ArticlePDF Available

Expression profiling of the Arabidopsis annexin gene family during germination, de-etiolation and abiotic stress

Authors:

Abstract and Figures

Annexins are a multigene family in most plant species and are suggested to play a role in a wide variety of essential cellular processes. In Arabidopsis thaliana there are eight different annexins (AnnAt1-8), which range from 29% to 83% in deduced amino acid sequence identity. As a first step toward clarifying the individual functions of these annexins, in this study we have used quantitative real time reverse transcription PCR to assess their differential expression in different tissues or after different stimuli. We determined which annexins are expressed during germination and early seedling growth by assaying annexin expression levels in dry and germinating seeds and in 7-day-old light-grown seedlings. Our results indicate that transcripts for all eight annexins are present in germinating seeds and that transcript levels for all the annexins increase by 7 days of normal growth. We assayed transcript levels in dark grown roots, cotyledons, and hypocotyls and found that the relative abundance of each annexin varied in these dark-grown tissues. We also examined the effects of red and far red light treatments on annexin expression in 5.5-day-old etiolated seedlings. Light treatments significantly altered transcript levels in hypocotyls and cotyledons for only two members of the gene family. Finally, we monitored annexin expression changes in response to a variety of abiotic stresses. We found that the expression of most of the Arabidopsis annexin genes is differentially regulated by exposure to salt, drought, and high- and low-temperature conditions, indicating a likely role for members of this gene family in stress responses.
Content may be subject to copyright.
Research paper
Expression profiling of the Arabidopsis annexin gene family
during germination, de-etiolation and abiotic stress
A. Cantero
a,1
, S. Barthakur
b,1
, T.J. Bushart
a
, S. Chou
c
, R.O. Morgan
d
, M.P. Fernandez
d
,
G.B. Clark
a
, S.J. Roux
a,*
a
Department of Molecular Cell and Developmental Biology, University of Texas, Austin, Texas 78713 USA
b
National Research Center on Plant Biotechnology, Indian Agricultural Research Institute, New Delhi 110 012 India
c
Department of Molecular and Cell Biology, University of California at Berkeley, Berkeley, California 94720-3200 USA
d
Department of Biochemistry and Molecular Biology, Edificio Santiago Gascon, Faculty of Medicine, University of Oviedo, 33006 Oviedo, Spain
Received 25 August 2005
Available online 28 February 2006
Abstract
Annexins are a multigene family in most plant species and are suggested to play a role in a wide variety of essential cellular processes. In
Arabidopsis thaliana there are eight different annexins (AnnAt1-8), which range from 29% to 83% in deduced amino acid sequence identity. As a
first step toward clarifying the individual functions of these annexins, in this study we have used quantitative real time reverse transcription PCR
to assess their differential expression in different tissues or after different stimuli. We determined which annexins are expressed during germina-
tion and early seedling growth by assaying annexin expression levels in dry and germinating seeds and in 7-day-old light-grown seedlings. Our
results indicate that transcripts for all eight annexins are present in germinating seeds and that transcript levels for all the annexins increase by
7 days of normal growth. We assayed transcript levels in dark grown roots, cotyledons, and hypocotyls and found that the relative abundance of
each annexin varied in these dark-grown tissues. We also examined the effects of red and far red light treatments on annexin expression in 5.5-
day-old etiolated seedlings. Light treatments significantly altered transcript levels in hypocotyls and cotyledons for only two members of the gene
family. Finally, we monitored annexin expression changes in response to a variety of abiotic stresses. We found that the expression of most of the
Arabidopsis annexin genes is differentially regulated by exposure to salt, drought, and high- and low-temperature conditions, indicating a likely
role for members of this gene family in stress responses.
© 2006 Elsevier SAS. All rights reserved.
Keywords: Annexin; Arabidopsis thaliana; Real time PCR; Expression; Abiotic stress
1. Introduction
Calcium is a universal signal in eukaryotic cells, and diverse
signaling pathways in plants include an increase in the concen-
tration of cytosolic Ca
2+
([Ca
2+
]
cyt
) in at least one of the steps.
Included among the calcium-binding proteins that transduce
these calcium signals into adaptive responses is the annexin
gene family [8,11]. Plant annexins have been localized at the
cell periphery of highly secretory cell types where they are
hypothesized to play a role in the Golgi-mediated secretion of
new wall materials and plasma membrane that occurs during
growth and development. Additionally, plant annexins have
been found in many other cellular locales including the cyto-
plasm, the vacuole and nucleus. They appear to be a multifunc-
tional gene family whose members play roles in a number of
diverse physiological processes. For example, the cellular dis-
tribution of certain annexins is changed in response to touch
[33], gravity [7] and diurnal cycles [19], and in Medicago an
annexin is involved in Nod signaling [4]. Plant annexins have
also been implicated in imparting tolerance to various abiotic
stresses [15,22]. Recent analyses of two Arabidopsis annexin
T-DNA insertional mutants provided evidence that these an-
nexins play an important role in germination during salt and
osmotic stress [26].
www.elsevier.com/locate/plaphy
Plant Physiology and Biochemistry 44 (2006) 1324
Abbreviations: RT, reverse transcription; MD, monocotdicot; APT1,
adenosine phosphoribosyl transferase.
*
Corresponding author. Tel.: +1 512 471 4238; fax: +1 512 232 3402.
E-mail address: sroux@uts.cc.utexas.edu (S.J. Roux).
1
These authors contributed equally to this work.
0981-9428/$ - see front matter © 2006 Elsevier SAS. All rights reserved.
doi:10.1016/j.plaphy.2006.02.002
There are eight different annexins in Arabidopsis, some of
which are likely to have unique individual functions [9]. Data
from individual annexin studies as well as transcriptome stu-
dies have established that plant annexin expression is con-
trolled by a number of environmental and developmental sig-
nals. Previously, we have shown that two of these annexins,
AnnAt1 and AnnAt2, have differential in situ RNA localization
patterns that are temporally and developmentally regulated [9].
We also noted differences in predicted functional motifs be-
tween the different Arabidopsis annexins.
Expression profiling has become an important tool for gain-
ing insights into the partitioning of functions and redundancy
within a multigene family [30]. Global surveys by microarray
analyses have provided a rough glimpse of how transcript
abundance of most Arabidopsis genes, including members of
the annexin gene family, changes under different circumstances
of development or environmental stimuli. However, recent re-
sults point to an especially important role for annexins in
growth control and in certain stress responses. Since none of
the global surveys carried out thus far were specifically fo-
cused on the role of annexins in these processes, they do not
provide adequate detail to resolve whether specific members of
the annexin family are differentially regulated in rapidly grow-
ing tissues or in tissues under abiotic stress. To address this
question we decided to use the quantitative real time RT-PCR
technique.
Quantitative real time RT-PCR allows quantification of the
transcript levels of different members of a conserved gene fa-
mily, all sharing a high percentage of identity, with both high
sensitivity and specificity [14]. Use of this technique facilitates
a determination of the absolute and relative level of any tran-
script of any type. Even transcripts with very high or extremely
low levels of expression can be quantified accurately. The ac-
curacy and the reliability of the real time PCR have previously
been demonstrated in studies on human genotyping and patho-
gen detection [17,12].
In plants the quantitative real time PCR technique has been
used to determine the number of T-DNA insertions in trans-
genic plants [20,37], to detect the presence of genetically mod-
ified organisms in food [18], and to quantify the transcript le-
vels in plant organs [25,28]. Particularly relevant to the present
studies, this technique has been successfully used to analyze
closely related genes in many different families of plant en-
zymes [31,38,39] and for expression profiling of signal trans-
duction protein families such as the entire Arabidopsis Shaggy-
like kinase multigene family [5], the CTR1 gene family in-
volved in ethylene signaling [1], and over 1400 Arabidopsis
transcription factor transcripts [10]. Quantitative real time RT-
PCR is also well suited to performing expression profiling in
response to environmental stimuli. For example, it was used to
analyze the expression of an Arabidopsis gene family encoding
plasma membrane aquaporins in response to abiotic stresses,
which allowed the authors to determine the stress-related bio-
logical function of certain gene members of this family [21].
In the experiments reported here we use quantitative real
time RT-PCR to profile how the transcript abundance of all
eight annexin genes change during seed germination, seedling
growth and development, and stress responses. Our results pro-
vide a clearer picture of which annexin genes respond most
strongly in each of these scenarios, and provide an important
database for directing future mutant and localization studies.
2. Results
2.1. Genomic, structural and phylogenetic analysis
of annexins in Arabidopsis
The schematic in Fig. 1 depicts the gene structures for all
eight annexins, and highlights functionally important features.
Notably, AnnAt3 and AnnAt4 are located on chromosome 2 in
tandem and AnnAt6 and AnnAt7 are located on chromosome 5
in tandem. AnnAt3 and AnnAt4 may share a 5promoter region
and this could result in dual regulation of their transcripts. An-
other interesting observation arising from the gene structure
analyses is that four of the Arabidopsis annexins, AnnAt3,An-
nAt4,AnnAt5, and AnnAt8, have congruent gene structures,
each having six exons. In contrast, analysis of the gene struc-
ture for the other four annexins shows the common feature of
intron losses. AnnAt1 has lost the final three introns, AnnAt2
lacks a single final intron, and both the members of the tandem
pair, AnnAt6 and AnnAt7, have lost their final two introns.
Phylogenetic analysis of the eight Arabidopsis (dicot) an-
nexins (identical to diploid cotton, not shown), 10 Oryza
(monocot) annexins, and two Ceratopteris annexins as the out-
group resulted in a fully resolved tree (Fig. 2). All bifurcations
were supported with at least 70% maximum likelihood or boot-
strap values. Branch lengths are proportional to the amount of
evolution (i.e. the number of amino acid replacements per site).
The presence of direct orthologs in Arabidopsis and rice sug-
gests that several annexins are ancient, predating the monocot-
dicot (MD) separation 200 million years ago (Fig. 2). For ex-
ample, Arabidopsis and rice share an orthologous tandem pair,
with AnnAt3 and AnnAt4 being direct orthologs of ANXD14
and ANXD16 on chromosomes 2 and 5 in the respective spe-
cies. Several more recent lineage-specific gene duplications
have occurred in Arabidopsis and rice following the MD se-
paration. One such lineage-specific amplification in Arabidop-
sis resulted in paralogs AnnAt1,AnnAt2,AnnAt6, and AnnAt7
and at least four recent gene duplications within the rice line-
age created additional paralogs with distinct subfamily num-
bering.
