80PIP2

283

PIP2 2

80�283

Lp r

0.357

�0.378

0.208

�0.099

�0.328

�0.219

0.013

�0.153

0.492

�0.401

0.399

�0.524

0.084

�0.509

0.0219*

0.0125*

0.1860

0.5342

0.0417*

0.1869

0.9418

0.3603

0.0013*

0.0070*

0.0089*

0.0002*

0.5831

0.0004*

Effects of AM fungal backgrounds on olive plants 2509

© 2016 John Wiley & Sons Ltd, Plant, Cell and Environment, 39, 2498–2514

between leaf P and K contents and gs (P< 0.05). Furthermore,no significant correlation was found between any leaf nutrientcontent and Lpr, except a negative correlation (P> 0.05) withK leaf contents, because plants fromFL soil had lessK contentsbut higher Lpr values.

After 4weeks of drought treatment, there was a generalreduction in growth and gs in all treatments. These are typicalresponses of plants exposed to severe drought stress. However,there were some responses of plants that reinforce the previousidea of fungal and soil chemistry action on roots.We found thatplants lost less leaves when growing in theGZ soils in combina-tion with GZ-inocula, in agreement with the previous idea ofthe advantages of certain communities in soils of a certainchemistry and structure (Burns et al. 2015). Olive trees areknown to have a tight control of gs under different wateringregimes (Torres-Ruiz et al. 2013), as we found under droughtstress conditions regardless of the soil or fungal communityapplied. This control of gs did not hamper the differentresponses in Lpr, as these plants seem to have a suitablehydraulic efficiency to yield water potentials that maintainsthe photosynthetic apparatus hydrated under different waterdemands (Raimondo et al. 2009). An increase of Lpr underdrought conditions was only found in the FL soil, and in plantsinoculated with the GZ inocula, being this time the dominantfungi present in the roots the OTU1 (Dominikia sp.) in

conjunction with OTU5 (Funneliformis sp.). Furthermore,plants growing on FL soil had a higher OeTIP1;3 expressionand higher phosphorylated PIP2Ser280 and PIP2Ser280/283. Plantand fungal aquaporin expression, abundance and phosphoryla-tion state can play a major role in root water transport(El-Mesbahi et al. 2012). Plant aquaporins are usuallyresponsible for the majority of radial root water transportunder severe drought conditions (Barzana et al. 2012). Thereare many studies trying to understand how the presence ofAM fungi regulates plant aquaporins (Barzana et al. 2014). Itis known that AM symbiosis results in altered rates of watertransfer in and out of the host plants (Auge 2001), and that alsomodifies Lpr (Barzana et al. 2014, Calvo-Polanco et al. 2014a,Sanchez-Romera et al. 2016). Aquaporins provide a low-resistance pathway for the movement of water across mem-branes. Furthermore, as aquaporins can be gated, this providesgreater control for the movement of water along plant tissues(Nyblom et al. 2009, Maurel et al. 2015). PIP and TIP isoformshave been recognized as central pathways for transcellular andintracellular water transport (Maurel et al. 2015). Thus, itseems likely that mycorrhizal symbiosis causes significantchanges in aquaporin activity of host plants by changes in theirphosphorylation status (Calvo-Polanco et al. 2014a, Sanchez-Romera et al. 2016). On the other hand, it has also been sug-gested that water could be absorbed by the external AM

Figure 6. Quantification of PIP2 proteins (A), PIP2 phosphorylated proteins at Ser-280 (B), Ser-283 (C) and both Ser-280 and 283 (D) in roots ofOlea europaea cv. Picual plants growing into two different types of soils (Grazalema-GZ and Freila-FL), and inoculated with different mycorrhizalcommunities from GZ and FL soils. The plants were either cultivated under well-watered conditions or subjected to drought stress for four weeksDifferent letters means significant differences (P< 0.05) among treatments after ANOVA and Tukey’s test (n= 3).

