Plant Science 187 (2012) 113–126
Contents lists available at SciVerse ScienceDirect
Plant Science
journal homepage: www.elsevier.com/locate/plantsci
Chromium-induced physiological and proteomic alterations in roots of
Miscanthus sinensis
Shamima Akhtar Sharmin 1 , Iftekhar Alam 1 , Kyung-Hee Kim, Yong-Goo Kim, Pil Joo Kim,
Jeong Dong Bahk, Byung-Hyun Lee ∗
Division of Applied Life Science (BK21 program), IALS, PMBBRC, Gyeongsang National University, Jinju 660-701, Republic of Korea
a r t i c l e
i n f o
Article history:
Received 13 December 2011
Received in revised form 31 January 2012
Accepted 2 February 2012
Available online 9 February 2012
Keywords:
Abiotic stress
Chromium
Heavy metal
Proteome
Miscanthus sinensis
a b s t r a c t
Despite the widespread occurrence of chromium toxicity, its molecular mechanism is poorly documented
in plants compared to other heavy metals. To investigate the molecular mechanisms that regulate the
response of Miscanthus sinensis roots to elevated level of chromium, seedlings were grown for 4 weeks
and exposed to potassium dichromate for 3 days. Physiological, biochemical and proteomic changes
in roots were investigated. Lipid peroxidation and H2 O2 content in roots were significantly increased.
Protein profiles analyzed by two-dimensional gel electrophoresis revealed that 36 protein spots were
differentially expressed in chromium-treated root samples. Of these, 13 protein spots were up-regulated,
21 protein spots were down-regulated and 2 spots were newly induced. These differentially displayed
proteins were identified by MALDI-TOF and MALDI-TOF/TOF mass spectrometry. The identified proteins
included known heavy metal-inducible proteins such as carbohydrate and nitrogen metabolism, molecular chaperone proteins and novel proteins such as inositol monophosphatase, nitrate reductase, adenine
phosphoribosyl transferase, formate dehydrogenase and a putative dihydrolipoamide dehydrogenase
that were not known previously as chromium-responsive. Taken together, these results suggest that Cr
toxicity is linked to heavy metal tolerance and senescence pathways, and associated with altered vacuole
sequestration, nitrogen metabolism and lipid peroxidation in Miscanthus roots.
© 2012 Elsevier Ireland Ltd. All rights reserved.
1. Introduction
Heavy metal contamination is a cause of major environmental hazards worldwide, leading to losses in agricultural yields and
harmfully affecting human health when contaminants enter the
food chain. Chromium (Cr) is the seventh most abundant element
on earth. It exists in nature in both trivalent (Cr III) and hexavalent (Cr VI) forms, of which the latter is more toxic [1]. Cr
compounds cause environmental pollution as a result of a large
number of industrial operations, including mining, pigment manufacturing, petroleum refining, leather tanning, wood preserving,
textile manufacturing, pulp processing and fungicide development
[2]. In India, about 2000–3200 tones of elemental Cr leak to the
environment annually with a Cr concentration ranging between
Abbreviations:
2-DE, two-dimensional gel electrophoresis; MALDI-TOF,
matrix-assisted laser desorption ionization time-of-flight; PMF, peptide mass
fingerprinting; ROS, reactive oxygen species; SDS-PAGE, sodium dodecylsulfate
polyacrylamide gel electrophoresis; TBARS, thiobarbituric acid reactive substance;
V-ATPase, vacuolar-type H+ -ATPase; UDP-GlcDH, UDP-glucose dehydrogenase.
∗ Corresponding author. Tel.: +82 55 772 1882; fax: +82 55 772 1889.
E-mail address:
[email protected] (B.-H. Lee).
1
S.A. Sharmin and I. Alam contributed equally to this work.
0168-9452/$ – see front matter © 2012 Elsevier Ireland Ltd. All rights reserved.
doi:10.1016/j.plantsci.2012.02.002
2000 and 5000 mg L−1 [3]. Very high level of Cr(VI) contamination
(14,600 mg kg−1 in ground water and 25,900 mg kg−1 in soil) has
been reported in some sites of Oregon state, USA [4]. Generally,
most Cr (VI) added to soil is promptly reduced to the inert form
Cr (III) by several agents. However, re-oxidation of Cr (III) to Cr
(VI) occur by microorganisms and, therefore, both states should
be regarded hazardous for the environment and for humans [5].
Both forms cause serious damage to plant tissues and organs at
differing concentrations. Cr phytotoxicity can result in the inhibition of seed germination, pigment degradation, disturbances in
the nutrient balance and the generation of reactive oxygen species
(ROS), which induces oxidative stress and alterations in antioxidant enzyme activities [6]. In the cell, free system reactivity of Cr
is generally considered by its interaction with glutathione (GSH),
NADH and H2 O2 -generating hydroxyl radicals (OH− ) [7]. Both Cr III
and VI react with cellular H2 O2 , generating highly reactive hydroxyl
radicals.
Industrial chromium wastes are generally treated with physicochemical processes before they are released into the environment.
Following primary treatments, the methods of removal of residual chromium (polishing) are expensive and the efforts are often
insufficient [8]. Consequently, residual Cr are released to the environment and accumulated in agricultural products through water,
114
S.A. Sharmin et al. / Plant Science 187 (2012) 113–126
air and polluted soils [6]. Soils from numerous sites in the USA are
contaminated with Cr at levels ranging from 1 to 1500 mg Cr kg−1
[9]. Soil pollution generates extra costs for soil management and
pollution control. The uses of plants for soil phytoremediation by
means of degradation (phytodegradation), adsorption (rhizofiltration) and absorption (phytoextraction) are efficient, renewable,
and natural processes that are leading competitors in the search
for solutions to these contamination issues. Unfortunately, most
known hyperaccumulator plants have very low biomass and/or
slow growth rates, are difficult to cultivate on a commercial scale
and have very few commercial uses. Therefore, attention has been
focused on several biomass crops that have fast growth rates
and high biomass and are able to accumulate moderate to large
amounts of heavy metals without sacrificing biomass gain. Most
studies involving Cr overaccumulation have focused on extreme
examples, representing plants native to highly Cr-rich environments [10,11]. Very little attention has been paid to commercially
important biomass-producing crops. Miscanthus sinensis, a perennial rhizomatous C4 grass, is a potentially efficient, sustainable
carbon-neutral producer of lignocellulosic biomass, making it very
suitable and promising for the production of biofuels and fiber [12].
Miscanthus sequesters higher amount of Cr to the aerial part at
extremely toxic levels, whereas the overall ability of this species to
remove Cr from the solution is higher at moderate toxicities [13].
These suggest that M. sinensis is a potential bioaccumulator of Cr
and other heavy meals.
Heavy metal-accumulating plants have expansive, advanced
antioxidant defense systems and other important features that
enable them to acquire tolerance [14,15]. Unfortunately, little is
known about the molecular basis of excess heavy metal tolerance.
Unlike other heavy metals, such as As, Cu, Pb and Cd, the partitioning of Cr by phytochelatin synthesis has not been observed;
therefore, the detoxification mechanism for this metal is poorly
understood [6]. Molecular events underlying Cr toxicity and the
defense-related signal transduction process have been only partially elucidated. A number of genes potentially involved in Cr
tolerance and accumulation were assessed by cDNA-AFLP and
reported [16]. Recently, the combination of genome-wide transcriptome profiling and metabolome analysis has been reported
in Cr-stressed rice plant [17].
Proteomics, the comprehensive and quantitative analysis of
proteins that are expressed in a given organ, tissue or cell line,
provides unique insights into biological systems that cannot be
acquired from genomic or transcriptomic approaches. Proteomics
has been used extensively to investigate the protein expression pattern under abiotic stresses. Expression pattern of maize proteins in
response to high concentrations of Cr (340–1019 M) have been
described for the first time by [18]. However, no proteomic study
has been carried on M. sinensis in response to Cr stress. Therefore,
we carried out a proteomic analysis of M. sinensis roots subjected
to Cr stress to identify proteins or primary targets, hoping to gain
a more thorough understanding of the molecular basis of heavy
metal tolerance in this species.
2. Materials and methods
2.1. Plant growth and treatments
M. sinensis (cv. Kosung) seeds were planted on commercial potting mix in plastic trays and allowed to germinate in a growth
chamber. Three weeks after germination, the seedlings were transferred to hydroponic cultures supplied with half strength Hoagland
nutrient solution (H2395, Sigma, USA). pH of the medium was
adjusted to 5.8. To ensure proper growth, the solutions were
aerated with aquarium aerators. Following a 1-week hydroponic
adaptation, the seedlings were subjected to treatments of 0, 50, 100,
200, 300, 500, 750 and 1000 M potassium dichromate (K2 Cr2 O7 ).
After a 3-day treatment, the roots were excised from untreated
(control) and treated seedlings and used for proteomic and physiological analyses. The entire experiment was conducted under light
conditions (500 mol m−2 s−1 , 16/8 h light/dark period) at 25 ◦ C
and 65% humidity.
2.2. Determination of Cr accumulation in roots
After 3 days of treatment, root samples were washed five
times with deionized water to remove surface Cr salts. The samples were dried in an incubator at 60 ◦ C for 72 h, weighed, and
then ground to a fine powder. Approximately 1 g of fine powder
from each treatment group was digested, using a ternary solution
(HNO3 /H2 SO4 /HClO4 , 10:1:4 v/v), and the total Cr in the digestion
solution was determined with a graphite furnace atomic absorption
spectrophotometer (GFAAS) (PerkinElmer SIMAA 6000, Norwalk,
CT, USA) [19]. Three different biological replicate root samples were
used for the analysis.
