You are on page 1of 13

Physiologia Plantarum 2010 Copyright © Physiologia Plantarum 2010, ISSN 0031-9317

Functional characterization of drought-responsive


aquaporins in Populus balsamifera and Populus simonii ×
balsamifera clones with different drought resistance
strategies
Adriana M. Almeida-Rodrigueza , Janice E.K. Cookeb , Francis Yeha and Janusz J. Zwiazeka∗
a Department of Renewable Resources, University of Alberta, Edmonton, Alberta, Canada T6E 2E3
b Department of Biological Sciences, University of Alberta, Edmonton, Alberta, Canada T6G 2E9

Correspondence We have characterized poplar aquaporins (AQPs) to investigate their possible


*Corresponding author, functions in differential drought responses of Populus balsamifera and Populus
e-mail: janusz.zwiazek@ualberta.ca
simonii × balsamifera leaves. Plants were exposed to mild and severe levels
Received 23 April 2010; of drought stress and to drought stress recovery treatment, and their responses
revised 20 July 2010 were compared with well-watered controls. Compared with P. balsamifera,
P. simonii × balsamifera used drought avoidance as the main drought
doi:10.1111/j.1399-3054.2010.01405.x resistance strategy, and rapidly reduced stomatal conductance in response to
stress. This strategy is correlated with growth rate reductions. Eleven AQPs
were transcriptionally profiled in leaves from these experiments and five
were functionally characterized for water channel activity. PIP1;3 and PIP2;5
were among the most highly expressed leaf AQPs that were responsive to
drought. Expression of PIP1;3 and five other AQPs increased in response to
drought in the leaves of P. simonii × balsamifera but not in P. balsamifera,
suggesting a possible role of these AQPs in water redistribution in the leaf
tissues. PIP2;5 was upregulated in P. balsamifera, but not in P. simonii
× balsamifera, suggesting that this AQP supports the transpiration-driven
water flow. Functional characterization of five drought-responsive plasma
membrane intrinsic proteins (PIPs) demonstrated that three PIP2 AQPs (PIP2;2,
PIP2;5, PIP2;7) functioned as water transporters in Xenopus laevis oocytes,
while the two PIP1 AQPs (PIP1;2 and PIP1;3) did not, consistent with the
notion that they may be functional only as heterotetramers.

Introduction (drought-tolerant) plants that rapidly reduce leaf water


potential (ψleaf ) in response to drought, isohydric
Water availability is among the most limiting factors for
(drought-avoidant) plants exercise tight control over
plant growth (Bogeat-Triboulot et al. 2007, Touchette
water loss mainly through stomatal movements (Tardieu
et al. 2007). Therefore, considerable research efforts
and Simonneau 1998, McDowell et al. 2008). Since
have focused on understanding drought resistance
isohydric plants rely on stomatal conductance as the
mechanisms in plants. As opposed to anisohydric
main drought resistance mechanism, the water content

Abbreviations – AQP, aquaporin; CNI, close neighbor interchange; LPI, leaf plastochron index; MIPs, major intrinsic proteins;
MP, maximum parsimony; NJ, neighbor joining; NIP, nodulin-like intrinsic protein; PIP, plasma membrane intrinsic protein;
qRT-PCR, quantitative reverse transcription polymerase chain reaction; RH, relative humidity; SIP, small basic intrinsic protein;
SWC, soil water content; TDR, time domain reflectometry; TIP, tonoplast intrinsic protein; XIP, X-intrinsic protein.

Physiol. Plant. 2010


of these plants fluctuates less and they tend to be withholding water, although AtPIP1;4 and AtPIP2;5 were
more susceptible to xylem cavitation compared with upregulated (Alexandersson et al. 2005). Expression of
anisohydric plants (McDowell et al. 2008). Although AQP s in roots and aerial parts of A. thaliana plants
hybrid poplars are generally considered to be relatively responded differently to mannitol-induced osmotic stress
isohydric (Tardieu and Simonneau 1998), they widely (Jang et al. 2004) or gradual drought stress (Alexanders-
vary in their stomatal sensitivity and susceptibility to son et al. 2005). These differential responses of AQP s to
xylem cavitation (Awad et al. 2010). drought stress are likely because of distinct functions that
Hydraulic conductivity of plant tissues plays an AQPs may have in different tissues and the functional
important role in drought resistance (Ogasa et al. 2010). differences between various tissues in plants exposed
Water transport involves a combination of apoplastic and to stress. In some cases, there can be discrepancies in
cell-to-cell (symplastic and transmembrane) pathways. AQP expression data reported in different studies (e.g.
Transmembrane movement of water is facilitated by Alexandersson et al. 2005, Jang et al. 2004). Variation
aquaporins (AQPs) (Maurel et al. 2008), a functional in simulating drought stress under laboratory conditions
class of channel proteins that belong to the ancient family and the fact that plants may have diverse resistance
of the major intrinsic proteins (MIPs) (Chaumont et al. strategies under different drought levels (Siemens and
2001, Johanson et al. 2001, Maurel et al. 2008, Siefritz Zwiazek 2003) can confound the interpretation of AQP
et al. 2002). This gene family has been highly conserved expression patterns reported by different studies.
during evolution, probably because of the strict structural In the present study, we compared a Populus simonii ×
requirements of water channels (Shapiguzov 2004). Five Populus balsamifera hybrid poplar clone, which, similar
subfamilies of MIPs have been identified in angiosperms to other hybrid poplars, has been characterized as having
[plasma membrane intrinsic protein (PIPs), tonoplast high stomatal sensitivity (Stettler et al. 1996, Tardieu and
intrinsic proteins (TIPs), nodulin-like intrinsic proteins Simonneau 1998), and a P. balsamifera clone with stom-
(NIPs), small basic intrinsic proteins (SIPs) and X-intrinsic ata that exhibits less sensitivity to drought (Pearce et al.
proteins (XIPs); Johanson et al. 2001, Chaumont et al. 2006). We hypothesized that the expression of leaf AQPs
2001, Sakurai et al. 2005, Danielson and Johanson would respond differently in these two poplar clones
2008]. The latter subfamily has been reported recently when the plants are exposed to drought conditions in
in the moss Physcomitrella patens, Populus trichocarpa accordance with their stomatal responses. Therefore, in
and in a wide variety of dicots, but not in monocots the present study, we subjected the plants of these two
or Arabidopsis thaliana (Danielson and Johanson clones to mild drought, severe drought and drought
2008, Gupta and Sankararamakrishnan 2009). AQPs recovery treatments and we compared their physiolog-
participate in water dynamics as facilitators of water and ical responses and AQP expression with well-watered
small solutes movement across cell membranes (Maurel plants.
et al. 2009). Their differential expression between plant
tissues and in response to abiotic stresses suggests that Materials and methods
they play roles in maintaining water balance in plants
Plant material
(Luu and Maurel 2005, Parent et al. 2009, Suga et al.
2001, Tyerman et al. 2002). However, the role of AQP s P. balsamifera (L.) (clone AP1004, gender unknown) and
in leaf water transport under drought conditions remains P. simonii (Carr.) × P. balsamifera (L.) (clone P38P38,
unclear (Heinen et al. 2009). As the expression of plant female) shoot cuttings were provided from single trees
AQP s is affected by drought (Aharon et al. 2003), it by Alberta-Pacific Forest Industries (Boyle AB, Canada).
could be expected that leaf AQP s that are involved in Ten-centimeter stem segments from 1-year-old trees
transpiration-driven water transport are strongly linked were re-hydrated for two days before planting in
to stomatal movements. 45 × 340 cm3 styroblocks (Stuewe and Sons, Tangent,
AQP expression in plants subjected to water deficit OR) containing peat moss, vermiculite and Turface cal-
stress varies depending on the species, developmen- cined clay (Profile Products, Buffalo Grove, IL) (2:1:1/2)
tal stage, growth conditions and severity of drought with 2.75 g l−1 of Nutricote Total controlled-release
(Alexandersson et al. 2005, Cocozza et al. 2010, Oono fertilizer (13–13–13; American Horticultural Supplies,
et al. 2003, Seki et al. 2001, 2002). For example, in San Marcos, CA). Rooted cuttings were replanted in
Nicotiana tabacum, drought stress reduced expression 4-l pots with the same soil composition and grown
of NtPIP1;1 and NtPIP2;1 while increasing expres- in the greenhouse under semi-controlled conditions
sion of NtAQP1 (Mahdieh et al. 2008). In A. thaliana [22/20◦ C day/night, 18/6 h light/dark, 60% relative
leaves, expression of several PIP genes was downregu- humidity (RH)], with watering six times per week until
lated in response to gradual drought stress induced by runoff and fertilization (2 g l−1 15–30–15) once per