2510 M. Calvo-Polanco et al.

© 2016 John Wiley & Sons Ltd, Plant, Cell and Environment, 39, 2498–2514

mycelium and delivered to the cortical apoplast, at the symbi-otic interfaces, where it would join water taken up via the rootapoplastic pathway (Smith et al. 2010; Barzana et al. 2012). Hy-phal water uptake and transfer to the host plants has been dem-onstrated in several studies (Marulanda et al. 2003; Khalvatiet al. 2005; Ruth et al. 2011). The increased water uptake by hy-phae will be critical when soil dries, and water is retained in thesmaller pores where fungal hyphae can grow (Marulanda et al.2003), although the role of this hyphae in relation to the inter-nal fungal development is not very well known yet under se-vere drought stress. However, a close relation between Lprand mycorrhizal hyphal growth in both soil and roots was ob-served. Aroca et al. (2009) observed that the expression ofone AM fungal aquaporin gene was higher in anti-sense to-bacco plants with lower expression of tobacco aquaporins thanin wild-type plants. This compensatory mechanism could alsoexplain the negative correlation found here between Lpr andthe expression and phosphorylation state of some olive aqua-porins, where the AM fungal aquaporins could be the mainpathway for water uptake.Most recently, Tataranni et al. (2015) found that Lpr in

olive trees subjected to drought stress was close correlatedwith the amount of suberin and root-cell density. It is knowthat AM fungi may increase lignin contents of colonized roots,especially in endodermis cells (Dehne and Schonbeck 1979).Furthermore, most recently, Almeida-Rodríguez et al. (2016)found that AM symbiosis modifies xylem vessels diameter inSalix purpurea plants, modifying also Lpr values under copperstress conditions. Anatomical changes caused by AM symbio-sis in olives plants cannot be ruled out, and they deserve fur-ther investigations. At the same time, gs explain much of thevariation observed in Lpr, and a significant (P< 0.05) nega-tive correlation between gs and Lpr was found. Such negativecorrelation could be indicative of a preferential watertransport through the cell-to-cell path in olive trees, becausethis path is not dependent on leaf transpiration (Steudle &Peterson 1998).The contribution of the different aquaporins studied was ad-

dressed with their mRNA expression and protein abundance.The combined expression of different PIP1, PIP2 and TIP1aquaporins (Fig. 5a) gave as a combined Lpr response of theplant to the different treatments, and each of them respondingdifferently to the several factors that were affecting the plants(Table 4). Different aquaporins are either up-regulated ordown-regulated by the same stress and those in the same sub-group may have distinct expression patterns under differentstress conditions (Guo et al. 2006; Barzana et al. 2014). Thecontribution of all these aquaporins toLpr allowed a fine regu-lation of water uptake under theGZ soils (Fig. 3b) and contrib-uted to the increase of Lpr under drought conditions in FL soilwith GZ inocula. Among all studied aquaporins, the one thatclearly increased under drought stress was OeTIP1;3. The in-terest on the role of TIPS in water transport has increased inthe last few years. Wang et al. (2014) showed that the expres-sion of TsTIP1;2 increased in response to various stresses inThellungiella salsuginea, and ectopic over expression ofTsTIP1;2 enhanced tolerance to drought, salt and oxidativestresses in transgenic Arabidopsis. Most interesting are the

results of Henry et al. (2012) who found a good correlation be-tween TIP expression and Lpr in rice roots subjected todrought stress and byKuwagata et al. (2012) in rice roots underdifferent air humidity conditions. Furthermore, Boursiac et al.(2005) found that the decrease of Lpr caused by salt stress inArabidopsis was correlated with a down regulation of PIPand TIP expression as well to a relocalization to internalmembranes. On the other hand, the effect of the phosphoryla-tion of PIP2 proteins at Ser283 residue may have contributed tothe increase of water uptake in olive plants. Aquaporinphosphorylation is considered as one of the major processesaffecting water transport (Suga and Maeshima 2004) and mayhave also contribute to the adaptation of plant to differentstresses, including drought.

Our results show negative correlations between the expres-sion and abundance of some aquaporins and Lpr. Similarresults were found by Sutka et al. (2011) analysing the naturalvariation in Lpr of different accessions of Arabidopsis. Mostrecently, Li et al. (2016) also found a negative correlationbetween Lpr and the gene expression of some ArabidopsisPIPs (AtPIP1;4 and AtPIP2;6). It is known that aquaporinscan interact each other with positive or negative results interms of membrane water permeability, mostly becauseinternalization processes (Bellati et al. 2010; Chaumont andTyerman 2014). Recently, Hachez et al. (2014) found thatthe plasma membrane amount of ZmPIP2;7 proteins is regu-lated by autophagic degradation when interact withtryptophan-rich sensory protein/translocator. Furthermore, ithas been shown that aquaporin mRNA expression does notalways matches aquaporin protein abundance (Marulandaet al. 2010).