In terms of protein evolution within the Arabidopsis annex-
in clade, AnnAt1,AnnAt2,AnnAt6, and AnnAt7 represent rela-
tively recent lineage-specific duplications and as such are ex-
pected to show the least divergence in protein structure and
function. The other tandem pair of AnnAt3 and AnnAt4 has a
longer history of divergence occurring before MD separation,
but possible concerted evolution due to their physical proxi-
mity may have limited the extent of their divergence.
In general, the primary sequences for plant annexin genes
are fairly divergent from each other. AnnAt2,AnnAt6, and An-
nAt7 are the most similar to each other, showing 8086% iden-
A. Cantero et al. / Plant Physiology and Biochemistry 44 (2006) 132414
tity at the nucleotide level and 7683% identity at the deduced
amino acid level. AnnAt3 and AnnAt4 appear to be the most
divergent at the amino acid level. The predicted protein se-
quences for the gene family range from 316322 amino acids
with molecular masses between 34 and 39 kDa. When the core
repeat regions for all eight of the Arabidopsis annexins are
compared to the annexin consensus sequences taken from
many annexin core repeat regions as analyzed by Barton et
al. [2], it is evident that the structural repeats that help define
this gene family are present in Arabidopsis annexins. Overall it
is clear that the first and fourth repeats show the highest level
of conservation while the second and third repeats are the most
divergent.
In annexins the type II calcium-binding site is determined
by the conserved glycine-X-glycinethreonine loop followed
42 amino acids downstream of the first glycine residue by a
glutamic acid or an aspartic acid residue. Regarding predicted
type II calcium-binding sites, many plant annexins have a
Fig. 1. Schematic diagram of the genomic structures of the annexin gene family in Arabidopsis. Exons are depicted as white boxes and introns as lines. Untranslated
regions (UTRs) are represented as black boxes and non-coding regions between tandem genes as gray boxes.
Fig. 2. Molecular evolution of annexins in the completed genomes of Arabidopsis thaliana (thale cress) and Oryza sativa japonica (rice). The phylogenetic tree was
determined by maximum likelihood analysis with release 3.0 of the IQPNNI program [34]. Known fern annexins were used as outgroup for the comparison of
aligned, full-length protein sequences using eight gamma rates to account for site-specific variation. The tree was fully resolved, with confidence values for the
branching topology all above 90 and congruency with the neighbor-joining, bootstrap tree analysis by MEGA 3.1 [23] (not shown). Bifurcations containing solid
circles represent common gene duplications, while open circles signify intra-lineage gene duplications. Nomenclature for members of the plant annexin family
(ANXD) includes taxon names based on phylogenetic analysis of all known plant annexins (unpublished) together with the systematic names given by the
corresponding genome sequencing projects. Finished sequence data are available from The Arabidopsis Information Resource (http://www.arabidopsis.org/), the
International Rice Genome Sequencing Project (http://rgp.dna.affrc.go.jp/IRGSP/) and NCBI-GenBank for the fern sequences (AF308588, AF308589).
A. Cantero et al. / Plant Physiology and Biochemistry 44 (2006) 1324 15
fairly well conserved site in either the first or fourth repeat.
AnnAt3 also appears to have a binding loop in the first and
fourth repeats, but its more divergent, tandem duplicate An-
nAt4 is essentially devoid of type II calcium binding sites
(Fig. 3).
2.2. Relative quantitation of annexin transcripts under normal
growth conditions
Raw C
T
values for the constitutively expressed housekeep-
ing gene, adenosine phosphoribosyl transferase (APT1), were
compared to each of the eight annexins in 7-day-old seedlings
(data not shown). Each of the eight genes shows significant
expression, indicating that all members of the family are tran-
scriptionally active under normal growth conditions.
In Fig. 4 we determined the relative quantitation of annexin
transcripts under normal growth conditions in dry seeds, ger-
minating seeds, and 7-day-old seedlings by using quantitative
real time reverse transcription polymerase chain reaction. For
this comparison, annexin levels were normalized to a grand
calibrator of AnnAt1 in dry seeds in order to allow for propor-
tional comparisons of annexins within a single time point or of
an annexin across time points. We were able to clearly distin-
guish quantitative differences in the relative transcript level of
the eight annexin family members. Every annexin was detect-
able in whole seeds and seedlings regardless of the time point.
With the exception of AnnAt4 transcript, all annexin transcript
levels decrease 26 hours after sowing, and this is followed by
an increase in transcript levels for all eight annexins after 7 days
of growth.
2.3. Relative quantitation of annexin transcripts in dark grown
tissues
Fig. 5 shows the RNA levels of all eight annexins as ana-
lyzed in the steady-state condition of 5.5-day-old dark-grown
seedlings. Generally, highest expression levels were seen in
roots and hypocotyls, while cotyledons had lower expression
levels. RNA levels for AnnAt4 are fairly even throughout the
plant while AnnAt5, 6, and 7 show the largest disparities be-
tween root tissue as compared to cotyledon (27.0-, 45.7-, 11.2-
fold decreases, respectively).
2.4. Red and Far-red light effects on annexin expression
Most of the light effects on annexin expression in germinat-
ing seedlings are not significant. However, the general trends
of expression changes are consistent, showing an increase in
Fig. 3. Sequence alignment of Arabidopsis annexin repeats. The amino acids highlighted at the 21 numbered positions are in agreement with the residues that have
conserved amino acid properties in each of the four annexin repeats of 22 annexin sequences as analyzed by Barton et al. [2]. The two asterisks (in repeat 1 of
AnnAt4 and in repeat 2 of AnnAt3) indicate locations of short amino acid insertions in these two annexin repeats that are not shown in this alignment. The amino
acids in the insertion for repeat 1 of AnnAt4 are EERAFEKCHDH and for repeat 2 of AnnAt3 are KKSLE. According to Barton et al. [2] in the 88 animal repeats
analyzed, the following amino acids occurred in the following sequence positions (P#) of repeats: P#11: A V; P#14: I L; P#17: A S; P#23: G; P#24: T V S; P#29:L I
W F; P#30: V I L T N; P#32: I L V N; P#33: M L I V F; P#34: T A C G V; P#36:R; P#37: S A T N; P#44: I V L T; P#48: Y F; P#56: M I L; P#60: I L V M; P#65: V
L I F T; P#67: G F; P#69: L F M Y I; P#72: V L A I M C G T; P#73: F L M I V.
A. Cantero et al. / Plant Physiology and Biochemistry 44 (2006) 132416
annexin expression after light exposure in cotyledons, and a
decrease in annexin levels in hypocotyls (Fig. 6). Within hypo-
cotyls, AnnAt5 is the only annexin that shows a change in RNA
levels significant to the 95% confidence level (Fig. 6A). An-
nAt5 shows a significant up-regulation in hypocotyls (sixfold)
that is reversed by far-red light treatment (sevenfold) (Fig. 6A).
AnnAt2 and AnnAt6 show a trend of decreased RNA levels
under red light with possible photoreversibility by far-red light
treatment (Fig. 6A). Within cotyledon tissues, AnnAt6 is the
only annexin that has statistically significant light-induced
changes (P< 0.05) (Fig. 6B). It shows significant up-regulation
after red light exposure as compared to dark grown seedlings,
and this up-regulation is reversed by far-red light treatment. In
contrast, AnnAt7 is up-regulated (10-fold) but this change is
not statistically significant and is not reversed by far-red light
treatment. Although it is true that physiological significance is
not based on magnitude of expression, AnnAt6 does show the
largest up-regulation (35-fold) compared to all other annexins,
which tend towards smaller (< 10-fold) expression changes.
2.5. Differential expression of annexin genes in whole
seedlings under various abiotic stresses
Comprehensive expression profiling of the eight members
of the annexin family in 7-day-old seedlings was carried out
under a series of abiotic treatments. Quantitative measurements
were done comparing expression of each gene under normal
growth conditions. The series of stresses included modifica-
tions of temperature down to 4 °C and up to 37 °C, dehydrat-
ing the plants for 2 hours, and modifying the salt concentration
Fig. 4. Changes in annexin RNA levels during germination. Real time RT-PCR
results are shown comparing RNA levels for all eight annexins within and
across three time points. Seeds were treated with cold stratification at 4 °C for
3 days before sowing on plates. Dry seed corresponds to unimbibed seeds,
germinated to whole seedlings collected at 26 hours post-sowing on nutrient
media plates, and 7-day-old seedlings to whole seedlings collected at 7 days
post-sowing on nutrient media plates. Total RNA was isolated from whole
seedlings collected at 26 hours post-sowing or after 7 days of growth. Results
are normalized in two directions, within treatment condition to AnnAt1 and
across time points to dry seed, giving a grand calibrator of AnnAt1 in dry seeds
(value of 1). With AnnAt1-dry seed as the grand calibrator, comparisons are
proportional both within any single time point, as well as with a single annexin
across time points. In order to keep both increases and decreases in relative fold
differences in RNA levels proportional the Y axis is presented in log scale.
Error bars represent standard error of the mean.
Fig. 5. Relative levels of annexins in dark grown tissues. Real time RT-PCR
results are shown comparing RNA levels for each annexin across three tissue
types. Roots, hypocotyls, and cotyledons were harvested for RNA isolation
from 5.5-day-old etiolated seedlings. Results are normalized to annexin levels
in roots (value of 1). In order to keep both increases and decreases in relative
fold differences in RNA levels proportional the Y axis is presented in log scale.
Error bars represent standard error of the mean.
Fig. 6. Effects of Red and far-red light on annexin RNA levels in cotyledons
and hypocotyls. A. Real time RT-PCR results are shown comparing RNA
levels of each annexin in hypocotyls 30 min after three different light
treatments, which were: Dark, plants grown for 5.5 days in complete darkness;
Red, dark grown plants exposed to red light for 2 min; Red + far-red, dark
grown plants exposed to red light for 2 min followed by far-red light for 3 min.
Results are normalized within each annexin to levels in dark treated plants
(value of 1). In order to keep both increases and decreases in relative fold
differences in RNA levels proportional the Y axis is presented in log scale.
Error bars represent standard error of the mean. B. Real time RT-PCR results as
in panel A, but showing changes in annexin RNA levels from cotyledon tissue.
A. Cantero et al. / Plant Physiology and Biochemistry 44 (2006) 1324 17
by adding 250 mM NaCl for 2 hours. Fig. 7 shows the relative
expression changes for each annexin gene family member and
the range of the transcript levels observed in the conditions
described above. The error bars represent the 95% confidence
intervals for each gene under different treatments. All the
changes noted below are statistically significant. Table 1 shows
a summary of the relative fold differences in the transcript le-
vels for the eight annexin genes after various stress treatments.
