ArticlePDF Available

A genome-wide transcriptional analysis using Arabidopsis thaliana Affymetrix gene chips determined plant responses to phosphate deprivation

Authors:

Abstract and Figures

Phosphorus, one of the essential elements for plants, is often a limiting nutrient because of its low availability and mobility in soils. Significant changes in plant morphology and biochemical processes are associated with phosphate (Pi) deficiency. However, the molecular bases of these responses to Pi deficiency are not thoroughly elucidated. Therefore, a comprehensive survey of global gene expression in response to Pi deprivation was done by using Arabidopsis thaliana whole genome Affymetrix gene chip (ATH1) to quantify the spatio-temporal variations in transcript abundance of 22,810 genes. The analysis revealed a coordinated induction and suppression of 612 and 254 Pi-responsive genes, respectively. The functional classification of some of these genes indicated their involvement in various metabolic pathways, ion transport, signal transduction, transcriptional regulation, and other processes related to growth and development. This study is a detailed analysis of Pi starvation-induced changes in gene expression of the entire genome of Arabidopsis correlated with biochemical processes. The results not only enhance our knowledge about molecular processes associated with Pi deficiency, but also facilitate the identification of key molecular determinants for improving Pi use by crop species.
Content may be subject to copyright.
A genome-wide transcriptional analysis using
Arabidopsis thaliana
Affymetrix gene chips
determined plant responses to phosphate deprivation
Julie Misson*, Kashchandra G. Raghothama
, Ajay Jain
, Juliette Jouhet
, Maryse A. Block
, Richard Bligny
,
Philippe Ortet
§
, Audrey Creff*, Shauna Somerville
, Norbert Rolland
, Patrick Doumas
, Philippe Nacry
,
Luis Herrerra-Estrella**, Laurent Nussaume*, and Marie-Christine Thibaud*
††
*Laboratoire de Biologie du De´ veloppement des Plantes, Unite Mixte de Recherche 6191, Centre National de la Recherche Scientifique–Commissariat a`
l’Energie Atomique, Aix-Marseille II, 13108 Saint-Paul-lez-Durance, France;
Department of Horticulture and Landscape Architecture, Purdue University,
West Lafayette, IN 47907-1165;
Laboratoire de Physiologie Cellulaire Ve´ge´ tale, Unite Mixte de Recherche 5168, Centre National de la Recherche
ScientifiqueCommissariat a´ l’Energie AtomiqueUniversite´ Joseph FourierInstitut National de la Recherche Agronomique, 38054 Grenoble, France;
§
De´ partement d’Ecophysiologie Ve´ge´ tale et de Microbiologie, 13108 Saint-Paul-lez-Durance, France;
Department of Plant Biology, Carnegie Institute,
Stanford, CA 94305;
Laboratoire de Biochimie et Physiologie Mole´ culaires des Plantes, Unite Mixte de Recherche 5004, Institut National de la Recherche
ArgonomiqueCentre National de la Recherche ScientifiqueE
´
cole Nationale Supe´ rieure d’Arts et Me´ tiersUniversite´ Montpellier II, 2 Place Viala, 34060
Montpellier, France; and **Departamento de Ingenieria Genetica de Plantas, Centro de Investigacio´ n de Estudios Avanzados del Instituto Politecnico
Nacional, Unidad Irapuato, Apartado Postal 629, 36500 Irapuato, Guanajuato, Mexico
Contributed by Luis Herrera-Estrella, June 23, 2005
Phosphorus, one of the essential elements for plants, is often a
limiting nutrient because of its low availability and mobility in
soils. Significant changes in plant morphology and biochemical
processes are associated with phosphate (Pi) deficiency. However,
the molecular bases of these responses to Pi deficiency are not
thoroughly elucidated. Therefore, a comprehensive survey of
global gene expression in response to Pi deprivation was done by
using Arabidopsis thaliana whole genome Affymetrix gene chip
(ATH1) to quantify the spatio-temporal variations in transcript
abundance of 22,810 genes. The analysis revealed a coordinated
induction and suppression of 612 and 254 Pi-responsive genes,
respectively. The functional classification of some of these genes
indicated their involvement in various metabolic pathways, ion
transport, signal transduction, transcriptional regulation, and
other processes related to growth and development. This study is
a detailed analysis of Pi starvation-induced changes in gene ex-
pression of the entire genome of Arabidopsis correlated with
biochemical processes. The results not only enhance our knowl-
edge about molecular processes associated with Pi deficiency, but
also facilitate the identification of key molecular determinants for
improving Pi use by crop species.
P
hosphate (Pi) is an essential macronutriment required for plant
growth and development. Its low availability to plants in many
soils re sults not only from limiting amounts, but also from its
association with cations and organic compounds creating insoluble
complexe s. Thus, Pi has become one of the major plant nutrition
problems limiting growth in both acidic and calcareous soils (1).
Applications of large quantities of fertilizers to correct this problem
are not economically sustainable and also lead to environmental
pollution. Therefore, efforts have been directed to understanding
the molecular basis of plants responses to Pi deficiency and to
identifying Pi-responsive genes whose expression can be manipu-
lated to enable plant growth in low-Pi environments. Although a
host of genes representing Pi transporters, phosphatase s, RNase s,
and others (2–5) have been identified and characterized by tradi-
tional expression studies, knowledge of global changes in the
expre ssion of Pi-responsive genes is still lacking. This technical
limitation has now been circumvented by recently developed mi-
croarray technology, which has been used successfully to study the
effects of different abiotic stresse s to large number of plant genes
in parallel (6). To date, relatively small microarrays containing
probes for less than one-third of all genes have been used in
Arabidopsis, rice, and white lupin to monitor molecular response s to
Pi deprivation (7–10). In Arabidopsis, the expression of 29% of
6,172 genes examined changed in response to Pi starvation for up
to 3 days (10). In another study with 8,100 genes of Arabidopsis,
differential expression of a group of 61 genes was observed after
100 h of Pi starvation (7). These studies represented the evaluation
of only a part of the Arabidopsis genome for Pi-responsive genes.
The advent of new microarray technique of Affymetrix ATH1
GeneChip, containing 22,810 Arabidopsis probe sets, has now made
it feasible to evaluate the expression of almost all of the genes in the
Arabidopsis genome with a sensitivity of one transcript per cell (6).
Here, we used the ATH1 GeneChip for a global evaluation of
genes that are spatio-temporally regulated in response to short-,
medium-, and long-term Pi deprivation. The sensitivity of this
technique was corroborated by expre ssion analysis. Identification of
differentially expre ssed genes revealed the coordinated activation
and repression of genes involved in many biochemical pathways that
are closely associated with plant responses to Pi deficiency. The se
genes could serve as potential candidates to decipher the compo-
nents of Pi-sensing mechanisms and developing strategies to im-
prove P efficiency in crops.
Materials and Methods
Supporting Information. For further details, see Figs. 5 and 6 and
Tables 1–9, which are published as supporting information on the
PNAS web site.
Plant Material and Growth Conditions. Arabidopsis thaliana (L.)
plants were raised in liquid culture and transferred in a medium
with or without Pi as described (11) for evaluating the short-term
(3, 6, and 12 h pooled) and medium-term (1 and 2 days pooled)
effects of Pi deficiency on the gene expression. For studying
long-term effects of Pi deficiency, surface-sterilized seeds were
sown in square (12 12 cm) Petri dishes on Murashige and Skoog
(MS)10 medium, 0.5% sucrose, 0.8% agar, and supplemented with
either 500
M(P)or5
M(P) Pi; plants were grown vertically
(12). Because some Arabidopsis genes are regulated by diurnal
rhythm and circadian clocks (13), the roots and the leaves were
harvested separately at the beginning and at the end of the
photoperiod and pooled. Samples were rinsed with distilled water,
blot-dried, and frozen in liquid nitrogen.
RNA Extraction and cRNA Preparation. Total RNA from shoot and
root (long-term) and from the whole plant (short- and medium-
Abbreviations: Pi, phosphate; qRT-PCR, quantitative RT-PCR.
††
To whom correspondence should be addressed. E-mail: mcthibaud.cea.fr.
© 2005 by The National Academy of Sciences of the USA
11934–11939
PNAS
August 16, 2005
vol. 102
no. 33 www.pnas.orgcgidoi10.1073pnas.0505266102
term) was extracted as described (11, 12). cRNA was prepared
following the manufacturer’s instructions (www.affymetrix.com
supporttechnicalmanualexpressionmanual.affx).
Microarray Hybridization and Data Analysis. The Affymetrix micro-
arrays (Arabidopsis ATH1 genome array) contain 22,810 probe
sets representing 80% of the gene sequence s on a single
array. Labeling and hybridization on the ATH1 microarrays (one
sample per chip) were performed according to the manufacturer’s
instr uctions (www.af fy metrix.comsupporttechnicalmanual
expre ssionmanual.affx). The probe arrays were scanned and fur-
ther analyzed with
GENESPRING software (version 5.0; Silicon
Genetics). Normalization per gene and per chip of the log
2
values
was performed to allow the comparison of the three independent
replicates performed for each set of experiment. In addition,
normalization was performed separately for each experiment and
plant tissue for all measurements using the flags (‘‘present,’’ ‘‘mar-
ginal,’’ or ‘‘absent’’) assigned by Affymetrix treatment of the arrays.