Physiol. Plant. 2010


week. Plants were 130–160-cm tall at the outset of the (PMS Instruments, Corvallis, OR) was used to measure
drought experiments. leaf water potential (ψleaf ) for LPI 7 as described by
A preliminary experiment was used to empirically Wan and Zwiazek (1999). For both the gas exchange
define mild drought stress, severe drought stress, and and pressure chamber measurements, one block of
stress recovery treatments. Soil apparent permittivity four replicates per treatment was analyzed per day,
(Ka ) was measured using a time domain reflectome- for a total of n = 8. Within each block, the order
ter (TDR; Tektronix 1502B Cable TDR Cable Tester; in which the plants were measured was randomized.
Tektronix, Scarborough, ON, Canada), and volumet- For the molecular experiment, complete leaves (with
ric water content (θ) or soil water content (SWC) was petioles removed) from LPI 7 to LPI 10 were harvested
determined based on the equation for organic soils from each of the six plants and immediately frozen in
(Robinson et al. 2003, Toop et al. 1980). Measurements liquid nitrogen. Samples were stored at −80◦ C until
of stomatal conductance (gs ) were carried out with an processed.
Li-1600 steady-state porometer (LI-COR, Lincoln, NE).
SWC of 20–25% was defined as mild stress, based on
Sequence analysis
the observation of changes in gs when compared with
well-watered plants. SWC lower than 20% was desig- P. trichocarpa (Torr. & Gray) AQP gene models were
nated as severe stress because, at this point, plants were identified in version 2.0 of the assembled, annotated
wilted and they showed low gs . Three days of re-watering genome (http://www.phytozome.net/poplar.php) using
following the point of severe stress was designated as P. trichocarpa AQP sequences reported by Gupta and
recovery, because at this point gs was comparable to Sankararamakrishnan (2009) as BLAST queries (Altschul
that in well-watered plants. et al. 1997) (see Appendix S1). AQP-deduced amino acid
Two duplicate drought experiments were conducted sequences from P. trichocarpa, A. thaliana (L.) Heynh.
in this study. One was used for physiological measure- (Johanson et al. 2001; www.arabidopsis.org) and repre-
ments (n = 8) (referred to as the physiological experi- sentative XIPs from Ph. patens (Danielson and Johanson
ment) and one for the gene expression analyses (n = 6) 2008) were aligned using CLUSTALW in MEGA v4 (Tamura
(referred to as the molecular experiment). In each exper- et al. 2007). Phylogenetic trees were constructed using
iment, trees for each of the two poplar clones were maximum parsimony (MP) in MEGA v4. A bootstrap
randomly assigned to the different drought treatments, consensus tree of 2000 replicates was obtained, using
i.e. mild drought, severe drought, recovery from severe close neighbor interchange (CNI) with search level of
drought and well-watered plants (controls). Because of 3, a random addition of 10 replicates, and a complete
the diurnal cycle of AQP activity, physiological measure- deletion for gaps and missing data. No out group was
ments (physiological experiment) and tissue harvesting included in the analysis.
(molecular experiment) were performed between 08:00
and 11:00 h. In the case of the physiological experi-
Quantitative reverse transcription polymerase
ment, the plants were measured within two days. For
chain reaction
both experiments, drought treatments were initiated on
different days so that mild drought, severe drought, and Total RNA was extracted using the hexadecyltrimethy-
drought recovery treatments ended on the same day. lammonium bromide (CTAB) extraction protocol of
For logistical reasons, the physiological experiment was Chang et al. (1993). Three micrograms of total RNA were
divided into two blocks of four independent replicates treated with DNase I (New England Biolabs, Ipswich,
per treatment per block; the treatments for the two MA) and used as template for first strand cDNA synthesis
blocks were initiated one day apart so that the plants using oligo(dT)23 VN and Moloney murine leukemia virus
from block 1 could be analyzed on one day, and the reverse transcriptase, following manufacturer’s instruc-
plants from block 2 could be analyzed the following tions (New England Biolabs).
day. Gene-specific qRT-PCR primers were designed in the
Leaves for analysis were selected according to the leaf 3 untranslated (UTR)-region using Primer Express v3
plastochron index (LPI; Larson and Isebrands 1971). For (Applied Biosystems, Foster City, CA) (Appendix S2).
the physiological experiment, single measurements of Populus AQP genes were selected based on phylo-
net photosynthesis (Pn ) and gs were measured for LPI 7 genetic proximity to genes associated with drought
on eight plants per treatment with an LI-6400 portable responses in other species, or by EST abundance pat-
infrared gas analyzer (LI-COR) with an incorporated terns, suggesting relatively high expression in leaves or
light source. All measurements were taken between in abiotic stress responses (Sterky et al. 2004). Several
08:00 and 11:00 h. A Scholander pressure chamber potential reference genes were tested for stable levels of