AnnAt1 and AnnAt2 responded quite differently to these
four abiotic stresses. AnnAt1 message levels were increased
by all four treatments with the highest induction observed in
response to NaCl treatment (42-fold) and dehydration treat-
ment (39-fold). In contrast, AnnAt2 message levels were
down-regulated in response to salt and drought treatments as
well as cold treatment. Under high temperature for 2 h at
37 °C, AnnAt2 message level was significantly up-regulated
(43-fold) while AnnAt1 message level increased only margin-
ally (less than twofold).
AnnAt3 was up-regulated by cold temperature (sevenfold)
and dehydration (13-fold) while remaining almost unchanged
in its transcript level under NaCl stress. In contrast it was down
regulated 19-fold by heat exposure. AnnAt4 responded strongly
at the transcriptional level to NaCl treatment showing a very
high rate (383-fold) compared to control expression under dis-
tilled water treatment controls, while changing little in response
to cold treatment and decreasing in response to both heat and
dehydration treatments (eightfold and fivefold, respectively).
AnnAt5 showed higher transcript levels after all four abiotic
stress treatments with highest transcript abundance in response
to saline treatment (64-fold).
Qualitatively, AnnAt6 and AnnAt8 showed similar response
patterns, being positively affected by high temperature, dehy-
dration and NaCl, and relatively unaffected by low tempera-
ture. Quantitatively, AnnAt6 showed a moderate increase in ex-
pression, and AnnAt8 showed high increases in transcript
abundance: 434 times more under NaCl stress and 175 times
under dehydration.
There was not a significant change in expression of AnnAt7
under dehydration stress, but 42 times more expression under
NaCl stress. Like all the other annexins except AnnAt2 and
AnnAt3,AnnAt7 was positively modulated under NaCl stress
when exposed to 250 mM NaCl solution for 2 hours. AnnAt7
and AnnAt5 were the only two genes that were strongly in-
duced by both high temperature and low temperature.
In this study we also used reference genes known to be spe-
cifically involved in abiotic stress tolerance and these genes
were quantified under similar experimental conditions in order
to validate our stress treatment induced annexin gene expres-
sion changes. Results indicate that Rd29 was up-regulated un-
der all the four different stress treatments, being positively
modulated by 630-fold by dehydration and 223-fold by NaCl
treatment (data not shown). Proline synthesizing enzyme P5CS
was induced by dehydration (512-fold) and by NaCl (79-fold),
whereas proline catabolizing enzyme PDH showed marginal
induction under the stress treatments as expected (data not
shown).
3. Discussion
The temporal and spatial expression of genes and the sub-
cellular localization of the proteins encoded by those genes
provide important clues to their functions. As expected for a
multifunctional multigene family, annexins showed differential
expression during growth and development and in response to
specific environmental stimuli. Previously, we have noted that
for AnnAt1 and AnnAt2, mRNA and protein localization pat-
terns are almost identical in seedlings, possibly indicating little
or no post-translational regulation of expression for these two
annexins during normal growth and development [5]. Post-
Fig. 7. Stress effects on annexin levels. Real time RT-PCR results are shown
comparing annexin RNA levels under stress conditions as compared to
untreated controls. All plants were 7-day-old with various treatments ending at
the time of RNA collection. Treatments and durations are as follows: 4 °C,
plants held at an ambient temperature of 4 °C for 2 hours; 37 °C, plants held at
an ambient temperature of 37 °C for 2 hours; Dehyd, dehydration on bench for
2 hours; NaCl, plants transferred to nutrient plates containing 250 mM NaCl for
2 hours. Error bars represent 95% confidence intervals calculated as described
in Section 4. Data is presented as the fold change in RNA levels as normalized
to Control untreated 7-day-old plants (value of 1). In order to keep both
increases and decreases in relative fold differences in RNA levels proportional
the Y axis is presented in log scale. All changes due to treatment condition are
statistically significant at the 95% confidence level and as such control data is
omitted for visual simplicity.
Table 1
Relative fold differences for each annexin in response to various stress
treatments
Gene 4 °C 37 °C Dehydration 250 mM NaCl
AnnAt1 13.3 2 39.2 44.2
AnnAt2 7.2 43.5 11.4 11
AnnAt3 718.9 13.1 1.3
AnnAt4 28.1 5.5 383.2
AnnAt5 5.8 8.4 4.4 63.7
AnnAt6 2.4 11.1 25.8 13.7
AnnAt7 5.3 20.8 1.5 42.2
AnnAt8 2.3 2.3 175.4 433.5
A. Cantero et al. / Plant Physiology and Biochemistry 44 (2006) 132418
translational regulation of active protein levels through signal-
mediated protein degradation via proteasomes has recently
been shown to be an important and commonly used level of
gene control in plant cells. In fact, it has been suggested that
altered protein levels of AnnAt1 in seeds in response to stress
may be due to proteasomal protein degradation [26]. In this
study we focus only on mRNA expression levels for the Ara-
bidopsis annexin gene family.
Although the results of many plant microarray studies that
include information on annexin expression patterns are avail-
able, the high throughput methodology of microarray analysis
often does not distinguish between related members of a gene
family.
Quantitative real time RT-PCR is more accurate at detecting
relative expression over a larger linear range because the scan-
ning and processing involved in microarray analysis often
eliminates the lowest and highest ends of its detectable range
[10]. For these reasons we used quantitative real time RT-PCR
as a precise, sensitive and reproducible technique to measure
the expression profiles of the annexin multigene family.
Gene structure and phylogenetic analyses of the Arabidopsis
annexins reveal genome-based differences within this gene fa-
mily. For example, these analyses show that the feature of in-
tron loss in the family is a shared characteristic of the more
recent, evolutionary clade. Sequence analysis of the core re-
peats indicate that second and third repeats may be sources of
functional differences between annexins as these repeat regions
have diverged more than the first and fourth repeats.
All eight annexins are detectable in dry seeds. This may be
indicative that germinating seeds use their RNA stores for in-
itial translation needs. A germinating seed has to start growth
rapidly despite its limited resources, so, as with other messages
and proteins, annexins may be part of a pool of pre-synthesized
material allowing for a quicker initiation of growth.
Increase in annexin transcript levels between a dry seed and
a germinated plant could be expected for a gene encoding a
growth-related protein, because transitioning from a dormant
mature embryo to a metabolically active seedling requires sig-
nificant growth. Surprisingly, seven of the eight annexins show
a decrease in RNA levels 26 hours after sowing as compared to
a dry seed. The decreased levels of annexin messages 26 hours
after sowing seems counterintuitive, since growth is not slow-
ing down during this period. Perhaps levels of annexin protein
optimal for growth have been reached at this time, reducing the
need to maintain high message levels. An alternative explana-
tion for this trend of down-regulation of annexin message le-
vels during germination would be that increases in annexin
message levels do occur in certain cells or tissues but we are
missing these changes by averaging message levels over all the
tissues in the germinating seed. In situ RNA localization of
each annexin, an extension of our previous report on cell-type
specific expression patterns in germinating seedlings for An-
nAt1 and AnnAt2 [9], would provide a rigorous test of this al-
ternative explanation.
It appears that all eight annexin transcripts are fairly abun-
dant after 7 days of growth in the light. This may indicate that
the whole annexin gene family contributes importantly to the
diverse cellular functions needed for seedling growth. These
results also show that the quantitative real time RT-PCR meth-
od is suitable for the identification and quantitative comparison
of the relative transcript levels of different annexin encoding
genes.
Localization of RNA across a plant can give clues as to the
roles and importance of various members of a gene family. In
order to survey the distribution of annexin transcripts, RNA
levels were examined with real time RT-PCR on tissues taken
from dark grown seedlings. All of the annexins exhibit the
trend of having higher levels in roots than in cotyledons. This
fits nicely with previous histological investigations indicating
that AnnAt1, for example, is found primarily in the roots of
young germinating seedlings [6,9]. The observed trend also
meshes with the general notion that annexins are involved in
growth, since in an etiolated plant, the majority of cell expan-
sion is occurring in the roots and hypocotyls, while the cotyle-
dons exhibit much less growth. We have previously used stan-
dard RT-PCR to assay annexin transcript levels in various light
grown tissues of older plants, but not many comparisons be-
tween the previous data and the present data can be made [9].
AnnAt4 transcript levels are similar in all three tissues ex-
amined and show the weakest changes during the germination
process with only a slow increase over time. In contrast to An-
nAt4,AnnAt5,AnnAt6, and AnnAt7 are not expressed evenly in
all three tissues within young etiolated seedlings. Each of these
genes shows strongest expression in roots and hypocotyls with
markedly decreased levels in cotyledons. Potential roles for
these annexins are discussed in relation to light effects on their
expression.
Light triggers global changes in gene expression in young
seedlings, and recent microarray studies have begun to elucidate
some of the specific transcript changes regulated by phyto-
chrome during seedling de-etiolation [32]. Because of potential
involvement of annexins in mediating growth and development
we examined phytochrome-mediated changes in annexin ex-
pression in cotyledons and hypocotyls, two tissues with differ-
ent growth responses to light. The growth changes induced by
red light are very rapid, so we examined the early gene expres-
sion changes that could be detected after only 30 min, on the
assumption that these changes would be the most influential in
mediating the light effects. At this early time point the RNA
levels of AnnAt5 in hypocotyls changed in a significant manner,
increasing sevenfold after red light, and this change was re-
versed by far-red light, indicating that AnnAt5 is among the
early-responsegenes rapidly regulated by phytochrome in this
tissue. This result suggests that AnnAt5 could be participating in
cellular changes leading to growth inhibition.
In cotyledons, where red light induces cell division and cell
expansion, red light induces a statistically significant 35-fold
increase in transcript abundance for AnnAt6, and far-red light
reverses this effect, making AnnAt6 a second annexin gene ra-
pidly regulated by phytochrome in seedlings. However, be-
cause this light-induced expression change happens in cotyle-
dons, AnnAt6, unlike AnnAt5 is associated with growth
A. Cantero et al. / Plant Physiology and Biochemistry 44 (2006) 1324 19
promotion rather than growth inhibition. Interestingly, AnnAt6
expression is down-regulated eightfold by red light in hypoco-
tyls. This change is not statistically significant because of the
large standard error, but if additional tests show this trend to be
significant they would support the conclusion that rapid up-
and down-regulation of growth in different seedling tissues
by red light closely parallels its rapid up- and down-regulation
of AnnAt6 transcript levels, suggesting an important role for
AnnAt6 in light-regulated growth.
Both red light and far-red light treatments of seedlings have
only minor effects on the transcript abundance of all other an-
nexin genes in cotyledons and hypocotyls, suggesting that
most annexins are not among the genes rapidly regulated by
phytochrome. Whether any of these other annexins are among
the late-responsegenes that are regulated by phytochrome
remains to be tested.