In conclusion, we have confirmed that the different charac-teristics of the soil affect the development of plants and theirresponses to drought stress. Furthermore, we have establishedthat the fitness among AM fungal communities, plant speciesand soil origin is crucial for the development of a properplant-fungus symbiosis. In our case, the inocula from GZ soilshad a higher impact in olive trees under well-wateredconditions and also these inocula had a higher impact in FLsoils under drought treatment. So, the AM fungi originatedfrom a wet climate had a better performance with olive plantsunder all conditions than the fungi coming from a more aridenvironment. The presence of these fungi may have beencritical to modulate plant aquaporin expression, phosphoryla-tion and abundance and hence, for root water transport. Therole of OeTIP1;3 under water stress may not be related withthe bulk transport of water in plants but may play a crucial rolein root osmoregulation of plants, being part of the fineregulated system of plants water balance.

ACKNOWLEDGMENTS

We would like to thank Sonia Molina, Jose Luis Manella andOlga M. López for their technical support. The study wassupported by the Ministry of Economy and Competitivenessof Spain (Juan de la Cierva Program) and Junta de Andalucia(P10-CVI-5920 project) for research funding.

Effects of AM fungal backgrounds on olive plants 2511

© 2016 John Wiley & Sons Ltd, Plant, Cell and Environment, 39, 2498–2514

REFERENCES

Abbaspour H., Saeidi-Sar S., Afshari H. & Abdel-Wahhab M.A. (2012)Tolerance of mycorrhiza infected Pistachio (Pistacia vera L.) seedling todrought stress under glasshouse conditions. Journal of Plant Physiology 169,704–709.

Abbott L.K., Robson A.D. & De Boer G. (1984) The effect of phosphorous onthe formation of hyphae in soil by the vesicular-arbuscular mycorrhizal fungus,Glomus faciculatum. New Phytologist 97, 437–446.

Al-Karaki G.N. (2006) Nursery inoculation of tomato with arbuscular mycorrhi-zal fungi and subsequent performance under irrigation with saline water.Scientia Horticulturae 109, 1–7.

Almeida-RodríguezA.M., GómesM.P., Loubert-HudonA., Joly S. & LabrecqueM. (2016) Symbiotic association between Salix purpurea L. and Rhizophagusirregularis: modulation of plant responses under copper stress. Tree Physiology36, 407–420.

Aroca R., Bago A., Sutka M., Paz J.A., Cano C., Amodeo G. & Ruiz-Lozano J.M. (2009) Expression analysis of the first arbuscularmycorrhizal fungi aquapo-rin described reveals concerted gene expression between salt-stressed andnonstressed mycelium.Molecular Plant-Microbe Interaction 22, 1169–1178.

Aroca R., Porcel R. & Ruiz-Lozano J.M. (2007) How does arbuscular mycorrhi-zal symbiosis regulate root hydraulic properties and plasma membraneaquaporins in Phaseolus vulgaris under drought, cold or salinity stresses?New Phytologist 173, 808–816.

Aroca R., Porcel R. & Ruiz-Lozano J.M. (2012) Regulation of root water uptakeunder abiotic stress conditions. Journal of Experimental Botany 63, 43–57.

Auge R.M. (2001)Water relations, drought and vesicular-arbuscular mycorrhizalsymbiosis.Mycorrhiza 11, 3–42.

Azcon R., Gomez M. & Tobar R. (1996) Physiological and nutritional responsesby Lactuca Sativa L. to nitrogen sources and mycorrhizal fungi under droughtconditions. Biology and Fertility of Soil 22, 156–161.

Barzana G., Aroca R., Paz J.A., Chaumont F., Martinez-Ballesta M., Carvajal M.& Ruiz-Lozano J.M. (2012) Arbuscular mycorrhizal symbiosis increasesrelative apoplastic water flow in roots of the host plant under both well-watered and drought stress conditions. Annals of Botany 109, 1009–1017.