Physiol. Plant. 2010


transcript abundance across samples to be compared. threshold) for analysis, and dissociation curves were
A member of the EF1α gene family was chosen as the verified for each of the genes.
reference gene, as transcript abundance corresponding
to this gene did not show significant differences across Osmotic water permeability assay
treatments when expressed as 2−Ct (P = 0.514 for the
Full-length PtdPIP1;2, PtdPIP1;3, PtdPIP2;2, PtdPIP2;5
drought experiment, and P = 0.257 for the tissue com-
and PtdPIP2;7 cDNAs were amplified using the Expand
parison experiment; Appendix S3).
High Fidelity PCR System (Roche, Indianapolis, IN),
Transcript abundance was quantified using standard using gene-specific primers incorporating Bgl II sites on
curves for both target and reference genes. Standard both ends (Appendix S2). Full-length cDNAs were sub-
curves were constructed from amplicons of full-length cloned into the Bgl II site of the pAWPLA (pXLII) vector
or near full-length P. trichocarpa × deltoides cDNAs for [dephosphorylated using calf intestinal alkaline phos-
conducting a tissue survey analysis (data not shown) phatase (Invitrogen)], carrying the 5 - and 3 -UTR regions
and for the drought experiment described in this study. of the beta-globin gene from Xenopus laevis. Plasmids
P. trichocarpa × deltoides cDNAs were either obtained containing the full-length poplar AQPs in the correct
from Université Laval (www.arborea.ca) or, in the orientation were determined by restriction mapping and
case of PtdSIP1;2 and PtdEF1α, cloned via RT-PCR sequencing. Plasmids were linearized downstream of
with primers described in Appendix S2 (Sambrook and the 3 -UTR beta-globin sequence using FastDigest SacI
Russell 2001). Plasmids were purified using GeneJET (Fermentas) for PtdPIP1;2 and PtdPIP1;3, NgoMIV (New
Plasmid Miniprep Kit (Fermentas, Burlington, ON, England Biolabs) for PtdPIP2;7 and Bpu10I (New Eng-
Canada) according to the manufacturer’s protocol. Each land Biolabs) for PtdPIP2;2 and PtdPIP2;5. Capped RNA
plasmid (30 ng μl−1 ) was used as template in a PCR (cRNA) was synthesized in vitro by T3 RNA polymerase
reaction using universal primers. PCR conditions were (mMESSAGE mMACHINE T3 kit, Ambion, Austin, TX)
as follows: 1 cycle at 94◦ C for 5 min, followed by utilizing the T3 promoter of the vector. DNA was
40 cycles at 94◦ C for 40 s, 50◦ C for 40 s, and 72◦ C removed using TURBO DNase, and the cRNA was pre-
for 1.5 min, and finishing with an extension step at cipitated for 2 h at −20◦ C using lithium chloride. cRNA
72◦ C for 5 min. Amplicons were purified using the quality and quantity were determined using the Nan-
QIAquick PCR purification kit (Qiagen, Mississauga, odrop and the 2100 Bioanalyser (Agilent, Santa Clara,
ON, Canada), and diluted 1:10 with autoclaved CA). cRNA was aliquotted and stored at −80◦ C until use.
RNase-free water. Concentration was determined using X. laevis oocytes were prepared as previously
a Nanodrop ND-1000 spectrophotometer (Thermo described (Cao et al. 1992, Daniels et al. 1996). Fifty
Scientific, Wilmington, DE). Target and reference genes nanoliters AQP cRNA (1 mg ml−1 ) or nuclease-free
were diluted individually to a final concentration of water were injected to the oocytes using a 10-μl microin-
4 × 109 molecules of DNA. Afterward, a pooled serial jector (Drummond Scientific, Broomall, PA). Injected
dilution series was made that included all target oocytes were randomly assigned to two separated six-
genes plus EF1α, comprising 4 × 107 , 4 × 106 , 4 × 104 , well cell culture plates, and incubated as described by
4 × 103 and 4 × 101 copies of DNA molecules for each Cao et al. (1992). One oocyte was added to one well of
gene. a two-well thick-slide filled with regular Barth’s solution
Real-time PCR was performed on a 7500 Fast Real- (200 mosmol), and an initial image of the cell was taken.
Time PCR system (Applied Biosystems). Three biological Then an oocyte was quickly transferred to the other
replicates, each with three technical replicates, were well of the slide containing hypotonic Barth’s solution
assayed for each sample. The reference gene was (Barth’s solution diluted five times with distilled water
included on each plate. cDNA from a single, pooled for a final concentration of 40 mosmol). Changes in cell
RT reaction representing one sample from the control volume were tracked via microscopy images taken at 5-s
(well-watered) plants from the drought experiment was intervals for 5 min. Images were analyzed using IMAGEJ
used as a calibrator to enable plate-to-plate comparison. (version 1.38X, Abramoff et al. 2004). The surface area
PCR was carried out in a volume of 10 μl including a from the oocyte picture was obtained automatically with
final concentration of 50 ng cDNA, 1× master mix con- IMAGEJ after a known scale was set. Oocyte surface area
taining 0.2 mM dNTPs, 0.3 U Platinum Taq Polymerase and volume were calculated based on the sphere surface
(Invitrogen), 0.25× SYBR Green and 0.1× ROX. PCR area and volume geometric formulas correspondingly.
conditions were as follows: 1 cycle at 95◦ C for 2 min, The relative volume (V/Vo ), the initial transmembrane
40 cycles at 95◦ C for 15 s, 60◦ C for 1 min, and a disso- volume flux (Jv ) and the osmotic water permeability
ciation stage including 2 cycles of 95◦ C for 15 s, 60◦ C coefficient (Pf ) were calculated based on Zhang and
for 1 min. Samples were subjected to auto Ct (cycle Verkman (1991).

Physiol. Plant. 2010


Statistical analysis Effects of PtdPIP s on the osmotic water permeability
coefficient for the oocytes exposed to hypotonic solu-
Data were analyzed using SAS version 9.1 (SAS Institute,
tion were evaluated by a nested analysis of variance. The
Cary, NC). Effects of drought on gs , Pn , height and
alternative hypothesis of determining the statistical differ-
ψleaf were evaluated by a complete randomized block
ences among different oocytes exposed to a hypotonic
design analysis of variance. For every treatment, at
solution that were injected individually with different
least six replicates were used for each poplar clone.
PtdPIP s. Pairwise comparisons were tested using the
The experimental design included two blocks, four
least-square means differences statement. P -values were
treatments, and four replicates per treatment in each
adjusted with Tukey–Kramer to a P -value of 0.05.
block. The alternative hypothesis of determining the
statistical differences among treatments for each of the
physiological measurements in each of the poplars Results
was evaluated. Preplanned comparisons were tested
using the least-square means differences statement Effects of drought treatments on Pn , gs , leaf and
height growth
with Tukey–Kramer P -values adjustment, and/or with
contrast statement, adjusting the alpha value to 0.03. It took two days for P. balsamifera to reduce SWC
Effects of drought on AQP transcript abundance were to 20–25% (mild drought). In contrast, it took six
evaluated by a randomized factorial analysis of variance. days for P. simonii × balsamifera to reach a similar
The alternative hypothesis of determining the statistical level. In P. simonii × balsamifera, this mild stress treat-
differences among treatments for each of the 11 poplar ment resulted in a reduction of Pn by about 60% and
AQP genes in each of the poplars was evaluated. Pre- gs by about 33% compared with well-watered plants
planned comparisons were tested using the least-square (Fig. 1A,B). In contrast, no significant differences were
means differences statement. P -values were adjusted observed in Pn and gs between control and mild drought
with Tukey–Kramer or LSD adjusted to a P -value of 0.03. stress treatment in P. balsamifera. In plants treated with

Fig. 1. Drought responses in two contrasting Populus clones, Populus balsamifera and Populus simonii × P. balsamifera. (A) Photosynthesis (Pn ),
(B) stomatal conductance (gs ) associated with the percentage of SWC, (C) plant height and (D) leaf water potential (ψleaf ). Ten-week-old plants were
well watered (control) or subjected to mild water stress, severe water stress, or stress recovery following severe water stress. Significant differences
among treatments indicated by letters above each bar (P ≤ 0.05, Tukey-Kramer adjustment). Means ± SE are shown (n = 6 minimum).