2.3. Measurement of lipid peroxidation and hydrogen peroxide
Lipid peroxidation was estimated by measuring the concentrations of 2-thiobarbituric acid-reactive substances (TBARS) as
described previously [20]. Briefly, 300 mg of powdered tissue were
homogenized in 20% trichloroacetic acid (TCA), containing 0.5% 2thiobarbituric acid, and heated at 95 ◦ C for 30 min [21]. The TBARS
concentrations were measured as the malondialdehyde (MDA;
ε = 155 mM−1 cm−1 ) concentrations, which were determined at
A532 and corrected for nonspecific turbidity at A600 . The hydrogen
peroxide (H2 O2 ) concentrations were measured spectrophotometrically as described by [22]. Briefly, H2 O2 was extracted by
homogenizing 300 mg of tissue samples with 3 mL of phosphate
buffer (50 mM, pH 6.8), containing the catalase inhibitor hydroxylamine (1 mM). The homogenate was centrifuged at 6000 × g for
25 min. A mixture comprised of 3 mL of extracted solution and
1 mL of 0.1% titanium sulfate in 20% (v/v) H2 SO4 was centrifuged at
6000 × g for 15 min. The intensity of the yellow color of the supernatant was measured at 410 nm. The H2 O2 level was calculated,
using the extinction coefficient 0.28 mol−1 cm−1 .
2.4. Protein extraction and 2-D electrophoresis
Proteins were extracted from the root sample using a phenol
extraction method according to our previous paper [23]. Briefly,
750 mg of tissue was homogenized with a Mg/NP-40 extraction
buffer [0.5 M Tris–HCl, pH 8.3, 2% (v/v) NP-40, 20 mM MgCl2 ,
1 mM phenyl methyl sulfonyl fluoride, 2% (v/v) -mercaptoethanol
and 1% (w/v) polyvinyl polypyrrolidone] and fractionated with
water-saturated phenol, followed by centrifugation at 12,000 × g
for 15 min. The proteins were recovered from the supernatant
by precipitation with ammonium acetate in methanol. The protein samples were then quantified using the Lowry method [24]
and subjected to two-dimensional gel electrophoresis (2-DE) using
a standard procedure. The protein samples were dissolved in
a reswelling buffer [8 M urea, 1% CHAPS, 0.5% (v/v) IPG buffer
pH 4–7, 20 mM dithiothreitol (DTT), and a trace of bromophenol
blue]. A total of 500 g of dissolved protein sample was applied
to the immobilized pH gradient (IPG) dry strip (pH 4–7, 18 cm)
for 13–14 h, followed by focusing for 47,500 V-h using an IPGphor (Amersham Bioscience, Uppsala, Sweden). After isoelectric
focusing (IEF), the IPG strips were equilibrated for 15 min in an
equilibration buffer [50 mM Tris–HCl, pH 8.8, 6 M urea, 30% (v/v)
glycerol, 2% (w/v) SDS, and a trace of bromophenol blue] containing 10 mg/mL DTT, followed by 15 min in an equilibration buffer
S.A. Sharmin et al. / Plant Science 187 (2012) 113–126
containing 25 mg/mL iodoacetamide. Second dimension SDS-PAGE
was carried out using a 12% polyacrylamide gel, and the gels were
stained with colloidal Coomassie brilliant blue (CBB).
2.5. Gel documentation and analysis
Images of CBB-stained gels, which were acquired using a
high-resolution scanner (GS-800 Calibrated Imaging Densitometer; Bio-Rad, Hercules, CA, USA), were used for analysis. Spots were
detected, quantified and then matched using the Bio-Rad PDQuest
software (Version 7.2; Bio-Rad). To compensate for the variability in gel staining, the volume of each spot (spot abundance) was
normalized as a relative volume. After automated detection and
matching, manual editing was performed. A minimum of three gels
were generated for each sample. Only spots that showed significant and reproducible changes of at least 1.5-fold were considered
to be differentially expressed proteins. The standard error (SE) was
calculated from three spots in replicated gels.
2.6. In-gel digestion, MALDI-TOF MS and database search
Selected protein spots were excised manually from the CBBstained gels, washed with 50% (v/v) acetonitrile (ACN) in a 0.1 M
NH4 HCO3 solution and then vacuum-dried. The gel fragments
were reduced for 45 min at 55 ◦ C in a solution of 10 mM DTT in
0.1 M NH4 HCO3 . After cooling, the DTT solution was immediately
replaced with 55 mM of iodoacetamide in 0.1 M NH4 HCO3 . After
washing with 50% ACN in 0.1 M NH4 HCO3 , the dried gel pieces
were left to swell in a minimum volume of 10 L of digestion
buffer (25 mM NH4 HCO3 and 12.5 ng/L trypsin, Promega, WI,
USA). Following overnight digestion at 37 ◦ C, the peptides were
dried. The samples were analyzed using a Voyager-DE STR MALDITOF mass spectrometer (PerSeptive Biosystems, Framingham, MA,
USA). Parent ion masses were measured in the reflectron/delayed
extraction mode with an accelerating voltage of 20 kV, a grid voltage of 76.000%, a guide wire voltage of 0.01%, and a delay time
of 150 ns. A two-point internal standard for calibration was used
with des-Arg1-Bradykinin (m/z 904.4681) and neurotensin (m/z
1672.9175). The software Data Explorer® (PerSeptive Biosystems,
Inc.,USA) was used to view and process data files. The peptide
mass fingerprintings (PMFs) obtained from each digested protein
were compared with PMFs in the non-redundant National Center for Biotechnology Information database (NCBInr, 2011/01/01)
using the ProFound program (http://prowl.rockefeller.edu/prowlcgi/profound.exe). The search was performed within all green
plants (Viridiplantae) using the following parameters: the maximum number of missed cleavages was set at one, the complete
carbamidomethylation of cysteines and variable oxidation of
methionines was assumed, monoisotopic masses were used and
a mass tolerance of 100 ppm was allowed. Only significant hits, as
defined by the ProFound ‘expectation value’ of <5e−2 (i.e. p < 0.05)
were chosen. The estimated experimental Mr /pI was applied to
increase the confidence of identification (Table 1).
2.7. MS/MS analysis
MS and MS/MS analyses were performed as described earlier
[25]. Mass spectra were acquired with an ABI 4800 Plus TOF–TOF
Mass Spectrometer (Applied Biosystems, Framingham, MA, USA),
which uses a 200 Hz ND:YAG laser operating at 355 nm. The ten
most and least intense ions per MALDI spot, with signal/noise ratios
>25, were selected for subsequent MS/MS analysis in 1 kV mode and
800–1000 consecutive laser shots. During MS/MS analysis, air was
used as the collision gas. Data were subjected to a Mass Standard Kit
for the 4700 Proteomics Analyzer (Calibration mixture 1). MS/MS
spectra were searched against the NCBInr database by ProteinPilot
115
v.3.0 (with MASCOT as the database search engine) with peptide
and fragment ion mass tolerance of 50 ppm. Carbamidomethylation
of cysteines and oxidation of methionines were allowed during the
search of the peptides. One missing trypsin cleavage was allowed.
Peptide mass tolerance and fragment mass tolerance of the selected
95 proteins were set to 50 ppm. High confidence identifications had
statistically significant search scores (greater than 95% confidence,
equivalent to MASCOT expect value p < 0.05), were consistent with
the protein’s experimental pI and Mr , and accounted for the majority of ions present in the mass spectra.
2.8. Statistical analysis
Results of the physiological parameters and spot intensity were
statistically analyzed by using analysis of variance (ANOVA) and
Duncan’s multiple range test (DMRT) to determine significant differences among group means. Significant differences from control
values were determined at p < 0.05 levels. All the results are represented as means ± SE of at least three independent replicates. The
statistical program SAS, version 9.1 (SAS Institute, Cary, NC, USA)
was used for the statistical analyses.
2.9. Multivariate analysis
To visualize patterns in abundance, multivariate analysis was
performed to the data from the differentially expressed proteins.
The 36 protein spots that had a 1.5-fold or greater variation
(p ≤ 0.05) in at least one point were used for principal component
analysis (PCA) and cluster analysis. The datasets from the three
replicates were grouped with components showing differences
between the individual treatments and the differentially expressed
spots. The PCA analysis was performed using XLSTAT software
(Addinsoft SARL). Clustering of the dataset was performed using
MeV software. For the calculation settings, six clusters were defined
and Euclidean distance was selected as the similarity measure.
3. Results
3.1. Cr-induced morphological changes
Our objective was to understand the possible mechanisms that
might be activated by the M. sinensis cell in detoxification during
Cr stress using proteomic approach. M. sinensis is a heavy metal
tolerant species [26,27]. However, a preliminary study was necessary to define the concentrations of Cr that induce cellular response
without leading to immediate cell death. A short exposure to low
concentration of Cr (50–300 M) did not exhibit any sign of growth
reduction (Supplementary Fig. 1A). In low level of Cr, Miscanthus plants continued to grow for several weeks. MDA content in
roots were almost similar to non-treated plants (Supplementary
Fig. 1B), suggesting a short exposure to low level Cr cause negligible oxidative damage. However, growth suppression was observed
starting from 500 M of Cr with the highest inhibition occurring at 1000 M. Under control condition, new roots were being
developed, while their formation was suppressed over 500 M. At
1000 M, new root formation was severely affected. These observations suggest that M. sinensis is relatively tolerant to Cr like maize.
[28] reported that low concentration of Cr have positive effect on
root growth of Miscanthus. Thus, based on the growth pattern and
earlier reports, 500–750 M could be considered moderate to toxic,
while 1000 M or higher concentrations are acutely toxic. Labra
et al. [18] also carried out a proteomic analysis of maize seedling
subjected to 340 and 1019 M Cr based on growth suppression.
A short exposure to moderate to acute toxic Cr may reveal the
proteins involved in altered metabolic homeostasis.
116
S.A. Sharmin et al. / Plant Science 187 (2012) 113–126
Table 1
Chromium stress-responsive differentially expressed proteins in Miscanthus sinensis roots identified by MALDI-TOF MS.
Spot no.