Barzana G., Aroca R., Bienert G.P., Chaumont F. & Ruiz-Lozano J.M. (2014)New insights into the regulation of aquaporins by the arbuscular mycorrhizalsymbiosis in maize plants under drought stress and possible implications forplant performance.Molecular Plant-Microbe Interaction 27, 349–363.

Bellati J., Alleva K., Soto G., Vitali V., Jozefkowicz C. & Amodeo G. (2010)Intracellular pH sensing is altered by plasma membrane PIP aquaporinco-expression. Plant Molecular Biology 74, 105–118.

BenabdellahK., Ruiz-Lozano J.M.&ArocaR. (2009)Hydrogen peroxide effectson root hydraulic properties and plasma membrane aquaporin regulation inPhaseolus vulgaris. Plant Molecular Biology 70, 647–661.

Binet M.N., Lemoine M.C., Martin C., Chambon C. & Gianinazzi S. (2007)Micropropagation of olive (Olea europaea L.) and application of mycorrhizato improve plantlet establishment. In Vitro Cellular & DevelopmentalBiology-Plant 43, 473–478.

Bonfante P. & Genre A. (2010) Mechanisms underlying beneficial plant-fungusinteractions in mycorrhizal simbiosis. Nature Communications 1, 48.

BompadreM.J., Rios deMolinaM.C., ColomboR.P., Bidondo L.F., Silvani V.A.,PardoA.G., Ocampo J.A.&GodeasA.M. (2013)Differential efficiency of twostrains of the arbuscular mycorrhizal fungus Rhizophagusirregularis on olive(Olea europaea) plants under two water regimes. Symbiosis 61, 105–112.

Boursiac Y., Chen S., LuuD.-T., SorieulM., van denDries N. &Maurel C. (2005)Early effects of salinity on water transport inArabidopsis roots. Molecular andcellular features of aquaporin expression. Plant Physiology 139, 790–805.

Burns J.H., Anacker B.L., Strauss S.Y. & Burke D.J. (2015) Soil microbialcommunity variation correlates most strongly with plant species identity,followed by soil chemistry, spatial location and plant genus. AoB Plants 7,plv030.

Calvo-Polanco M., Molina S., Zamarreno A.M., Garcia-Mina J.M. & Aroca R.(2014a) The symbiosis with the arbuscular mycorrhizal fungus Rhizophagusirregularis drives root water transport in flooded tomato plants. Plant and CellPhysiology 55, 1017–1029.

Calvo-Polanco M., Sanchez-Romera B. & Aroca R. (2014b) Mild salt stressconditions induce different responses in root hydraulic conductivity ofPhaseolus vulgaris over-time. PLoS One 9, e90631.

Calvo-Polanco M., Señorans J. & Zwiazek J.J. (2012) Role of adventitious rootsin water relations of tamarack (Larix laricina) seedlings exposed to flooding.BMC Plant Biology 12, 99.

CarrM.K.V. (2013)Thewater relations and irrigation requirements of olive (Oleaeuropaea L.): a review. Experimental Agriculture 49, 597–639.

Chartzoulakis K., Bosabalidis A., Patakas A. & Vemmos S. (2000) Effects ofwater stress on water relations, gas exchange and leaf structure of olive tree.Acta Horticulturae 537, 241–247.

Chaumont F. & Tyerman S.D. (2014) Aquaporins: Highly regulated channelscontrolling plant water relations. Plant Physiology 164, 1600–1618.

Connor D.J. (2005) Adaptation of olive (Oleaeuropea L.) to water-limitedenvironment. Australian Journal of Agricultural Research 56, 1181–1189.

Dehne H.W. & Schonbeck F. (1979) Influence of endotrophic mycorrhiza onplant diseases. II. Phenol metabolism and lignifications. Journal ofPhytophatology 95, 210–216.

El-Mesbahi M.N., Azcon R., Ruiz-Lozano J.M. & Aroca R. (2012) Plantpotassium content modifies the effects of arbuscular mycorrhizal symbiosison root hydraulic properties in maize plants. Mycorrhiza 22, 555–564.