However, only those transcripts that were declared present or
marginal in at least three of six chips were taken into account. Such
a procedure facilitated the elimination of transcripts with very low
signals in both treatments (declared ‘‘absent’’). This elimination was
achieved by selecting the genes that were absent in at least four
microarrays (four, five, or six arrays) and on this basis, comple-
mentary lists of genes were generated. Data were then analyzed for
each of the experiments comprising plants deprived of Pi for
different lengths of time and results were compared with corre-
sponding P() plants (control). The genes that revealed significant
changes in their expression in P() plants were selected by applying
a t test (one-way ANOVA Welch t te st, P 0.05). Moreover, a
cutoff value of 2-fold change, which is commonly used for microar-
ray analysis (10), was adopted to discriminate expression of gene s
that were differentially altered in re sponse to Pi deficiency. Anno-
tation of the genes represented on the microarray to genomic ORFs
was done with ‘‘gene description’’ and ‘‘gene ontology’’ programs of
GENESPRING (based on information from the international Arabi-
dopsis Genome Initiative sequencing project in collaboration with
The Institute for Genome Research). To test the hybridization
quality, ‘‘Arabidopsis control genes’’ coding for GAPDH, actin,
tubulin, ubiquitin, and several ribosomal RNAs (25S, 5S), spotted
by the manufacturer, were verified. The expression ratios [P()
P()] of the control genes were consistently in the range of
0.81–1.29.
Real-Time Quantitative RT-PCR (qRT-PCR) and Northern Analysis. A
few differentially regulated Pi-responsive genes identified from the
microarray analysis were selected for validation of the results by
qRT-PCR and Northern analysis. cDNA was used for performing
qRT-PCR (iCycler Real-Time PCR Detection System, Bio-Rad).
Specific primers (T
m
, 5863°C) were designed to generate PCR
products between 150 and 350 bp (Table 1). qRT-PCR of GAPDH
C (At3g04120) was performed for standardization. Platinium
Quantitative PCR SuperMix-UDG (Invitrogen) was used for the
PCRs according to the manufacturer’s protocol with a minor
modification (0.33
M of each primer in a final volume of 15
l).
Northern analysis was performed as described (11).
S6K2 Promoter Fusion. The gene AtS6K2 (At3g08720), coding for a
ribosomal protein S6 kinase is induced during Pi deficiency (Table
7). A 520-bp fragment of its promoter located upstream of the
transcription start was PCR amplified and cloned into the modified
binary vector pBIN-
35
S-mgfp4 at HindIIIXbaI replacing the
35
S
promoter (kindly provided by J. Haseloff, University of Cambridge,
Cambridge, U.K.). Transgenic plants were generated by vacuum
infiltration of inflorescence with Agrobacterium as described (14).
Quantification of Lipids, Anthocyanins, Macro- and Microelements,
and Pi. Lipids, fatty acids, and anthocyanins were quantified as
described (15–17). Soluble inorganic Pi was measured in plants
from short- and medium-term Pi deficiency treatments (12),
whereas samples from long-term treatment were mineralized in
14% HNO
3
in a microwave system (MarsX, CEM) for the deter-
mination of macro- and microelements by ICP (ICP OES Vista
MPX, Varian).
31
P- and
13
C-NMR Spectroscopy. Metabolite analysis was performed
by NMR as described (18) on leaf extracts from the plants grown
in Petri dishes in a medium without sucrose for 20 days. The plants
were enriched with
13
C in a growth chamber that allowed accurate
regulation of the atmospheric gas composition and environmental
parameters. The growth chamber conditions were: 10-h photope-
riod with 100
molm
2
s
1
light, 2218°C daynight temperature,
and 90% humidity. CO
2
(containing 10% of
13
CO
2
purchased from
Euriso-Top, Saint-Aubin, France) concentration in the chamber
was maintained at 350
lliter
1
during the light period by auto-
matic injections to compensate for photosynthetic assimilation.
Results and Discussion
Phenotypic, Physiological, and Biochemical Modifications in Pi-
Starved Plants. The plants were grown on low Pi (5
M) to reduce
the nonspecific effects of complete nutrient deficiency on growth.
The effects of Pi deficiency treatments on some of the morpho-
logical, biochemical, and physiological traits were evaluated. A
significant decline of 55% and 68% in soluble Pi content after 12
and 48 h of Pi deficiency, respectively (Fig. 1A) indicated a rapid
effect of Pi withdrawal from the growth medium. This response to
Pi deprivation could be due to small seed size of Arabidopsis with
low Pi re serves. Generally, the amount of metabolic reserve in a
seed is correlated with its size, which could be critical for seedling
survival during various environmental stresse s (19). However, a
decline in Pi content did not result in a significant accumulation of
anthocyanins (data not shown), which is one of the traits associated
with Pi deficiency (1). Although many of the well characterized Pi
starvation-induced genes including Pht1;4 (10, 11), are induced
during short- and medium-term Pi deficiency, the degree of Pi stress
may not be sufficient enough to elicit anthocyanin accumulation.
Accumulation of anthocyanins was observed in the leaves only
when the plants were subjected to long-term Pi deficiency (Fig. 1B).
In addition, there was a reduction in leaf size and a modification in
the root architecture (i.e., higher densities of lateral roots and root
hairs), typical responses of plants to Pi deficiency (20). Pi-deficient
plants also exhibited an early arrest of the primary root growth (0.3
cmday
1
and2cmday
1
in low and high Pi, respectively), whereas
secondary roots continued to grow (Fig. 1C). Long-term Pi defi-
ciency-induced modifications in morphological and biochemical
traits could be attributed to a decline in the total Pi content in the
leaves and the roots (Fig. 1D). A significant decline in the concen-
tration of K in leaves, and an appreciable increase in the concen-
tration of S in both the roots and the leaves, was also observed (Fig.
1D). Likewise, higher accumulation of some micronutrients (i.e.,
Fe, Zn) was observed during long-term Pi deprivation (Fig. 1E).
This observation suggests that, during Pi deficiency, the activity
andor utilization of other nutrients was altered. This finding is not
surprising considering the importance of Pi in numerous energy-
requiring metabolic and transport processes (21). Effects of long-
term Pi deficiency on various water-soluble metabolite s involved in
P and C metabolism were also examined.
13
C-NMR spectroscopy
analysis showed a reduction in the concentration of fumarate,
whereas glutamine and arginine increased in Pi-deficient plants
(Fig. 1F). The level of fumarate, a storage form of C, is known to
be affected during Pi deficiency (22). Accumulation of polyami-
nated glutamine and arginine in Pi-deficient plants could be an
adaptive response toward meeting the demand for N for protein
synthesis. The
31
P-NMR analysis also revealed a decrease in the
Misson et al. PNAS
August 16, 2005
vol. 102
no. 33
11935
PLANT BIOLOGY
concentrations of inorganic Pi and phosphorylcholine (results not
shown). Reductions in both inorganic Pi and soluble phosphory-
lated compounds in plants grown under Pi-deficiency have been
correlated with an inhibition of growth (23).
Microarray Analysis of the Spatio-Temporally Regulated Pi-Respon-
sive Genes.
ATH1 microarrays were used to evaluate the spatio-
temporal regulation of genes in response to Pi deficiency. During
short-term Pi deficiency (Table 2), 72 genes were induced, whereas
only four genes were suppressed (Fig. 2). These numbers increased
significantly (291 genes induced, 34 genes suppressed) during
medium-term Pi starvation (Table 3). At these two time points,
16% of the induced genes had overlapping expression, whereas only
1 gene was suppressed (Fig. 2). Furthermore, the induction (91
genes) or suppression (22 genes) of some genes was only transient.
This pattern of gene expression points to a very rapid but transient
change occurring even during short periods of Pi deficiency.
Modulation in the expression of the Pi-responsive genes correlated
with a decline of soluble Pi content during early stages of Pi
deficiency treatments (Fig. 1 A) as reported earlier (10). Long-term
Pi deprivation resulted in the differential regulation of 732 genes of
which 501 were induced [228 in the roots (Table 4) and 404 in the
leaves (Table 5)] and 231 were suppressed, 74 in the roots (Table
4) and 169 in the leaves (Table 5). Expression of 26.1% of the
induced and 4.8% of the suppressed genes overlapped in both leaves
and roots. Nevertheless, most of the genes were specific for either
roots or leaves, suggesting that different plant organs respond to Pi
deficiency by activating distinct sets of gene s. Comparison of the
microarray data from all of the three time points showed the
common induction of 48 genes and suppression of only one gene
(Fig. 2). These results are in agreement with results from the smaller
microarrays with Arabidopsis, rice, and white lupin showing similar
patterns of gene expre ssion (7–10). The differential expre ssion of
Pi-responsive genes is considered an adaptive response by plants to
Pi deficiency, which facilitate acquisition of sparingly available Pi
and concurrent attenuation of some of the energy-requiring met-
abolic pathways (21). Furthermore, differential regulation of the
closely related members of the purple acid phosphatase gene family
was found in the roots and leaves of long-term Pi-starved plants.