Protein
Organism
Accession no.a
Theoretical
Observed
1
2
5
7
9
10
11
12
13
Predicted protein
Vacuolar H+-ATPase subunit B
Os06g0136600
Hypothetical protein SORBIDRAFT
Vacuolar ATP synthase catalytic subunit A
Unknown
Inositol monophosphatase
Vacuolar ATP synthase catalytic subunit A
SAM-2(S-adenosylmethionine synthetase 2);
copper ion binding
S-adenosyl methionine synthetase
Cell division control protein 2 homolog C
Tetratricopeptide-like helical
PsHSP71.2
Hypothetical protein OsJ 02626
CPK31; ATP binding/calcium ion
binding/calmodulin-dependent protein kinase
Predicted protein
Glutamine synthetase
Os11g0229200
Hypothetical protein
Unknown
Hypothetical protein
Hypothetical protein SORBIDRAFT 02g044060
Hypothetical protein SORBIDRAFT 10g016920
Hypothetical protein
Glyceraldehyde-3-phosphate dehydrogenase C
subunit (GapC)
Hypothetical protein
Unknown protein
Hypothetical protein SORBIDRAFT 01g043060
Hypothetical protein
ATP synthase F0 subunit 1
Unknown protein
Hypothetical protein
UDP-glucose 6-dehydrogenase
Hypothetical Protein SORBIDRAFT 03g013290
Micromonas pusilla
Zostera marina
Oryza sativa
Sorghum bicolor
Zea mays
Zea mays
Ostreococcus tauri
Zea mays
Arabidopsis thaliana
226456463
118721470
115466256
242054033
195658441
223973319
116055491
195658441
15234354
90.50/4.9
54.47/5.2
48.15/5.4
58.42/5.9
68.69/5.3
72.82/5.5
77.03/5.6
68.69/5.3
43.64/5.7
Oryza rufipogon
Antirrhinum majus
Medicago truncatula
Pisum sativum
Oryza sativa
Arabidopsis thaliana
100801628
5921446
92870988
562006
125571194
42570056
Populus trichocarpa
Saccharum officinarum
Oryza sativa
Arabidopsis thaliana
Zea mays
Zea mays
Sorghum bicolor
Sorghum bicolor
Oryza sativa
Arabidopsis thaliana
Vitis vinifera
Arabidopsis thaliana
Sorghum bicolor
Oryza sativa
Oryza sativa
Arabidopsis thaliana
Oryza sativa
Zea mays
Sorghum bicolor
14
18
21
22
26
27
30
32
36
37
39
48
50
55
56
57
62
67
68
69
70
72
73
75
78
a
b
c
d
SC (%)b
PMc
Expectd
54/5.0
56/5.2
54/5.1
55/5.7
70/5.6
70/5.6
70/5.7
70/5.7
45/6.1
14
24
21
35
25
22
21
21
23
8
9
7
13
13
12
10
9
7
2.0e−2
2.8e−5
7.2e−3
6.5e−5
8.9e−4
2.1e−3
1.5e−2
1.0e−3
4.2e−3
42.99/5.7
34.51/6.8
87.67/6.9
71.55/5.2
27.50/5.5
55.08/6.0
45/5.9
47/6.1
80/6.8
70/5.2
35/5.7
34/5.7
25
28
13
28
17
25
8
6
5
11
4
7
2.0e−3
2.0e−2
2.2e−2
9.6e−3
8.7e−3
4.9 e−2
222834292
56681315
115484821
7268210
194701624
226507242
242051414
242095836
47900451
21593240
58.00/5.8
39.57/5.5
38.89/5.9
23.86/5.5
19.50/5.1
38.78/6.3
27.26/5.2
41.64/6.2
40.92/6.5
37.09/6.6
40/6.2
42/5.7
30/5.2
30/5.4
25/5.1
39/6.8
27/5.7
43/6.5
43/6.6
40/6.8
27
18
22
26
44
22
49
21
24
35
8
8
7
5
7
5
8
5
7
7
2.0e−2
1.6e−2
3.7e−2
2.6 e−2
1.1 e−2
1.9 e−2
4.8 e−4
2.7 e−3
1.2 e−2
9.9 e−3
225448323
15222614
242041787
222631421
194033257
79474381
35215055
195623986
242057247
41.80/5.3
44.84/5.4
58.41/6.1
55.45/7.3
55.64/5.8
58.34/8.4
34.34/5.5
53.51/5.7
53.39/6.5
44/5.5
55/5.5
56/6.1
55/6.2
55/6.2
35/6.3
36/5.8
55/6.4
58/6.7
44
19
35
15
36
23
33
22
25
12
6
13
6
18
6
6
7
7
4.4 e−4
1.1e−2
1.9 e−4
9.9 e−3
8.6 e−7
2.6 e−4
3.7 e−2
2.7 e−2
5.3 e−3
Mr /pI
Accession number in NCBI database.
SC, sequence coverage by PMF using MALDI-TOF MS.
PM, number of peptides matched.
ProFound expectation value, a value of <5e−2 indicates p < 0.05.
Fig. 1. Effect of Cr treatment on morphology (A), root growth (B) and Cr content (C) in M. sinensis roots after treated with indicated concentrations of K2 Cr2 O7 for 3 days in
hydroponics in a Hoagland medium. Dry weights were calculated from pooled root samples of 12 plants. The root length and Cr amounts represent the mean values and SE.
Different letters above the bars indicate statistically significant differences (p < 0.05).
S.A. Sharmin et al. / Plant Science 187 (2012) 113–126
When M. sinensis seedlings were subjected to 500–1000 M
of Cr for 3 days, a considerable reduction in root growth was
observed in parallel with the doses of Cr in the medium. The
decreases in root growth following Cr treatment were characterized by reductions in root lengths and dry weight with increasing
Cr concentrations (Fig. 1A and B). Chromium accumulation in M.
sinensis roots increased with increasing concentrations of potassium dichromate added to the solution (Fig. 1C). After 3 days of
exposure to 1000 M, the root tissue accumulated 1308 mg kg−1
on a dry weight basis. No Cr accumulation could be detected in the
control plants.
3.2. Accumulation of H2 O2 and lipid peroxidation
To investigate whether growth inhibition was associated with
oxidative stress, the amount of H2 O2 and malondialdehyde (MDA)
were examined. As shown in Fig. 2, H2 O2 accumulation was
much higher in Cr-treated root samples compared to the control.
Although physiological concentrations of ROS have important functions in stress signaling, excess amount can cause oxidative stress,
leading to cell death, if they are not detoxified. Membrane lipids are
the main cellular targets that are susceptible to damage, and lipid
peroxidation is believed to be a free radical-mediated process [29].
Thus, we estimated lipid peroxidation in roots by the thiobarbituric
acid (TBA) method, in which the quantified TBA-reactive substance
was malondialdehyde, an end product of lipid peroxidation. MDA
concentrations were increased markedly in 750 and 1000 M of
Cr, indicating increased lipid peroxidation. The occurrence of lipid
peroxidation induced by Cr was further validated by a histochemical assay, using Schiff’s reagent. As shown in Fig. 3A, an intense
coloration was detected by Schiff’s stain with increasing concentrations of Cr. By contrast, control roots had very small stain. Together
with the quantitative estimation, the histochemical detection provides additional advantages for localizing TBA-reactive products
in situ in roots with high sensitivity [30]. These results suggest
that like other heavy metals, Cr toxicity also generated ROS, which
resulted in oxidative stress in the Miscanthus roots. Increased lipid
peroxidation induced by heavy metals such as aluminum [31], lead
[32] and arsenic toxicity [23] have been reported in various plants.
Evans blue staining indicated that cell death occurred earlier and
more robustly with increasing Cr concentrations (Fig. 3B). These
results are consistent with those from the lipid peroxidation assay,
indicating that Miscanthus suffers from Cr-induced oxidative stress
at high concentrations. Cr (VI)-mediated • OH radical generation in
cells has been reported [33]. Taken together, these results indicated
that plants exposed to Cr treatment generate ROS, which resulted
oxidative stress and cell death in roots.
3.3. Proteomic alteration of Miscanthus roots under Cr stress
To investigate differentially expressed proteins from the Miscanthus root in response to excess levels of Cr, proteins were
extracted from control and Cr-treated roots and separated by 2DE. A high resolution of 2-DE gel pattern with a pI range of 4–7
was detected by CBB staining (Fig. 4). More than 1150 protein
spots were reproducibly detected in each CBB-stained gel by 2-DE
analysis. Among the well-resolved spots, a densitometric analysis of the replicated gels revealed 36 proteins showed at least
1.5-fold increase or decrease in expression in at least one treatment (Figs. 4 and 5 ). Several regions of the gels are enlarged in
Supplementary Fig. 2. Average spot volumes were compared for
the individual spots across the three treatments. The relative abundance of protein spots on the gel is shown in Fig. 5. Two spots
(spots 11 and 12) were hardly detectable in the control sample and
117
were induced after treatment, while three spots nearly disappeared
(spots 21, 67 and 72) due to Cr treatments.
3.4. Identification of M. sinensis root proteins induced by Cr stress
To identify differentially expressed proteins, spots were excised
from the preparative gels, in-gel digested by trypsin and analyzed using MALDI-TOF or MALDI-TOF/TOF MS. The identity of
34 differentially expressed protein spots was obtained by PMF
of MALDI-TOF MS (Table 1). Two additional proteins not recognized by MALDI-TOF MS were identified by MALDI-TOF/TOF MS
and the sequences were determined (Table 2). Relatively small differences were observed between the theoretical and the predicted
molecular masses. The molecular mass is robust toward amino acid
changes. The pI values, however, vary more substantially, probably
due to occurrence of isoforms and amino acid changes between
species. Some of the identified proteins were annotated either as
unknown and hypothetical proteins or as proteins without a specific function in the database. To gain functional information about
these proteins, we searched them against their known homologs
with BLASTP algorithm (www.ncbi.nlm.nih.gov/BLAST/) using their
amino acid sequences as queries. Thirteen corresponding homologues with the highest homology are shown in Table 3. Most spots
except spot 26, 30 and 48 shared more than 95% positives with
homologues at the amino acid level, indicating that they might
have similar function. Differential expression levels of the protein
spots revealed that 13 proteins were up-regulated, 21 were downregulated and two were newly induced (Fig. 5, Supplementary Fig
2). Despite the progress being made in plant proteomics, the power
of proteomics in non-model species has not been assessed thoroughly. As few nucleotide sequences are available from Miscanthus,
cross-species protein identification is used. In the present experiment, more than half of the proteins were matched with rice,
maize and sorghum sequences. Compared to nucleotides, proteins
are generally better-conserved, making the identification of nonmodel gene products quite efficient when they are compared to
well-known orthologous proteins. Therefore, cross-species identification is the only option for studying gene expression when
analyzing poorly characterized genomes. Our results indicate that
proteomic techniques can be successfully applied to plant systems
that are not well-represented in nucleic acid and protein databases.