Ehrenfeld J.G., Ravit B.&ElgersmaK. (2005) Feedback in the plant–soil system.Annual Review of Environmental Resources 30, 75–115.

Ennajeh M., Tounekti T., Vadel A.M., Khemira H. & Cochard H. (2008) Waterrelations and drought-induced embolism in olive (Olea europaea) varieties‘Meski’ and ‘Chemlali’ during severe drought. Tree Physiology 28, 971–976.

Epstein E. (1972) Mineral Nutrition of Plants: Principles and Perspectives. JohnWiley and Sons, Inc., New York.

Estrada B., Aroca R., Azcón-Aguilar C., Barea J.M. & Ruiz-Lozano J.M. (2013)Importance of native arbuscular mycorrhizal inoculation in the halophyteAsteriscus maritimus for successful establishment and growth under salineconditions. Plant and Soil 370, 175–185.

Farzaneh M., Vierheilig H., Lössl A. & Kaul H.P. (2011) Arbuscular mycorrhizaenhances nutrient uptake in chickpea.Plant, Soil andEnvironment 57, 465–470.

Flores R.A., Borges B.M.M.N., Almeida H.J. & Prado R.D.M. (2015) Growthand nutritional disorders of eggplant cultivated in nutrients solutions withsuppressed macronutrients. Journal of Plant Nutrition 38, 1097–1109.

Giovannetti M. & Mosse B. (1980) Evaluation of techniques for measuringvesicular arbuscular mycorrhizal infection in roots. New Phytologist 84,489–500.

Gollotte A., van Tuinen D. & Atkinson D. (2004) Diversity of arbuscularmycorrhizal fungi colonising roots of the grass species Agrostis capilaris andLolium perenne in a field experiment. Mycorrhiza 14, 111–117.

Gómez-del-Campo M. (2007) Effect of water supply on leaf area development,stomatal activity, transpiration, and dry matter production and distribution inyoung olive trees. Australian Journal of Agricultural Research 58, 385–391.

Guerfel M., Baccouri O., BoujnahD., ChaïbiW. &ZarroukM. (2009) Impacts ofwater stress on gas exchange, water relations, chlorophyll content and leafstructure in the two main Tunisian olive (Olea europaea L.) cultivars. ScientiaHorticulturae 119, 257–263.

Guo L., Wang Z.Y., Lin H., Cui W.E., Chen J., Liu M.H., … Gu H.Y. (2006)Expression and functional analysis of the rice plasma-membrane intrinsicprotein gene family. Cell Research 16, 277–286.

Gupta A.B. & Sankararamakrishnan R. (2009) Genome-wide analysis of majorintrinsic proteins in the tree plant Populus trichocarpa: characterization ofXIP subfamily of aquaporins from evolutionary perspective. BMC PlantBiology 9, 134.

Hachez C., Veljanovski V., Reinhardt H., et al (2014) The Arabidopsis abioticstress-induced TSPO-related protein reduces cell-surface expression of theaquaporin PIP2;7 through protein-protein interactions and autophagicdegradation. The Plant Cell 26, 4974–4990.

Hachez C., Moshelion M., Zelazny E., Cavez D. & Chaumont F. (2006)Localization and quantification of plasma membrane aquaporin expression inmaize primary root: a clue to understanding their role as cellular plumbers.Plant Molecular Biology 62, 305–323.

Henry A., Cal A.J., Batoto T.C., Torres R.O. & Serraj J. (2012) Root attributesaffecting water uptake of rice (Oryza sativa) under drought. Journal ofExperimental Botany 63, 4751–4763.

Innes L., Hobbs P.J. & Bardgett R.D. (2004) The impacts of individual plantspecies on rhizosphere microbial communities in soils of different fertilities.Biology and Fertility of Soils 40, 7–13.

Jahromi F., Aroca R., Porcel R. & Ruiz-Lozano J.M. (2008) Influence of salinityon the in vitro development of Glomus intraradices and on the in vivophysiological and molecular responses of mycorrhizal lettuce plants.MicrobialEcology 55, 45–53.