Physiol. Plant. 2010


severe drought (SWC lower than 20%), both Pn and gs wilted in response to severe drought while only a few
were reduced but still higher in P. balsamifera compared leaves in some P. simonii × balsamifera plants showed
with P. simonii × balsamifera (Fig. 1A,B). Following signs of wilting. Similarly, young leaves of P. balsamifera
stress recovery, Pn and gs in both poplars returned displayed a relatively high extent of necrosis and leaf
to the levels measured in well-watered control plants abscission which occurred in only a few P. simonii ×
(Fig. 1A,B). balsamifera plants (data not shown).
Drought stress treatments did not affect heights of
P. balsamifera plants. However, P. simonii × balsam-
ifera plants were shorter following stress recovery Analysis of the P. trichocarpa AQP family
treatment compared with the control and mild stress Fifty-five full-length AQP gene models were identified
treatments (Fig. 1C). The leaf exhibited trends similar to in version 2.0 of the P. trichocarpa genome (Appendix
Pn and gs in both clones (Fig. 1D). Mild drought did not S4), consistent with the findings of Gupta and Sankarara-
significantly affect leaf in P. balsamifera, but decreased makrishnan (2009), who reported on an earlier version
leaf in P. simonii × balsamifera (Fig. 1D). Both poplars of the P. trichocarpa genome. In agreement with Gupta
showed a significant reduction of leaf when exposed and Sankararamakrishnan (2009), phylogenetic analy-
to severe drought and returned to near control levels sis (Fig. 2) demonstrated that the P. trichocarpa AQP s
in re-watered plants (Fig. 1D). Leaves of P. balsamifera fall into five subfamilies: the PIPs, TIPs, NIPs, SIPs and

Fig. 2. Dendrogram obtained using MP illustrating relationships between members of the AQP family in Populus trichocarpa. P. trichocarpa AQPs
are denoted with black diamonds. The AQP family from Arabidopsis thaliana (AtAQPs) (Johanson et al. 2001) and two XIPs previously reported in
the moss Physcomitrella patens (PpAQPs) (Danielson and Johanson 2008) are included for comparison. Arrows indicate aquaporins selected for gene
expression analysis.

Physiol. Plant. 2010


the recently described XIPs (Danielson and Johanson Functional analysis of five PIPs
2008, Gupta and Sankararamakrishnan 2009). Further
Five PIP s cloned from P. trichocarpa × deltoides were
subgroups of proteins could be identified in the TIP
tested for their ability to transport water using a X. laevis
(TIP1, TIP2, TIP3, TIP4, TIP5) and PIP (PIP1 and PIP2)
oocyte swelling assay (Fig. 4). Relative volumes of
subfamilies. For consistency, the same gene names are
oocytes injected with PtdPIP1;2 and PtdPIP1;3 were
used in this paper as those proposed by Gupta and
similar to those injected with water (controls). In con-
Sankararamakrishnan (2009).
trast, oocytes injected with one of three PtdPIP2s showed
All 55 AQP s were assigned to chromosomes in v2.0
higher relative volumes compared to the control oocytes.
of the P. trichocarpa genome, compared to only 46 in
The greatest changes in oocyte volumes were observed
earlier versions (Gupta and Sankararamakrishnan 2009).
for PtdPIP2;5 followed by PtdPIP2;2 and PtdPIP2;7.
While 15 of 19 linkage groups (chromosomes) exhib-
All oocytes injected with any of the PIP2s ruptured
ited 5 AQP genes or less, linkage group 9 contained
under hypotonic conditions during the 5-min assay.
9 AQP s, including 5 of the 6 PtXIP s. Several closely
The osmotic water permeability coefficient (Pf ) was
related gene pairs could be identified based on dendro-
calculated based on its relationship with the initial trans-
gram location and locations within homologous genome
membrane volume flux (Jv ). Pf values for oocytes injected
blocks described by Tuskan et al. (2006), which were
with cRNAs corresponding to PtdPIP2;5, PtdPIP2;7 and
based on the v1.1 genome release: PtPIP1;4/PtPIP1;5,
PtdPIP2;2 were respectively 9-, 3.2- and 5.2-fold higher
PtPIP1;3/PtPIP1;2, PtPIP2;4/PtPIP2;3, PtPIP2;1/PtPIP2;2,
than those of control oocytes. For the oocytes injected
PtTIP1;3/PtTIP1;4, PtTIP1;1/PtTIP1;2, PtTIP2;1/PtTIP2;2,
with the cRNA corresponding to PtdPIP1;3 and Ptd-
PtNIP3;3/PtNIP3;4, PtNIP3;1/PtNIP3;2, PtSIP1;1/
PIP1;2, the Pf values were the same or lower than for the
PtSIP1;2, PtSIP1;3/PtSIP1;4 and PtSIP2;2/PtSIP2;1.
control (the ratios between the Pf values of the putative
PtdPIP1 and the control were 1 and 0.64, respectively;
Expression profiling of Populus AQPs Fig. 4).
A subset of AQP s was selected for a qRT-PCR survey
of transcript abundance in leaves of P. balsamifera and Discussion
P. simonii × balsamifera subjected to drought conditions
P. balsamifera and P. simonii × balsamifera exhibit
identical to the experiments described above. As PIPs
contrasting strategies for coping with drought
and TIPs have been associated with drought responses
in A. thaliana (Alexandersson et al. 2005, Jang et al. P. balsamifera and its descendant P. simonii × bal-
2004), eight PIP s (PIP1;2, PIP1;3, PIP2;1, PIP2;2, PIP2;3, samifera hybrid are phreatophytic poplars (Amlin and
PIP2;4, PIP2;5 and PIP2;7 ) and two TIP s (TIP1;6 and Rood 2003). We have shown that these two poplar
TIP2;1) were chosen for gene expression analysis (Fig. 3). clones have different strategies to resist drought stress.
One SIP (SIP1;2) was also selected for the analysis. The responses of P. simonii × balsamifera to drought
PIP1;2, PIP1;3, PIP2;1, PIP2;2, PIP2;7, TIP1;6 and TIP2;1 can be referred to as drought avoidant, characterized
showed statistically significant responses to drought by rapid stomatal closure. This type of drought strat-
in P. simonii × balsamifera, but not in P. balsamifera. egy has been also previously reported for Populus
PIP2;1 and PIP2;2 showed increased transcript abun- euramericana cv. (P. deltoides × nigra) (Tardieu and
dance in leaves during mild and severe drought and then Simonneau 1998). It is interesting that despite the dif-
reduced transcript abundance following re-watering in ferences in stomatal conductance, leaf water potentials
P. simonii × balsamifera. PIP1;2 and PIP2;7 displayed were similar in both clones, suggesting possible differ-
increased transcript abundance during severe drought, ences in osmotic adjustment. This response, combined
with the latter also showing a decrease in transcript abun- with decreased stomatal conductance, could have con-
dance following re-watering. TIP1;6 exhibited decreased tributed to less extensive leaf abscission and necrosis in
transcript abundance under severe drought; PIP2;3 and severely drought-stressed P. simonii × balsamifera com-
PIP2;4 showed similar trends, although the difference in pared with P. balsamifera. In contrast, P. balsamifera
transcript abundance is not statistically significant. Inter- maintained relatively high gs and Pn during mild drought.
estingly, only PIP2;5 showed a statistically significant The capacity of P. balsamifera to maintain relatively
response to drought in P. balsamifera, exhibiting high rates of gas exchange, during sudden drought
increased transcript abundance at mild and severe stress. events, compared with P. simonii × balsamifera is also
In contrast, PIP2;5 showed decreased transcript abun- associated with a lower vulnerability to xylem cavita-
dance in P. simonii × balsamifera under mild and severe tion (Arango-Velezet al., unpublished data), as reported
stress, although this was not statistically significant. for other tree species (Maherali et al. 2006), allowing