Among the identified proteins, two enzymes involved in glycolysis pathway, enolase (spot 5) and glyceraldehyde-3-phosphate
dehydrogenase (GAPDH, spot 57), were significantly downregulated, except for GAPDH, which was up-regulated under
1000 M Cr. This could be due to post-translational modification.
On the other hand, proteins involved in mitochondrial respiration
such as ATP synthase (spot 70) and dihydrolipoamide dehydrogenase (spot 78) were increased. A putative mitochondrial processing
peptidase (MPP; spot 68) is up-regulated; the protein plays an
essential role in mitochondrial protein import. Novel accumulation
of inositol monophosphatase (IMPase; spot 11) was observed in the
treated plants. Several spots representing components of vacuolar
transporters (spots 2, 9 and 12) were up-regulated by 2- to 25fold. Among the differentially accumulated nitrogen metabolism
related proteins, both glutamine synthetase (GS; spot 32) and
nitrate reductase (spot 30) were down-regulated. Defense-related
proteins, such as chitinase (spots 34 and 35) and the NB-LRR
protein (spot 36) were highly increased. Adenine phosphoribosyltransferase (APRT; spot 39), formate dehydrogenase (spot 55)
and two spots representing S-adenosyl-l-methionine synthetase
(SAMS; spots 13 and 14) were down-regulated; these are involved
in biosynthesis of mugineic acid (MA), an iron chelating components exclusively in grasses. Multiple spots could be isoforms or
same proteins with different post-translational modifications. The
cell wall polysaccharide biosynthesis-related enzyme UDP-glucose
118
S.A. Sharmin et al. / Plant Science 187 (2012) 113–126
Fig. 2. Physiological responses of M. sinensis roots subjected to treatment. H2 O2 (A) and MDA (B) concentration in control and Cr-treated roots. The data represent the mean
values and SE of three independent experiments. Different letters above the bars indicate statistically significant differences (p < 0.05).
Fig. 3. Histochemical localization of lipid peroxidation and loss of plasma membrane integrity. (A) Differentially stained M. sinensis roots by Schiff’s reagent under different
concentrations of Cr. A more intense pink color indicates more TBA-reactive products. (B) Loss of plasma membrane integrity detected by Evan’s blue staining. Higher
concentration of Cr accumulates more frequent and intense pigmentation as a result of greater damage compared to control. (For interpretation of the references to color in
this figure legend, the reader is referred to the web version of the article.)
Table 2
Chromium stress-responsive differentially expressed proteins in Miscanthus sinensis roots identified by MS/MS analysis.
Spot no.
Protein (organism)
Accession no.a
SC (%)b
Scorec
Peptide hitd
Sequence identified
34
Chitinase II (Hordeum vulgare
subsp. vulgare)
9501334
25
210
3
R.ELAAFFGQTSHETTGGTR.G
R.GAADQFQWGYCFK.E
K.ATSPPYYGR.G
35
Chitinase II (Hordeum vulgare
subsp. vulgare)
563487
25
118
3
R.ELAAFFGQTSHETTGGTR.G
R.GAADQFQWGYCFK.E
K.ATSPPYYGR.G
a
b
c
d
NCBI accession number.
SC, sequence coverage.
Score is the protein score based on combined MS and MS/MS spectra.
Peptide hit is the unique number of MS/MS spectra matched to the trypsin peptide.
S.A. Sharmin et al. / Plant Science 187 (2012) 113–126
119
Fig. 4. A 2-DE analysis of M. sinensis root proteins under 500 (B), 750 (C) and 1000 M (D) of Cr compared to control (A). The arrows indicate differentially expressed proteins
in response to the Cr stress. A total of 500 g of protein was separated by 2-DE as described in Section 2 and visualized with colloidal CBB staining.
dehydrogenase (UDP-GlcDH, spot 75) was up-regulated. In addition, several proteins were identified as unknown/hypothetical
proteins. Overall, the proteins can be broadly classified into
several groups according to their putative physiological functions: (1) energy- and metabolism-related proteins (2) vacuolar
ATPases (3) defense related proteins such as heat shock
proteins (HSPs) (4) nitrogen metabolism proteins, (5) cell division
and (6) stress signaling proteins associated with metal detoxification. Ion transporters, HSPs and energy metabolism-related
proteins were the largest functional categories, suggesting that
energy metabolism pathways are disrupted, and ion transporters
and HSPs may play important roles in protecting the cells from
Table 3
The homologs of unknown proteins. BLASTP (http://www.ncbi.nlm.nih.gov/BLAST/) was used to search for homologs of the unknown proteins.
Spot no.
Accessiona
Homolog protein
Organism
Accessionb
Identities
Positives
Expect
5
7
10
26
30
36
39
48
50
55
62
67
68
69
72
78
115466256
242054033
223973319
125571194
222834292
115484821
194701624
226507242
242051414
242095836
225448323
15222614
242041787
222631421
79474381
242057247
Enolase1
Mitochondrial F1-ATPase beta subunit
Heat shock 70 kDa protein
Protein phosphatase 2C
Assimilatory nitrate reductase (NADH) small subunit
NBS-LRR type resistance protein - barley (fragment)
Adenine phosphoribosyl transferase 1
Putative r40c1 protein
Cytosolic Ascorbate Peroxidase
Formate dehydrogenase 1
Actin
Heat shock protein 70
Mitochondrial-processing peptidase beta subunit
Putative cytochrome P450
Salt-inducible protein homolog
Putative dihydrolipoamide dehydrogenase precursor
Zea mays
Oryza sativa
Zea mays
Oryza sativa
Cupriavidus metallidurans
Oryza sativa
Zea mays
Oryza sativa
Zea mays
Zea mays
Persea americana
Planctomycete str. 140
Zea mays
Oryza sativa
Arabidopsis thaliana
Oryza sativa
CAA3944
218147
226500540
4339763
94313742
62732752
226498252
AAN64997.1
195643366
195640660
281485191
2644998
195628546
47777427
2244775
13365781
96
93
98
79
88
100
99
89
98
98
99
100
97
95
97
93
99
96
99
91
90
100
99
94
99
99
99
100
99
95
97
97
0.0
0.0
0.0
3e−115
1e−154
0.0
4e−96
5e−173
7e−142
0.0
0.0
0.0
0.0
0.0
0.0
0.0
a
b
The accession number of the unknown proteins in Table 1.
The accession number of the homologs identified by BLAST.
120
S.A. Sharmin et al. / Plant Science 187 (2012) 113–126
30
Spot 1
a
14
a
Spot 2
a
25
45
Spot 5
b
b
10
c
20
c
5
0
500
45
750
a Spot 10
40
b
35
0
1000
b
C
500
750
35
c
25
500
750
a
20
10
10
10
5
5
0
5
b
0
C
500
750
1000
90
Spot 18
a
80
C
500
750
b
0
1000
C
500
750
Spot 21
a
140
8
100
6
80
40
180
750
a
140
b
b
c
80
80
60
60
40
40
20
0
20
0
500
750
Spot 26
bc
10
0
c
C
500
C
1000
120
250
a
Spot 30
100
b
b
500
750
b
40
150
b
b
500
750
1000
0
0
500
750
1000
b
10
5
0
750
60
1000
Spot 56
a
50
b
40
500
750
a
100
150
60
100
40
b
c
c
0
10
0
0
0
500
500
750
1000
Spot 48
a
d
0
1000
Spot 35 a
a
0
250
500
a
200
750
1000
Spot 36
b
b
b
750
1000
0
0
500
120
750
1000
Spot 50
a
120
120
100
c
60
c
20
0
0
0
0
500
750
1000
500
1000
Spot 55
a
45
750
40
35
60
25
b
b
c
b
20
b
b
750
1000
15
10
5
0
500
750
1000
Spot 62
a
a
a
b
40
40
0
50
30
b
80
b
0
80
20
Spot 57 a
140
c
b
100
140
160
20
1000
0
b
10
750
c
20
50
20
c
40
10
60
20
b
30
c
20
80
500
C
40
20
100
30
0
0
30
b
b
500
b
30
15
C
b
b
50
40
25
a
80
10
Spot 34
60
30
b
1000
80
Spot 39
50
20
750
200
70
a
Spot 37
a
35
500
a
50
60
40
1000
Spot 27
100
50
20
750
90
150
b
c
0
b
100
c
500
b
50
120
250
b
40
C
Spot 32
80
b
40
10
300
a
200
60
0
1000
b
70
c
20
0
750
a
C
1000
a
20
40
2
1000
60
4
b
750
Spot 14
120
100
30
20
500
160
100
C
C
180
140
b
50
60
30
c
0
1000
a Spot 13
160
60
b
50
500
Spot 22
120
60
C
70
a
10
70
0
1000
160
12
30
20
120
15
15
10
1000
a Spot 12
a
25
15
b
50
10
C
30
a
b
b
5
35
20
20
b
40
a Spot 11
a
30
60
b
c
5
0
1000
25
30
b
10
2
0
Spot 9
a
40
15
4
a
80
70
15
25
c
6
Spot 7
a
30
8
15
25
20
35
10
20
0
a
40
12
0
0
500
750
1000
750
90
1000
Spot 67
a
80
0
60
60
50
50
40
40
30
30
20
20
0
b
b
b
500
750
1000
500
Spot 68
a
80
70
0
0
90
70
10
0
500
b
c
d
10
0
0
500
750
1000
Fig. 5. The expression levels of the identified proteins compared to those of control. Bars indicate the expression level of control, 500, 750 and 1000 M of Cr consecutively.