Likewise, the microarray discriminated among closely related mem-
bers of Pht1 family (Pht1;2 and Pht1;3), which share 95% se-
quence homology (2). This finding clearly demonstrates one ad-
vantage conferred by the use of Affymetrix ATH1 microarray (6).
Validation of the ATH1 Microarray Data. The expression levels
determined from ATH1 arrays were confirmed by a combination
of qRT-PCR, Northern analysis, and promoter-reporter gene fu-
sion studies. Although qRT-PCR results were in general agreement
with the microarray data, quantitative difference s in the modifica-
tion of expression level of some of the genes (At5g05340,
At2g02990, and At5g56870) were observed (Fig. 3A). Extreme
expre ssion ratios of some of the genes that were also barely
expre ssed in one of the conditions (declared absent by the Af-
fymetrix analysis) have a poor quantitative significance (raw values
of microarray re sults are provided in Tables 2–5). Earlier microar-
ray analysis had also shown value s derived from qPCR generally
exceeded those from the array (24). Affymetrix technique was
found to be more sensitive than Northern analysis (e.g., At5g56870
transcripts in the leaves, Fig. 3B). In addition, GFP fused to the
promoter of AtS6K2 [induced in roots of P() plants] revealed in
Fig. 1. Effects of Pi deficiency on
some of the traits of Arabidopsis.
(A) Soluble Pi in whole plant after
transfer to low (open bars) or high
(filled bars) Pi. (B) Anthocyanin ac-
cumulation in the leaves in re-
sponse to Pi deficiency. (C) second-
aryprimary root length ratio in
low-Pi (open bars) or high-Pi (filled
bars). Macroelements (D) and mi-
croelements (E) in the high Pi (black
bars) and low Pi (white bars) leaves
and high Pi (dark gray bars) and low
Pi (light gray bars) roots. (F)
13
C-
NMR analysis. fum, fumarate; Glu,
glutamate; Gln, glutamine; Arg, ar-
ginine; ref, reference (500
mol
maleate). Upper traces correspond
to the enlargement of the corre-
sponding spectrum in the rectangle
in the lower trace.
Fig. 2. Number of genes induced (A) or repressed (B) in low Pi. Comparison of
short-term (diamonds), medium-term (squares), and long-term (circles) experi-
ments. The whole plant was analyzed to monitor the regulation of Pi-responsive
genes during short- and medium-term experiments. Results from leaves and roots
analyzed separately were mixed as long term experiment.
11936
www.pnas.orgcgidoi10.1073pnas.0505266102 Misson et al.
vivo gene expre ssion restricted only to the stele and emerging lateral
roots of Pi-deprived plants (Fig. 3C). This result illustrates the high
sensitivity of the Affymetrix technique, which allowed the evalua-
tion of Pi-responsive genes that are either expre ssed at low levels or
restricted to specific cell type s and substantiates the robustness of
transcript profiling with Affymetrix arrays (6).
Functional Classification of Pi-Responsive Genes with Altered Expres-
sion Patterns.
Pi-responsive genes were categorized into different
functional groups (Table 7). Approximately 84% and 75% of the
induced and suppressed genes, re spectively, had known functions.
The first group contained genes that are related. In addition to
genes harboring general metabolic functions, this study highlights
the involvement of genes related to the uptake and transport of Pi
and other inorganic ions and to the Pi salvage systems. Among the
genes coding the Pht1 family of Pi transporters (2), Pht1;4 was
induced during short-, medium-, and long-term treatments, indi-
cating a rapid and sustained induction of this gene in response to
Pi deprivation, which is consistent with its proposed role in both
acquisition and mobilization of Pi (11, 12). Some of the members
of this family were induced only during medium- andor long-term
Pi-deficiency (Fig. 5A). Spatio-temporal regulation of the gene s of
Pht1 family indicates an apparent lack of functional redundancy
among family members (2). Induction of Pht3;2, encoding a mito-
chondrial ATPADP antiporter, during medium- and long-term Pi
deprivation, was shown (Fig. 5A). Genes involved in Pi mobilization
from organic compounds such as the gene encoding glucose-6PPi
translocator were induced only upon prolonged Pi deficiency. These
‘‘late’’ genes are thought to play a role in promoting the efficient use
of Pi by the plant (25). It is therefore evident from the microarray
results that there is Pi-deficiency-induced spatio-temporal regula-
tion of genes involved in Pi acquisition and its mobilization; complex
mechanisms necessary for plants to thrive under Pi deficiency. Late
induction of PHO1;H1, which is involved in the loading of Pi into
the xylem vessels (26), supports this view. Furthermore, there was
induction of genes coding sulfate transporters (SULTR 1;3, SULTR
3;4) that may have facilitated the higher uptake of S during Pi
deficiency (Fig. 1D), possibly to meet the demand for increased
sulfolipid synthe sis and for balancing the anioncation ratio in the
absence of Pi ions (27). Mechanisms involved in maintenance of Fe
homeostasis (28), such as the differential and coordinated suppres-
sion of the iron transporter IRT1 in the roots and induction of
AtFER1 (encoding a protein involved in iron storage in the chlo-
roplast) in leaves, reflected the plant response to Fe overload
induced by Pi-deficiency. These expression data correlated with the
elevated concentrations of Fe, and other metals observed in
Pi-deficient plants (Fig. 1E) may be linked to increased availability
in the medium in absence of Pi. This theory was confirmed by
running the chemical equilibrium model (
VISUAL MINTEQ version
2.30). In addition, genes for several metal and ATP-binding cassette
transporters were induced in Pi–deprived plants (Tables 6 and 7).
This finding suggests that one adaptive response to Pi deficiency by
the plant is to enhance its capacity to absorb and scavenge Pi-
complexing metals to release sequestered Pi from the medium.
Another group of genes identified by our experiment is involved
in triggering Pi-salvage via the conversion of organic phosphorus
into available Pi (4, 29). Of the 29 genes that encode different
purple acid phosphatases in Arabidopsis, 27 were pre sent on the
ATH1 microarray analysis, and 11 of them were induced by Pi
deficiency (Fig. 4B). In addition, one of the ribonuclease genes
(RNS1) was induced early in response to Pi deprivation, suggesting
that ribonucleases could play a role in the remobilization of P during
Pi deprivation (4). Induction of pyrophosphate-dependent phos-
phofructo-1-kinase and nucleotide pyrophosphatase genes during
medium- and long-term Pi deficiency treatments also repre sent an
important mechanism to reduce Pi demand in plant organs (21).
The distinct induction of the expression of genes for glycerol-3-P
permease s also points toward the complexity of the various adap-
tive modifications that occur in the plant for optimal utilization of
Pi under limiting conditions.
Fig. 3. Validation of ATH1 results. (A) Comparison of
chip results and q-PCR. P()P() ratio in leaves and
roots for some selected genes are shown. Two differ-
ent scales were used for the genes that were up- or
down-regulated. Q-PCR and ATH1 results are means
and SD of three assays performed on triplicates. (B)
Northern blot analysis and chip results [ratio P()
P()] of Pi-responsive genes in leaves (L) and roots (R).
Plants were harvested after 5, 10, and 15 days of
transfer in low [P()] or high Pi [P()], and 10
gof
total RNA, isolated from the whole plant, was hybrid-
ized with
32
P-labeled cDNA fragments of the genes.
Equivalence of RNA loading in all of the lanes is shown
by
32
P-labeled tubulin hybridization and ethidium bro-
mide-stained rRNA (Lower). (C) Expression of GFP
fused to AtS6K2 (At3g08720) promoter in the stele and
emerging lateral roots of transgenic plants grown un-
der high [P()] and low Pi [P()]. (Scale bar, 50
m.)
Misson et al. PNAS
August 16, 2005
vol. 102
no. 33
11937
PLANT BIOLOGY
Detailed analysis of Pi-re sponsive genes also revealed that 7%
(44 genes) are involved in lipid biosynthetic pathways (Table 6). Of
these genes, only two were suppressed. About 50% of the lipid
related genes were induced within 2 days of Pi deprivation. Induced
genes largely represented those coding for enzyme s involved in
phospholipid degradation and galacto- and sulfolipid synthesis (Fig.