Jang J.Y., Lee S.H., Rhee J.Y., Chung G.C., Ahn S.J. & Kang H. (2007)Transgenic Arabidopsis and tobacco plants overexpressing an aquaporinrespond differently to various abiotic stresses. Plant Molecular Biology 64,621–632.

Kay R., Chan A., Daly M. & McPherson J. (1987) Duplication of CaMV-35Spromoter sequences creates a strong enhancer for plant genes. Science 236,1299–1302.

2512 M. Calvo-Polanco et al.

© 2016 John Wiley & Sons Ltd, Plant, Cell and Environment, 39, 2498–2514

Khalvati M.A., Hu Y., Mozafar A. & Schmidhalter U. (2005) Quantification ofwater uptake by arbuscular mycorrhizal hyphae and its significance for leafgrowth, water relations, and gas exchange og barley subjected to drought stress.Plant Biology 7, 706–712.

Kuwagata T., Ishikawa-Sakurai J., Hayashi H., Nagasuga K., Fukushi K.,Ahamed A., … Murai-Hatano M. (2012) Influence of low-air humidity andand low-root temperature on water uptake, growth and aquaporin expressionin rice plants. Plant and Cell Physiology 53, 1418–1431.

Lee B.-R.,Muneer S., Avice J.-C., JinW. &KimT.-H. (2012)Mycorrhizal coloni-zation and P supplement effects on N uptake and N assimilation in perennialryegrass under well-watered and drought-stressed conditions. Mycorrhiza 22,525–534.

Li G., Tillard P., Gojon A. & Maurel C. (2016) Dual regulation of root hydraulicconductivity and plasma membrane aquaporins by plant nitrate accumulationand high-affinity nitrate transporter NRT2.1. Plant and Cell Physiology 57,733–742.

Lichtenthaler H.K. (1987) Chlorophyll and carotenoids: pigments ofphotosynthetic biomembranes. In Methods in Enzymology. Plant CellMembranes (eds Packer L. & Douce R.), pp. 350–382. Academic Press,New York.

Livak K.J. & Schmittgen T.D. (2001) Analysis of relative gene expression datausing real-time quantitative PCR and the 2(T)(-ΔΔ Ct) method. Methods 25,402–408.

Lovisolo C., Secchi F., Nardini A., Salleo S., Buffa R. & Schubert A. (2007)Expression of PIP1 and PIP2 aquaporins is enhanced in olive dwarf genotypesand is related to root and leaf hydraulic conductance. Physiologia Plantarum130, 543–551.

Luu D.T. &Maurel C. (2005) Aquaporins in a challenging environment: molecu-lar gears for adjusting plant water status. Plant, Cell and Environment 28,85–96.

Marulanda A., Azcón R., Chaumont F., Ruiz-Lozano J.M. & Aroca R. (2010)Regulation of plasma membrane aquaporins by inoculation with Bacillusmegaterium strain in maize (Zea mays L.) plants under unstressed and salt-stressed conditions. Planta 232, 533–543.

Marulanda A., Azcón R. & Ruiz-Lozano J.M. (2003) Contribution of sixarbuscular mycorrhizal fungal isolates to water uptake byLactuca sativa plantsunder drought stress. Physiologia Plantarum 119, 526–533.

Maurel C., Boursiac Y., Luu D.T., Santoni V., Shahzad Z. & Verdouq L. (2015)Aquaporins in plants. Physiological Reviews 95, 1321–1358.

Maurel C., Kado R.T., Guern J. & Chrispeels M.J. (1995) Phosphorylationregulates the water channel activity of the seed-specific aquaporin α-TIP.EMBO Journal 14, 3028–3035.

Maurel C., Verdoucq L., Luu D.T. & Santoni V. (2008) Plant aquaporins:membrane channels withmultiple integrated functions.Annual ReviewofPlantBiology 59, 595–624.

Mekahlia M.N., Beddiar A. & Chenchouni H. (2013)Mycorrhizal dependency inthe olive tree (Olea europea) across a xeric climatic gradient. Advances inEnvironmental Biology 7, 2166–2174.

Muñoz-Mérida A., González-Plaza J.J., Cañada A., Blanco A.M., García-LópezM.D., Rodríguez J.M., … Beuzon C.R. (2013) De novo assembly andfunctional annotation of the olive (Olea europea) transcriptiome. DNAResearch 20, 93–18.