Spot intensities were measured using a densitometer and compared to those of the control. The average values of the relative increase levels of three replicate samples are
shown in the histograms. The data represent the mean values and SE of three independent experiments. Different letters above the bars indicate statistically significant
differences (p < 0.05).
S.A. Sharmin et al. / Plant Science 187 (2012) 113–126
90
50
Spot 69
a
45
a
b
40
70
35
60
30
50
160
a Spot 70
a
80
a
a
Spot 72
140
120
100
80
25
40
20
c
b
60
30
15
10
5
0
121
20
40
10
20
0
0
0
500
40
750
1000
0
120
Spot 73 a
a
35
500
a
100
750
1000
Spot 75
a
a
0
b
b
b
500
750
1000
120
a
100
a Spot 78
a
30
80
25
20
15
10
60
60
40
40
20
20
b
c
5
0
80
b
b
0
500
750
1000
0
0
500
750
1000
0
0
500
750
1000
Fig. 5. (Continued).
damage following Cr toxicity. The broad category view is shown in
Fig. 6.
3.5. Multivariate analysis
Principal component analysis was conducted to statistically
classify protein spots with differential expression patterns and
exhibit the difference in proteomes across the treatments. The twodimensional PCA plots show that samples are positioned differently
(Fig. 7). The variation in expression pattern appears to be correlated to Cr concentrations. Energy metabolism and ion transporters
are closely grouped. We also applied hierarchical clustering to the
proteome dataset. Hierarchical clustering method uses pair wise
average-linkage algorithm and constructs a dendrogram by which
all expression patterns assemble in a single tree whose branch
length reflects the degree of similarity (Fig. 8A). Pearson correlation
coefficient was applied to define the similarity and the averagelinkage to assemble the items. K-mean cluster analysis was used
to place the protein spots with differing abundance for Cr treatments into six clusters. K-means clustering was also applied to
Miscellaneous
and unknown
Energy metabolism
Ion transportation
Stress signalling and
metal detoxification
Defense and
detoxification
Cell division
Nitrogen metabolism
Fig. 6. Pie chart illustrating the assignment of the identified proteins to functional
categories.
categorize the differentially expressed proteins and showed more
clearly the abundance relationship with Cr treatments. Thirty-six
Cr-responsive proteins were categorized in six expression groups
(Fig. 8B).
4. Discussion
Cr has a complex chemistry and hence the detailed mechanism
of toxicity of Cr is yet to be clearly explained in higher plants. Cr
(VI) is a strong oxidant with a high redox potential. Higher H2 O2
production and lipid peroxidation observed in the present study
indicates that extensive oxidative damage could have occurred to
the root cells under Cr stress. To investigate molecular mechanism
behind the stress response, protein- and RNA-based measurements are complementary, because each technique focuses on
a subset of genes/proteins. Each technique has its advantages
and disadvantages. However, a 2-DE approach will result in a
better characterization when a species is poorly represented in
sequence databases [34]. We used multivariate approach to express
of expression patterns of the protein spots. The PCA reduces the
dimensionality of the multidimensional analysis to display the two
principal components that distinguish between two largest sources
of variation within the dataset. When the abundance patterns for
the protein species of different Cr treatments were analyzed by
PCA, the spots are positioned differently (Fig. 7). However, several
proteins with similar functions tend to group together. Such application of PCA had been successfully used before by other authors
[35]. The cluster analysis summarizes major protein expression
patterns, possible resistance, adaptation and sustained tolerance,
following exposure to Cr (Fig. 8A and B). For instance, cluster-2
proteins were upregulated following Cr treatment, and maintained
their level at higher concentrations. These proteins are involved
in vacuolar transportation, ATP production and protein stabilization during stress condition. Regardless of Cr concentrations, its
increased level might indicates their primary role in Cr tolerance.
On the other hand, some other proteins increased (cluster 4) or
decreased (cluster 3) at moderate Cr concentration but maintained
similar level at higher concentrations indicate their role in Cr adaptation. Our results suggest that the application of hierarchical and
122
S.A. Sharmin et al. / Plant Science 187 (2012) 113–126
Fig. 7. Representation of the samples by PCA. 2-D plot of main principal components (PC1 and PC2) of: (A) all spots and (B) differentially expressed spots. Each data point in
PCA plots (B) describes the expression values for the subset of proteins whose ratios varied 1.5-fold or more.
nonhierarchical clustering methods is useful in presenting proteomic data as shown by others also [36]. In the following sections,
we discussed the possible role of the Cr-induced proteins involved
in a wide range of plant processes.
4.1. Energy metabolism
Our proteomic data showed that the two key enzymes of the
glycolysis pathway were strongly affected following Cr treatment.
Fig. 8. Clustering analysis of the differentially expressed proteins under Cr treatments. (A) Dendrogram of the spots clustering is showed in the left. Relative expression
values of individual proteins displayed as heat map. All quantitative information is transmitted using a color scale in which the color ranges from green for the highest
down-regulation (−1.5) to red for the highest up-regulation (1.5). Black boxes indicate no changes in expression pattern compared to control condition (0 M Cr). Each
row of colored boxes is representative of a single spot and each treatment is represented using a single column (indicated). (B) K-means clustering showing the expression
patterns for individual protein spots in the six main Cr-responsive clusters. (For interpretation of the references to color in this figure legend, the reader is referred to the
web version of the article.)
S.A. Sharmin et al. / Plant Science 187 (2012) 113–126
Enolase (spot 5) and GAPDH (spot 57) were significantly downregulated by Cr. Interestingly, GAPDH was up-regulated at 1000 M
Cr. Photosynthesis and respiration are negatively affected by Cr
due to damage of the photosynthetic apparatus, inhibited redox
reactions and oxidative stress damage [37–39]. Several glycolytic
enzymes and heat-shock proteins are specifically prone to Crdependent oxidative damage in Cr-treated yeast cells [40]. Thus,
along with the evidence of oxidative stress, it could be speculated that Cr treatment may inhibit carbon flux in glycolysis in
M. sinensis plants. Consequently, a lower reducing power (NADH)
and reductions in ATP, carbon skeletons and pyruvate could be
expected. Growth inhibition may lead to an accumulation of carbohydrates or direct inclusion into the mitochondrial respiration
process (simultaneously) to compensate for the higher energy
demands. In addition, Cr-induced oxidative stress requires a high
reducing power to cope with the stress. These could be provided
by the up-regulation of mitochondrial respiration. The higher abundance of a mitochondrial ATP synthase (spot 70), which catalyzes
the formation of ATP from ADP in the membranes of mitochondria,
indicated that mitochondrial respiration may be increased under Cr
stress. In addition, dihydrolipoamide dehydrogenase (E-3 component of pyruvate dehydrogenase; spot 78) is up-regulated, which is
essential for ATP production. The up-regulation of these enzymes
has been reported under arsenic [41] or aluminum stress [42], but
not in Cr. Because of its prime role in energy transduction, increased
abundance under stress presumably reflects altered patterns of carbon flux in response to reduced photosynthesis and increased need
for energy. Our proteomic data are in agreement with metabolite
analysis showing inhibition of glycolysis enzymes by heavy metal
in poplar [43].
Less usage of photoassimilates due to growth suppression and
a decreased breakdown of carbohydrates may trigger an increased
accumulation of stored carbohydrates, as observed in Cr-treated
bush bean [38] and Cd-treated poplar [44]. These molecules may
play roles in osmotic adjustments or protection of cell constituents.
Cr has been shown to induce osmotic stress in bush bean [45]. We
identified inositol monophosphatase (IMPase; spot 11) that catalyzes de novo inositol synthesis from glucose-6-phosphate and is
required for the breakdown of inositol trisphosphate. Free inositol
can act as an osmolyte. Accumulation of free inositol was reported
in Cd-treated poplar plants [44]. Thus, in our proteomics experiment, novel accumulation of an IMPase could further strengthen
the previous hypothesis proposed from sugar analysis [38,44].
4.2. Ion transporters for excess metal management
A number of up-regulated spots were identified as components of vacuolar (spots 2, 9 and 12) and mitochondrial (spot 7)
transporters. Vacuolar-type H+ -ATPase (V-ATPase) energizes plant
endomembranes. The differential resistances of a single species,
such as barley, to various heavy metals involve unique capabilities for vacuolar compartmentation [46]. Metal tolerances versus
metal sensitivities of closely related species or genotypes also
appear to depend on additional membranes, including the tonoplast, and transporters [47,48]. Thus, in absence of phytochelatin
synthesis [6,49], a very important role could be reserved for the
transmembrane transport of toxic Cr ions by V-ATPases. Therefore,
it is reasonable to assume that V-ATPase is affected by Cr treatment.
Little information is available on the influence of heavy metals on
either the structure or the activity of V-ATPase. The antiporter activity depends on the presence of a proton gradient across the vacuolar
membrane and thus, indirectly, on the V-ATPase. Metal–proton
antiport activity has been reported for several metals, such as Cd,
Zn, and Mn in oat roots [50,51] and Zn in Silene vulgaris [48]. By
contrast, other authors could not detect such activity for Cd, either
in oat or Silene [52]. Until now, Cr-specific transporters were not
123
known. Therefore, the large differential expression of several vacuolar and mitochondrial ATPases found in our proteomic study may
provide important clues for further investigations.
4.3. Proteins associated with defense and detoxification
mechanisms
Biotic and abiotic stresses often share multiple nodes of the
same response signaling pathways, and their outputs may have
significant functional overlaps [53]. Cr treatment resulted in highly
up-regulated chitinase (spots 34 and 35) and NB-LRR protein (spot
36). Chitinases are components of plant defense against pathogens
and heavy metals as well at toxic level [54]. The chitinases activity was reported to increase in barley and rape by Cr(VI) but not
by Cr(III) [55]. Increased chitinase activities in plants subjected to
abiotic stresses may result from the induction of cross-tolerance
via cross-talking signaling pathways. Chitinase profiling in pea,
bean, soybean, barley and maize under As, Pb or Cd stress [56] suggests isoform/specific expression. Chitinase appears to counteract
oxidative stress as shown in transgenic tobacco plants expressing
a fungal chitinase [57]. The plants were not only more resistant
to fungal infection but also to salt and metal ion stress. Although
elucidating the relationships between chitinase isoforms and metal
tolerances would be speculative at this moment, the study of metalspecific chitinases may provide very promising insight into the
mechanisms of Cr detoxification processes. NB-LRR proteins are key
molecules in signaling cascades that often culminate in the activation of programmed cell death (PCD) [58]. Here, large increases in
one of these proteins (spot 36) may be associated with PCD signaling under toxic concentrations of Cr.