Prior studies have demonstrated that AnnAt1 and AnnAt2
expression and localization are positively correlated with secre-
tion and growth activities [6], and the results discussed above
could indicate that AnnAt6 expression may also be positively
correlated with growth. Testing this hypothesis would require
developing specific antibody tools, as has been done for An-
nAt1 and AnnAt2 [6]. However, the negative correlation of An-
nAt5 transcript expression with hypocotyl growth seems to re-
fute the conclusion that all annexins are engaged in activities
that favor growth, and lends further evidence that annexin
functions are not redundant.
One emerging theme from recent plant annexin studies is
that annexins participate in abiotic stress responses, therefore
we wanted to examine expression patterns for the entire Arabi-
dopsis annexin gene family in response to a variety of abiotic
stresses. After 2 hours exposure to different abiotic stress con-
ditions most of the annexin genes were differentially regulated.
The rapidity of these responses makes it less likely that they
arose from secondary effects. There is evidence supporting a
role for calcium signaling in the responses to the different abio-
tic stresses we used [35], thus the differential regulation of ex-
pression for the annexin family of calcium-binding proteins is
probably a meaningful component of these responses.
As discussed earlier Lee et al. have described evidence sup-
porting a role for two Arabidopsis annexins, AnnAt1 and An-
nAt4, in germination under stress conditions of high osmoti-
cum and salt [26]. In their study, treatment with NaCl
appeared to decrease levels of immunodetectable AnnAt1 and
induced cytoplasmic AnnAt1 to move to the membrane frac-
tion. Their RNA gel blot analysis of AnnAt1 transcript levels
in roots of 2-week-old seedlings that were responding to
250 mM NaCl indicated no change in AnnAt1 transcript levels
at 2 hours or later time points. Based on these results, the
authors suggest that salt treatment may cause AnnAt1 to be
targeted to the proteasome for degradation.
In contrast to the Lee et al. study [26], our results show
more than a 40-fold increase in AnnAt1 transcript levels in
whole 7-day-old seedlings 2 hours after treatment with
250 mM NaCl. The difference between our results may be
due to the use of different developmental stages, different tis-
sue sources, or different techniques. However, in order to un-
derstand the potential role of AnnAt1 in response to salt stress,
this discrepancy in reports on its expression will need to be
resolved.
AnnAt1 has previously been implicated in stress responses
based on its ability to complement the oxy5 bacterial strain.
This ability was apparently due to inherent peroxidase activity,
and, indeed, AnnAt1 does have a catalase motif in the first re-
peat of its protein structure [15]. A second report described
additional experiments supporting the ability of this protein to
protect against oxidative cellular damage [24]. More recently,
the property of inherent peroxidase activity for AnnAt1 has
been confirmed [16], and since H
2
O
2
is involved in many
stress and defense responses, this feature of AnnAt1 could al-
low it to play a key role in stress signaling.
Although we separately tested the effects of each abiotic
stress on annexin expression, often these individual stresses
can occur together and there is cross-talk between the various
stress pathways. Drought and salinity can both result in osmo-
tic stress, and both involve calcium and gene expression
changes in their signal transduction pathways. Drought and salt
treatments appear to affect the level of certain annexin mes-
sages similarly. For example, drought and salt treatments in-
crease the message level for AnnAt1 (39-fold and 44-fold, re-
spectively) and decrease the message levels for AnnAt2 (11.4-
fold and 11-fold, respectively). AnnAt5,AnnAt6, and AnnAt8
are all up-regulated by both salt and drought treatments,
although for these annexins there are differences in the level
of induction by the two treatments.
Transcript levels for all the annexins except AnnAt2 and
AnnAt3 increased following NaCl treatment. Two of them, An-
nAt4 and AnnAt8, showed a particularly high positive response
that was comparable to that of Rd29 and P5Cs (two marker
genes for salt stress in Arabidopsis [36]) under similar condi-
tions (data not shown). This result suggests an important role
for annexin gene family members during salt stress. Presum-
ably their function would be to help transduce the signaling
effects of Ca
2+
, which has at least two documented roles in salt
tolerance: to mediate adaptation to higher salt levels in the
growth medium or soil, and to regulate Na
+
transport systems
[13,40,41]. Transcriptional induction of annexin genes in re-
sponse to salt and other abiotic stresses open a new area of
investigation on the roles played by these proteins in the trans-
duction of stress signals.
Interestingly, the message levels of AnnAt3 and AnnAt4,
which share a common 2.5 kb 5region in their gene structure,
do not appear to be regulated in the same fashion by the dif-
ferent abiotic stress treatments. Analysis of the regulatory ele-
ments found in the shared 5region reveals some differences
within the 300 bp region immediately upstream of the respec-
tive start sites for these two genes. For example, the proximal
promoter region of AnnAt4 has a salt responsive sequence mo-
tif which is not found in this region for AnnAt3. This motif
difference might explain our results showing salt-induced in-
crease in AnnAt4 message levels and lack of salt regulation
for AnnAt3 message levels.
A. Cantero et al. / Plant Physiology and Biochemistry 44 (2006) 132420
Previous results obtained in wheat provide evidence for a
potential role of annexins in cold responses. Breton et al. [3]
found that two annexins redistributed from the cytoplasm to
the plasma membrane within 30 min after cold treatment. Ad-
ditionally, the protein levels for these two wheat annexins gra-
dually increased in response to cold treatment reaching a peak
1 day after cold treatment. In this study, we observed that cold
treatment leads to a fairly substantial increase in the transcript
levels for four of the annexins, AnnAt1,AnnAt3,AnnAt5, and
AnnAt7, possibly indicating a role for these annexins in cold
responses in Arabidopsis.
The results presented and discussed here provide clues to-
wards elucidating the function of Arabidopsis annexin gene
family during growth and development and under adverse en-
vironmental conditions. Differential expression pattern collec-
tively suggest annexin genes do not serve completely redun-
dant function and have both distinct and overlapping function
in certain signal transduction pathway. In general the expres-
sion patterns also suggest a role for this gene family in seedling
growth and development and response to abiotic stress. This
information will be a valuable starting point for future gene
knockout and other genetic studies that will be needed to clar-
ify the precise function of each individual annexin and to re-
solve which annexins have overlapping or distinct functions.
These genetic studies plus the expression results discussed here
will help reveal just how important role these proteins play in
plant Ca
2+
signaling.
4. Methods
4.1. Plant material and growth conditions
Arabidopsis thaliana ecotype Columbia (Col) and Wassi-
lewskija (Ws) seeds were sown on Murashige and Skoog med-
ium [27], supplemented with 2% sucrose and solidified with
0.7% agar. Ws seeds were used for all experiments except
stress studies which used Col seeds. All seeds were treated
with cold stratification at 4 °C for 3 days prior to sowing. Nor-
mal growth was at 22 °C under continuous light (140 μmol/m
2
per s). Seven-day-old seedlings were harvested after germina-
tion and frozen in liquid nitrogen before storage at 80 °C.
4.2. Stress treatments
To test the affect of various abiotic stresses 7-day-old seed-
lings were exposed to the following stress treatments for
2 hours each: 250 mM NaCl, dehydration on the bench,
37 °C, and 4 °C. Wild-type plants without any stress treatments
were used as controls, substituting normal growth condition for
temperature stress, distilled water for NaCl treatment, and nor-
mal growth temperature in closed media as control for dehy-
dration stress.
4.3. Light treatments
To test the effects of light on annexin gene expression, 5.5-
day-old etiolated seedlings were either exposed to 2 min red
light (76 μmol/m
2
per s) followed by 30 min dark, 2 min red
light followed immediately by 3 min far-red light
(112 μmol/m
2
per s) and then 30 min dark, or no light treat-
ment before harvesting hypocotyls and cotyledons.
4.4. RNA isolation and cDNA synthesis
Total RNA was isolated from whole seedlings by using the
Plant RNeasy extraction kit (Qiagen USA, Valencia, CA,
USA). Total RNA was isolated from dry seeds by using the
method described by Salmi et al. [29] for RNA isolation from
dry spores with slight modifications. Briefly, the modifications
include a precipitation of carbohydrates using 20% isopropanol
followed by 100% isopropanol at room temperature instead of
95% ethanol at 80 °C. To eliminate residual genomic DNA,
the RNA was treated with RNase free DNase (Invitrogen,
Carlsbad, CA, USA) and quantified in a spectrophotometer
(Nanodrop). cDNA was synthesized using Superscript II re-
verse transcriptase (Invitrogen) according to the manufacturers
instructions. The efficiency of cDNA synthesized was assessed
by end point RT-PCR using constitutive gene specific primers
encoding APT1 (Adenosine phosphoribosyl transferase).
4.5. Oligonucleotide probes and primers for real time RT-PCR
To facilitate real time RT-PCR measurement of transcripts
of the eight annexins, oligonucleotide primers were required to
meet a stringent set of criteria to ensure maximum specificity
and efficiency. For designing the individual annexin specific
primers sequences were selected within the divergent structural
repeats of individual annexins or, if distinct sequence was not
available within a particular gene, the UTRs were used.
LUXfluorogenic labeled primers were designed using the
web-based LUXdesigner software of Invitrogen according
to the specifications of 1824 nucleotides in length, product
size ranging between 80 and 150 bp, and a melting temperature
range of 6080 °C. Table 2 shows primer sequences for annex-
in and APT1 genes. The forward experimental primers were
labeled by the fluorophore FAM and forward control house-
keeping gene primers by the fluorophore JOE. Global align-
ments of the selected primer sequences for individual annexins
with genomic sequences and transcript sequences of all the an-
nexin genes were performed to retain and maintain specificity
for individual annexins.
We also used several positive control genes known to be
involved in abiotic stress tolerance. Specifically, gene specific
primers were designed and synthesized for Rd29 (Accession
no. D13044) (shown to respond to a variety of environmental
stresses including dehydration, low temperature, saline treat-
ments and osmotic changes) and two components of proline
biosynthetic pathway P5CS (Accession no. NM123232) (delta
A. Cantero et al. / Plant Physiology and Biochemistry 44 (2006) 1324 21
1-pyrroline-5-carboxylate synthetase) and PDH (Accession no.
NM113981) (proline dehydrogenase).
Gene specific primers were designed and synthesized for the
constitutive housekeeping gene APT1 (Accession no.
AF325045) which was used as an internal control. The internal
control gene was chosen on the basis of its constant expression
levels over experimental condition (time/treatment), which is
ideally a constitutive housekeeping gene. For use with the
comparative C
T
method, the control gene should have similar
amplification efficiency to experimental genes, which was con-
firmed after running a validation experiment. Efficiencies be-
tween the control gene APT1 and each of the eight annexins
were compared through analysis of serial template dilutions.