Physiol. Plant. 2010


Fig. 3. Transcript abundance corresponding to 11 AQPs in leaves of Populus balsamifera and Populus simonii × balsamifera subjected to four
different water availability treatments. Transcript abundance for each AQP was quantified by standard curve, and is expressed relative to transcript
abundance corresponding to EF1-α. Mean EF1-α transcript abundance across treatments for each clone was not statistically significantly different, and
thus EF1-α was considered to be constitutively expressed within the context of these experiments (see Appendix S3). n = 3 biological replicates; each
biological replicate was assayed in triplicate. Error bars = SE. Letters above bars designate statistical differences at P = 0.05 in transcript abundance
for a given gene within a clone; significance levels are shown with asterisks (∗ P = 0.03; ∗∗ P < 0.01; ∗∗∗ P < 0.008).

P. balsamifera to maintain a comparable level of carbon in stomatal responses of P. balsamifera and P. simonii ×
gain. However, while P. balsamifera may be able to balsamifera are likely linked to cavitation resistance.
maintain higher photosynthetic activity and growth rates Poplars often cope with drought by reducing their
under short-term mild drought conditions, sluggish stom- transpiration area through leaf and branch abscission
atal responsiveness might compromise its survival under (Amlin and Rood 2003). Although in our study, ψleaf , gs
severe drought (Street et al. 2006). and Pn of severely drought-stressed P. balsamifera and
Although all poplars are relatively susceptible to P. simonii × balsamifera returned to their pre-drought
cavitation (Cochard et al. 2007a, 2007b), the differences levels within three days following re-watering, leaf

Physiol. Plant. 2010


stress, expression levels of PIP2;5 increased by almost
fourfold in P. balsamifera and decreased by several-fold
in P. simonii × balsamifera, although the latter was
not significant because of variability in gene expression
across replicates. This supports the notion that PIP2;5
may be involved in the regulation of leaf hydraulic con-
ductance with a possible link to stomatal responses.
The putative ortholog to PIP2;5 was the main leaf AQP
found in P. tremula × tremuloides (Marjanović et al.
2005), bur oak (Quercus macrocarpa; Voicu et al. 2009)
and walnut (Juglans regia; Cochard et al. 2007a, 2007b).
Expression of these putative PIP2;5 orthologs has been
Fig. 4. Pf values of oocytes injected with five cRNAs from Populus shown to be affected by light conditions in these lat-
trichocarpa × deltoides PIPs or with water (control). These values were ter two species (Cochard et al. 2007a, 2007b; Voicu
obtained based on the surface area measurements and the relative
volume calculations of a swelling assay in which the injected oocytes
et al. 2009), providing additional evidence to support
were exposed to hypotonic conditions. Pf values presented are means the hypothesis that PIP2;5 may play a role in leaf water
± SE of at least nine replicates. Significant differences between the relations in poplar. We confirmed water channel activity
genes and the control are indicated by lowercase letters (P ≤ 0.05, for PIP2;5. PIP2;5 corresponds to PoptrPIP2.5 in Secchi
Tukey-Kramer adjustment). et al. (2009), who also demonstrated that this gene prod-
uct functions as a water channel. PIP2s from many other
senescence and abscission of older leaves still occurred species function as water channels, including OsPIP2;7
in plants recovering from drought. (Oryza sativa; Li et al. 2008), OePIP2;1 (Olea europaea
L.; Secchi et al. 2007), SsAQP2 (Samanea saman; Moshe-
lion et al. 2002), ZmPIP2;5 (Zea mays; Chaumont et al.
A subset of AQPs are drought responsive in leaves, 2000), ZmPIP2;1, ZmPIP2;4 (Fetter et al. 2004), JrPIP2;1,
and show different responses to drought in
JrPIP2;2 (J. regia; Sakr et al. 2003), HvPIP2;1 (Hordeum
P. balsamifera vs P. simonii × balsamifera
vulgaris; Katsuhara et al. 2002), NtPIP2;1 (N. tabacum;
Based on these contrasting stomatal responses, we Mahdieh et al. 2008) and PM28A (Spinacea oleracea;
hypothesized that AQP s expressed in leaves would show Johansson et al. 1998).
different responses in P. balsamifera and P. simonii × Contrary to PIP2;5, PIP1;3 expression was drought
balsamifera when the plants were exposed to drought responsive in P. simonii × balsamifera leaves rather
conditions. Expression profiling of 11 AQP s representing than P. balsamifera leaves. The role of this AQP in
3 of the 5 AQP subfamilies revealed that a number of leaf drought responses is less clear. PIP1;3 is closely
AQP s were drought responsive, and that some of these related to PIP1;2, yet shows markedly different pat-
AQP s showed contrasting patterns of responsiveness in terns and levels of expression, particularly across tissues
P. balsamifera and P. simonii × balsamifera, supporting (data not shown). This observation supports the notion
our hypothesis (Fig. 3). The higher level of expression for that subfunctionalization and/or neofunctionalization
several AQP s in control leaves of P. simonii × balsam- has occurred in the Populus AQP gene family. Sim-
ifera compared with P. balsamifera suggests that, in the ilar to PIP1;1 in P. trichocarpa (Secchi et al. 2009)
absence of significant differences in leaf hydraulic con- and other plants (Chaumont et al. 2000, Fetter et al.
ductance between the two poplars (data not shown), the 2004, Mahdieh et al. 2008, Moshelion et al. 2002,
cell-to-cell pathway may play a greater role in the leaf Secchi et al. 2007), PtdPIP1;3 and PtdPIP1;2 had low Pf
water transport of P. simonii × balsamifera compared when expressed individually in heterologous system of
with P. balsamifera when soil water is not a limiting Xenopus oocytes. PIP1 functionality in cell membranes
factor. may depend on their interactions with PIP2 isoforms.
PIP1;3 and PIP2;5 present interesting candidates for For example, co-expression of maize ZmPIP1;2 and
further investigation into potential roles of these AQP s ZmPIP2;5 in Xenopus oocytes improved the targeting
in regulating cell-to-cell movement of water during of ZmPIP1;2 into the plasma membrane, raising the
water deficit conditions, because these genes are not Pf values higher than for ZmPIP2;5 expressed alone
only drought responsive but also show substantially (Fetter et al. 2004). In maize mesophyll protoplasts,
greater transcript abundance in leaves relative to other when expressed alone, ZmPIP1;2 was retained in the
AQP s that were examined (data not shown). When endoplasmic reticulum and ZmPIP2;1 was found in the
plants were subjected to mild and severe drought plasma membrane. However, when co-expressed, they

Physiol. Plant. 2010


formed hetero-oligomers and ZmPIP1;2 was relocalized for assistance with qRT-PCR, and Xing-Zhen Chen and
to the plasma membrane (Zelazny et al. 2007). However, Jung Woo Yang for assistance with X. laevis oocytes cRNA
we cannot rule out the possibility that PIP1 isoforms were injections. This research was funded by the Natural Sciences
not correctly expressed in Xenopus oocytes. and Engineering Research Council of Canada (NSERC).