We also identified three spots as heat shock proteins (HSP70
family; spots 10, 22 and 67) and another chaperone protein (spot
21). Members of the HSP70 family are up-regulated as a result
of thermal and oxidative stress, including exposure to Cd, As and
other heavy metals [59]. However, HSP has not been reported to be
up-regulated following Cr exposure in the published proteomics
reports. HSPs have a broad range of functions, including protein
folding, assembly, translocation and degradation [59]. In Arabidopsis, Cd treatment induces the up-regulation of genes involved in
protein folding [60], which demonstrates that Cd toxicity is in part
due to the induction of protein denaturation, probably by oxidative
modifications [61]. Thus, HSPs and chaperone like proteins were
up-regulated probably to protect cells against damages induced by
Cr. Cytochrome P450s are one of the bioindicator of altered cellular
metabolism caused by environmental pollution. Cytochrome P450
(spot 69), was previously reported as responsive to Cd and other
stresses [62] and are probably functioning in detoxification of cytotoxic product. In addition, one of the key ROS scavenging enzymes,
ascorbate peroxidase (APX) is down-regulated (spot 50). Antioxidant enzymes are generally increased in activity during metal
toxicity. Indeed, decreased APX activity has been observed at very
high heavy metal toxicities [63].
4.4. Nitrogen metabolism
Glutamine synthetase (GS) is the key enzyme in ammonia
assimilation and catalyzes the ATP-dependent condensation of
ammonium ions with glutamate to produce glutamine. The GS is
prone to degradation under oxidative stress conditions [64]. Oxidation of GS was reported in GS-enriched soybean root extract by
metal-catalyzed oxidation system (producing hydroxyl radicals).
The oxidized GS is inactive and more susceptible to degradation
than the non-oxidized form [65]. Cr stress causes free radical generation via the Fenton reaction. Thus, the co-occurrence of increased
oxidative stress and a down-regulated GS (spot 32) in our experiment is reasonable. Aside from transcriptional regulation, GS in
124
S.A. Sharmin et al. / Plant Science 187 (2012) 113–126
plants may also be regulated at the level of protein turnover.
Decreased GS activity has been shown in leaves subjected to water
stress and exposed to excess Cu [66]. A decline in GS activity may
result, at least in part, in an accumulation of ammonium ions. Under
physiological conditions, nitrate reductase [NAD(P)H:nitrate oxidoreductase EC 1.6.6.2] is reversibly converted into an inactive
enzyme upon the addition of ammonia. In the presence of methionine sulfoximine, nitrate reductase is no longer inactivated by
ammonia when GS activity is lost. The addition of ammonia to cell
suspensions drastically prevents the reduction of nitrate to nitrite
in the blue-green alga Agmenellum [67] and in the red alga Cyanidium [68]. Here, we identified a nitrate reductase protein (spot 30)
that was down-regulated by Cr stress for the first time. Thus, a
down-regulated nitrate reductase protein is consistent with previous physiological data.
4.5. Cell division related proteins
Cr-induced disruption in microtubule organization and subsequent micronuclei formation has been reported [69]. However,
proteins involved with these processes are not well described.
Microtubules play key roles in both nuclear division and cytokinesis in plants and are one of the main subcellular targets of Cr (VI)
toxicity. We identified increased levels of an actin protein (spot 62),
which is major cytoskeletal components that play roles in gene
transcription and signal transduction events in plants [70]. The
functional complexity of actins makes them likely targets of oxidative stress. Al-induced damaging effect of F-actin in root stele cells
and subsequent inhibition of root elongation has been shown in
maize [71]. In addition, cell division control protein 2 homologue C
(CDC2; spot 18), a protein kinase, is conserved throughout eukaryotes and acts as a key regulator of the cell cycle, acting through
cyclin-dependent phosphorylation. Mutations in cdc2/cdc28 result
in arrest at the G1/S or G2/M phase of the cell cycle in fission
and budding yeast, respectively [72]. Cr treatment perturbed the
alignment of microtubules in a concentration-dependent manner
in onion roots [69]. Thus, proteomic identification of cell divisionassociated proteins may provide the biochemical basis of the
cytotoxicity and/or stress signaling.
4.6. Stress signaling and metal detoxification
SAMS synthesizes SAM from l-methionine, which is the major
methyl group donor in the transmethylation of proteins and many
other substances. The existence and differential expression of
different SAM homologues has been generally connected to the
metabolic importance of SAM. SAM is the key precursor in the
ethylene biosynthesis pathway. Downregulation of two isoforms
of SAMS imply a decrease in SAM and ACC levels. It has been
demonstrated that Cr inhibits ethylene biosynthesis by reducing
availability of ACC for ethylene synthesis in plants and delays senescence [73]. By contrast, other heavy metals such as Cd were shown
to induce senescence by accelerated ethylene production in root
tissues [74].
The Ca2+ and calmodulin messenger systems have been recognized to be involved in plant-environment interactions [75].
Calmodulin-dependent protein kinases (CPK31; spot 27) are primarily regulated by the Ca2+ /calmodulin complex. Some members
of this protein family could act as metal transporters. A calmodulinbinding tobacco plasma membrane protein (NtCBP4) was identified
to be similar in structure to cyclic-nucleotide-gated non-selective
cation channels. A 2–3-fold overexpression of NtCBP4 enhanced
Pb2+ uptake and improved hypersensitivity to Pb2+ , attenuated Ni2+
uptake and improved Ni2+ tolerance in transgenic plants [75]. Thus,
the precise roles of specific CPK proteins under Cr toxicity need to
be further investigated.
Adenine phosphoribosyltransferase (APRT) is a ubiquitous
enzyme that functions to specifically salvage adenine by converting
it to AMP. The balance of ATP in the cell is maintained by de novo
adenine biosynthesis and purine salvage. This protein was downregulated in our experiment (spot 39). Earlier studies demonstrated
that the single step APRT pathway is the predominant pathway
for the salvaging of adenine in plants under stress conditions [76].
However, APRT has never been reported before to be involved in Cr
stress. The increased level of transcript (HvAPT1) and enzyme activity was reported in barley roots following iron deficiency [77]. The
authors explained possible role of APRT, formate dehydrogenase
(spot 55) and SAMS (spots 13 and 14) in biosynthesis of mugineic
acid that chelate iron exclusively in grasses. The biosynthesis of MA
begins with the activation of methionine molecules by ATP to form
SAM [78]. APRT may function to salvage large amount of adenine
during biosynthesis of MA [77]. AMP synthesized by APRT would
finally be converted to ATP, which again takes part in the methionine cycle through S-adenylation of methionine. MA chelates both
iron and other heavy metals, such as copper, zinc and cobalt [79].
The overall decreases in the expressions of these proteins may be
associated with reduced ability of Cr chelation. The inability of
plants to sequester Cr by phytochelation has been reported earlier
[6,49] and need to be further investigated.
4.7. Miscellaneous and unknown proteins
The enzyme UDP-GlcDH converts UDP-glucose into UDPglucuronic acid (UDP-GlcA). Roughly 50% of the cell wall biomass
is metabolically derived from UDP-GlcA in Arabidopsis [80]. An
up-regulation of UDPGlcDH (spot 75) may be involved in modification of cell wall to prevent metal uptake or block plasmodesmata,
inhibiting the symplastic transfer of solutes, increase root apoplastic barriers under Cr toxicity. Increased cell wall elasticity was
observed during Cr (VI) treatment [45]. An R40c1 protein (spot 48)
was down-regulated, which plays a role in the adaptative response
of roots to an hyper-osmotic environment [81]. Transcript level of
Osr40cl was increased in response to ABA, conversely negatively
regulated by JA and SA. This protein was also down-regulated in rice
roots following exposure to As [23]. Since, metal toxicity induces an
increased amount of jasmonic acid (JA) and salicylic acid (SA) production in plants [74]. Thus, decreased expression of the Osr40c1
protein in response to Cr may be involved in metal-induced JA
production in Miscanthus roots. A putative protein phosphatase 2C
(spot 26) is involved in stress signal transduction. A salt-inducible
protein homolog (spot 72) was down-regulated. However, the precise role of these proteins is not clear under Cr stress. In addition
to the proteins described here, several proteins (spots 1, 37, 56
and 73) were identified as protein of unknown functions. We were
unable to correlate their activities with Cr stress. Further studies
are needed to address their possible roles in relation to this heavy
metal stress.
5. Conclusions
In this study, for the first time we investigated chromium stressinduced physiological and biochemical responses, and proteomic
changes in M. sinensis roots. A total of 36 proteins were identified
that were differentially expressed in chromium-treated root samples. The majority of these proteins were ion transporters, energy
and nitrogen metabolism-related proteins, oxidative stress-related
regulatory proteins that might work together to establish a new
homeostasis in response to chromium stress. The identification of
some chromium-responsive proteins might provide new insights
to the heavy metal homeostasis as well as helpful to improve this
commercially important biomass crop.
S.A. Sharmin et al. / Plant Science 187 (2012) 113–126
Acknowledgments
This work was supported by the National Research Foundation
of Korea (NRF) Grant (NRF-2011-F00013). SA Sharmin, KH Kim and
YG Kim are supported by a scholarship; I Alam is supported by
a postdoctoral grant from BK21 program at Gyeongsang National
University, Republic of Korea.
Appendix A. Supplementary data
Supplementary data associated with this article can be found, in
the online version, at doi:10.1016/j.plantsci.2012.02.002.