4). Interestingly, only a few of the genes coding for phospholipases
D (At3g05630) and C (At3g03540) were induced during Pi defi-
ciency. These results suggest a role for these genes in the lipid
metabolic pathway during Pi deficiency. Genes involved in the
subsequent utilization of DAG to synthesize galactolipids (MGDG,
DGDG) were strongly up-regulated at early stages of Pi depriva-
tion, which is consistent with previous data (30). Genes coding for
MGDG synthases (MGD2 and MGD3) were induced 4- to 10-fold
during short-term Pi deprivation; whereas expression of DGD1 and
DGD2, coding for DGDG synthases, was enhanced during medi-
um- and long-term Pi deficiency, re spectively. Similarly, the genes
enc oding UDP glucose-4-epimerase and UDP galactose-4-
epimerase, which convert UDP-glucose to UDP-galactose (galac-
tolipid precursor), were induced during medium- and long-term Pi
deficiency. This induction could facilitate the production of galac-
tose required for galactolipid synthesis. Comparatively, the genes
coding for UDP-sulfoquinovose synthase and UDP-sulfoquinovo-
syl:DAG sulfoquinovosyltransferase exhibited early and sustained
induction during Pi deficiency treatments; this was reflected by a
4-fold increase in the level of SQDG in P() leaves during
long-term Pi deficiency. Although SQDG is not considered essen-
tial for plant development (31), under Pi deficiency it could possibly
replace PG and may allow photosynthesis to continue despite a
reduction in phospholipid content. Arabidopsis mutants defective in
sulfolipid synthase show impaired growth during Pi deprivation
(31). These modulations of lipid biosynthetic pathways indicate a
complex mechanism to replace membrane phospholipids with
nonphosphorus galacto- and sulfolipids that may have evolved to
scavenge and conserve Pi in plants under Pi-limiting conditions
(30). These results correlated with variations in phospholipids,
sulfolipids and glycosylglyceride s (Fig. 4). Alteration of the lipid
content became apparent within 2 days, whereby a decrease of PG
and PC was compensated by an increase of SQDG and DGDG. In
leaves of plants grown in Pi-deficient medium, reduction in the
levels of all phospholipids except DPG was observed. Interestingly
in the P() roots no significant difference was detected in any of
the phospholipid species including PC, but there was a substantial
increase in the level of DGDG. This finding suggests that lipid
composition is more sensitive to Pi deficiency in leaves than in roots,
as indicated earlier (16). Despite an early induction of MGD2 and
MGD3, there was no significant increase in MGDG level, even
during long-term Pi deficiency; this may be due to rapid conversion
of MGDG into DGDG by DGD1 and DGD2, whose activity
increased during long-term Pi deficiency. Furthermore, the genes
DGD1 and DGD2 exhibited differential regulation in roots and
leaves. Earlier studie s have also shown the preferential biogenesis
of DGD2 and DGD1 outside of plastids and in the chloroplast
membrane, respectively (30, 32). The pre sent microarray analysis
revealed an early, sustained, and coordinated induction of a host of
Pi-responsive genes involved in Pi acquisition and conversion of
organic phosphorus into available Pi. These experiments also
indicated that Pi deprivation is perceived at the molecular level as
soon as Pi is withdrawn from the medium, suggesting that the plant
is able to sense decrease of Pi concentration either in the medium
or in cells.
Because accumulation of anthocyanin is a characteristic response
of plants to long-term Pi deficiency (Fig. 1B), microarray data were
also evaluated for the differential regulation of genes involved in its
biosynthesis (Fig. 6). At each step of the anthocyanin biosynthetic
pathway leading to the synthesis of cyanidin, pelargonidin, and
flavonoids, at least one gene was found to be induced. Although the
majority of these genes were induced only during long-term Pi
deficiency treatment and more specifically in the leaves, a few of
them encoding flavonol 3-O-glucosyltransferase were significantly
up-regulated after 2 days of Pi deficiency. On average, genes were
induced 3- to 4-fold, with the notable exceptions of dihydroflavo-
nol-4-reduct ase, anthoc yanindin synthase, and flavonone-3-
hydroxylase, which showed 5- to 6-fold induction. Moreover, two
genes coding for anthocyanin 5-aromatic acyltransferase and one
for anthocyanin2 were induced (Table 7). Interestingly, in Pi-
deficient leaf samples, there was an induction of a gene coding for
chalcone synthase, which is responsible for the conversion of
4-coumarate to naringenin in flavonoid biosynthetic pathway (Fig.
6). Accumulation of naringenin has been shown to affect the
transport of auxin to roots (33), which could be re sponsible for an
altered root architecture, a hallmark of plant response s to Pi
deficiency (34). In addition, the expression of some of the genes
associated with phytohormone responses was also altered during Pi
deficiency. For instance, genes involved in auxin response were
modulated during medium-and long-term Pi-deprivation (Table 7).
Fig. 4. Transcriptional regulation in
pathway of glycosylglyceride biosyn-
thesis in short-term (squares), medium-
term (diamonds), and long-term
[leaves (circles) and roots (triangles)
analyzed separately] experiments.
Fold change: red, 10; orange, 4–10;
yellow, 2– 4; white, 0.5–2; green, 0.25–
0.5; pale blue, 0.1– 0.25; dark blue,
0.1. *, Significant, one-way ANOVA,
P 0.05. (Insets) Metabolite quantifi-
cation [% of total lipids in P() and
P() and ratio P()P()] in short-,
medium- and long- (leaf, root) term
treatments. DPG, diphosphatidylglyc-
erol; PG, phosphatidylglycerol; PI,
phosphatidylinositol; PE, phosphati-
dylethanolamine; PC, phosphatidyl-
choline; SQDG, sulfoquinovosyldiacyl
glycerol; MGDG, galactosyl-1,2-diacyl-
glycerol; DGDG, digalactosyl-diacyl-
glycerol.
11938
www.pnas.orgcgidoi10.1073pnas.0505266102 Misson et al.
A gene for an auxin response element, which was induced in
Pi-deficient roots, is known to be involved in transcriptional regu-
lation (35). Likewise, ethylene-response genes also showed differ-
ential regulation in response to Pi deprivation. The role of auxin and
ethylene in the Pi-starvation response is well established and thus
it is not surprising to see the changes in the expre ssion of genes
involved in the response pathways for these hormone s.
Pi deficiency is known to cause significant stress to the plant and
this was reflected in modulation of an array of stress-related genes
(Tables 6 and 7). A large number of them are related to disease or
pathogen resistance (e.g., chitinases, PR-proteins) and toxin catab-
olism (GSTs). General features of oxidative stress [superoxide
dismutase (SOD), peroxidase, GST, cytochrome P450] are also
strongly induced in Pi-deficient samples mainly in long-term ex-
periments (Table 7). This finding suggests interactions among
several stre ss-related pathways that may have functional implica-
tions in plant survival under Pi deficiency. Pi starvation also
resulted in spatio-temporal expre ssion of the genes (10 induced and
two suppressed) involved in controlling the level of reactive oxygen
species (ROS) (Table 7). They encode different ROS-scavenging
enz y mes such as SOD, monodehydroasc orbate reduct ase
(MDAR), glutathione peroxidase (GPX). One of the genes coding
for NADPH oxidase-like enzyme that is responsible for generating
ROS showed a transient suppression during medium-term Pi
deprivation. Recent microarray analysis of Arabidopsis, subjected to
different abiotic stre sses, had also demonstrated differential regu-
lation of 152 genes coding for different enzymes involved in
scavenging and generation of ROS (36).
In Pi-limiting conditions, genes coding for enzyme s involved in
protein degradation (six genes) and protein biosynthesis (21 genes)
are induced and suppressed, respectively (Table 7), sugge sting that
initiation of Pi recycling processe s. Furthermore, gene s coding for
protein phosphatase s and kinase s were up-regulated during Pi
deficiency. Some of the protein kinase genes were also found to be
down-regulated in leaves and roots of Pi-deficient plants (Table 7).
During medium-and long-term Pi deficiency, modulation in the
expre ssion of the genes encoding various enzymes involved in cell
wall metabolism was observed, and this finding is in conformity with
earlier microarray analysis (9). Majority of these induced genes (2-
to 10-fold) encoded enzymes like xyloglucan endo-1,4-
-D-
glucanase, polygalacturonase inhibiting protein-1, put ative
pectinesterase, (1,4)-
-mannan endohydrolase precursor, and po-
lygalacturonase inhibiting protein. A few genes for xylosidase or
cellulose synthase-like protein were found to be down-regulated.
Aside from the production of galactose to support increased
synthesis of MGDG and DGDG, increased galactose may also have
a critical role in cell wall biosynthesis of the modified root archi-
tecture, as shown by transcript analysis of lateral root induction (37).
Analysis of long-term Pi-deprived plants revealed that genes for
these enzymes were mainly modulated in roots.
Because Pi deficiency responses are known to be regulated at the
transcription level (1), microarray data were further analyzed for
the Pi-deficiency-induced genes encoding transcription regulatory
elements. The expre ssion of a total of 80 genes, presumed to be
associated with transcriptional regulation of gene expression, was
altered during Pi deficiency (Table 6). A few of them were
up-regulated during short-term (five gene s) and medium-term (10
genes) Pi deficiency. However, their induction was more pro-
nounced (47 genes) during long-term Pi deprivation. Interestingly,
the induction of only a small number of transcription factor genes
overlapped during different stage s of Pi deficiency. This finding
suggests that specific sets of transcription factors are involved in
regulating early and late response s of plants to Pi deficiency. To gain
further insight into the mode of regulation of Pi-responsive genes,
the conserved sequences located upstream (1to2,000 bp) of the
ATG start codon were analyzed (Tables 8 and 9). Promoters of the
Pi-responsive genes coding for Pi transporters, phosphatase s, and
those involved in protein synthesis were found to be significantly
enriched with the PHR1 binding sequence, which is recognized by
a MYB-domain-containing transcription factor (5). Furthermore,
deduced protein sequence of some of the Pi-responsive gene s
harbored an SPX domain, which has been identified in proteins
involved in either transport or sensing of Pi (26).