NyblomM., FrickA.,WangY., EkvallM., HallgreenK., Hedfalk K.,…Tornroth-Hoersefield S. (2009) Structural and functional analysis of SoPIP2;1 mutantsadds insight into plant aquaporin gating. Journal of Molecular Biology 387,653–668.

Navarro-Fernández C.M., Aroca R. & Barea J.M. (2011) Influence of arbuscularmycorrhizal fungi and water regime on the development of endemic Thymusspecies in dolomitic soils. Applied Soil Ecology 48, 31–37.

Öpik M., Metsis M., Daniell T.J., Zobel M. & Moora M. (2009) Large-scaleparallel 454 sequencing reveals host ecological group specificity ofarbuscular mycorrhizal fungi in a boreonemoral forest. New Phytologist184, 424–437.

Park W., Scheffler B.E., Bauer P.J. & Campbell B.T. (2010) Identification of thefamily of aquaporin genes and their expression in upland cotton (Gossypiumhirsitum L.) BMC Plant Biology 10, 142.

Parodi L.R. (1978) Enciclopedia Argentina de Agricultura y JardineríaTomo 1,Volumen 2, pp. 1114. Editorial ACME, Buenos Aires.

Porcel R., Aroca R., Azcón R. & Ruiz-Lozano J.M. (2006) PIP aquaporingene expression in arbuscular mycorrhizal Glycine max and Lactuca sativaplants in relation to drought stress tolerance. Plant Molecular Biology 60,389–404.

Porcel R., Azcón R. & Ruiz-Lozano J.M. (2004) Evaluation of the role of genesencoding for Δ

1-pyrroline-5-carboxylate synthetase (P5CS) during drought

stress in arbuscular mycorrhizal Glycine max and Lactuca sativa plants.Physiological and Molecular Plant Pathology 65, 211–221.

Raimondo F., Trifilò P., Lo Gullo M.A., Buffa R., Nardini A. & Salleo S. (2009)Effects of reduced irradiance on hydraulic architecture and water relations oftwo olive clones with different growth potentials. Environmental andExperimental Botany 66, 249–256.

Renker C., Weisshuhn K., Kellner H. & Buscot F. (2006) Rationalizingmolecular analysis of field collected roots for assessing diversity of arbuscularmycorrhizal fungi: to pool, or not to pool, that is the question. Mycorrhiza16, 525–531.

Ruiz-Lozano J.M., Porcel R., Azcón R. & Aroca R. (2012) Regulation byarbuscular mycorrhizae of the integrated physiological response to salinity inplants: new challenges in physiological and molecular studies. Journal ofExperimental Botany 63, 4033–4044.

Ruth B., Khalvati M. & Schmidhalter U. (2011) Quantification of mycorrhizalwater uptake via high-resolution on-line water content sensors. Plant and Soil342, 459–468.

Sanchez-Romera B., Ruiz-Lozano J.M., Zamarreño A.M., García-Mina J.M. &Aroca R. (2016) Arbuscular mycorrhizal symbiosis and methyl jasmonateavoid the inhibition of root hydraulic conductivity caused by drought.Mycorrhiza 26, 111–122.

Schachtman D.P. & Goodger J.Q. (2008) Chemical root to shoot signaling underdrought. Trends in Plant Sciences 13, 281–287.

Secchi F., Lovisolo C., Uehlein N., Kandenhoff R. & Schubert A. (2007) Isolationand functional characterization of three aquaporins from olive (Olea europaeaL.) Planta 225, 381–392.

Shannon C.E. & Weaver W. (1963) The mathematical theory of communica-tion1949. Urbana, Univ. Illinois Press, Urbana.

Sheng M., Tang M., Chen H., Yang B.W., Zhang F.F. & Huang Y.H. (2008)Influence of arbuscular mycorrhizae on photosynthesis and water status ofmaize plants under salt stress. Mycorrhiza 18, 287–296.

Smith S.E., Facelli E., Pope S. & Smith F.A. (2010) Plant performance in stressfulenvironments: interpreting new and established knowledge of the roles ofarbuscular mycorrhizas. Plant and Soil 326, 3–20.