References
[1] S.K. Panda, H.K. Patra, Physiology of chromium toxicity in plants—a review,
Plant Physiol. Biochem. 24 (1997) 10–17.
[2] Y.T. Wang, C. Xiao, Factors affecting hexavalent chromium reduction in pure
cultures of bacteria, Water Res. 29 (1995) 2467–2474.
[3] P. Chandra, S. Sinha, U.N. Rai, Bioremediation of chromium from water and soil
by vascular aquatic plants, in: Phytoremediation of Soil and Water Contaminants, American Chemical Society, 1997, pp. 274–282.
[4] KrishnamurthyF S., M. Wilkens, Environmental chemistry of Cr, Northeast.
Geol. 16 (1994) 14–17.
[5] N. Sethunathan, M. Megharaj, L. Smith, S.P.B. Kamaludeen, S. Avudainayagam,
R. Naidu, Microbial role in the failure of natural attenuation of chromium(VI) in
long-term tannery waste contaminated soil, Agric. Ecosyst. Environ. 105 (2005)
657–661.
[6] A.K. Shanker, M. Djanaguiraman, B. Venkateswarlu, Chromium interactions in
plants: current status and future strategies, Metallomics 1 (2009) 375–383.
[7] J. Aiyar, H.J. Berkovits, R.A. Floyd, K.E. Wetterhahn, Reaction of chromium(VI)
with glutathione or with hydrogen peroxide: Identification of reactive intermediates and their role in chromium(VI)-induced DNA damage, Environ. Health
Perspect. 92 (1991) 53–62.
[8] C. Namasivayam, W.H. Höll, Chromium(III) removal in tannery waste waters
using Chinese Reed (Miscanthus sinensis), a fast growing plant, Eur. J. Wood
Prod. 62 (2004) 74–80.
[9] C.D. Palmer, P.R. Wittbrodt, Processes affecting the remediation of chromiumcontaminated sites, Environ. Health Perspect. 92 (1991) 25–40.
[10] Y. Su, F.X. Han, B.B.M. Sridhar, D.L. Monts, Phytotoxicity and phytoaccumulation
of trivalent and hexavalent chromium in brake fern, Environ. Toxicol. Chem. 24
(2005) 2019–2026.
[11] X.-H. Zhang, J. Liu, H.-T. Huang, J. Chen, Y.-N. Zhu, D.-Q. Wang, Chromium
accumulation by the hyperaccumulator plant Leersia hexandra Swartz, Chemosphere 67 (2007) 1138–1143.
[12] I. Lewandowski, U. Schmidt, Nitrogen, energy and land use efficiencies of miscanthus, reed canary grass and triticale as determined by the boundary line
approach, Agric. Ecosyst. Environ. 112 (2006) 335–346.
[13] I. Arduini, A. Masoni, L. Ercoli, Effects of high chromium applications on miscanthus during the period of maximum growth, Environ. Exp. Bot. 58 (2006)
234–243.
[14] N. Rascio, F. Navari-Izzo, Heavy metal hyperaccumulating plants: how and why
do they do it? And what makes them so interesting? Plant Sci. 180 (2011)
169–181.
[15] P. Pongrac, F.-J. Zhao, J. Razinger, A. Zrimec, M. Regvar, Physiological responses
to Cd and Zn in two Cd/Zn hyperaccumulating Thlaspi species, Environ. Exp.
Bot. 66 (2009) 479–486.
[16] S. Quaggiotti, G. Barcaccia, M. Schiavon, S. Nicolé, G. Galla, V. Rossignolo,
M. Soattin, M. Malagoli, Phytoremediation of chromium using Salix species:
cloning ESTs and candidate genes involved in the Cr response, Gene 402 (2007)
68–80.
[17] S. Dubey, P. Misra, S. Dwivedi, S. Chatterjee, S. Bag, S. Mantri, M. Asif, A. Rai,
S. Kumar, M. Shri, P. Tripathi, R. Tripathi, P. Trivedi, D. Chakrabarty, R. Tuli,
Transcriptomic and metabolomic shifts in rice roots in response to Cr (VI) stress,
BMC Genomics 11 (2010) 648.
[18] M. Labra, E. Gianazza, R. Waitt, I. Eberini, A. Sozzi, S. Regondi, F. Grassi, E. Agradi,
Zea mays L. protein changes in response to potassium dichromate treatments,
Chemosphere 62 (2006) 1234–1244.
[19] J.B. Jones Jr., B. Wolf, H.A. Mills, Plant Analysis Handbook. A Practical Sampling,
Preparation, Analysis, and Interpretation Guide, Micro-Macro Publishing, Inc.,
Athens, Georgia, 1991.
[20] I. Alam, S.A. Sharmin, K.-H. Kim, J. Yang, M. Choi, B.-H. Lee, Proteome analysis
of soybean roots subjected to short-term drought stress, Plant Soil 333 (2010)
491–505.
[21] R.L. Heath, L. Packer, Photoperoxidation in isolated chloroplasts: I. Kinetics and
stoichiometry of fatty acid peroxidation, Arch. Biochem. Biophys. 125 (1968)
189–198.
[22] I. Alam, D.-G. Lee, K.-H. Kim, C.-H. Park, S. Sharmin, H. Lee, K.-W. Oh, B.-W. Yun,
B.-H. Lee, Proteome analysis of soybean roots under waterlogging stress at an
early vegetative stage, J. Biosci. 35 (2010) 49–62.
125
[23] N. Ahsan, D.G. Lee, I. Alam, P.J. Kim, J.J. Lee, Y.O. Ahn, S.S. Kwak, I.J. Lee, J.D.
Bahk, K.Y. Kang, J. Renaut, S. Komatsu, B.H. Lee, Comparative proteomic study
of arsenic-induced differentially expressed proteins in rice roots reveals glutathione plays a central role during As stress, Proteomics 8 (2008) 3561–3576.
[24] O.H. Lowry, N.J. Rosebrough, A.L. Farr, R.J. Randall, Protein measurement with
the folin phenol reagent, J. Biol. Chem. 193 (1951) 265–275.
[25] I. Alam, S.A. Sharmin, K.-H. Kim, Y.-G. Kim, J.J. Lee, J.D. Bahk, B.-H. Lee, Comparative proteomic approach to identify proteins involved in flooding combined
with salinity stress in soybean, Plant Soil 346 (2011) 45–62.
[26] J.R. Stewart, Y.O. Toma, F.G. FernÁNdez, A.Y.A. Nishiwaki, T. Yamada, G. Bollero,
The ecology and agronomy of Miscanthus sinensis, a species important to bioenergy crop development, in its native range in Japan: a review, GCB Bioenergy 1
(2009) 126–153.
[27] B. Ezaki, E. Nagao, Y. Yamamoto, S. Nakashima, T. Enomoto, Wild plants, Andropogon virginicus L. and Miscanthus sinensis Anders, are tolerant to multiple
stresses including aluminum, heavy metals and oxidative stresses, Plant Cell
Rep. 27 (2008) 951–961.
[28] I. Arduini, L. Ercoli, A. Masoni, Response of Miscanthus sinensis to long term
application of low Cr levels in hydroponics, Agrochimica 50 (2006) 187–199.
[29] M. Thompson, T. Thorpe, Metabolic and non metabolic role of carbohydrates, in:
J. Bonga, D. Durzan (Eds.), Cell and Tissue Culture in Forestry, Martinus Nijhoff,
Dordrecht, 1987, pp. 89–112.
[30] R. Sperotto, T. Boff, G. Duarte, J. Fett, Increased senescence-associated gene
expression and lipid peroxidation induced by iron deficiency in rice roots, Plant
Cell Rep. 27 (2008) 183–195.
[31] Y. Yamamoto, Y. Kobayashi, H. Matsumoto, Lipid peroxidation is an early symptom triggered by aluminum, but not the primary cause of elongation inhibition
in pea roots, Plant Physiol. 125 (2001) 199–208.
[32] S. Verma, R.S. Dubey, Lead toxicity induces lipid peroxidation and alters the
activities of antioxidant enzymes in growing rice plants, Plant Sci. 164 (2003)
645–655.
[33] S.K. Panda, Chromium-mediated oxidative stress and ultrastructural changes in
root cells of developing rice seedlings, J. Plant Physiol. 164 (2007) 1419–1428.
[34] S.C. Carpentier, B. Panis, A. Vertommen, R. Swennen, K. Sergeant, J. Renaut, K.
Laukens, E. Witters, B. Samyn, B. Devreese, Proteome analysis of non-model
plants: a challenging but powerful approach, Mass Spectrom. Rev. 27 (2008)
354–377.
[35] B. Sghaier-Hammami, L. Valledor, N. Drira, J.V. Jorrin-Novo, Proteomic analysis of the development and germination of date palm (Phoenix dactylifera L.)
zygotic embryos, Proteomics 9 (2009) 2543–2554.
[36] P.M. Lonosky, X. Zhang, V.G. Honavar, D.L. Dobbs, A. Fu, S.R. Rodermel, A proteomic analysis of maize chloroplast biogenesis, Plant Physiol. 134 (2004)
560–574.
[37] J.L. Hall, Cellular mechanisms for heavy metal detoxification and tolerance, J.
Exp. Bot. 53 (2002) 1–11.
[38] M.D. Vázquez, C. Poschenrieder, J. Barceló, Chromium VI induced structural and
ultrastructural changes in bush bean plants (Phaseolus vulgaris L.), Ann. Bot. 59
(1987) 427–438.
[39] F.A. Austenfeld, The effect of Ni, Co and Cr on net photosynthesis of primary and
secondary leaves of PhaseoIus vulgaris L., Photosynthetica 13 (1979) 434–438.
[40] E.R. Sumner, A. Shanmuganathan, T.C. Sideri, S.A. Willetts, J.E. Houghton, S.V.
Avery, Oxidative protein damage causes chromium toxicity in yeast, Microbiology 151 (2005) 1939–1948.