Induction of phosphatase s and kinase s further suggests the
involvement of numerous posttranslational modifications, which
remain to be identified. This study thus presents a global analysis of
plant transcriptomic responses to Pi deficiency and physiological
and biochemical correlations with observed phenotypes. The list of
putative targets established will significantly add to our knowledge
about the complex molecular processes associated with Pi nutrition.
We thank P. Auroy and P. Richaud for the analysis of ion ic content; M.
Pean, S. Boiry, and the GRAP team for plant growth and
13
C enrichment
experiments; D. Varadarajan for Arabidopsis culture; K. Ramonell and
L. Rose for the chip hybridization and initiation to Genespring; M. H.
Montane and A. Nublat for their help for the qPCR technique; and A. M.
Boisson and E. Gout for the NMR analysis. This project was supported
in part by a grant from Commissariat a` l’Energie Atomique and region
Provence-Alpes-Coˆte d’Azur (to J.M.) and U.S. Department of Agri-
culture (to K.G.R.).
1. Raghothama, K. G. (1999) Annu. Rev. Plant Physiol. Plant Mol. Biol. 50, 665–693.
2. Mudge, S. R., Rae, A. L., Diatlof f, E. & Smith, F. W. (2002) Plant J. 31, 341–353.
3. Baldwin, J. C., Karthikeyan, A. S. & Raghothama, K. G. (2002) Plant Physiol. 125, 728–737.
4. Bariola, P. A., Howard, C. J., Taylor, C. B., Verburg, M. T., Jaglan, V. D. & Green, P. J.
(1994) Plant J. 6, 673–685.
5. Rubio, V., Francisco, L., Roberto, S., Ana, C., Martin, J. I., Antonio, L. & Paz-Ares, J. (2001)
Gene Dev. 15, 2122–2133.
6. Redman, J. C., Haas, B. J., Tan imoto, G. & Town, C. D. (2004) Plant J. 38, 545–561.
7. Hammond, J. P., Bennett, M. J., Bowen, H. C., Broadley, M. R., Eastwood, D. C., May, S. T.,
Rahn, C., Swarup, R., Woolaway, K. E. & White, P. J. (2003) Plant Physiol. 132, 578–596.
8. Uhde-Stone, C., Zinn, K. E., Ramirez-Yanez, M., Li, A., Vance, C. P. & Allan, D. L. (2003)
Plant Physiol. 131, 1064–1079.
9. Wasaki, J., Yonetani, R., Kuroda, S., Shinano, T., Yazaki, J., Fujii, F., Shimbo, K.,
Yamamoto, K., Sakata, K., Sasaki, T., et al . (2003) Plant Cell Environ. 26, 1515–1523.
10. Wu, P., M, a L., Hou, X., Wang, M., Wu, Y., Liu, F. & Deng, X. W. (2003) Plant Physiol.
132, 1260–1271.
11. Karthikeyan, A. S., Varadarajan, D. K., Mukatira, U. T., D’Urzo, M. P., Damsz, B. &
Raghothama, K. G. (2002) Plant Physiol. 130, 221–233.
12. Misson, J., Thibaud, M. C., Bechtold, N., Raghothama, K. G. & Nussaume, L. (2004) Plant
Mol. Biol. 55, 727–741.
13. Schaffer, R., Landg raf, J., Accerbi, M., Simon, V., Larson, M. & Wisman, E. (2001) Plant
Cell 13, 113–123.
14. Bechthold, N., Elis, J. & Pelletier, G. (1993) C. R. Acad . Sci. Paris 316, 1194–1199.
15. Folch, J., Lees, M. & Stanley, G. H. S. (1957) J. Biol. Chem. 226, 497–509.
16. Jouhet, J., Marechal, E., Bligny, R., Joyard, J. & Block, M. A. (2003) FEBS Lett. 544, 63– 68.
17. Lange, H., Shropshire, W., Jr., & Mohr, H. (1971) Plant Physiol. 47, 649655.
18. Aubert, S., Bligny, R. & Douce, R. (1996) in Current Topics in Plant Physiology, eds.
Shashar-Hill, Y. & Pfeffer, P. (Am. Soc. Plant Physiol., Rockv ille, MD), pp. 109–154.
19. Westoby, M., Leishman, M. & L ord, J. (1997) in Plant Life Histor ies: Ecology, Phylogeny, and
Evolution, eds. Silvertown, J. W., Franco, M. & Harper, J. L. (Cambridge Univ. Press,
Cambridge, U.K.), pp. 143–162.
20. Ma, Z., Bask in, T. I., Brown, K. M. & Lynch, J. P. (2003) Plant Physiol. 131, 1381–1390.
21. Plaxton, W. C. & Carswell, M. C. (1999) in Plant Responses to Environmental Stresses: From
Phytohormones to Genome Organization, ed. Lerner, H. R. (Dekker, New York), pp. 349–372.
22. Chia, D. W., Yoder, T., Reiter, W.-D. & Gibson, S. (2000) Planta 211, 743–751.
23. Gout, E., Boisson, A. M., Aubert, S., Douce, R. & Bligny, R. (2001) Plant Physiol. 125, 912–925.
24. Wang, R., Okamoto, M., Xing, X. & Crawford, N. M. (2003) Plant Physiol. 132, 556–567.
25. Hammond J. P., Broadley, M. R. & White, P. J. (2004) Ann. Bot. 94, 323–332.
26. Wang, Y., Ribot, C., Rezzonico, E. & Yves Poirier, Y. (2004) Plant Physiol . 135, 400 411.
27. Essigmann, B., Guler, S., Narang, R. A., Linke, D. & Benning, C. (1998) Proc. Natl . Acad.
Sci. USA 95, 1950–1955.
28. Petit, J. M., Briat, J.-F. & L obre´aux, S. (2001) Biochem J. 359, 575–582.
29. Duff, S. M. G., Sarath, G. & Plaxton, W. C. (1994) Physiol. Plant. 90, 791–800.
30. Benning, C. & Otha, H. (2005) J. Biol. Chem. 280, 2397–2400.
31. Yu, B., Xu, C. & Benning, C. (2002) Proc. Natl. Acad. Sci. USA 99, 5732–5737.
32. Jouhet, J., Mare´chal, E., Baldan, B., Bligny, R., Joyard, J. & Block, M. A. (2004) J. Cell Biol.
167, 863–874.
33. Brown, D. E., Rashotte, A. M., Murphy, A. S., Normanly, J., Tague, B. W., Peer, W. A., Taiz,
L. & Muday, G. K. (2001) Plant Physiol. 126, 524–535.
34. Lo´pez-Bucio, J., Herna´ndez-Abreu, E., Sa´nchez-Caldero´n, L., Nieto-Jacobo, M. R., Simp-
son, J. & Herrera-Estrella L. (2002) Plant Physiol. 129, 244–256.
35. Ulmasov, T., Hagen, G. & Guilfoyle, T. J. (1999) Plant J. 19, 309 –319.
36. Mittler, R., Vanderauwera, S., Gollery, M. & Van Breusegem, F. (2004) Trends Plant Sci. 9, 490 498.
37. Brinker, M., van Zyl, L., Liu, W., Craig, D., Sederoff, R. R., Clapham, D. H. & von Arnold,
S. (2004) Plant Physiol. 135, 1526–1539.
Misson et al. PNAS
August 16, 2005
vol. 102
no. 33
11939
PLANT BIOLOGY
... In addition, other physiological modifications controlled by phosphate starvation responses (PSRs) can be observed. They trigger multiple processes related to Pi recovery, transport, and recycling (Misson et al. 2005;Thibaud et al. 2010). Known as the systemic Pi response, these reactions are controlled by a regulatory module involving two conserved protein families : (i) the PHR1 master transcription factor gene family Rubio et al. 2001) and (ii) their inhibitors, SPX proteins (Lv et al. 2014;Puga et al. 2014;Wang et al. 2014a, b). ...
... Alternative pathways are also activated. For instance, phosphate starvation conditions induce several enzymes that bypass various steps of glycolysis (Misson et al. 2005;Plaxton and Tran 2011;Thibaud et al. 2010). These enzymes utilize inorganic pyrophosphate (PPi), such as nucleoside diphosphate kinase or PPi-phosphofructokinase, for phosphorylation reactions instead of ATP. ...
... In some cases, they even divert ATP-requiring reactions, as evidenced in the case of nonphosphorylating NADP glyceraldehyde 3-Pi dehydrogenase. Transcriptional regulation of these enzymes occurs rapidly, typically within one to two days following phosphate starvation (Misson et al. 2005). Within the same timeframe, a major metabolic shift occurs as phospholipids are converted into galactolipids (Cruz-Ramirez et al. 2006;Dormann and Benning 2002;Gaude et al. 2008;Hartel et al. 2000;Jouhet et al. 2004) and sulfolipids (Hammond et al. 2003;Misson et al. 2005;Thibaud et al. 2010;Yu et al. 2002). ...