Smith S.E. & Read D.J. (2008) Mycorrhizal symbiosis. Academic Press, SanDiego.

Steudle E. & Peterson C.A. (1998) How does water gets through roots? Journalof Experimental Botany 49, 775–788.

Suárez M.P., López-Rivares E.P., Lavee S. & Troncoso A. (1999) Rootingcapability of olive cuttings, cv. Gordal: influence of the presence of leavesand buds. Acta Horticulturae 474, 39–42.

Suga S. & Maeshima M. (2004) Water channel activity of radish plasmamembrane aquaporins heterologously expressed in yeast and their modifica-tion by site directed mutagenesis. Plant and Cell Physiology 45, 823–830.

SutkaM., LiG.W., Boudet J., BoursiacY.,Doumas P.&Maurel C. (2011)Naturalvariation of root hydraulics in Arabidopsis grown in normal and salt-stressedconditions. Plant Physiology 155, 1264–1276.

Tamura K., Dudley J., NeiM.&Kumar S. (2007)MEGA4:Molecular Evolution-ary Genetics Analysis (MEGA) software version 4.0. Molecular Biology andEvolution 24, 1596–1599.

Tataranni G., Santarcangelo M., Sofo A., Xiloyannis C., Tyerman S.D. & DichioB. (2015) Correlations between morpho-anatomical changes and radialhydraulic conductivity in roots of olive trees under water deficit and rewatering.Tree Physiology 35, 1356–1365.

Torres-Ruiz J.M.,Diaz-EspejoA.,Morales-SilleroA.,Martín-PalomoM.J.,MayrS., Beikircher B. & Fernández J.E. (2013) Shoot hydraulic characteristics, plantwater status and stomatal response in olive trees under different soil waterconditions. Plant and Soil 373, 77–87.

Tugendhaft Y., Eppel A., Kerem Z., Barazani O., Ben-Gal A., Kadereit J.W. &Dag A. (2016) Drought tolerance of three olive cultivars alternativelyselected for rain fed or intensive cultivation. Scientia Horticulturae 199,158–162.

Varela-Cevero S., Vasar M., Davison J., Barea J.M., Öpik M. &Azcón-Aguilar C. (2015) The composition of arbuscular mycorrhizal fungalcommunities differs among the roots, spores and extraradical myceliaassociated with five Mediterranean plant species. Environmental Microbiol-ogy 17, 2882–2895.

Wang L.-L., Chen A.-P., Zhong N.-Q., Liu N., Wu X.-M., Wang F. & Xia G.-X.(2014) The Thellungiella salsuginea tonoplast aquaporin TsTIP1;2 functionsin protection against multiple abiotic stresses. Plant and Cell Physiology 55,148–161.

Received 30 October 2015; received in revised form 14 July 2016;accepted for publication 15 July 2016

Effects of AM fungal backgrounds on olive plants 2513

© 2016 John Wiley & Sons Ltd, Plant, Cell and Environment, 39, 2498–2514

SUPPORTING INFORMATION

Additional Supporting Information may be found in the onlineversion of this article at the publisher’s web-site:

Supplemental File S1. Primers for aquaporin expressionSupplemental File S2. Multiple alignment C-terminal regionsOlPIP2;2, OlPIP2;3, OlPIP2;4, OlPIP2;5, and OlPIP2;6proteins, with Phaseolus vulgaris PvPIP2;1, respectively. Theconsensus amino acids are underlined. The PvPIP2;1sequences correspond to the peptide used to make therespective antibody.

Supplemental File S3. Aquaporins expression in roots of O.europaea cv. Picual seedlings growing into two different typesof soils (Grazalema-GZ and Freila-FL), and inoculated withdifferent mycorrhizal communities from GZ and FL soils.The plants were either cultivated under well-wateredconditions or subjected to drought stress for four weeks.Different lettersmeans significant differences (p< 0.05) amongtreatments after ANOVA and Tuckey’s test. Asterisks abovebars indicate significant differences (p< 0.05) between well-watered and drought treatments after t-student test (n= 6).

2514 M. Calvo-Polanco et al.

© 2016 John Wiley & Sons Ltd, Plant, Cell and Environment, 39, 2498–2514