[41] N. Ahsan, D.G. Lee, K.H. Kim, I. Alam, S.H. Lee, K.W. Lee, H. Lee, B.H. Lee, Analysis of arsenic stress-induced differentially expressed proteins in rice leaves by
two-dimensional gel electrophoresis coupled with mass spectrometry, Chemosphere 78 (2010) 224–231.
[42] T. Fukuda, A. Saito, J. Wasaki, T. Shinano, M. Osaki, Metabolic alterations proposed by proteome in rice roots grown under low P and high Al concentration
under low pH, Plant Sci. 172 (2007) 1157–1165.
[43] K. Stobrawa, G. Lorenc-Plucinska, Changes in carbohydrate metabolism in fine
roots of the native European black poplar (Populus nigra L.) in a heavy-metalpolluted environment, Sci. Total Environ. 373 (2007) 157–165.
[44] P. Kieffer, S.b. Planchon, M. Oufir, J. Ziebel, J. Dommes, L. Hoffmann, J.-F. Hausman, J. Renaut, Combining proteomics and metabolite analyses to unravel
cadmium stress-response in poplar leaves, J. Proteome Res. 8 (2008) 400–417.
[45] J. Barceló, C. Poschenrieder, B. Gunsé, Water relations of chromium VI treated
bush bean plants (Phaseolus vulgaris L. cv. contender) under both normal and
water stress conditions, J. Exp. Bot. 37 (1986) 178–187.
[46] A. Brune, W. Urbach, K.-J. Dietz, Differential toxicity of heavy metals is partly
related to a loss of preferential extraplasmic compartmentation: a comparison
of Cd-, Mo-, Ni- and Zn-stress, New Phytol. 129 (1995) 403–409.
[47] J. Verkleij, P. Koevoets, M. Blake-Kalff, A. Chardonnens, Evidence for an important role of the tonoplast in the mechanism of naturally selected zinc tolerance
in Silene vulgaris, J. Plant Physiol. 153 (1998) 188–191.
[48] A.N. Chardonnens, P.L.M. Koevoets, A. van Zanten, H. Schat, J.A.C. Verkleij, Properties of enhanced tonoplast zinc transport in naturally selected zinc-tolerant
Silene vulgaris, Plant Physiol. 120 (1999) 779–786.
[49] S.K. Panda, S. Choudhury, Chromium stress in plants, Braz. J. Plant Physiol. 17
(2005) 95–102.
[50] D.E. Salt, G.J. Wagner, Cadmium transport across tonoplast of vesicles from
oat roots. Evidence for a Cd2+ /H+ antiport activity, J. Biol. Chem. 268 (1993)
12297–12302.
[51] A. Gonzalez, V. Koren’Kov, G.J. Wagner, A comparison of Zn, Mn, Cd, and Ca
transport mechanisms in oat root tonoplast vesicles, Physiol. Plant 106 (1999)
203–209.
126
S.A. Sharmin et al. / Plant Science 187 (2012) 113–126
[52] K.S. Schumaker, H. Sze, Calcium transport into the vacuole of oat roots.
Characterization of H+ /Ca2+ exchange activity, J. Biol. Chem. 261 (1986)
12172–12178.
[53] S. AbuQamar, H. Luo, K. Laluk, M.V. Mickelbart, T. Mengiste, Crosstalk between
biotic and abiotic stress responses in tomato is mediated by the AIM1 transcription factor, Plant J. 58 (2009) 347–360.
[54] L.S. Graham, M.B. Sticklen, Plant chitinases, Can. J. Bot. 72 (1994) 1057–1083.
[55] S. Jacobsen, M.Z. Hauschild, U. Rasmussen, Induction by chromium ions of chitinases and polyamines in barley (Hordeum vulgare L.) and rape (Brassica napus
L. ssp. oleifera), Plant Sci. 84 (1992) 119–128.
[56] B. Békésiová, Š. Hraška, J. Libantová, J. Moravčíková, I. Matušíková, Heavy-metal
stress induced accumulation of chitinase isoforms in plants, Mol. Biol. Rep. 35
(2008) 579–588.
[57] M.d.l.M. Dana, J.A. Pintor-Toro, B. Cubero, Transgenic tobacco plants overexpressing chitinases of fungal origin show enhanced resistance to biotic and
abiotic stress agents, Plant Physiol. 142 (2006) 722–730.
[58] J.D.G. Jones, J.L. Dangl, The plant immune system, Nature 444 (2006) 323–329.
[59] W. Wang, B. Vinocur, O. Shoseyov, A. Altman, Role of plant heat-shock proteins
and molecular chaperones in the abiotic stress response, Trends Plant Sci. 9
(2004) 244–252.
[60] N. Suzuki, N. Koizumi, H. Sano, Screening of cadmium-responsive genes in
Arabidopsis thaliana, Plant Cell Environ. 24 (2001) 1177–1188.
[61] M.C. Romero-Puertas, J.M. Palma, M. Gómez, L.A. Del Río, L.M. Sandalio, Cadmium causes the oxidative modification of proteins in pea plants, Plant Cell
Environ. 25 (2002) 677–686.
[62] M. González-Fernández, T. García-Barrera, A. Arias-Borrego, J. Jurado, C.
Pueyo, J. López-Barea, J.L. Gómez-Ariza, Metallomics integrated with proteomics in deciphering metal-related environmental issues, Biochimie 91
(2009) 1311–1317.
[63] L.M. Sandalio, H.C. Dalurzo, M. Gómez, M.C. Romero-Puertas, L.A. del Río,
Cadmium-induced changes in the growth and oxidative metabolism of pea
plants, J. Exp. Bot. 52 (2001) 2115–2126.
[64] J.F. Palatnik, N. Carrillo, E.M. Valle, The role of photosynthetic electron transport in the oxidative degradation of chloroplastic glutamine synthetase, Plant
Physiol. 121 (1999) 471–478.
[65] J.L. Ortega, D. Roche, C. Sengupta-Gopalan, Oxidative turnover of soybean root
glutamine synthetase. In vitro and in vivo studies, Plant Physiol. 119 (1999)
1483–1496.
[66] L.-M. Chen, C. Huei Kao, Relationship between ammonium accumulation and
senescence of detached rice leaves caused by excess copper, Plant Soil 200
(1998) 169–173.
[67] S.E. Stevens, C. Van Baalen, Control of nitrate reductase in a blue-green alga:
the effects of inhibitors, blue light, and ammonia, Arch. Biochem. Biophys. 161
(1974) 146–152.
[68] C. Rigano, V. Di Martino Rigano, V. Vona, A. Fuggi, G. Aliotta, Studies in vivo on
the control by ammonia of nitrate reduction to nitrite in the unicellular alga
Cyanidium caldarium, Plant Sci. Lett. 13 (1978) 301–307.
[69] E. Eleftheriou, I.-D. Adamakis, P. Melissa, Effects of hexavalent chromium on
microtubule organization, ER distribution and callose deposition in root tip
cells of Allium cepa L., Protoplasma (2011) 1–16.
[70] P. Percipalle, N. Visa, Molecular functions of nuclear actin in transcription, J.
Cell Biol. 172 (2006) 967–971.
[71] M. Amenós, I. Corrales, C. Poschenrieder, P. Illéš, F. Baluška, J. Barceló, Different
effects of aluminum on the actin cytoskeleton and brefeldin A-sensitive vesicle
recycling in root apex cells of two maize varieties differing in root elongation
rate and aluminum tolerance, Plant Cell Physiol. 50 (2009) 528–540.
[72] J. Hindley, G.A. Phear, Sequence of the cell division gene CDC2 from Schizosaccharomyces pombe; patterns of splicing and homology to protein kinases, Gene
31 (1984) 129–134.
[73] C. Poschenrieder, B. Gunsé, J. Barceló, Chromium-induced inhibition of ethylene
evolution in bean (Phaseolus vulgaris) leaves, Physiol. Plant 89 (1993) 404–408.
[74] M. Rodríguez-Serrano, M.C. Romero-Puertas, A. Zabalza, F.J. Corpas, M. Gómez,
L.A. Del Río, L.M. Sandalio, Cadmium effect on oxidative metabolism of pea
(Pisum sativum L.) roots. Imaging of reactive oxygen species and nitric oxide
accumulation in vivo, Plant Cell Environ. 29 (2006) 1532–1544.
[75] T. Arazi, B. Kaplan, R. Sunkar, H. Fromm, Cyclic-nucleotide-and
Ca2+ /calmodulin-regulated channels in plants: targets for manipulating
heavy-metal tolerance, and possible physiological roles, Biochem. Soc. Trans.
28 (2000) 471.
[76] B. Moffatt, C. Pethe, M. Laloue, Metabolism of benzyladenine is impaired in
a mutant of Arabidopsis thaliana lacking adenine phosphoribosyltransferase
activity, Plant Physiol. 95 (1991) 900–908.
[77] R. Itai, K. Suzuki, H. Yamaguchi, H. Nakanishi, N.K. Nishizawa, E. Yoshimura,
S. Mori, Induced activity of adenine phosphoribosyltransferase (APRT) in irondeficient barley roots: a possible role for phytosiderophore production, J. Exp.
Bot. 51 (2000) 1179–1188.
[78] J.F. Ma, G. Kusano, S. Kimura, K. Nomoto, Specific recognition of mugineic
acid–ferric complex by barley roots, Phytochemistry 34 (1993) 599–603.
[79] T. Iwashita, Y. Mino, H. Naoki, Y. Sugiura, K. Nomoto, High-resolution proton
nuclear magnetic resonance analysis of solution structures and conformational
properties of mugineic acid and its metal complexes, Biochemistry 22 (1983)
4842–4845.
[80] E. Zablackis, J. Huang, B. Muller, A.G. Darvill, P. Albersheim, Characterization of
the cell-wall polysaccharides of Arabidopsis thaliana leaves, Plant Physiol. 107
(1995) 1129–1138.
[81] A. Moons, J. Gielen, J. Vandekerckhove, D.V.D. Straeten, G. Gheysen, M.V. Montagu, An abscisic-acid- and salt-stress-responsive rice cDNA from a novel plant
gene family, Planta 202 (1997) 443–454.