The slope of the log template versus ΔC
T
was found to be less
than 0.1 for each comparison, validating the use of the com-
parative C
T
method. Melting curves on products from each
primer set gave discrete peaks, indicating single target amplifi-
cation. Conventional agarose gel electrophoresis also showed
single bands of expected size for each primer set.
4.6. Quantitative real time RT-PCR conditions
Reverse transcription polymerase chain reactions were per-
formed using an ABI Prism 7700 or 7900HT Sequence Detec-
tion System (Applied Biosystems, Foster City, CA, USA). The
thermal cycling profile was performed with either a two-step or
three-step cycle. The two-step program was: 50 °C for 2 min,
95 °C for 2 min, followed by 40 cycles of 95 °C for 15 s and
60 °C for 1 min. The three-step program was: 50 °C for 2 min,
95 °C for 2 min, followed by 40 cycles of 95 °C for 15 s,
55 °C for 30 s and 72 °C for 30 s. A dissociation curve was
also run to check for presence of artifacts. On the 7700 this
was done by running the samples at 95 °C for 20 s, 60 °C
for 20 s and 95 °C for 15 s with a Ramp of 19 min and 59 s
from 60 °C to 95 °C, on the 7900HT the automated dissocia-
tion program was used.
PCR reactions were performed in 96- or 384-well polypro-
pylene plates sealed with transparent adhesive covers. Plati-
num® Quantitative PCR SuperMix-UDG (Invitrogen) was
used for 20 μl final volume reactions with 1 μl of ROX refer-
ence dye. Primers were used at 100 nM each with the equiva-
lent of 75200 ng of reverse-transcribed RNA template per re-
action. In all experiments, appropriate negative controls
containing no template RNA were subjected to the same pro-
cedure to exclude or detect any possible contamination or carry
over. Before proceeding with the actual experiments, a series
of primer dilutions and template dilution were carried out to
determine optimum primer and template concentrations to be
used in the experiments for optimum amplification of target.
Each sample was analyzed at least three times and three repli-
cates were used in each case. Reactions were repeated as ne-
cessary to give a minimum of six useable data points based on
a minimum requirement of two technical replicates for each of
three biological replicates. Actual data averaged an nof 9.6
across all experiments, with a range of 721.
4.7. Data analysis
C
T
values were obtained using the manufacturer supplied
SDS 1.91 or SDS 2.2.1 software. Baselines and thresholds
were set manually based on amplification plots, with typical
baselines running from cycles 3 through 15 and threshold va-
lues selected within the early exponential portion of the plot.
C
T
values were exported for expression calculations with Mi-
crosoft Excel. Data points with irregular amplification or melt-
ing curves were excluded from further computations.
Relative fold expression changes were calculated using the
comparative C
T
method, where fold change is calculated as
2
ΔΔ CT
. Validation experiments for the comparative C
T
meth-
od were performed as described above.
Calculations were performed in the following manner. ΔC
T
values were calculated as the difference between the mean C
T
Table 2
Real time RT-PCR primers
Gene Sequence Strand
AnnAt1 5-CATCTTTCGCACTTCTCGGTGAAGA[FAM]G-3Forward
5-GAGCAGCTTATGTTTCTCTGTGGA-3Reverse
AnnAt2 5-GACTAATGACGCCATTGTTGGGATTAG[FAM]C-3Forward
5-ACGTGCGTCTGCTTCGTCTC-3Reverse
AnnAt3 5-CACCAGGCATGGACAATGCTATTACTGG[FAM]G-3Forward
5-CAAGTCACAAGCAAAGCCAAGAAT-3Reverse
AnnAt4 5-CAACTCGGGATGGGAATGGGAG[FAM]TG-3Forward
5-CGCAGTATTGAAGCGGGAGAA-3Reverse
AnnAt5 5-GTATCGTTCAAAGGAAGAGGCTGCGA[FAM]AC-3Forward
5-TTGTGTTGCATTGCGATGAG-3Reverse
AnnAt6 5-GACGAGAGACAGTGTGCATGTTCCTCG[FAM]C-3Forward
5-GCCATTCTCCGACCACTGAA-3Reverse
AnnAt7 5-CACCATCCATTGCTAAAGACACTCATGG[FAM]G-3Forward
5-TGCAAGTCCAACCACAACTG-3Reverse
AnnAt8 5-CATCAACTTGACAAGCCGCCTTGA[FAM]G-3Forward
5-CCACCATTGTTTCTCCTCCACA-3Reverse
APT1 5-CAACGTGGCCCTCCTATTGCG[JOE]TG-3Forward
5-CCGAAATAACCTTCCCAGGTAGC-3Reverse
A. Cantero et al. / Plant Physiology and Biochemistry 44 (2006) 132422
value of APT over three replicates and each individual C
T
va-
lue of the corresponding annexins. For each treatment condi-
tion and annexin, ΔC
T
values were then averaged and standard
deviations calculated across all biological and technical repli-
cates.
Standard error of the means was estimated using:
S:E:¼s
ffiffi
n
p
sis the standard deviation of nnumber of data points.
Where indicated, 95% confidence intervals were determined
using the two-tailed distribution formula:
ΔΔCT±tα½n1
s
ffiffi
n
p
ΔΔC
T
values are naturally normalized to a value of 1 based
on the treatment used as the calibrator. AnnAt1 or specific
treatments or conditions were arbitrarily used as the calibrator
in order to give reasonable standardization across graphs.
Acknowledgements
We would like to thank Dr. Stuart Reichler and Andy Al-
verson with help in preparation of the manuscript. Sharmistha
Barthakur was funded by Department of Science and Technol-
ogy (DST), Government of India under BOYSCAST fellow-
ship. This work was supported by a NASA grant (NAGW
1519) to G.B. Clark and S.J. Roux.
References
[1] L. Adams-Phillips, C. Barry, P. Kannan, J. Leclercq, M. Bouzayen, J.
Giovannoni, Evidence that CTR1-mediated ethylene signal transduction
in tomato is encoded by a multigene family whose members display dis-
tinct regulatory features, Plant Mol. Biol. 54 (2004) 387404.
[2] G.J. Barton, R.H. Newman, P.S. Freemont, M.J. Crumpton, Amino acid
sequence analysis of the annexin super-gene family of proteins, Eur. J.
Biochem. 198 (1991) 749760.
[3] G. Breton, A. Vazquez-Tello, J. Danyluk, F. Sarhan, Two novel intrinsic
annexins accumulate in wheat membranes in response to low tempera-
ture, Plant Cell Physiol. 41 (2000) 177184.
[4] F. de Carvalho-Niebel, A.C.J. Timmers, M. Chabaud, A. Defaux-Petras,
D.G. Barker, The Nod factor-elicited annexin MtAnn1 is preferentially
localised at the nuclear periphery in symbiotically activated root tissues
of Medicago truncatula, Plant J. 32 (2002) 343352.
[5] B. Charrier, A. Champion, Y. Henry, M. Kreis, Expression profiling of
the whole Arabidopsis Shaggy-like kinase multigene family by real-time
reverse transcriptase-polymerase chain reaction, Plant Physiol. 130
(2002) 577590.
[6] G.B. Clark, D.W. Lee, M. Dauwalder, S.J. Roux, Immunolocalization
and histochemical evidence for the association of two different Arabidop-
sis annexins with secretion during early seedling growth and develop-
ment, Planta 220 (2005) 621631.
[7] G.B. Clark, D.S. Rafati, R.J. Bolton, M. Dauwalder, S.J. Roux, Redistri-
bution of annexin in gravistimulated pea plumules, Plant Physiol. Bio-
chem. 38 (2000) 937947.
[8] G.B. Clark, S.J. Roux, Annexins of plant cells, Plant Physiol. 109 (1995)
11331139.
[9] G.B. Clark, A. Sessions, D.J. Eastburn, S.J. Roux, Differential expres-
sion of members of the annexin multigene family in Arabidopsis, Plant
Physiol. 126 (2001) 10721084.
[10] T. Czechowski, R.P. Bari, M. Stitt, W.R. Scheible, M.K. Udvardi, Real-
time RT-PCR profiling of over 1400 Arabidopsis transcription factors:
unprecedented sensitivity reveals novel root- and shoot-specific genes,
Plant J. 38 (2004) 366379.
[11] D.P. Delmer, T.S. Potikha, Structures and functions of annexins in
plants, Cell. Mol. Life Sci. 53 (1997) 546553.
[12] J.L.E. Ellingson, J.L. Anderson, S.A. Carlson, V.K. Sharma, Twelve
hour real-time PCR technique for the sensitive and specific detection of
Salmonella in raw and ready-to-eat meat products, Mol. Cell. Probes 18
(2004) 5157.
[13] T.J. Flowers, Improving crop salt tolerance, J. Exp. Bot. 55 (2004) 307
319.
[14] C. Gachon, A. Mingam, B. Charrier, P.C.R. Real-time, what relevance to
plant studies?, J. Exp. Bot. 55 (2004) 14451454.
[15] X. Gidrol, P.A. Sabelli, Y.S. Fern, A.K. Kush, Annexin-like protein from
Arabidopsis thaliana rescues delta oxyR mutant of Escherichia coli from
H
2
O
2
stress, Proc. Natl. Acad. Sci. USA 93 (1996) 1126811273.
[16] K.M. Gorecka, D. Konopka-Postupolska, J. Hennig, R. Buchet, S. Piku-
la, Peroxidase activity of annexin 1 from Arabidopsis thaliana, Biochem.
Biophys. Res. Com. 336 (2005) 868875.
[17] O. Greiner, P.J.R. Day, P.P. Bosshard, F. Imeri, M. Altwegg, D. Nadal,
Quantitative detection of Streptococcus pneumoniae in nasopharyngeal
secretions by real-time PCR, J. Clin. Microbiol. 39 (2001) 31293134.
[18] M. Hernandez, A. Rio, T. Esteve, S. Prat, M. Pla, A rapeseed-specific
gene, Acetyl-CoA carboxylase, can be used as a reference for qualitative
and real-time quantitative PCR detection of transgenes from mixed food
samples, J. Agric. Food Chem. 49 (2001) 36223627.
[19] D. Hoshino, A. Hayashi, Y. Temmei, N. Kanzawa, T. Tsuchiya, Bio-
chemical and immunohistochemical characterization of Mimosa annexin,
Planta 219 (2004) 867875.
[20] D.J. Ingham, S. Beer, S. Money, G. Hansen, Quantitative real-time PCR
assay for determining transgene copy number in transformed plants, Bio-
tech. 31 (2001) 132141.