Leaf AQPs and contrasting stomatal behavior in References


Populus
Abramoff MD, Magelhaes PJ, Ram SJ (2004) Image
It has been suggested that AQP s might play a role in plant processing with ImageJ. Biophoton Int 11: 36–42
water balance by determining isohydric or anisohydric Aharon R, Shahak Y, Wininger S, Bendov R, Kapulnik Y,
behavior in tomato (Solanum lycopersicum; Sade et al. Galili G (2003) Overexpression of a plasma membrane
2009). However, the role of AQPs in leaf water trans- aquaporin in transgenic tobacco improves plant vigor
port and the contribution of the cell-to-cell pathway to under favorable growth conditions but not under
leaf hydraulic conductance remain contentious (Heinen drought or salt stress. Plant Cell 15: 439–447
et al. 2009). Although temperature responses (Sack et al. Alexandersson E, Fraysse L, Sjövall-Larsen S, Gustavsson
2004) and chemical inhibition of leaf hydraulic conduc- S, Fellert M, Karlsson M, Johanson U, Kjellbom P (2005)
tance (Nardini et al. 2005) suggest that AQPs play a role Whole gene family expression and drought stress
in bulk leaf water transport, other studies suggest that regulation of aquaporins. Plant Mol Biol 59: 469–484
this may not be true for all plants. In Q. macrocarpa, Altschul SF, Madden TL, Schäffer AA, Zhang J, Zhang
Z, Miller W, Lipman DJ (1997) Gapped BLAST and
there was no significant change in expression of AQP
PSI-BLAST: a new generation of protein database search
genes in leaves with light-induced high hydraulic con-
programs. Nucleic Acids Res 25: 3389–3402
ductance compared to leaves with low hydraulic con-
Amlin NM, Rood SB (2003) Drought stress and recovery of
ductance (Voicu et al. 2009). Similarly, in Arabidopsis,
riparian cottonwoods due to water table alteration along
leaf hydraulic conductance was not altered in double- Willow Creek, Alberta. Trees 17: 351–358
antisense plants with reduced expression of PIP1 and Awad H, Barigah T, Badel E, Cochard H, Herbette S (2010)
PIP2 aquaporins (Martre et al. 2002), while in tobacco, Poplar vulnerability to xylem cavitation acclimates to
leaf hydraulic conductance was similar in wild-type and drier soil conditions. Physiol Plant 139: 280–288
transgenic plants constitutively overexpressing PIP2;5 Bogeat-Triboulot MB, Brosché M, Renaut J, Jouve L, Le
and PIP1;4 AQP s (Lee et al. 2009). Therefore, the sig- Thiec D, Fayyaz P, Vinocur B, Witters E, Laukens
nificance of AQP -mediated leaf water transport may K, Teichmann T, Altman A, Hausman JF, Polle
vary between plants and under different environmental A, Kangasjärvi J, Dreyer E (2007) Gradual soil water
conditions. depletion results in reversible changes of gene
In conclusion, we have identified a number of AQP s expression, protein profiles, ecophysiology, and growth
that show distinct responses to drought in leaves of performance in Populus euphratica, a poplar growing in
P. simonii × balsamifera and P. balsamifera, two poplar arid regions. Plant Physiol 143: 876–892
clones that exhibit contrasting drought response strate- Cao Y, Anderova M, Crawford NM, Schroeder JI (1992)
gies. PIP2;5 and PIP1;3 are the most highly expressed of Expression of an outward-rectifying potassium channel
these drought-responsive AQP s. The water-transporting from maize mRNA and complementary RNA in Xenopus
PIP2;5 may be linked to hydraulic and stomatal oocytes. Plant Cell 4: 961–969
responses, while the functions of PIP1;3 are less clear. Chang S, Puryear J, Cairney J (1993) A simple method for
The generally higher levels of AQP expression in non- isolating RNA from pine trees. Plant Mol Biol Rep 11:
113–116
stressed leaves of P. simonii × balsamifera compared to
Chaumont F, Barrieu F, Jung R, Chrispeels MJ (2000)
P. balsamifera suggest that the cell-to-cell pathway may
Plasma membrane intrinsic proteins from maize cluster
constitute a greater role in water transport of leaves for
in two sequence subgroups with differential aquaporin
this drought-avoidant hybrid. Further studies are needed
activity. Plant Physiol 122: 1025–1034
to delineate the in planta function of these two AQPs in
Chaumont F, Barrieu F, Wojcik E, Chrispeels MJ, Jung R
poplars, particularly with respect to leaf hydraulics. (2001) Aquaporins constitute a large and highly
divergent protein family in maize. Plant Physiol 125:
Acknowledgements – We thank Alfons Weig and François 1206–1215
Chaumont for the pAWPLA (pXLII) vector, François Cochard H, Venisse JS, Barigah TS, Brunel N, Herbette S,
Chaumont for comments on the manuscript, Leonardo Guilliot A, Tyree MT, Sakr S (2007a) Putative role of
Galindo-Gonzalez for assistance with sequence analysis, aquaporins in variable hydraulic conductance of leaves
Walid El Kayal for assistance with cloning, Troy Locke in response to light. Plant Physiol 143: 122–133