Article
Full-text available
Adapting to varying phosphate levels in the environment is vital for plant growth. The PHR1 phosphate starvation response transcription factor family, along with SPX inhibitors, plays a pivotal role in plant phosphate responses. However, this regulatory hub intricately links with diverse biotic and abiotic signaling pathways, as outlined in this review. Understanding these intricate networks is crucial, not only on a fundamental level but also for practical applications, such as enhancing sustainable agriculture and optimizing fertilizer efficiency. This comprehensive review explores the multifaceted connections between phosphate homeostasis and environmental stressors, including various biotic factors, such as symbiotic mycorrhizal associations and beneficial root-colonizing fungi. The complex coordination between phosphate starvation responses and the immune system are explored, and the relationship between phosphate and nitrate regulation in agriculture are discussed. Overall, this review highlights the complex interactions governing phosphate homeostasis in plants, emphasizing its importance for sustainable agriculture and nutrient management to contribute to environmental conservation.
... Once the electron transport to ferredoxin-NADP + oxidoreductase is blocked, oxygen may serve as an alternative electron acceptor, running the risk of massive O 2 − generation 28 28 , such as KatG (Fig. 7c). In fact, SOD induction in response to P limitation could be a common strategy amongst photosynthetic organisms since a similar phenomenon is widely observed in cyanobacteria 42 , a marine diatom 43 , and land plants 44 . It should be noted that SOD induction could be extremely important for cyanobacteria that grow under Fedeficient conditions because O 2 − damage occurs primarily at PSI 45 . ...
... Nutrient amendment experiments performed in the Fepoor western North Atlantic Ocean showed that the alkaline phosphatase activity of the community responded positively to Fe additions, implicating a role of Fe in P acquisition 47 . The phenomenon of increased Fe content under P deficiency, shown here ( Supplementary Fig. 18), is widely reported in photosynthetic organisms 24,44,48 . Thus, regulation of a Fe uptake gene by PhoB is physiologically necessary, as it allows cyanobacteria to meet their increased Fe requirement under P deficiency. ...
Article
Full-text available
Iron and phosphorus are essential nutrients that exist at low concentrations in surface waters and may be co-limiting resources for phytoplankton growth. Here, we show that phosphorus deficiency increases the growth of iron-limited cyanobacteria (Synechocystis sp. PCC 6803) through a PhoB-mediated regulatory network. We find that PhoB, in addition to its well-recognized role in controlling phosphate homeostasis, also regulates key metabolic processes crucial for iron-limited cyanobacteria, including ROS detoxification and iron uptake. Transcript abundances of PhoB-targeted genes are enriched in samples from phosphorus-depleted seawater, and a conserved PhoB-binding site is widely present in the promoters of the target genes, suggesting that the PhoB-mediated regulation may be highly conserved. Our findings provide molecular insights into the responses of cyanobacteria to simultaneous iron/phosphorus nutrient limitation.
... The reduction in the level of P results in the inhibition of root elongation that is restored under Fe-deficient conditions without any interference in P availability. Contrary to this, an increase in Fe accumulation was reported in the shoots and roots of Arabidopsis following growth under P-starved conditions [235,247]. In plants, a higher Fe concentration observed under P-deficient conditions is attributed to the expression of the Fe-responsive genes [232]. ...
... In phosphate-deficient plants, a transcriptomic analysis revealed that the expression of genes associated with the regulation of excess iron is enhanced, while a decrease in gene expression associated with iron deficiency was observed [247,248]. The expression of the gene AtFER1 (encode ferritin protein) that regulates iron storage was induced in plants with a P deficiency. ...
Article
Full-text available
Plants being sessile are exposed to different environmental challenges and consequent stresses associated with them. With the prerequisite of minerals for growth and development, they coordinate their mobilization from the soil through their roots. Phosphorus (P) and iron (Fe) are macro- and micronutrient; P serves as an important component of biological macromolecules, besides driving major cellular processes, including photosynthesis and respiration, and Fe performs the function as a cofactor for enzymes of vital metabolic pathways. These minerals help in maintaining plant vigor via alterations in the pH, nutrient content, release of exudates at the root surface, changing dynamics of root microbial population, and modulation of the activity of redox enzymes. Despite this, their low solubility and relative immobilization in soil make them inaccessible for utilization by plants. Moreover, plants have evolved distinct mechanisms to cope with these stresses and coregulate the levels of minerals (Fe, P, etc.) toward the maintenance of homeostasis. The present study aims at examining the uptake mechanisms of Fe and P, and their translocation, storage, and role in executing different cellular processes in plants. It also summarizes the toxicological aspects of these minerals in terms of their effects on germination, nutrient uptake, plant–water relationship, and overall yield. Considered as an important and indispensable component of sustainable agriculture, a separate section covers the current knowledge on the cross-talk between Fe and P and integrates complete and balanced information of their effect on plant hormone levels.
... Expression analyses of the various tissues of A. thaliana have allowed the likely location of the different AtPHT1 proteins to be mapped and putative roles identified . Nearly all of the AtPHT1 genes are repressed by Pi (Lapis-Gaza et al. 2014;Misson et al. 2005). Studies examining Pi responsiveness of PHT1 gene expression in plants such as A. thaliana and crops such as Oryza sativa (rice), Zea mays (maize), Setaria italica (foxtail millet) and Triticum aestivum (wheat) often include a high-P treatment ranging from 200 µM to 500 µM Pi (Ai et al. 2009;Ceasar et al. 2014;Misson et al. 2004;Nagy et al. 2006;Teng et al. 2017). ...
... (Barragán-Rosillo et al. 2021;Bustos et al. 2010;Lapis-Gaza et al. 2014;Misson et al. 2005;Morcuende et al. 2007;Scheible et al. 2023). ...
Article
Full-text available
Background and aims Phosphorus (P) is an essential plant nutrient and integral for crop yield. However, plants adapted to P-impoverished environments, such as Hakea prostrata (Proteaceae), are often sensitive to P supplies that would be beneficial to other plants. The strategies for phosphate uptake and transport in P-sensitive species have received little attention. Methods Using a recently-assembled transcriptome of H. prostrata, we identified 10 putative members of the PHOSPHATE TRANSPORTER1 (PHT1) gene family, which is responsible for inorganic phosphate (Pi) uptake and transport in plants. We examined plant growth, organ P concentrations and the transcript levels for the eight PHT1 members that were expressed in roots of H. prostrata at Pi supplies ranging from P-impoverished to P-excess. Key results Hakea prostrata plants suppressed cluster root growth above ecologically-relevant Pi supplies, whilst non-cluster root mass ratios were constant. Root P concentrations increased with increasing Pi supply. Of the eight H. prostrata PHT1 genes tested, four had relatively high transcript amounts in young roots suggesting important roles in Pi uptake; however, a maximum five-fold difference in expression between P-impoverished and P-excess conditions indicated a low P-responsiveness for these genes. The HpPHT1;8 and HpPHT1;9 genes were paralogous to Pi-responsive Arabidopsis thaliana PHT1;8 and PHT1;9 orthologues involved in root-to-shoot translocation of P, but only HpPHT1;9 was P responsive. Conclusions An attenuated ability of H. prostrata to regulate PHT1 expression in response to Pi supply is likely responsible for its low capacity to control P uptake and contributes to its high P sensitivity.
... Zinc is a crucial element for plant metabolism, enzymatic activity, and ion transport. Previous studies have demonstrated an antagonistic relationship between Zn and P, as a deficiency in phosphorus can stimulate the expression of genes related to Zn and Fe homeostasis [52]. Our study found that the HS solution had a higher concentration of P compared to the other treatments, but it resulted in lower absorption of Zn. ...
Article
Full-text available
Organic products are gaining popularity due to their positive impact on human health and the environment. While hydroponics is commonly used in vegetable production, it relies on mineral fertilizers derived from limited and non-renewable resources. As a result, farmers are actively seeking sustainable farming solutions. This study comprehensively evaluated the effectiveness of vermi-liquids (organic nutrient solutions) as a replacement for conventional inorganic nutrient solutions in promoting growth and nutrient acquisition in Diplotaxis muralis plants in a controlled environment. The results showed that plant biomass and SPAD values of D. muralis grown in Hoagland solution and enhanced vermitea (vermitea having relatively low pH and high EC) were higher compared to standard vermitea (high pH and low EC). The findings also revealed improved nutrient assimilation of phosphorus, potassium, calcium, iron, manganese, copper, and zinc in the enhanced vermitea plants. The heavy metal contents in D. muralis leaves were evaluated, too, and they were found to fall significantly below the safe threshold, rendering them safe for human consumption. However, the standard vermitea, with its high pH and low EC, performed poorly as a hydroponic solution. This research suggests that enhanced vermitea can completely replace chemical nutrient solutions in hydroponic agriculture. This substitution could lead to reduced production costs and improved product quality.
... Pearson's correlation coefficient showed an antagonism among P, and the microelements Fe, Cu, and Mn which agrees with numerous references in the literature [51,[69][70][71][72]. Sánchez-Rodríguez et al. [73,74] in studies conducted in calcareous soils with species with different sensitivities to Fe chlorosis, revealed that phosphate fertilization modifies the availability of Fe in the soil and aggravates Fe chlorosis in sensitive plants. ...