[21] J.Y. Jang, D.G. Kim, Y.O. Kim, J.S. Kim, H.S. Kang, An expression
analysis of a gene family encoding plasma membrane aquaporins in re-
sponse to abiotic stresses in Arabidopsis thaliana, Plant Mol. Biol. 54
(2004) 713725.
[22] I. Kovacs, F. Ayaydin, A. Oberschall, I. Ipacs, S. Bottka, D. Dudits, E.C.
Toth, Immunolocalization of a novel annexin-like protein encode by a
stress and abscisic acid responsive gene in alfalfa, Plant J. 15 (1998)
185197.
[23] S. Kumar, K. Tamura, M. Nei, MEGA3: Integrated software for Molecu-
lar Evolutionary Genetics Analysis and sequence alignment, Brief. Bioin-
form. 5 (2004) 150163.
[24] A. Kush, K. Sabapathy, Oxy5, a novel protein from Arabidopsis thali-
ana, protects mammalian cells from oxidative stress, Int. J. Bioch. Cell
Biol. 33 (2001) 591602.
[25] P.J. Lammers, J. Jun, J. Abubaker, R. Arreola, A. Gopalan, B. Bago, C.
Hernandez-Sebastia, J.W. Allen, D.D. Douds, P.E. Pfeffer, Y. Shachar-
Hill, The glyoxylate cycle in an arbuscular mycorrhizal fungus. Carbon
flux and gene expression, Plant Physiol. 127 (2001) 12871298.
[26] S. Lee, E.J. Lee, E.J. Yang, J.E. Lee, A.R. Park, W.H. Song, O.K. Park,
Proteomic identification of annexins, calcium-dependent membrane bind-
ing proteins that mediate osmotic stress and abscisic acid signal transduc-
tion in Arabidopsis, Plant Cell 16 (2004) 13781391.
[27] T. Murashige, F. Skoog, A revised medium for rapid growth and bioas-
says with tobacco tissue cultures, Physiol. Plant. 15 (1962) 473497.
[28] B. Reintanz, A. Szyroki, N. Ivashikina, P. Ache, M. Godde, D. Becker,
K. Palme, R. Hedrich, AtKC1, a silent Arabidopsis potassium channel
alpha-subunit modulates root hair K
+
influx, Proc. Natl. Acad. Sci.
USA 99 (2002) 40794084.
[29] M.L. Salmi, T.J. Bushart, S.C. Stout, S.J. Roux, Profile and analysis of
gene expression changes during early development in germinating spores
of Ceratopteris richardii, Plant Physiol. 138 (2005) 17341745.
A. Cantero et al. / Plant Physiology and Biochemistry 44 (2006) 1324 23
[30] P.G. Sappl, J.L. Heazlewood, A.H. Millar, Untangling multi-gene fa-
milies in plants by integrating proteomics into functional genomics, Phy-
tochem. 65 (2004) 15171530.
[31] B.C. Tan, L.M. Joseph, W.T. Deng, L.J. Liu, Q.B. Li, K. Cline, D.R.
McCarty, Molecular characterization of the Arabidopsis 9-cis epoxycaro-
tenoid dioxygenase gene family, Plant J. 35 (2003) 4456.
[32] J.M. Tepperman, M.E. Hudson, R. Khanna, T. Zhu, S.H. Chang, X.
Wang, P.H. Quail, Expression profiling of phyB mutant demonstrates
substantial contribution of other phytochromes to red-light-regulated
gene expression during seedling de-etiolation, Plant J. 38 (2004) 725
739.
[33] C. Thonat, C. Mathieu, M. Crevecoeur, C. Penel, T. Gaspar, N. Boyer,
Effects of a mechanical stimulation on localization of annexin-like pro-
teins in Bryonia dioica internodes, Plant Physiol. 114 (1997) 981988.
[34] L.S. Vinh, A. von Haeseler, IQPNNI: Moving fast through tree space and
stopping in time, Mol. Biol. Evol. 21 (2004) 15651571.
[35] L.M. Xiong, J.K. Zhu, Abiotic stress signal transduction in plants: Mole-
cular and genetic perspectives, Physiol. Plant. 112 (2001) 152166.
[36] K. Yamaguchishinozaki, K. Shinozaki, Characterization of the expres-
sion of a desiccation-responsive Rd29 gene of Arabidopsis thaliana and
analysis of its promoter in transgenic plants, Mol. Gener. Genet. 236
(1993) 331340.
[37] L.T. Yang, J.Y. Ding, C.M. Zhang, J.W. Jia, H.B. Weng, W.X. Liu, D.B.
Zhang, Estimating the copy number of transgenes in transformed rice by
real-time quantitative PCR, Plant Cell Rep. 23 (2005) 759763.
[38] R. Yokoyama, K. Nishitani, A comprehensive expression analysis of all
members of a gene family encoding cell-wall enzymes allowed us to pre-
dict cis-regulatory regions involved in cell-wall construction in specific
organs of Arabidopsis, Plant Cell Physiol. 42 (2001) 10251033.
[39] R. Yokoyama, J.K.C. Rose, K. Nishitani, A surprising diversity and
abundance of xyloglucan endotransglucosylase/hydrolases in rice. Classi-
fication and expression analysis, Plant Physiol. 134 (2004) 10881099.
[40] J.K. Zhu, Genetic analysis of plant salt tolerance using Arabidopsis, Plant
Physiol. 124 (2000) 941948.
[41] J.K. Zhu, Regulation of ion homeostasis under salt stress, Curr. Opin.
Plant Biol. 6 (2003) 441445.
A. Cantero et al. / Plant Physiology and Biochemistry 44 (2006) 132424
... Group I was further sub-grouped into Ia and Ib and contained 11 SbAnns. In Group Ia, Number 41/2024 ROMANIAN AGRICULTURAL RESEARCH Arabidopsis ANNAT1, ANNAT2, ANNAT6, and ANNAT7 annexins formed a speciesspecific subgroup, which was consistent with the results of (Cantero et al., 2006). Group II contained 3 SbAnns and also ANNAT4 as its outgroup. ...
... Furthermore, the intron number in SbAnn genes ranged from 0 to 5, similar in wheat and barley (Xu et al., 2016), though 3 to 5 were observed in Arabidopsis (Cantero et al., 2006). Also in barley, two HvAnns did not possess introns like SbAnn13. ...
... Also, SbAnn5 and SbAnn13 orthologs in O. sativa, Os08g32970 is expressed during heat stress (Jami et al., 2012). Lastly in Arabidopsis, AnnAt6 the lone ortholog of SbAnn12 is upregulated during far red light exposure at the seedling stage (Cantero et al., 2006). ...
Article
Full-text available
The annexin family is found in animals and plants and responds to a wide range of environmental stressors. In this study, 19 annexin proteins were identified based on their putative annexin repeats in Sorghum bicolor and further grouped on a phylogenic tree. These genes were randomly distributed across seven chromosomes. The amino acid length of sorghum annexins (SbAnn) varied from 241 (SbAnn3) to 370 (SbAnn2). Additionally, selective pressure analysis suggests the role of purification selection in conserving the annexin gene family in sorghum. Most SbAnns were located in the nucleus, cytoplasm, chloroplast, and mitochondrion. Also, higher syntenic relationships were observed between S. bicolor and monocots (S. italica, Z. mays, and O. sativa) than in A. thaliana. Cis-acting elements in the SbAnns gene promoters were grouped into response to light, hormone and environmental stress, and development-related elements. A number of miRNAs related to abiotic stress response targeted eleven SbAnns. Furthermore, five (SbAnn 4, 5, 14, 15, and 18) of the seven SbAnn genes utilized had relatively higher fold change under high-temperature treatment. This research provides insights into the characterization and functions of annexins under high temperatures.
... Normally, there are multiple members in the annexin family in each plant species, with different members having different subcellular localizations and physiological functions. Data obtained from plants that have been investigated so far showed that annexin family members are involved in seed germination and early seedling growth [39][40][41], the transition from vegetative to the reproductive phase [42], pollen germination and tube growth [43], etc. Biotic stresses (fungal or viral pathogen attack) and abiotic stresses (cold, heat, salt, and drought) promoted the expression of some members of the annexin gene family, and the expression levels of most annexin members were positively correlated with the tolerance of plant cells to, or their defense against, multiple environmental stresses [39,40,[44][45][46][47][48][49]. ...
... Normally, there are multiple members in the annexin family in each plant species, with different members having different subcellular localizations and physiological functions. Data obtained from plants that have been investigated so far showed that annexin family members are involved in seed germination and early seedling growth [39][40][41], the transition from vegetative to the reproductive phase [42], pollen germination and tube growth [43], etc. Biotic stresses (fungal or viral pathogen attack) and abiotic stresses (cold, heat, salt, and drought) promoted the expression of some members of the annexin gene family, and the expression levels of most annexin members were positively correlated with the tolerance of plant cells to, or their defense against, multiple environmental stresses [39,40,[44][45][46][47][48][49]. ...
Article
Full-text available
Extracellular ATP (eATP) plays multiple roles in plant growth and development, and stress responses. It has been revealed that eATP suppresses growth and alters the growth orientation of the root and hypocotyl of Arabidopsis thaliana by affecting auxin transport and localization in these organs. However, the mechanism of the eATP-stimulated auxin distribution remains elusive. Annexins are involved in multiple aspects of plant cellular metabolism, while their role in response to apoplastic signals remains unclear. Here, by using the loss-of-function mutations, we investigated the role of AtANN3 in the eATP-regulated root and hypocotyl growth. Firstly, the inhibitory effects of eATP on root and hypocotyl elongation were weakened or impaired in the AtANN3 null mutants (atann3–1 and atann3–2). Meanwhile, the distribution of DR5-GUS and DR5-GFP indicated that the eATP-induced asymmetric distribution of auxin in the root tips or hypocotyl cells occurred in wild-type control plants, while in atann3–1 mutant seedlings, it was not observed. Further, the eATP-induced asymmetric distribution of PIN2-GFP in root-tip cells or that of PIN3-GFP in hypocotyl cells was reduced in atann3–1 seedlings. Finally, the eATP-induced asymmetric distribution of cytoplasmic vesicles in root-tip cells was impaired in atann3–1 seedlings. Based on these results, we suggest that AtANN3 may be involved in eATP-regulated seedling growth by regulating the distribution of auxin and auxin transporters in vegetative organs.