Physiol. Plant. 2010


Cochard H, Casella E, Mencuccini M (2007b) Xylem Luu DT, Maurel C (2005) Aquaporins in a challenging
vulnerability to cavitation varies among poplar and environment: molecular gears for adjusting plant water
willow clones and correlates with yield. Tree Physiol status. Plant Cell Environ 28: 85–96
27: 1761–1767 Mahdieh M, Mostajeran A, Horie T, Katsuhara M (2008)
Cocozza C, Cherubini P, Regier N, Saurer M, Frey B, Drought stress alters water relations and expression of
Tognetti R (2010) Early effects of water deficit on two PIP-type aquaporin genes in Nicotiana tabacum plants.
parental clones of Populus nigra grown under different Plant Cell Physiol 49: 801–813
environmental conditions. Func Plant Biol 37: 244–254 Maherali H, Moura CF, Caldeira MC, Willson CJ,
Daniels MJ, Chaumont F, Mirkov TE, Chrispeels MJ (1996) Jackson RB (2006) Functional coordination between leaf
Characterization of a new vacuolar membrane gas exchange and vulnerability to xylem cavitation in
aquaporin sensitive to mercury at a unique site. Plant temperate forest trees. Plant Cell Environ 29: 571–583
Cell 8: 587–599 Marjanović Z, Uehlein N, Kaldenhoff R, Zwiazek JJ, Weiß
Danielson JÅH, Johanson U (2008) Unexpected M, Hampp R, Nehls U (2005) Aquaporins in poplar:
complexity of the Aquaporin gene family in the moss what a difference a symbiont makes! Planta 222:
Physcomitrella patens. BMC Plant Biol 8: 1–15 258–268
Fetter K, Van Wilder V, Moshelion M, Chaumont F (2004) Martre P, Morillon R, Barrieu F, North G, Nobel P,
Interactions between plasma membrane aquaporins Chrispeels M (2002) Plasma membrane aquaporin play
modulate their water channel activity. Plant Cell 16: a significant role during recovery from water deficit.
215–228 Plant Physiol 130: 2101–2110
Gupta AB, Sankararamakrishnan R (2009) Genome-wide Maurel C, Verdoucq L, Luu D-T, Santoni V (2008) Plant
analysis of major intrinsic proteins in the tree plant aquaporins: membrane channels with multiple
Populus trichocarpa: characterization of XIP subfamily integrated functions. Annu Rev Plant Biol 59: 595–624
of aquaporins from evolutionary perspective. BMC Plant
Maurel C, Santoni V, Luu D-T, Wudick MM, Verdoucq L
Biol 9: 134
(2009) The cellular dynamics of plant aquaporin
Heinen RB, Ye Q, Chaumont F (2009) Role of aquaporins
expression and functions. Plant Biol 12: 690–698
in leaf physiology. J Exp Bot 60: 2971–2985
McDowell N, Pockman WT, Allen CD, Breshears DD,
Jang JY, Kim DG, Kim YO, Kim JS, Kang H (2004) An
Cobb N, Kolb T, Plaut J, Sperry J, West A, Williams DG,
expression analysis of a gene family encoding plasma
Yepez EA (2008) Mechanisms of plant survival and
membrane aquaporins in response to abiotic stresses in
mortality during drought: why do some plants survive
Arabidopsis thaliana. Plant Mol Biol 54: 713–725
while others succumb to drought? New Phytol 178:
Johanson U, Karlsson M, Johansson I, Gustavsson S,
719–739
Sjövall S, Fraysse L, Weig A, Kjellbom P (2001) The
Moshelion M, Becker D, Biela A, Uehlein N, Hedrich R,
complete set of genes encoding major intrinsic proteins
Otto B, Levi H, Moran N, Kaldenhoff R (2002) Plasma
in Arabidopsis provides a framework for a new
nomenclature for major intrinsic proteins in plants. Plant membrane aquaporins in the motor cells of Samanea
Physiol 126: 1358–1369 saman: diurnal and circadian regulation. Plant Cell 14:
Johansson I, Karlsson M, Shukla VK, Chrispeels MJ, 727–739
Larsson C, Kjellbom P (1998) Water transport activity of Nardini A, Salleo S, Andri S (2005) Circadian regulation of
the plasma membrane aquaporin PM28A is regulated by leaf hydraulic conductance in sunflower (Helianthus
phosphorylation. Plant Cell 10: 451–459 annuus L. cv Margot). Plant Cell Environ 28: 750–759
Katsuhara M, Akiyama Y, Koshio K, Shibasaka M, Ogasa M, Miki N, Yoshikawa K (2010) Changes of
Kasamo K (2002) Functional analysis of water channels hydraulic conductivity during dehydration and
in barley roots. Plant Cell Physiol 43: 885–893 rehydration in Quercus serrata Thunb. and Betula
Larson PR, Isebrands JG (1971) The plastochron index as platyphylla var. japonica Hara: the effect of xylem
applied to developmental studies of cottonwood. Can structures. Tree Physiol 30: 608–617
J Forest Res 1: 1–11 Oono Y, Seki M, Nanjo T, Narusaka M, Fujita M, Satoh R,
Lee S, Chung GC, Zwiazek JJ (2009) Light decreases cell Satou M, Sakurai T, Ishida J, Akiyama K, Iida K,
hydraulic conductivity and turgor pressure in bundle Maruyama K, Satoh S, Yamaguchi-Shinozaki K,
sheath cells of tobacco (Nicotiana tabacum) leaves. Shinozaki K (2003) Monitoring expression profiles of
Plant Sci 176: 248–255 Arabidopsis gene expression during rehydration process
Li GW, Peng YH, Yu X, Zhang MH, Cai WM, Sun WN, after dehydration using ca. 7000 full-length cDNA
Su WA (2008) Transport functions and expression microarray. Plant J 34: 868–887
analysis of vacuolar membrane aquaporins in response Parent B, Hachez C, Redondo E, Simonneau T,
to various stresses in rice. J Plant Physiol 165: Chaumont F, Tardieu F (2009) Drought and abscisic
1879–1888 acid effects on aquaporin content translate into changes