Article
Full-text available
The European “Green Deal” policies are shifting toward more sustainable and environmentally conscious agricultural practices, reducing the use of chemical fertilizer and pesticides. This implies exploring alternative strategies. One promising alternative to improve plant nutrition and reinforce plant defenses is the use of beneficial microorganisms in the rhizosphere, such as “Plant-growth-promoting rhizobacteria and fungi”. Despite the great abundance of iron (Fe) in the Earth’s crust, its poor solubility in calcareous soil makes Fe deficiency a major agricultural issue worldwide. Among plant promoting microorganisms, the yeast Debaryomyces hansenii has been very recently incorporated, for its ability to induce morphological and physiological key responses to Fe deficiency in plants, under hydroponic culture conditions. The present work takes it a step further and explores the potential of D. hansenii to improve plant nutrition and stimulate growth in cucumber plants grown in calcareous soil, where ferric chlorosis is common. Additionally, the study examines D. hansenii’s ability to induce systemic resistance (ISR) through a comparative relative expression study by qRT-PCR of ethylene (ET) biosynthesis (ACO1), or ET signaling (EIN2 and EIN3), and salicylic acid (SA) biosynthesis (PAL)-related genes. The results mark a significant milestone since D. hansenii not only enhances nutrient uptake and stimulates plant growth and flower development but could also amplify induced systemic resistance (ISR). Although there is still much work ahead, these findings make D. hansenii a promising candidate to be used for sustainable and environmentally friendly integrated crop management.
Article
Full-text available
Post-anthesis effect of phosphorus (P)-deficit is less studied in plants. P-deficit analysis is rarely performed in different tissues of the plant together. The present study analyzed total-P, inorganic-P, acid phosphatases, ribonucleases, carbohydrates, and antioxidants in the roots, lower leaves, flag leaves, and spikelets at 0, 10, 20, and 30 days after anthesis and recorded yield components at harvest in rice cultivars CSR10 (low-P tolerant) and Pusa44 (low-P susceptible) under P-deprivation (P0) compared to P-application at 30 kg ha⁻¹ (P30). Total-P was deficient in the roots and leaves of Pusa44 compared to CSR10 under P0. Acid phosphatases and ribonucleases were enhanced by roots under P0, and the increase was exceptionally high in CSR10. Hexoses, sucrose, and starch levels remained low in the vegetative tissues of Pusa44 compared to CSR10 under P0. Starch and sucrose were mobilized from vegetative tissues especially from the roots of CSR10 compared to Pusa44 under P0. Sucrose and starch levels were high in the spikelets of CSR10 compared to Pusa44 under P0. Antioxidants were high in the vegetative tissues especially in the roots of CSR10 compared to Pusa44 under P0. At harvest, unfilled grains per panicle increased and grain yield reduced in Pusa44 under P0. Results concluded that post-anthesis roots metabolic activities, in the form of acid phosphatases/ribonucleases for P-acquisition and antioxidants to counteract stress and starch/sucrose mobilization toward spikelets, may play an important role toward low-P tolerance in rice.
Article
Full-text available
Introduction Phosphorus (P) deficiency in plants creates a variety of metabolic perturbations that decrease photosynthesis and growth. Phosphorus deficiency is especially challenging for the production of bioenergy feedstock plantation species, such as poplars (Populus spp.), where fertilization may not be practically or economically feasible. While the phenotypic effects of P deficiency are well known, the molecular mechanisms underlying whole-plant and tissue-specific responses to P deficiency, and in particular the responses of commercially valuable hardwoods, are less studied. Methods We used a multi-tissue and multi-omics approach using transcriptomic, proteomic, and metabolomic analyses of the leaves and roots of black cottonwood (Populus trichocarpa) seedlings grown under P-deficient (5 µM P) and replete (100 µM P) conditions to assess this knowledge gap and to identify potential gene targets for selection for P efficiency. Results In comparison to seedlings grown at 100 µM P, P-deficient seedlings exhibited reduced dry biomass, altered chlorophyll fluorescence, and reduced tissue P concentrations. In line with these observations, growth, C metabolism, and photosynthesis pathways were downregulated in the transcriptome of the P-deficient plants. Additionally, we found evidence of strong lipid remodeling in the leaves. Metabolomic data showed that the roots of P-deficient plants had a greater relative abundance of phosphate ion, which may reflect extensive degradation of P-rich metabolites in plants exposed to long-term P-deficiency. With the notable exception of the KEGG pathway for Starch and Sucrose Metabolism (map00500), the responses of the transcriptome and the metabolome to P deficiency were consistent with one another. No significant changes in the proteome were detected in response to P deficiency. Discussion and conclusion Collectively, our multi-omic and multi-tissue approach enabled the identification of important metabolic and regulatory pathways regulated across tissues at the molecular level that will be important avenues to further evaluate for P efficiency. These included stress-mediating systems associated with reactive oxygen species maintenance, lipid remodeling within tissues, and systems involved in P scavenging from the rhizosphere.
Article
Full-text available
To cope with phosphate (Pi) starvation, plants trigger an array of adaptive responses to sustain their growth and development. These responses are largely controlled at transcriptional levels. In Arabidopsis (Arabidopsis thaliana), PHOSPHATE RESPONSE 1 (PHR1) is a key regulator of plant physiological and transcriptional responses to Pi starvation. PHR1 belongs to a MYB-CC-type transcription factor family which contains 15 members. In this PHR1 family, PHR1/PHR1-like 1(PHL1) and PHL2/PHL3 form two distinct modules in regulating plant development and transcriptional responses to Pi starvation. PHL4 is the most closely related member to PHR1. Previously, using the phr1phl4 mutant, we showed that PHL4 is also involved in regulating plant Pi responses. However, the precise roles of PHL1 and PHL4 in regulating plant Pi responses and their functional relationships with PHR1 have not been clearly defined. In this work, we further used the phl1phl4 and phr1phl1phl4 mutants to perform comparative phenotypic and transcriptomic analyses with phr1, phr1phl1, and phr1phl4. The results showed that both PHL1 and PHL4 act redundantly and equally with PHR1 to regulate leaf senescence, Pi starvation induced-inhibition of primary root growth, and accumulation of anthocyanins in shoots. Unlike PHR1 and PHL1, however, the role of PHL4 in maintaining Pi homeostasis is negligible. In regulating transcriptional responses to Pi starvation at genomic levels, both PHL1 and PHL4 play minor roles when acts alone, however, they act synergistically with PHR1. In regulating Pi starvation-responsive genes, PHL4 also function less than PHL1 in terms of the number of the genes it regulates and the magnitude of gene transcription it affects. Furthermore, no synergistic interaction was found between PHL1 and PHL4 in regulating plant response to Pi starvation. Therefore, our results clarified the roles of PHL1 and PHL4 in regulating plant responses to Pi starvation. In addition, this work revealed a new function of these three transcription factors in regulating flowering time.
Article
Full-text available
From germination to maturity, crops face myriad stresses thereby threatening food security. The foundation of modern agriculture rests on the status of seed health and resilience. Hence, developing highly efficient, low-cost, farmer-friendly, and sustainable approaches for improving seed health and performance under both field and greenhouse conditions. Seed bio-priming with plant beneficial microorganisms (++; mutualistic) improves the physiological, molecular, and stress tolerance functions of the seeds. This process allows the microorganisms adhere to the seed coat and establish an early relationship with the radicle, thereby forming the first line of defense against any external threat. Seeds bio-primed by mutualistic rhizomicroorganisms stimulate plant immunity by inducing the biosynthesis of defense-related proteins, phytohormones, antioxidants, polyphenols, etc. This review maps the various functional and applied aspects of seed bio-priming on the overall plant health under stressed environments. Furthermore, it critically examines the modulation of biochemical and molecular mechanisms for establishing redox homeostasis.
Article
Full-text available
Photoassimilates are used by plants for production of energy, as carbon skeletons and in transport of fixed carbon between different plant organs. Many studies have been devoted to characterizing the factors that regulate photoassimilate concentrations in different plant species. Most studies examining photoassimilate concentrations in C3 plants have focused on analyzing starch and soluble sugars. However, work presented here demonstrates that a number of C3 plants, including the popular model organism Arabidopsis thaliana (L.) Heynh., and agriculturally important plants, such as soybean, Glycine max (L.) Merr., contain significant quantities of fumaric acid. In fact, fumaric acid can accumulate to levels of several milligrams per gram fresh weight in Arabidopsis leaves, often exceeding those of starch and soluble sugars. Fumaric acid is a component of the tricarboxylic acid cycle and, like starch and soluble sugars, can be metabolized to yield energy and carbon skeletons for production of other compounds. Fumaric acid concentrations increase with plant age and light intensity in Arabidopsis leaves. Moreover, Arabidopsis phloem exudates contain significant quantities of fumaric acid, raising the possibility that fumaric acid may function in carbon transport.