... Specifically, 15 genes directly participate in drought stress. For example, ADH1 and ADHIII are involved in the response to water deprivation [21]; ANN1 is differentially regulated under drought conditions [22]; CRY1 and CRY2 participate in the response to water deprivation [23]; ERD7 and ERD14 participate in the rapid response to dehydration [24]; DRPD participates in the response to desiccation and is abundantly expressed in dried leaves [25]; EDL3 participates in the response to drought stress via the plant hormones [26]; HVA22E plays a role in stress tolerance and is differentially regulated by dehydration stress [27]; PLAT1 acts as a positive regulator in response to abiotic stress [28]; ARP1 is involved in the response to water deprivation, affecting ABA-regulated seed germination [29]; ASPG1 is involved in drought avoidance through ABA signaling [30]; REM4.2 plays a role in various abiotic stresses and participates in the SnRK1-mediated signaling pathway [31]; and CDSP32 participates in the plastid defense against oxidative damage in response to water deprivation [32]. Based on the results of this study, we believe that these DEGs largely contribute to the seed germination under drought stress. ...
Article
Full-text available
Peganum harmala L. is a perennial herbaceous plant that plays critical roles in protecting the ecological environment in arid, semi-arid, and desert areas. Although the seed germination characteristics of P. harmala in response to environmental factors (i.e., drought, temperature, and salt) have been investigated, the response mechanism of seed germination to drought conditions has not yet been revealed. In this study, the changes in the physiological characteristics and transcriptional profiles in seed germination were examined under different polyethylene glycol (PEG) concentrations (0–25%). The results show that the seed germination rate was significantly inhibited with an increase in the PEG concentration. Totals of 3726 and 10,481 differentially expressed genes (DEGs) were, respectively, generated at 5% and 25% PEG vs. the control (C), with 1642 co-expressed DEGs, such as drought stress (15), stress response (175), and primary metabolism (261). The relative expression levels (RELs) of the key genes regulating seed germination in response to drought stress were in accordance with the physiological changes. These findings will pave the way to increase the seed germination rate of P. harmala in drought conditions.
... AtAnn1 and AtAnn2 double-mutant ann1ann2 showed more susceptibility to Botrytis cinerea [31], implicating the positive regulation effect of AtAnn1 and AtAnn2 in plant resistance against B. cinerea. Furthermore, the expression of AnnAt1 increased significantly during osmotic stress and powdery mildew infection [19,32,33]. In addition, AtAnn8 was a negative regulator of RPW8.1-mediated resistance to powdery mildew and cell death [20]. ...
Article
Full-text available
(1) Annexins are proteins that bind phospholipids and calcium ions in cell membranes and mediate signal transduction between Ca2+ and cell membranes. They play key roles in plant immunity. (2) In this study, virus mediated gene silencing and the heterologous overexpression of TaAnn12 in Arabidopsis thaliana Col-0 trials were used to determine whether the wheat annexin TaAnn12 plays a positive role in plant disease resistance. (3) During the incompatible interaction between wheat cv. Suwon 11 and the Puccinia striiformis f. sp. tritici (Pst) race CYR23, the expression of TaAnn12 was significantly upregulated at 24 h post inoculation (hpi). Silencing TaAnn12 in wheat enhanced the susceptibility to Pst. The salicylic acid hormone contents in the TaAnn12-silenced plants were significantly reduced. The overexpression of TaAnn12 in A. thaliana significantly increased resistance to Pseudomonas syringae pv. tomato DC3000, and the symptoms of the wild-type plants were more serious than those of the transgenic plants; the amounts of bacteria were significantly lower than those in the control group, the accumulation of Reactive Oxygen Species (ROS)and callose deposition increased, and the expression of resistance-related genes (AtPR1, AtPR2, and AtPR5) significantly increased. (4) Our results suggest that wheat TaAnn12 resisted the invasion of pathogens by inducing the production and accumulation of ROS and callose.
... The annexin D2-like pecan protein had a 3.2-fold increase in expression as the kernel begins maturation (p-value 0.016). Annexins are involved in calcium signal transduction and expression is induced during germination in Arabidopsis thaliana [47]. A protein spot containing the NAD-dependent malic enzyme and heat shock 70 kDa protein is also more abundant after the early shift to maturation (3.5-fold, p-value 0.009). ...
Article
Full-text available
Pecan (Carya illinoinensis) nuts are an economically valuable crop native to the United States and Mexico. A proteomic summary from two pecan cultivars at multiple time points was used to compare protein accumulation during pecan kernel development. Patterns of soluble protein accumulation were elucidated using qualitative gel-free and label-free mass-spectrometric proteomic analyses and quantitative (label-free) 2-D gel electrophoresis. Two-dimensional (2-D) gel electrophoresis distinguished a total of 1267 protein spots and shotgun proteomics identified 556 proteins. Rapid overall protein accumulation occurred in mid-September during the transition to the dough stage as the cotyledons enlarge within the kernel. Pecan allergens Car i 1 and Car i 2 were first observed to accumulate during the dough stage in late September. While overall protein accumulation increased, the presence of histones diminished during development. Twelve protein spots accumulated differentially based on 2-D gel analysis in the weeklong interval between the dough stage and the transition into a mature kernel, while eleven protein spots were differentially accumulated between the two cultivars. These results provide a foundation for more focused proteomic analyses of pecans that may be used in the future to identify proteins that are important for desirable traits, such as reduced allergen content, improved polyphenol or lipid content, increased tolerance to salinity, biotic stress, seed hardiness, and seed viability.
... Specifically, a total of 29 genes were directly associated with salt stress. For example, ALDH7B4 is activated by high salinity, dehydration, and ABA in a tissue-specific manner [48]; annexin (ANNs) in Arabidopsis is regulated by exposure to salt, drought, and extreme temperature conditions [49]; and RD22 is induced by salt stress and water deficit during the early and middle stages of seed development [50]. CAMBP25 is induced in Arabidopsis seedlings exposed to high salinity, dehydration, and low temperature [51]. ...
Article
Full-text available
Apocynum venetum is a semi-shrubby perennial herb that not only prevents saline-alkaline land degradation but also produces leaves for medicinal uses. Although physiological changes during the seed germination of A. venetum in response to salt stress have been studied, the adaptive mechanism to salt conditions is still limited. Here, the physiological and transcriptional changes during seed germination under different NaCl treatments (0-300 mmol/L) were examined. The results showed that the seed germination rate was promoted at low NaCl concentrations (0-50 mmol/L) and inhibited with increased concentrations (100-300 mmol/L); the activity of anti-oxidant enzymes exhibited a significant increase from 0 (CK) to 150 mmol/L NaCl and a significant decrease from 150 to 300 mmol/L; and the content of osmolytes exhibited a significant increase with increased concentrations, while the protein content peaked at 100 mmol/L NaCl and then significantly decreased. A total of 1967 differentially expressed genes (DEGs) were generated during seed germination at 300 mmol/L NaCl versus (vs.) CK, with 1487 characterized genes (1293 up-regulated, UR; 194 down-regulated, DR) classified into 11 categories, including salt stress (29), stress response (146), primary metabolism (287), cell morphogenesis (156), transcription factor (TFs, 62), bio-signaling (173), transport (144), photosynthesis and energy (125), secondary metabolism (58), polynucleotide metabolism (21), and translation (286). The relative expression levels (RELs) of selected genes directly involved in salt stress and seed germination were observed to be consistent with the changes in antioxidant enzyme activities and osmolyte contents. These findings will provide useful references to improve seed germination and reveal the adaptive mechanism of A. venetum to saline-alkaline soils.
... For example, after NaCl or ABA treatment, the expression level increase of the BGLU24, a member of the glycoside hydrolase family, which was reported to respond to salt tolerance, in the WT plants was more obvious than that in the mutants. Similarly, the expression levels of another five salt-responsive genes, AKR4C8, ANN1, GSTU17, SIGE and EGR1 [54][55][56][57][58], also increased with the same trends. The stress-responsive gene LEA3 was up-regulated in the two CgZFP1 transgenic lines that conferred the salinity and drought tolerance under drought stress conditions [59], but its expression was down-regulated in the Atwrky66 knockdown plants under the NaCl or ABA treatment (Table 4). ...
Article
Full-text available
Group Ⅲ WRKY transcription factors (TFs) play pivotal roles in responding to the diverse abiotic stress and secondary metabolism of plants. However, the evolution and function of WRKY66 remains unclear. Here, WRKY66 homologs were traced back to the origin of terrestrial plants and found to have been subjected to both motifs’ gain and loss, and purifying selection. A phylogenetic analysis showed that 145 WRKY66 genes could be divided into three main clades (Clade A–C). The substitution rate tests indicated that the WRKY66 lineage was significantly different from others. A sequence analysis displayed that the WRKY66 homologs had conserved WRKY and C2HC motifs with higher proportions of crucial amino acid residues in the average abundance. The AtWRKY66 is a nuclear protein, salt- and ABA- inducible transcription activator. Simultaneously, under salt stress and ABA treatments, the superoxide dismutase (SOD), peroxidase (POD) and catalase (CAT) activities, as well as the seed germination rates of Atwrky66-knockdown plants generated by the clustered, regularly interspaced, short palindromic repeats/CRISPR-associated 9 (CRISPR/Cas9) system, were all lower than those of wild type (WT) plants, but the relative electrolyte leakage (REL) was higher, indicating the increased sensitivities of the knockdown plants to the salt stress and ABA treatments. Moreover, RNA-seq and qRT-PCR analyses revealed that several regulatory genes in the ABA-mediated signaling pathway involved in stress response of the knockdown plants were significantly regulated, being evidenced by the more moderate expressions of the genes. Therefore, the AtWRKY66 likely acts as a positive regulator in the salt stress response, which may be involved in an ABA-mediated signaling pathway.
... Although upregulation of all Arabidopsis AtAnns except for Ann2 and Ann3 was observed in Arabidopsis salt induced AtAnns [77], we found that Ann2 in S. parvula was significantly upregulated in salt stress conditions. The divergent expression patterns of the two studied halophytes compared to A. thaliana as a glycophyte reveals that Ann genes in halophytes might change regulatory mechanisms that contribute to salt tolerance. ...
Article
Full-text available
Annexins (Anns) play an important role in plant development, growth and responses to various stresses. Although Ann genes have been characterized in some plants, their role in adaptation mechanisms and tolerance to environmental stresses have not been studied in extremophile plants. In this study, Ann genes in Schrenkiella parvula and Eutrema salsugineum were identified using a genome-wide method and phylogenetic relationships, subcellular distribution, gene structures, conserved residues and motifs and also promoter prediction have been studied through bioinformatics analysis. We identified ten and eight encoding putative Ann genes in S. parvula and E. salsugineum genome respectively, which were divided into six subfamilies according to phylogenetic relationships. By observing conservation in gene structures and protein motifs we found that the majority of Ann members in two extremophile plants are similar. Furthermore, promoter analysis revealed a greater number of GATA, Dof, bHLH and NAC transcription factor binding sites, as well as ABRE, ABRE3a, ABRE4, MYB and Myc cis-acting elements in compare to Arabidopsis thaliana. To gain additional insight into the putative roles of candidate Ann genes, the expression of SpAnn1, SpAnn2 and SpAnn6 in S. parvula was studied in response to salt stress, which indicated that their expression level in shoot increased. Similarly, salt stress induced expression of EsAnn1, 5 and 7, in roots and EsAnn1, 2 and 5 in leaves of E. salsugineum. Our comparative analysis implies that both halophytes have different regulatory mechanisms compared to A. thaliana and suggest SpAnn2 gene play important roles in mediating salt stress.