Physiol. Plant. 2010


in hydraulic conductivity and leaf growth rate: a Siefritz F, Tyree M, Lovisolo C, Schubert A, Kaldenhoff R
trans-scale approach. Plant Physiol 149: 2000–2012 (2002) PIP1 plasma membrane aquaporins in tobacco:
Pearce DW, Millard S, Bray DF, Rood SB (2006) Stomatal from cellular effects to function in plants. Plant Cell 14:
characteristics of riparian poplar species in a semi-arid 869–876
environment. Tree Physiol 26: 211–218 Siemens JA, Zwiazek JJ (2003) Effects of water deficit stress
Robinson DA, Jones SB, Wraith JM, Or D, Friedman SP and recovery on the root water relations of trembling
(2003) A review of advances in dielectric and electrical aspen (Populus tremuloides) seedlings. Plant Sci 165:
conductivity measurement in soils using time domain 113–120
reflectometry. Vadose Zone J 2: 444–475 Sterky F, Bhalerao RR, Unneberg P, Segerman B,
Sack L, Streeter CM, Holbrook NM (2004) Hydraulic Nilsson P, Brunner AM, Campaa L, Jonsson Lindvall J,
analysis of water flow through leaves of sugar maple and Tandre K, Strauss SH, Sundberg B, Gustafsson P,
red oak. Plant Physiol 134: 1824–1833 Uhlen M, Bhalerao RP, Nilsson O, Sandberg G,
Sade N, Vinocur BJ, Diber A, Shatil A, Ronen G, Nissan H, Karlsson J, Lundeberg J, Jansson S (2004) A Populus EST
Wallach R, Karchi H, Moshelion M (2009) Improving resource for plant functional genomics. Proc Natl Acad
plant stress tolerance and yield production: is the Sci USA 101: 13951–13956
tonoplast aquaporin SlTIP2;2 a key to isohydric Stettler RF, Zsuffa L, Wu R (1996) The role of hybridization
to anisohydric conversion? New Phytol 181: in the genetic manipulation of Populus. In: Stettler RF,
651–661 Bradshaw HD Jr, Heilman PE, Hinckley TM (eds)
Sakr S, Alves G, Morillon R, Maurel K, Decourteix M, Biology of Populus and its Implications for Management
Guilliot A, Fleurat-Lessard P, Julien JL, Chrispeels MJ and Conservation. NRC Research Press, Ottawa,
(2003) Plasma membrane aquaporins are involved in Canada, pp 87–112
Street NR, Skogström O, Sjödin A, Tucker J,
winter embolism recovery in walnut tree. Plant Physiol
Rodrı́guez-Acosta M, Nilsson P, Jansson S, Taylor G
133: 630–641
(2006) The genetics and genomics of the drought
Sakurai J, Ishikawa F, Yamaguchi T, Uemura M,
response in Populus. Plant J 48: 321–341
Maeshima M (2005) Identification of 33 rice aquaporin
Suga S, Imagawa S, Maeshima M (2001) Specificity of the
genes and analysis of their expression and function.
accumulation of mRNAs and proteins of the plasma
Plant Cell Physiol 46: 1568–1577
membrane and tonoplast aquaporins in radish organs.
Sambrook J, Russell DW (2001) Molecular Cloning: A
Planta 212: 294–304
Laboratory Manual, 3rd Edn. Cold Spring Harbor
Tamura K, Dudley J, Nei M, Kumar S (2007) MEGA4:
Laboratory Press, Cold Spring Harbor, NY
molecular evolutionary genetics analysis (MEGA)
Secchi F, Lovisolo C, Uehlein N, Kaldenhoff R, Schubert S
software version 4.0. Mol Biol Evol 24: 1596–1599
(2007) Isolation and functional characterization of three
Tardieu F, Simonneau T (1998) Variability among species
aquaporins from olive (Olea europaea L.). Planta 225:
of stomatal control under fluctuating soil water status
381–392 and evaporative demand: modeling isohydric and
Secchi F, Maciver B, Zeidel ML, Zwieniecki MA (2009) anisohydric behaviours. J Exp Bot 49: 419–432
Functional analysis of putative genes encoding the PIP2 Toop GC, Davis JL, Annan AP (1980) Electromagnetic
water channel subfamily in Populus trichocarpa. Tree determination of soil water content using TDR: I.
Physiol 29: 1467–1477 Applications to wetting fronts and steep gradients. Soil
Seki M, Narusaka M, Abe H, Kasuga M, Yamaguchi- Sci Soc Am J 46: 672–678
Shinozaki K, Carninci P, Hayashizaki Y, Shinozaki K Touchette BW, Iannacone LR, Turner GR, Frank AR (2007)
(2001) Monitoring the expression pattern of 1300 Drought tolerance versus drought avoidance: a
Arabidopsis genes under drought and cold stresses by comparison of plant-water relations in herbaceous
using a full-length cDNA microarray. Plant Cell 13: wetland plants subjected to water withdrawal and
61–72 repletion. Wetlands 27: 656–667
Seki M, Narusaka M, Ishida J, Nanjo T, Fujita M, Oono Y, Tuskan GA, DiFazio S, Jansson S, Bohlmann J, Grigoriev I,
Kamiya A, Nakajima M, Enju A, Sakurai T, Satou M, Hellsten U, Putnam N, Ralph S, Rombauts S,
Akiyama K, Taji T, Yamaguchi-Shinozaki K, Carninci P, Salamov A, Schein J, Sterck L, Aerts A, Bhalerao RR,
Kawai J, Hayashizaki Y, Shinozaki K (2002) Monitoring Bhalerao RP, Blaudez D, Boerjan W, Brun A, Brunner A,
the expression profiles of 7000 Arabidopsis genes under Busov V, Campbell M, Carlson J, Chalot M, Chapman J,
drought, cold and high-salinity stresses using a Chen G-L, Cooper D, Coutinho PM, Couturier J,
full-length cDNA microarray. Plant J 31: 279–292 Covert S, Cronk Q, Cunningham R, Davis J, Degroeve S,
Shapiguzov AY (2004) Aquaporins: structure, systematics, Déjardin A, dePamphilis C, Detter J, Dirks B, Dubchak I,
and regulatory features. Russ J Plant Physiol 51: Duplessis S, Ehlting J, Ellis B, Gendler K, Goodstein D,
127–137 Gribskov M, Grimwood J, Groover A, Gunter L,

Physiol. Plant. 2010


Hamberger B, Heinze B, Helariutta Y, Henrissat B, to regulate their subcellular localization. Proc Natl Acad
Holligan D, Holt R, Huang W, Islam-Faridi N, Jones S, Sci USA 104: 12359–12364
Jones-Rhoades M, Jorgensen R, Joshi C, Kangasjärvi J, Zhang R, Verkman AS (1991) Water and urea permeability
Karlsson J, Kelleher C, Kirkpatrick R, Kirst M, Kohler A, properties of Xenopus oocytes: expression of mRNA
Kalluri U, Larimer F, Leebens-Mack J, Leplé JC, from toad urinary bladder. Am J Physiol 260: C26–C34
Locascio P, Lou Y, Lucas S, Martin F, Montanini B,
Napoli C, Nelson DR, Nelson C, Nieminen K,
Nilsson O, Pereda V, Peter G, Philippe R, Pilate G, Supporting Information
Poliakov A, Razumovskaya J, Richardson P, Rinaldi C, Additional Supporting Information may be found in the
Ritland K, Rouzé P, Ryaboy D, Schmutz J, Schrader J,
online version of this article:
Segerman B, Shin H, Siddiqui A, Sterky F, Terry A,
Tsai C-J, Uberbacher E, Unneberg P, Vahala J, Wall K, Appendix S1. AQP sequences from Populus tri-
Wessler S, Yang G, Yin T, Douglas C, Marra M, chocarpa, Arabidopsis thaliana and Physcomitrella
Sandberg G, Van de Peer Y, Rokhsar D (2006) The patens that were included in this study.
genome of black cottonwood, Populus trichocarpa
(Torr. & Gray). Science 313: 1596–1604 Appendix S2. List of gene-specific primers used in this
Tyerman SD, Niemietz CM, Bramley H. 2002. Plant study.
aquaporins: multifunctional water and solute channels
with expanding roles. Plant Cell Environ 25: 173–194. Appendix S3. Expression levels of the reference gene
Voicu MC, Cooke JEK, Zwiazek JJ (2009) Aquaporin gene EF1α, a gene from the elongation factor-1 alpha gene
expression and apoplastic water flow in bur oak family.
(Quercus macrocarpa) leaves in relation to the light
Appendix S4. AQP gene models from version 2.0 of the
response of leaf hydraulic conductance. J Exp Bot 60:
4063–4075
Populus trichocarpa genome.
Wan X, Zwiazek J (1999) Mercuric chloride effects on root Please note: Wiley-Blackwell are not responsible for
water transport in aspen seedlings. Plant Physiol 121: the content or functionality of any supporting materials
939–946
supplied by the authors. Any queries (other than missing
Zelazny E, Borst JW, Muylaert M, Batoko H, Hemminga
material) should be directed to the corresponding author
MA, Chaumont F (2007) FRET imaging in living maize
for the article.
cells reveals that plasma membrane aquaporins interact

Edited by M. Uemura

Physiol. Plant. 2010

You might also like