Article
Polar transport of the plant hormone auxin controls many aspects of plant growth and development. A number of synthetic compounds have been shown to block the process of auxin transport by inhibition of the auxin efflux carrier complex. These synthetic auxin transport inhibitors may act by mimicking endogenous molecules. Flavonoids, a class of secondary plant metabolic compounds, have been suggested to be auxin transport inhibitors based on their in vitro activity. The hypothesis that flavonoids regulate auxin transport in vivo was tested in Arabidopsis by comparing wild-type (WT) and transparent testa (tt4) plants with a mutation in the gene encoding the first enzyme in flavonoid biosynthesis, chalcone synthase. In a comparison between tt4 and WT plants, phenotypic differences were observed, including three times as many secondary inflorescence stems, reduced plant height, decreased stem diameter, and increased secondary root development. Growth of WT Arabidopsis plants on naringenin, a biosynthetic precursor to those flavonoids with auxin transport inhibitor activity in vitro, leads to a reduction in root growth and gravitropism, similar to the effects of synthetic auxin transport inhibitors. Analyses of auxin transport in the inflorescence and hypocotyl of independent tt4 alleles indicate that auxin transport is elevated in plants with a tt4 mutation. In hypocotyls of tt4, this elevated transport is reversed when flavonoids are synthesized by growth of plants on the flavonoid precursor, naringenin. These results are consistent with a role for flavonoids as endogenous regulators of auxin transport.
Article
Phosphate (Pi) is one of the least available plant nutrients found in the soil. A significant amount of phosphate is bound in organic forms in the rhizosphere. Phosphatases produced by plants and microbes are presumed to convert organic phosphorus into available Pi, which is absorbed by plants. In this study we describe the isolation and characterization of a novel tomato (Lycopersicon esculentum) phosphate starvation-induced gene (LePS2) representing an acid phosphatase.LePS2 is a member of a small gene family in tomato. The cDNA is 942 bp long and contains an open reading frame encoding a 269-amino acid polypeptide. The amino acid sequence of LePS2 has a significant similarity with a phosphatase from chicken. Distinct regions of the peptide also share significant identity with the members of HAD and DDDD super families of phosphohydrolases. Many plant homologs of LePS2 are found in the databases. TheLePS2 transcripts are induced rapidly in tomato plant and cell culture in the absence of Pi. However, the induction is repressible in the presence of Pi. Divided root studies indicate that internal Pi levels regulate the expression of LePS2. The enhanced expression of LePS2 is a specific response to Pi starvation, and it is not affected by starvation of other nutrients or abiotic stresses. The bacterially (Escherichia coli)expressed protein exhibits phosphatase activity against the synthetic substrate p-nitrophenyl phosphate. The pH optimum of the enzyme activity suggests that LePS2 is an acid phosphatase.
Article
Plants respond to day/night cycling in a number of physiological ways. At the mRNA level, the expression of some genes changes during the 24-hr period. To identify novel genes regulated in this way, we used microarrays containing 11,521 Arabidopsis expressed sequence tags, representing an estimated 7800 unique genes, to determine gene expression levels at 6-hr intervals throughout the day. Eleven percent of the genes, encompassing genes expressed at both high and low levels, showed a diurnal expression pattern. Approximately 2% cycled with a circadian rhythm. By clustering microarray data from 47 additional nonrelated experiments, we identified groups of genes regulated only by the circadian clock. These groups contained the already characterized clock-associated genes LHY, CCA1, and GI, suggesting that other key circadian clock genes might be found within these clusters.
Article
Hydrolysis of phosphate esters is a critical process in the energy metabolism and metabolic regulation of plant cells. This review summarizes the characteristics and putative roles of plant acid phosphatase (APase). Although immunologically closely related, plant APases display remarkable heterogeneity with regards to their kinetic and molecular properties, and subcellular location. The secreted APases of roots and cell cultures are relatively non-specific enzymes that appear to be important in the hydrolysis and mobilization of Pi from extracellular phosphomonoesters for plant nutrition. Intracellular APases are undoubtedly involved in the routine utilization of Pi reserves or other Pi-containing compounds. A special class of intracellular APase exists that demonstrate a clear-cut (but generally nonabsolute) substrate selectivity. These APases are hypothesized to have distinct metabolic functions and include: phytase, phosphoglycolate phosphatase, 3-phosphoglycerate phosphatase, phosphoenolpyruvate phosphatase, and phosphotyrosyl-protein phosphatase. APase expression is regulated by a variety of developmental and environmental factors. Pi starvation induces de novo synthesis of extra- and intracellular APases in cell cultures as well as in whole plants. Recommendations are made to achieve uniformity in the analyses of the different APase isoforms normally encountered within and between different plant tissues.
Article
As most soil phosphates exist as insoluble inorganic phosphate and organic phosphates, higher plants have developed several strategies for adaptation to low phosphorus (P). These include the secretion of acid phosphatase and organic acids, induction of the inorganic phosphate (Pi) transporter and the substitution of some enzyme activities as alternative pathways to increase P utilization efficiency. It has been proposed that plants also have a ‘pho regulon’ system, as observed in yeast and Escherichia coli; however, the detail of the regulation system for gene expression on P status is still unclear in plants. To investigate the alteration of gene expression of rice roots grown under P-deficient conditions, a transcriptomic analysis was conducted using a cDNA microarray on rice. Based on the changes of gene expression under a –P treatment, the up-regulation of some genes due to P deficiency was confirmed . Some new important metabolic changes are suggested, namely: (1) acceleration of carbon supply for organic acid synthesis through glycolysis; (2) alteration of lipid metabolism; (3) rearrangement of compounds for cell wall; and (4) changes of gene expression related to the response for metallic elements such as Al, Fe and Zn.
Article
Two stimuli that have been associated with nutrient remobilization in plants are phosphate (P(i)) starvation and senescence. Little is known about how the nutrient remobilization machinery is induced at the molecular level, but in the case of P(i) starvation, ribonucleases are considered to play important roles in the remobilization process. Here, the control of two closely related ribonuclease genes of Arabidopsis, RNS1 and RNS3 is investigated. The RNS1 gene is sharply induced during starvation for P(i), an effect specific among the major macronutrients, whereas RNS3 transcript levels remain relatively constant. RNS1 and RNS3 produced in yeast co-migrate with Arabidopsis ribonuclease activities that exhibit the same induction properties as the transcripts in both wild-type plants and the pho1 mutant, which is defective in xylem loading of P(i). In contrast to what occurs during P(i) starvation, both RNS1 and RNS3 are modestly induced during senescence, indicating that the two stimuli could trigger different signal transduction pathways. The characterization of RNS1, in particular, provides an important first step towards elucidating the mechanisms by which plants sense and respond to P(i) limitation, a prominent condition in many soil types.
Article
Photosynthetic membranes of higher plants contain specific nonphosphorous lipids like the sulfolipid sulfoquinovosyl diacylglycerol in addition to the ubiquitous phospholipid phosphatidylglycerol. In bacteria, an environmental factor that drastically affects thylakoid lipid composition appears to be the availability of phosphate. Accordingly, we discovered an increase in the relative amount of sulfolipid and a concomitant decrease in phosphatidylglycerol in Arabidopsis thaliana grown on medium with reduced amounts of phosphate, as well as in the pho1 mutant of A. thaliana deficient in phosphate transport. To investigate the molecular basis of the observed change in lipid composition, we isolated a cDNA of A. thaliana, designated SQD1, that encodes a protein involved in sulfolipid biosynthesis as suggested by three lines of evidence. First, the cDNA shows high sequence similarity to bacterial sqdB genes known to be essential for sulfolipid biosynthesis; second, the SQD1 gene product is imported into chloroplasts where sulfolipid biosynthesis takes place; and third, transgenic plants expressing SQD1 in antisense orientation show a reduction in sulfolipid content. In the pho1 mutant as well as in wild-type plants grown under reduced phosphate availability, increased amounts of SQD1 mRNA and SQD1 protein are detected, suggesting that the increase in sulfolipid content under phosphate limitation is the result of an increased expression of at least one gene required for sulfolipid biosynthesis in A. thaliana. It is suggested that a certain amount of anionic thylakoid lipid is maintained by substituting sulfolipid for phosphatidylglycerol under reduced phosphate availability.
Article
Auxin response factors (ARFs) are transcription factors that bind with specificity to TGTCTC auxin response elements (AuxREs) found in promoters of primary/early auxin response genes. ARFs are encoded by a multi-gene family, consisting of more than 10 genes. Ten ARFs have been analyzed by Northern analysis and were found to be expressed in all major plant organs and suspension culture cells of Arabidopsis. The predicted amino acid sequences indicate that the 10 ARFs contain a novel amino-terminal DNA binding domain and a carboxyl-terminal dimerization domain, with the exception of ARF3 which lacks this dimerization domain. All ARFs tested bind with specificity to the TGTCTC AuxRE, but there are subtle variations in the sequence requirements at positions 5 (T) and 6 (C) of the AuxRE. While the amino-terminal domain of about 350 amino acids is sufficient for binding ARF1 to TGTCTC AuxREs, this domain is not sufficient for the binding of some other ARFs to palindromic AuxREs. Our results suggest that ARFs must form dimers on palindromic TGTCTC AuxREs to bind stably, and this dimerization may be facilitated by conserved motifs found in ARF carboxyl-terminal domains. Dimerization in at least some cases may dictate which ARF(s) are targeted to AuxREs.