Preprint
Full-text available
Extracellular ATP (eATP) exists in the apoplast and plays multiple roles in growth, development, and stress responses. eATP has been revealed to suppresses growth rate and alters growth orientation of root and hypocotyl of Arabidopsis thaliana seedlings by affecting auxin transport in these organs. However, the mechanism of eATP-stimulated auxin distribution remains unclear. Annexins are involved in multiple aspects of plant cellular metabolism, while the role of annexins in response to apoplast signal remains unclear. Here, by using loss-of-function mutants, we investigated the role of several annexins in eATP-regulated root and hypocotyl growth. Since mutants of AtANN3 did not respond to eATP sensitively, the role of AtANN3 in eATP regulated auxin transport was intensively investigated. Firstly, the inhibitory effect of eATP on root or hypocotyl elongation was weakened or impaired in AtANN3 null mutants ( atann3-1 and atann3-2 ). Meanwhile, single-, double- or triple-null mutant of AtANN1 , AtANN2 or AtANN4 responded to eATP in same manner and degree with Col-0. The distribution of DR5-GUS and DR5-GFP indicated that eATP-induced asymmetric distribution of auxin in root tip or hypocotyl cells, which appeared in wild type controls, were lacking in atann3-1 seedlings. Further, eATP-induced asymmetric distribution of PIN2-GFP in root tip cells or PIN3-GFP in hypocotyl cells were reduced in atann3-1 seedlings. Based on these results, we suggest that AtANN3 may be involved in eATP-regulated seedling growth through regulating auxin transport in vegetative organs.
Article
Full-text available
With its theoretical basis firmly established in molecular evolutionary and population genetics, the comparative DNA and protein sequence analysis plays a central role in reconstructing the evolutionary histories of species and multigene families, estimating rates of molecular evolution, and inferring the nature and extent of selective forces shaping the evolution of genes and genomes. The scope of these investigations has now expanded greatly owing to the development of high-throughput sequencing techniques and novel statistical and computational methods. These methods require easy-to-use computer programs. One such effort has been to produce Molecular Evolutionary Genetics Analysis (MEGA) software, with its focus on facilitating the exploration and analysis of the DNA and protein sequence variation from an evolutionary perspective. Currently in its third major release, MEGA3 contains facilities for automatic and manual sequence alignment, web-based mining of databases, inference of the phylogenetic trees, estimation of evolutionary distances and testing evolutionary hypotheses. This paper provides an overview of the statistical methods, computational tools, and visual exploration modules for data input and the results obtainable in MEGA.
Article
Full-text available
The arbuscular mycorrhizal (AM) symbiosis is responsible for huge fluxes of photosynthetically fixed carbon from plants to the soil. Lipid, which is the dominant form of stored carbon in the fungal partner and which fuels spore germination, is made by the fungus within the root and is exported to the extraradical mycelium. We tested the hypothesis that the glyoxylate cycle is central to the flow of carbon in the AM symbiosis. The results of 13C labeling of germinating spores and extraradical mycelium with 13C2-acetate and 13C2-glycerol and analysis by nuclear magnetic resonance spectroscopy indicate that there are very substantial fluxes through the glyoxylate cycle in the fungal partner. Full-length sequences obtained by polymerase chain reaction from a cDNA library from germinating spores of the AM fungus Glomus intraradices showed strong homology to gene sequences for isocitrate lyase and malate synthase from plants and other fungal species. Quantitative real-time polymerase chain reaction measurements show that these genes are expressed at significant levels during the symbiosis. Glyoxysome-like bodies were observed by electron microscopy in fungal structures where the glyoxylate cycle is expected to be active, which is consistent with the presence in both enzyme sequences of motifs associated with glyoxysomal targeting. We also identified among several hundred expressed sequence tags several enzymes of primary metabolism whose expression during spore germination is consistent with previous labeling studies and with fluxes into and out of the glyoxylate cycle.
Article
Mechanical stimulation exerted by rubbing a young internode of Bryonia dioica plants inhibits its growth. Previous cellular and biochemical studies showed that this growth inhibition is associated with Ca2+ redistribution and profound modifications of plasma membrane characteristics. We extracted and purified Ca2+-dependent phospholipid-binding proteins from B. dioica internodes. Two main proteins, p33 and p35, and other minor bands were isolated and identified as annexin-like proteins because of their biochemical properties and their cross-reactions with antibodies against maize (Zea mays L.) annexins. Rabbit antiserum was obtained by injection of B. dioica p35. This antiserum was used for the immunocytolocalization of annexin-like proteins in internode parenchyma cells. It appeared that the distribution of annexin-like proteins was different before and 30 min after the mechanical stimulation. Western analysis of proteins in membrane fractions after separation by free-flow electrophoresis showed that p35 was present in most fractions, whereas p33 appeared mainly in plasmalemma-enriched fractions after the mechanical stimulation. It is hypothesized that a subcellular redistribution of these proteins might be involved in growth inhibition by mechanical stress.
Article
Although in most plant species no more than two annexin genes have been reported to date, seven annexin homologs have been identified in Arabidopsis, Annexin Arabidopsis 1–7 (AnnAt1–AnnAt7). This establishes that annexins can be a diverse, multigene protein family in a single plant species. Here we compare and analyze these seven annexin gene sequences and present the in situ RNA localization patterns of two of these genes, AnnAt1 andAnnAt2, during different stages of Arabidopsis development. Sequence analysis of AnnAt1–AnnAt7 reveals that they contain the characteristic four structural repeats including the more highly conserved 17-amino acid endonexin fold region found in vertebrate annexins. Alignment comparisons show that there are differences within the repeat regions that may have functional importance. To assess the relative level of expression in various tissues, reverse transcription-PCR was carried out using gene-specific primers for each of the Arabidopsis annexin genes. In addition, northern blot analysis using gene-specific probes indicates differences in AnnAt1 and AnnAt2expression levels in different tissues. AnnAt1 is expressed in all tissues examined and is most abundant in stems, whereas AnnAt2 is expressed mainly in root tissue and to a lesser extent in stems and flowers. In situ RNA localization demonstrates that these two annexin genes display developmentally regulated tissue-specific and cell-specific expression patterns. These patterns are both distinct and overlapping. The developmental expression patterns for both annexins provide further support for the hypothesis that annexins are involved in the Golgi-mediated secretion of polysaccharides.
Article
Polymerase chain reaction (PCR) methods are very useful techniques for the detection and quantification of genetically modified organisms (GMOs) in food samples. These methods rely on the amplification of transgenic sequences and quantification of the transgenic DNA by comparison to an amplified reference gene. Reported here is the development of specific primers for the rapeseed (Brassica napus) BnACCg8 gene and PCR cycling conditions suitable for the use of this sequence as an endogenous reference gene in both qualitative and quantitative PCR assays. Both methods were assayed with 20 different rapeseed varieties, and identical amplification products were obtained with all of them. No amplification products were observed when DNA samples from other Brassica species, Arabidopsis thaliana, maize, and soybean were used as templates, which demonstrates that this system is specific for rapeseed. In real-time quantitative PCR analysis, the detection limit was as low as 1.25 pg of DNA, which indicates that this method is suitable for use in processed food samples which contain very low copies of target DNA.
Article
The annexins are a widespread family of calcium-dependent membrane-binding proteins. No common function has been identified for the family and, until recently, no crystallographic data existed for an annexin. In this paper we draw together 22 available annexin sequences consisting of 88 similar repeat units, and apply the techniques of multiple sequence alignment, pattern matching, secondary structure prediction and conservation analysis to the characterisation of the molecules. The analysis clearly shows that the repeats cluster into four distinct families and that greatest variation occurs within the repeat 3 units. Multiple alignment of the 88 repeats shows amino acids with conserved physicochemical properties at 22 positions, with only Gly at position 23 being absolutely conserved in all repeats. Secondary structure prediction techniques identify five conserved helices in each repeat unit and patterns of conserved hydrophobic amino acids are consistent with one face of a helix packing against the protein core in predicted helices a, c, d, e. Helix b is generally hydrophobic in all repeats, but contains a striking pattern of repeat-specific residue conservation at position 31, with Arg in repeats 4 and Glu in repeats 2, but unconserved amino acids in repeats 1 and 3. This suggests repeats 2 and 4 may interact via a buried saltbridge. The loop between predicted helices a and b of repeat 3 shows features distinct from the equivalent loop in repeats 1, 2 and 4, suggesting an important structural and/or functional role for this region. No compelling evidence emerges from this study for uteroglobin and the annexins sharing similar tertiary structures, or for uteroglobin representing a derivative of a primordial one-repeat structure that underwent duplication to give the present day annexins. The analyses performed in this paper are re-evaluated in the Appendix, in the light of the recently published X-ray structure for human annexin V. The structure confirms most of the predictions and shows the power of techniques for the determination of tertiary structural information from the amino acid sequences of an aligned protein family.
Article
We characterized the expression of genes that correspond to a cDNA clone, RD29, which is induced by desiccation, cold and high-salt conditions in Arabidopsis thaliana. Northern analysis of desiccation-induced expression revealed a two-step induction process. Early induction occurs within 20 min and secondary induction occurs 3 h after the start of desiccation. Exogenous abscisic acid (ABA) induces RD29 mRNA within 3 h. Two genes corresponding to RD29, rd29A and rd29B, are located in tandem in an 8 kb region of the Arabidopsis genome and encode hydrophilic proteins. Desiccation induces rd29A mRNA with two-step kinetics, while rd29B is induced only 3 h after the start of desiccation. The expression of both genes is stimulated about 3 h after application of ABA. It appears that rd29A has at least two cis-acting elements, one involved in the ABA-associated response to desiccation and the other induced by changes in osmotic potential. The beta-glucuronidase (GUS) reporter gene driven by the rd29A promoter was induced at significant levels by desiccation, cold, high-salt conditions and ABA in both transgenic Arabidopsis and tobacco. Histochemical analysis of GUS activity revealed that the rd29A promoter functions in almost all the organs and tissues of vegetative plants during water deficiency.